CVX4343 Course Material

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 123

OPEN FACULTY OF ENGINEERING TECHNOLOGY

UNIVERSITY
DIPLOMA IN TECHNOLOGY LEVEL 4
OF SRI LANKA
CVX4343 – SOIL MECHANICS

SOIL MECHANICS
Published by the
Open University of Sri Lanka
Course Team

Course Team Chair Language Editor


Ms. N.C. Kannangara

Course Team Manager Media Designer

Author Desk Top Publishing


Dr. H.G.P.A. Ratnaweera
Dr. H.G.P.A. Ratnaweera

Graphic Artist
Educational Technologist
Dr. H.G.P.A. Ratnaweera

Word processing
Content Editors
Dr. H.G.P.A. Ratnaweera

The Department of Civil Engineering acknowledges the contributions


made by Dr. (Ms.) P. Sivaprakasapillai Sivasegaram, Dr. A.U.
Gunasekara, Mr. S.U. Upasena, Dr. P.N. Wikramanayake and Dr.
H.G.P.A. Ratnaweera who served as course team members during
formulation of course objectives.

The author gratefully acknowledges the services rendered by Ms. T.N.


Kaludewage, who prepared certain diagrammes used in this course
material; Dr. B.G. Jayatilake, who gave valuable comments on the
document layout; and Ms. N.P.M. Rajaguru, who checked the
document for any errors.

First published [2014]


ISBN – 978-955-23-1487-2

Revised [2018]

© [2014] Open University of Sri Lanka


CVX4343 – Soil Mechanics

TABLE OF CONTENTS

Session 1: Why learn Geotechnical Engineering 01


Session 2: Knowing soils of engineering significance 12
Session 3: Soil and rock as a three-phase particulate system 20
Session 4: Towards classifying engineering soils 28
Session 5: Laboratory classification of engineering soils 39
Session 6: Classifying soils based on visual inspection and feel 44
Session 7: How soil layers carry load 50
Session 8: Stresses in soils due to imposed loads 57
Session 9: Computing stresses acting on a soil element 67
Session 10: Compacting soils 71
Session 11: Water movement through soil and rock 76
Session 12: Computing settlements in soils 84
Session 13: Computing time for settlement 95
Session 14: Soil strength 103
Session 15: Laboratory tests to assess soil strength 107

iii
Interpreting action-verbs …

An action verb is a verb that conveys to you your teacher’s expectations. In a


conventional classroom, a teacher does this through lectures and discussions while
learners are able to seek clarifications if needed. In an open-distance learning
environment, face-to-face communication is very limited. This makes the teacher a
facilitator, who assists learners to self-learn.

A teacher directs learners by conveying his expectations via learning aims and
learning objectives. These are listed at course level, unit level and at session level.
He uses action-verbs to communicate learning outcomes to you; hence it is
important that you pay attention to these verbs.

Perhaps you may recall, as kids, we were made to memorise the multiplication
table, i.e. from 1x1 to 12x12. This gave us the competency to recall multiplication
of two particular numbers when performing a mental calculation or when
performing a calculation on paper.

If you were asked to list even numbers between 0-10, your response would be
2,4,6,8. If you are asked to plot the variation of primary consolidation settlement
with time, you are required to indicate all data points on a graph sheet drawn to
scale; to draw a smooth curve through data points to show the variation. When
presenting this information one should not forget to name the graph, its variables,
and the two scales. If you are to sketch the same you could draw the trend line to an
approximate scale indicating the axes and title of graph.

Sometimes we are asked to compare apples and oranges. Both are considered
edible fruits. If you are asked to contrast, oranges belong to the citrus family while
apples do not. If you are to distinguish between steel and aluminium you could
state that aluminium is less dense and does not show magnetic properties.

Action verbs highlighted above instruct learners to follow a particular learning


path. These action verbs may suggest to you the competencies you need to develop
while following the course.

iv
CVX4343 – Soil Mechanics

On learning…

A teacher sometimes puts you through a learning experience that can be


vague or unclear to you at the time of learning. A Zen monk asks his
teacher what he should do to attain enlightenment; the teacher in turn
instructs him to wash his breakfast bowl. This perhaps makes us realise that
learning is an inner experience and the teacher can merely assist you to
achieve it.

During learning, learners are expected to discover why we learn, what we


learn and how we learn, and the teacher is expected to facilitate this.
Learning Aims and Learning Objectives may help you in the above quest.
During learning the teacher may expect us to improve knowledge – i.e.
cognitive skills, to change the way we see issues – i.e affective skills or to be
able to do certain physical activities – i.e. psychomotor skills.

The Course Outline gives an overall perspective of the course and explains
how its units are structured. Unit Outlines explain how sessions in each unit
are structured, and their expected learning outcomes. Self-Assessment
Questions, and the Summary, which appears at the end of each session, help
you to assess whether you have met the intended learning objectives.

Learning takes place through reflection, and hence you are able to relate
such experiences during a future application. Reflection also helps you to
gain new knowledge. Learners who practice deep learning are able to
process information to create knowledge. Deep learning requires
commitment and endurance and hence we all should strive to become deep
learners.

v
About this Course …

The course, Soil Mechanics (CVX4343) is the first Geotechnical Engineering


course offered in this programme. This is offered at Level 4 and expects learners
to assign 150 hours of their time for all requirements. It explains the use of soil
and rock in civil engineering construction, and discusses their physical and
mechanical properties. This course also serves as the foundation to learn
Engineering Geology (CVX4344), Geotechnics (CEX6444) and Geotechnical
Design (CEX6241).

The course expects you to know concepts that describe rigid body equilibrium,
stress-strain behaviour of engineering materials, energy concepts that makes water
to move in pipes and channels. These concepts were discussed in Strength of
Materials (CVX3442) and Hydraulics and Hydrology (CVX3340). These concepts
are used to describe soil and rock behaviour. The course curriculum and syllabus
were revised recently to make the course more interactive. Print material is
integrated with laboratory and home activities, with links to Internet sources.

The course material consists of fifteen sessions. Continuous assessment (CA)


encourages you to learn at a constant pace. CA for this course consists of three (3)
tutor marked assignments [TMAs], two (2) continuous assessment tests [CATs], a
laboratory activity [LAB] and reflective learning logs [PD]. The criterion for
continuous assessment is stated as:

OCAM >= 40% and [LAB] >= 40%


OCAM=Avg [TMA] x 25% +Avg [CAT] x 25% +[LAB] x 40% +[PD] x 10%
OCAM - Overall Continuous Assessment Mark.

The four day schools of this course are designed as interactive learning sessions,
and not as formal lectures; hence you should come prepared to discuss your doubts.

The laboratory activity is scheduled over 6 consecutive days. The laboratory


activity is a group-work. You are required to plan and execute laboratory
instructions as a group. At the end of each activity the group is given an
opportunity to assess each other. Learners are expected to present group
observations with their own calculations for assessment. The viva-voce
examination also assesses learner’s ability to relate observations and results to
principles and concepts learnt in the course.

The first session attempts to explain the need to design and construct soil and rock
structures. It discusses applications, ancient and modern, drawing your attention to
some familiar geotechnical issues and concepts. The session aims at raising your
awareness and interest to learn this subject. This exposure may also help you to
recall Geotechnical applications you are already familiar with. Case examples
cited also discuss and relate observations to engineering concepts. This session
also introduces new terminology and the learner is expected to pay particular
attention.

Session 2 introduces you to various soils types of Sri Lanka. You are expected to
familiarise yourself with the soil map of Sri Lanka, published by the Department of
Survey. This knowledge is important to interpret geology and soil maps of the
vi
CVX4343 – Soil Mechanics

region. The session also introduces you to soil groups namely, sandy, clayey,
residual soils and organic peat. The session expects learners to recall engineering
properties associated with these soil groups.

Soil consists of three phases: solid, liquid and gaseous. The interactions between
the three phases decide their engineering behaviour. Session 3 shows how volume
and mass quantities of the three phases are defined. These quantities are used to
derive parameters that are used to model engineering behaviour of a granular
medium.

Session 4 discusses the laboratory tests that are used to identify and classify
engineering soils. The sieve analysis test and hydrometer test are used to
determine the size distribution of particles of a soil sample. The Atterberg
consistency limits particularly the Liquid Limit and Plastic Limit quantify
`plasticity’ of fine-grained soils.

Session 5 discusses the Unified Soil Classification System, which was introduced
by researchers at Massachusetts Institute of Technology (MIT). This classification
system was later adopted by the British Standard Institution. The said system is
based on the analysis of particle size and its distribution, and Liquid and Plastic
Limit tests. The said method is also compared with the classification system
proposed by the American Society for Testing Materials (ASTM).

Session 6 explains how soils are classified visually. You will learn how fine-
grained soils are classified based on dry strength, dilatancy reaction, toughness of
soil thread at Plastic Limit and Plasticity. The course-grained soils are classified
based on their soil group. The physical observations made on natural soil samples
are also used to describe soils. Even though these methods are considered
rudimentary, this standardised procedure is practiced during field logging of
boreholes.

Granular soils are non-homogeneous and anistopic. We know that water is


homogeneous and isotropic and hence water pressure at a certain depth is given by
hg. This pressure is equal in all directions. Soil is made of three phases. A
saturated soil has the solid and liquid phases, which show different mechanical
characteristics.

Session 7 discusses how total vertical stress at a given depth is computed, assuming
a uniform soil density. The Principle of Effective Stress quantifies the load carried
by solid soil grains (i.e. as inter-granular stress or effective stress) and the load
carried by water (i.e. as pore-water pressure). The session explains the
computation of effective horizontal stress, hence the total stress as explained by the
Principal of Effective Stress. Session 1 also explains how the above principle is
modified to represent a partially saturated soil.

Session 8 analyses the stress effect on a soil mass due to an imposed surface load,
of a finite or an infinite extent. The analysis helps us to identify the zone of
influence due to a particular imposed load. The session also discusses the observed
variations in stress-strain behaviour when a dense sand specimen is axially
compressed. These results show that at very low axial strains material behaviour
can be considered as `elastic’. This enables us to define its Young’s modulus and

vii
Poisson’s ration, and hence normal strains in all three dimensions. The session
provides numerical and graphical outputs to compute stresses at various points in
the soil mass. Expressions for plane-strain and plane-stress conditions are derived.

Session 9 explains the Mohr’s Circle of stress concept used to determine stresses
acting at a point. When stresses in a soil mass are computed, we consider that
vertical and horizontal stresses remain as principal stresses before and during
loading. This simplifying assumption eliminates the need to know initial shear
stresses and shear stresses when a load is imposed. The session explores how
stresses at a given point changes for different loading and unloading situations.

Session 10 discusses the theory of compaction and the use of Standard Proctor
Compaction Test to establish the compaction curve for a particular soil. The use of
Sand Cone to determine in-situ dry unit weight is also discussed.

Session 11 discusses flow through a granular medium and in rock; expresses


Darcy’s Law and the Cubic Law. It recalls the use of hydraulic head, which
determines the hydraulic gradient. Constant Head and Falling Head tests are used
to determine the coefficient permeability (hydraulic conductivity) of a soil. The
session discusses the use of in-situ pumping tests to determine aquifer
characteristics.

Session 12 discusses a field situation where a soil layer is subjected to compression


in the vertical direction only. The session explains why the state of stress satisfies
`K0 – compression behaviour with zero lateral strain. The Oedometer test is
discussed in detail; laboratory results are interpreted; key indices that represent
total settlement are derived.

Session 13 identifies total settlement of a compressible soil stratum under a given


load. Total settlement is identified as due to initial compression, primary
consolidation and secondary compression. The session discusses the mathematical
model put forward by Carl Terzarghi to explain primary consolidation behaviour.
It explains how oedometer test results are used to predict the variation of primary
consolidation settlement with time.

Session 14 introduces learners to shear strength and stress – strain behaviour of


soils. It describes the observed behaviour based on a mechanism at particle level,
valid for granular materials. The session also presents Mohr-Coulomb failure
criterion and representation of two dimensional stresses on  - n space.

Session 15 discusses the Unconfined Compression (UC) Test and Conventional


Triaxial Loading Tests (i.e. CD, CU and UU tests). It also discusses the types of
shear strength parameters that can be obtained and their relevance in performing a
total stress analysis and an effective stress analysis.

viii
Session 1
Why learn Geotechnical Engineering

What you will learn in this session


This 2-hour session introduces you to Geotechnical applications in civil
construction, to create awareness and interest to learn this subject. Case examples
cited introduce you to geotechnical engineering concepts, which you will learn in
this course.
This session explains how soils support buildings and other structures (Section
1.2), use of earth embankments as bunds that retains water (Section 1.3), building
railways, highways and runways (Section 1.4), reclaiming land for development
(Section 1.5), making slopes stable (Section 1.6), designing walls to retain earth
(Section 1.7), preventing excess seepage through soil (Section 1.8), tunnelling
through rock (Section 1.9), and constructing safe municipal landfills (Section 1.10).

1.1 Introduction
Geotechnical Engineering is the branch of civil technology that deals with aspects
of `geo’ or the Earth. It is the art and science of performing changes to the geo-
environment. Changes to the geo-environment should be made after a careful
study. There are many examples where such changes had resulted in adverse
impacts to society, economy and ecology, at both local and global levels.
Geotechnical Engineering practice is inter-disciplinary, and hence should not be
viewed only as a discipline of civil engineering construction. It expects us to
understand natural processes concerning Earth’s crust, and socio-economic and
environmental issues resulting from man-made changes.
Geotechnical Engineering is also the study of the behaviour of soil and rock. Soil
is an aggregate of mineral grains while rock is a blocky mass made of minerals.
They differ from elements such as cement blocks, beams, columns etc, since they
do not have definitive boundaries or geometric shapes. Therefore when soil and
rock are used as an engineering material, we need to demarcate the region or zone
of influence. Soil and rock deform when subjected to external forces and reactions.
These give rise to internal stresses such as compressive, bearing and shear stresses.
We’ve learnt that stresses give rise to internal
strains, compressive strains and shear strains.
Elastic strains cause recoverable deformations
while plastic strains give rise to permanent
deformations.
The leaning tower of Pisa (refer Figure 1.1) has
been tilting over a few centuries. It was found
that ground water extraction had made the
underlying layers to compress. A recent study
reveals that the tilting rate has reduced
significantly, due to a rise in ground water level. Figure 1.1: Leaning tower of
The foundation soil has been grouted with a Pisa
cement-sand mix to control tilting. This in fact
had increased tilting.
Session 1: Why learn Geotechnical engineering

Soil and rock are excellent engineering materials when they are adequately
supported or confined. They allow water to flow through their pore spaces, hence
is considered permeable.
Soil is a non-homogeneous material, which implies that they are not made of the
same basic units. It is also considered anisotropic since its properties are found to
change with the direction or orientation.
Soil is a mixture of three phases consisting of solid soil particles, water and air.
The relative composition of each phase decides its response to stress. If a soil mass
is stressed further its grains will slide and rotate causing it to deform. Interlocking
grains make soil withstand external loads.
Rocks display spatial and directional variations similar to soils. A continuous rock
mass may not show such variations in properties. In jointed or fractured rock,
presence of joints, fractures and bedding planes decide their engineering behaviour.
Geotechnical Engineering explains how sub-surface materials respond to natural
and man-made changes. A Geotechnical Engineer designs and supervises such
construction to make them safe and to serve our needs.

1.2 Soil supporting buildings and other structures


When building a single story house we ensure that the load is safely transferred to
ground. In a house with load bearing walls, the wall load is transferred to the
ground via a plinth made of rubble masonry, and a strip footing (refer Figure 1.2).
In a framed structure the column loads are transferred to ground via pad footings.
If you recall the cross sectional sizes of these structural elements, we require a
larger footing area to disperse or spread the load evenly. Soil being a weaker
material, the load needs to be spread over a larger cross sectional area to reduce
stresses in soil.

Figure 1.2: A schematic showing a shallow strip footing.


Ancient pyramids, the Great Wall of China, ancient Thupas in Sri Lanka
demonstrate how well ancient builders understood the interaction between the
structure and the bearing soil. It is stipulated that they had acquired the art and
science of building through practice. Some still think that ancient builders adopted
a trial and error method. If a structural foundation worked well before, should it
perform well next time? Had it failed, should it be made wider and deeper? Or did

2
CVX4343 – Soil Mechanics

they possess a much deeper understanding on load bearing mechanisms and


construction technology?
`Building of the Maha Thupa’ (refer Figure 1.3) quotes:
The ground was dug to a depth of 10 ft., to lay a
massive foundation to withstand the weight of the
mass above ground, having taken the circle by
holding a staff and going around, with the centre
pole tied to the rope in the centre of the circle.
But, the king was compelled to reduce the size of
the Thupa by half. The foundation to the
MahaThupa was laid in eleven stages. Rock
stones brought to the site by the king's soldiers
were broken with hammers, and in order to make
the foundation firmer, the crushed stones were
stamped down by huge elephants and their feet Figure 1.3: Ruwanweli Seya
were bound with leather to prevent injury.
After settling the stones, a kind of clay called 'butter-clay' ('vendaru-meti') was
spread over the layer of stones, and bricks were laid over them. Above the bricks
rough cement was laid and over it cinnabar (sulphide of mercury) or
'kuruvindapashana', and over it a sheet of gravel ('marumba') six inches thick
AriyadasaRatnasinghewww.lankarama.using.net/temp_stuff/
building_of_the_maha_thupa.htm
This above account describes a particular construction process adopted by ancient
builders to ensure a greater service life. Present day civil construction practices
and challenges are different to those of ancient times. The industry emphasises the
need to use resources in an optimal way to meet more complex and sophisticated
needs.
Figure 1.2 is a schematic of a typical foundation used for a house. You may learn
in this course how a shallow foundation is designed to transfer the structural loads
(i.e. forces and moments) to the ground. The underlying soil and footing are
expected to be in harmony. This means that both should have the same deflections
(i.e. compatibility) and should satisfy equilibrium.
Figure 1.4 shows how the concept discussed
above is put to practice. It shows how
construction workers spread gravel prior to
placing a strip footing. Gravel is well compacted
and levelled to ensure continuity between the
foundation and the bearing soil.
Scarcity of land has forced us to construct
buildings on reclaimed lands. During such
construction if a proper bearing stratum is not
found at a shallow depth, the load has to be
transferred to a sound bearing stratum or bedrock.
This is done using concrete, steel or timber piles.
Piled made of solid materials such as steel, timber Figure 1.4: Preparing ground
or pre-stressed concrete are driven or hammered to lay a shallow strip
in to the soil. Reinforced concrete piles are cast foundation.
in-situ by auguring a cylindrical hole to the

3
Session 1: Why learn Geotechnical engineering

required depth; placing the reinforcement cage and filling with concrete.
Many recent constructions have been on reclaimed lands. Making a foundation
safe and economical requires us to predict how the structure behaves just after
construction and after many years of service. This requires us to learn more about
underlying soils and their properties. Geotechnical Engineer assesses the sub-
surface profile and its variation, and estimates its engineering properties. These
properties are used to assess the load bearing capacity of the soil. The soil strength
available to carry building loads is computed by allowing a safety margin of 2.5-3
times its capacity. Such margins of safety are required to account for soil
variability and sub-surface conditions.

1.3 Earth embankments retaining water bodies


Earth embankment dams were an important aspect of Sri Lanka’s hydraulic
civilisation. `A short history of damming’ states:
South Asia, too, has a long history of dam building. Long earthen embankments
were built to store water for Sri Lankan cities from the 4th century BC. One of
these early embankments was raised in 460 A.D. to a height of 34 metres and was
the world's highest dam for more than a millennium later. One old embankment he
(king Parakramabahu) enlarged to a height of 15 metres and an incredible length
of nearly 14 kilometres. No dam equalled it in volume until the early 20th century.
http://www.irn.org/index.asp?id=/basics/whatdamsdo.html
Ancient Tank builders had made the tank network system to be in harmony with
the eco-environment. The central highlands with its largely undeveloped forest
region served as a perennial water source, nourishing surface and sub-surface water
regimes. The waters feeding major water bodies were used to supplement minor
tanks experiencing shortages during dry-weather season.
The network of tanks that spread over North-central and southeast dry zones have
many earth embankments. These structures were sited at suitable locations on firm
ground and anchored to rock abutments.
Present-day reservoirs cater to hydropower generation and hence waterways are
dammed at upper reaches of the river (refer Figure 1.5). These also serve as
storage reservoirs, to control floods and for irrigation needs.

Figure 1.5: Victoria dam being constructed lcweb2.loc.gov/frd/cs/sri_lanka


/lk03_08a.jpg

4
CVX4343 – Soil Mechanics

They require strong abutments and a water-tight reservoir bed. The steep slopes
forming the reservoir are stabilised to protect earth slides. Micro-seismic activities
are also a concern when a large mass of water sits on an unfavourable rock
formation. De-silting of reservoir also becomes a concern when reservoir
catchments are not protected.
Ancient tank builders built their network of tanks in the lowest peneplain, hence
did not have to confront such issues. Earth embankments were made of soils
obtained locally. These were compacted well to make them strong and watertight.
It is believed that elephants were used to compact
soil (refer Figure 1.6a). A 2000kg elephant with
a foot imprint of 175cm2 exerts 3.8kg/cm2
pressure (assuming 3 legs at a given instance
exerting the pressure). The foot pressure of 1400-
7000kPa of a present-day sheeps-foot roller (refer
Figure 1.6b) is far superior to 375kPa foot
pressure of the elephant. Interestingly, we do not
know how such compactions were done during
ancient times even though the technology is (a) An elephant assisting
peasants to build a roadway,
considered to be far superior. Present day trials
Thailand.
have shown that an elephants tend to follow the
footpath of the one in front and as a result
compacted densities are much lower than
expected. Perhaps, it would be of interest to
know this issue was addressed then.
Earth embankments are built along riverbanks to
protect areas from flooding and these are known
as levies. Such levies are built along Gin-Ganga
and Nilwala Ganga to channel floodwaters to sea. (b) The use of machinery.
Nilwala Ganga Flood Protection Scheme has
Figure 1.6: Compaction (a) the
caused many social, economic and environmental
ancient way and (b) the
issues among certain farmer communities in modern way.
Matara District.
The alteration of the groundwater flow regime had increased soil-acidity, making
land unsuitable for paddy cultivation. Sand mining has made salt water to move
inland during low flow.

1.4 Building railways, highways and


runways
Sometime ago, the state had plans to reconstruct
the Palaly airstrip. The news item further stated
that larger aircrafts could not generate sufficient
speed to lift off due to poor surface conditions.
During rainy season, most roadways become
waterways; after a spell of heavy rain potholes
appear exposing sub-grade soils. They
disintegrate further if timely interventions are not
made.
The components of a road, from top asphalt Figure 1.7: A road cross-
section showing its
surface to sub-grade soil (refer Figure 1.7) are
components.

5
Session 1: Why learn Geotechnical engineering

meant to transmit vehicular traffic loads safely to the ground. These layered
components are subjected to repeated stresses. A road is designed to bear repeated
or cyclic loading. When a road goes through several million cycles of traffic load,
signs of distress due to fatigue become visible.
Road sections are designed to keep water away from its sub-base and sub-grade, by
providing proper road camber. Having properly maintained drains on either side of
the road is essential. Roads are elevated adequately to protect against rising
groundwater. The base, sub-base and sub-grade materials are compacted to a
specified standard. Standard specifications require that they be of acceptable
quality and well graded.
Highways such as the Colombo port access road, Colombo-Katunayake highway,
Southern highway passes through deposits of highly compressible organic soils and
peat. Such roads undergo settlements due to traffic load and embankment weight,
sometimes being of the order of 500-1500mm.
These settlements continue over a few decades unless engineering measures are
taken to speed up the process. Compression of soft peat hinders sub-surface water
movement. This becomes critical when highway embankments of significant
height are constructed over thick deposits of soft peat.
Speedy construction of embankments may soften sub-surface soils. This loss of
strength is caused by an increase in water pressure in soil pores. Modern
construction uses geotextiles to prevent migration of soil from the embankment to
soft compressible soils, while allowing the transfer of vertical stress to ground.

1.5 Reclaiming land for urban use


Low-lying lands are reclaimed and used for housing and commercial purposes. In
the past, such lands were used for dumping municipal refuse. Colombo residents
had their municipal refuse dumped at Orugodawatte and Mattakkuliya open dumps.
Other low-lying lands surrounding the city are being filled with municipal waste.
Urban land reclamation and development for residential, commercial and industrial
use commenced in the 80’s.
Sri Lanka Land Reclamation and Development
Corporation planned and carried out filling of
certain lands at Narahenpita, Nawala,
Kerawalapitiya and Papiliyana. Filling of low-
lying areas naturally reduces the capacity to store
water during a storm surge. In Greater Colombo
are, loss of land for storm water storage was
offset by rehabilitating the canal system.
A fill primarily raises land above flood level. A
well-compacted fill with a sufficient layer
thickness can safely bear footings. A poorly
engineered fill may cause underlying material to
loose strength or settle, causing damage to the
building and its service lines. Figure 1.8: Dynamic
compaction drops a weight
An engineered fill should safely disperse the load from a height (www.dgi-
to underlying weak layers without causing excess menard.com/
settlement or loss of strength. When settlement compaction.html).

6
CVX4343 – Soil Mechanics

continues for several years, interventions to accelerate settlement through enhanced


drainage are possible. Surcharging is the technique by which a temporary load is
placed on ground to reduce subsequent settlement during service.
Figure 1.8 shows the method of dynamic compaction used to enhance the bearing
strength of weak soils. This technique was used to improve soil at housing
complex for members of Parliament at Madiwela.

1.6 Designing stable slopes


Central hills and south-western hill slopes of Sri Lanka are susceptible to landslides
and local slope failures. Landslides are triggered by high intensity spells of rain.
Watawala landslide had disrupted up-country railway supply route while the slide
at Naketiya had made soil and rock debris to move several kilometres down-slope.
Local slope failures are common in unprotected road cuts (refer Figure 1.9) and in
hill slopes. During heavy rains, infiltrating waters increase the weight of soil mass.
When soil strength is insufficient to hold the soil mass sliding is initiated.
Sliding occurs along a surface that is planer or circular depending on soil type.
Slip surfaces with various shapes occur when material is non-homogeneous.
During slippage, soil resistance is further reduced. Sliding sometimes take place
over many years; process being accelerated during each rainy season. Tension
cracks and local settlements become visible during initial stages of sliding.
Stabilising slides can be done at their initial stages by diverting subsurface waters
away from the unstable mass. This may require grouting and embedding
horizontal drains to remove water. Such interventions have been effectively used
to stabilize the Watawala slide.

Figure 1.9: A typical highway slope failure.


Excavation of slopes during road widening has resulted in many active slides,
causing concerns among the public. Such slides are being monitored by local
government bodies and the Road Development Authority with technical assistance
from the National Building Research Organisation (NBRO).

7
Session 1: Why learn Geotechnical engineering

The potential risk of a slope failure depends on depth of soil cover, slope angle,
land management and use, bedrock characteristics and drainage characteristics of
catchment. NBRO uses a scheme to assess areas with high-risk and makes
recommendations on land use to local government bodies.

1.7 Structures that support earth


Earth retaining walls are built to support a soil
mass with a vertical face. Have you seen
retaining walls supporting vertical cuts adjoining
major roadways? These are usually constructed
with rubble masonry.
Soils cannot stand unsupported unless their grains
are held or cemented together. You may have
seen unsupported vertical cuts made on lateritic
soil formations, which support their own weight.
A granular soil has the tendency to form a heap of
debris, and such a heap is termed a talus. When
carrying out excavations in sandy soils vertical
sides are supported using timber shoring or sheet
piles. You may see these in gem pits (refer
Figure 1.10) or during laying of water mains Figure 1.10: Shoring of a gem
pit. E.J. Gubelin, Ekanite,
along main roads. Basements of high-rise Gemmologist, Vol. 31 (1962)
buildings have concrete walls that brace the pp. 142-157.
supporting soil against earth and water pressures.
Construction of Gabion retaining walls and Gabion blankets are becoming a
popular earth support option. You may have seen the Gabion retaining wall that
supports the main canal banks bordering Open University’s main campus (refer
Figure 1.11). Gabion walls are used to support roadside cuts. These flexible
structures allow for subsequent settlement and provide adequate drainage.
Reconstruction work of the three basement levels of New York’s World Trade
Centre is in progress. The debris of the previous basements was removed after
erecting a sheet pile wall to support near by high rise buildings

1.8 Seepage through soils


Seepage is the term that describes flow of water through soil. A depleted domestic
well gets replenished overnight due to seepage. During a rainy spell a domestic
well may have a high water level, which subsides with time. Seepage takes place
from a high elevation to a low elevation. This is similar to water flow in canals or
in conduits.
The amount of water that can be extracted from a well depends on how fast the
well is recharged. National Water Supply and Drainage Board supplies pumped
ground water to industrial facilities and communities. When implementing a new
supply scheme, water extraction rate or daily yield is obtained by performing a
pumping test. Excessive pumping can cause significant draw down levels in
nearby domestic wells, and hence it is required to determine a suitable pumping
rate.
In earth embankment dams, high flow velocities can cause migration of soil
particles, causing flow pipes or channels. These channels of seepage originate at
the upstream surface. An engineered earth embankment dam provides controlled

8
CVX4343 – Soil Mechanics

seepage through its body. Selection of embankment soils with suitable particle
sizes and gradation provides safe passage to water, intercepted by a downstream
filter. Many embankment dams have a semi-permeable or an impermeable core
constructed to reduce seepage flow rates.

(a)

(b)
Figure 1.11: (a) Gabion structures (b) a storm water outlet integrated to the
structure, adjoining main campus, Open University.

When designing an anicut seepage beneath its foundation need to be controlled.


Excessive seepage may lead to the formation of channels causing scour. This
opens up leaks, sometimes causing the structure to tilt or crack.

1.9 Tunnelling through rock


Railway tunnels cut across hilly terrain. Sri Lanka’s upcountry railway has many
tunnels built by British engineers. Tunnels are also used to convey water across

9
Session 1: Why learn Geotechnical engineering

basins or to convey water to underground hydro-power plants.Tunnels also provide


vehicular access across rivers. The tunnel across the English Channel was recently
completed.
Tunnels are `excavated’ through controlled blasting. When material is removed,
the compressive stresses surrounding the tunnel caused by its overburden are
released. This may cause instabilities in supporting rock mass, unless it is
sufficiently strong. If weaknesses are encountered such rock masses need to be
stabilised by driving `rock bolts’. Tunnels through fragmented formations may
have water seepage. Seepage pressures sometimes become too excessive to
control. In such instances measures are taken to arrest seepage temporarily until a
concrete lining is placed to secure the tunnel section.

1.10 Constructing landfills


Construction of solid and liquid waste containments is an emerging need.
Countries including Sri Lanka show a high quantity of generated solid waste. A
landfill is a large man-made heap of solid waste compacted and stored in cells or
compartments. Material closest to compacted solid waste is soil, and hence landfill
settlement and stability of side slopes are assessed and interpreted using concepts
described in soil mechanics. Figure 1.11 shows a regulated landfill
(www.metrokc.gov/ dnr/kidsweb/landfill.htm ).
During operation, leachate levels build up. Leachate is contaminated water
generated within the landfill. These tend to seep through foundation soils to reach
the groundwater table. Present landfills have a barrier made of plastic or clay to
prevent such leakages. A leachate collection system is used to collect
contaminated water and is treated and disposed safely. Leachate generation is also
minimised by reducing rainwater infiltration. Selecting an appropriate soil to cap
the landfill is important.

Figure 1.12: A schematic showing a regulated landfill


(www.metrokc.gov/ dnr/kidsweb/landfill.htm )

10
CVX4343 – Soil Mechanics

To sum up
Geotechnical engineering is the branch of Civil Engineering that specialises in
design and construction aspects related to soil and rock. This session makes you
appreciate various issues and challenges you may confront during civil engineering
practice.
Geotechnical engineering is also the study of engineering properties of soil and
rock and how they respond to various engineering situations. The series of courses
on Geotechnical Engineering, offered in the Diploma in Technology and Bachelor
of Technology programmes may give you opportunities to explore and interpret
principles of soil mechanics, rock mechanics and engineering geology.

11
Session 2
Knowing soils of engineering significance

What you will learn in this session


This 2-hour session introduces you to a system of classification that is used to
identify various soil types of Sri Lanka (Section 2.1). This is presented in the soil
map of Sri Lanka, published by the Department of Survey. You may find that this
system of classification is useful in identifying soil types, but is inadequate in
describing the engineering properties of soils.
Section 2.2 discusses sub-surface geology of Colombo, as described in Wadia
(1941). Section 2.3 discusses how borehole logs are used to identify the sub-
surface profile. Soils when used as engineering materials are identified as sandy,
clayey, soils with residual origin and organic soils. Section 2.4 attempts to relate
certain observations you are already familiar with to understand soil behaviour.

2.1 Introduction
Soils form during weathering of rock and hence consist of rock fragments and
minerals. Figure 2.1 below shows a typical soil profile you may observe in an
excavation or a cut slope.

Figure 2.1: A typical soil profile


The transformation from rock to soil and the formation of soil deposits depend on
climate and topography, assisted by vegetation and organisms. During
deforestation, soil erosion removes layers  and  (refer Figure 2.1). Alternate
wetting and drying in layer  makes its clays to harden resulting in a reddish
brown laterite crust. Leached chemicals and dissolved minerals move down and
gets deposited in layers  to . The rate of weathering is influenced by
infiltration and percolation influences the rate at which the soil overburden is
formed.
Soil deposits contain both organic and inorganic matter. The void spaces are filled
with air and water. A consolidated deposit is a deposit, which has its particles
packed densely, and hence has less void spaces in the soil matrix. These are
formed due to the weight of soil accumulated during deposition, which is
CVX4343 – Soil Mechanics

commonly known as the overburden. Such deposits are strong and can bear
structural loads. Poorly consolidated deposits are those with large void spaces.
Sands transported by wind, sediments deposited in lakes and marine environments,
organic soils in marshy areas are some examples.

2.2 Knowing sub-surface geology


Figure 2.2 is a map of Geology of Colombo area referred to in D.N. Wadia, 1941.

Figure 2.2: A sketch-map of geology of Colombo area (D.N. Wadia, 1941).

13
Session 2: Knowing soils of engineering significance

It identifies deposits (1) alluvium, (2) sand and semi-compact sandstone, (3) littoral
shelly sandstone, (4) crystalline rock outcrop, (5) laterite. Rivers, canals and lakes
are shown in solid black. The map shows the railway line and the main canal that
borders the main campus of the Open University. The map indicates that the main
campus region is underlain with laterite (belonging to red-yellow podzolic soil
group).To perform an engineering design, knowledge on soil beneath the surface,
up to sound bedrock is required. Geology maps show only the general geology of
the area and hence do not show enough detail.
Geotechnical investigations performed on nearby sites show the presence of sandy
and silty clays with organic material and organic soils with peat. The buildings of
the Colombo Regional Centre were constructed on an earth fill, which was placed
during mid 70’s.
Mapping sub-surface geology is done for major civil construction work. The
Geology Map of Colombo (Wadia, 1941) shows the general soil and rock profile of
the region. This map is useful to understand the origin of soil and rock formations,
prior to planning a detailed sub-surface investigation.

2.3 Logging sub-surface profile


As we have seen earlier, the type of foundation system that needs to support a
building depends on the subsurface profile. During a geotechnical investigation,
soil profile and its variation are recorded. Such information is contained in a
borehole log. Logs of two neighbouring boreholes are used to deduce the
subsurface profile between the two holes.
A few years ago, the Open University had proposed a new multi-storey building for
the Faculties of Natural Sciences and Engineering Technology. Figure 2.3 shows
the selected location. You may find the new Automobile Laboratory at this
location. The Geotechnical Division of National Building Research Organisation
performed the subsurface exploration. Figure 2.3 shows the layout of four
boreholes that were sunk.

Figure 2.3: Site layout, Soil investigation for proposed academic building for the
Open University of Sri Lanka.

14
CVX4343 – Soil Mechanics

Figure 2.4 shows the boreholes logs BH4 and BH1 drawn to the same elevation.
BH4 is drilled to bedrock while BH1 was stopped midway. The soils were
identified in the field and later classified in the laboratory. The duel symbols (e.g.
SM, MH etc.) are based on the Unified Soil Classification System, a universal
method used to classify engineering soils. Soil types are identified as Gravel (G),
Sand (S), Silt (M), Clay (C) and Peat (Pt), with many qualifiers being used to
describe soils. Medium-dense, silty sand is a sand; which is medium-dense sand
having some amount of silt. Medium dense to loose silty sand is sand, which is a
mixture of medium-dense and loose sand.

Figure 2.4: Vertical profile through boreholes BH4 and BH1.


The graph shows the variation of Standard Penetration Resistance, (i.e. the N
Value), which represents soil compactness and strength. You may observe that N
value drops with depth from 1 – 5m but gains in medium dense to dense sand
layers. The soft organic clay and peat occurs at shallow depths, i.e. 3 – 5m. These
layers are weak in strength and are highly compressible.
Figure 2.4 is an attempt to correlate soil strata across the two holes. The solid lines
indicate when predictions are made with a higher degree of certainty; dotted lines
are used when information is not sufficient to conclude with certainty.

2.4 Knowing different soil types


Soil grains being a weathered product have the same mineral constituents as its
parent rock. They vary in size, shape, texture and strength. Natural soils do not
occur in pure form; but as a mixture of more than one soil type.

15
Session 2: Knowing soils of engineering significance

Sandy Soils
Sand is a granular soil that is transported via waterways and deposited in beaches
and flood plains. Aoline deposits are formed when sand is transported by air.
Sands are identified as coarse, medium and fine based on particle sizes and their
distribution. The finest particles are identified as silt. Sand particles come in
different shapes, namely, rounded, sub-rounded and angular. Rounded shapes are
formed through abrasion when being transported. Angular particles can interlock
better than rounded particles. They may also differ in surface roughness.
Surface roughness and grain interlocking makes a soil form a heap. Have you seen
road aggregate being piled up along road shoulders or a sand heap in a construction
site? The slope angle maintained in such a heap is termed the `angle of repose’.
Sand with a high surface roughness maintains a greater angle of repose. If you
attempt to heap marbles you may find that it has a zero angle of repose.
Activity 2.1
Place a plank over a sand heap. Observe whether the heap could carry your weight
if you stand on top. The plank distributes your weight evenly. Picture how your
weight is transferred to the ground through the heap.
The frictional property of sand also depends on how compact it is. If a sandy soil
has particle sizes to fill most of its void spaces, we have a well-graded sand. A
well-graded soil has more particle contact points touching and hence can generate
more frictional resistance. A poorly graded soil is a soil with particles of a uniform
size range. Sometime you may find that certain sizes or size ranges are absent in
the mix. Such soils have more voids in the soil fabric and a lesser number of
contact points.
Sand is made of quartz, a mineral that has silica (SiO2) as its building block. You
may recall that two oxygen atoms form a stable structure with a silicon atom by
sharing its four electrons (two each) in the outer shell. This stable structure makes
sand an inert material.
O - 1s2 2s22p4 needs two electrons to be stable while Si – 1s2 2s22p6 3s23p2 requires
four electrons.
Activity 2.2
Make a dry sand heap in a bowl. Gradually fill the bowl with water until it tops the
sand heap. Observe whether the heap stays intact or it falls apart.
Sandy soils do not adsorb water hence drains off fast. However, when making a
sandcastle, moist sand can make walls to stand vertical. It is interesting to know
what makes sand sticky when moist; why dry sand and saturated sand heaps in a
similar manner, with no attraction between sand grains.
Adsorption is the process by which water molecules get attracted to the surface.
Absorption is the process by which molecules occupy void spaces within the
structure.
When void spaces between grains are sufficiently small, moisture forms a film of
water that holds grains together by surface tension. This force can hold sand
particles together making moist sand to be moulded to different shapes.

16
CVX4343 – Soil Mechanics

Activity 2.3
Mould a cylindrical shape using moist sand; place it in a bowl. Gradually fill the
container with water until it overtops the moulded object without disturbing it.
Describe what you see when sand becomes fully saturated.
Activity 2.4
Fill a tall glass with water until it is half full. Dip a transparent drinking straw and
mark the rise in water level. Fill the straw with very fine sand and record the
capillary rise in the straw. Does surface tension in soil capillaries cause water to
rise more?
Capillary rise above water table makes moisture available within the root zone.
Clayey Soils
Clay usage extends to prehistoric times, when earthenware was made.
Archaeological excavations in Pahiyangala rock cave have shown ancient pottery
dating 35,000 years. Makulumeti was used in ancient frescos to make paint
coatings. `Samara’ is a form of clay that is used to paint walls. China Clay
commonly known as `kirimeti’, has clay mineral Kaolinite. Kaolinite is used in
ceramic ware, electrical insulations and in sanitary ware. Building clays are used
to make bricks and tiles. These are rich in iron and hence have a reddish colour.
Such deposits occur in residual or alluvial deposits. Ball clay known as `Bola
meti’ is a clay used for ceramic ware. Ball clay is found to be more `plastic’
compared to Kaolinite hence has better bonding.
When a pat of moist clay is subjected to thumb pressure it deforms; when the
pressure is relieved it stays deformed. Plasticity is the property that makes clay
mouldable.
Plasticene shows plasticity and like clay it can be moulded to shapes. The stuff
that we played as kids is also named `clay’, even though it is made of plasticene.
Clay particles are smaller than sand particles. When inspected through an electron
microscope, these appear as platy or elongated particles. Clays are formed from
feldspathic minerals. The three main clay mineral types are Kaolinite, Illite and
Monmerillonite.
Clay particles aren’t inert, but carry negative charges on its surface and positive
charges at its edges. The surface charge makes water molecules to be adsorbed to
the surface. A lean-clay adsorbs less water while a fat-clay adsorbs more water. A
Montmerrilonite mineral grain has a higher negative charge that enables it to attract
more water molecules. It also has a high specific surface area (i.e. area per unit
mass of soil) thus has more sites for water molecules to attach. Such clays adsorb
more water and also show a high plasticity over a range of water contents. When
more water occupies void spaces, such soils become more compressible and are
weak in strength, hence do not possess favourable engineering properties.

17
Session 2: Knowing soils of engineering significance

Residual Soils
Residual soils are formed during in-situ weathering of parent rock. Sri Lanka,
being a hot and humid region with high rainfall, forms many residual soil types.
Latarite commonly known as `kabook’, is a secondary formation observed in the
wet zone of Sri Lanka. Figure 2.5 shows such a formation obtained from a coastal
hillock in Thalalla. Lateritic soil profiles can also be observed in roadside cuts or
on lands where lateritic soil mining has taken place.

Figure 2.5: A coastal lateritic hill.


During the formation of lateritic soils many mineral types are leached by water
moving under gravity. Its reddish brown colour is given by insoluble aluminium
and ferric ions that are retained in the soil.
Many ancestral homes were built using Kabook blocks. Presently partially
weathered lateritic soils are used in landfills and embankment fills. Such soils
should have a lesser clay content to be used as a good fill material. Humid tropical
countries such as Sri Lanka have residual soil deposits with varying degrees of
weathering. Highly weathered soils are considered unsuitable as a fill material.
Organic soils, Peat and Muck
The low-lying western and south-western coastal peneplain contains organic soils,
peat and muck. Organic soils are accumulation of highly organic material formed
in place by the growth and decay of plant life. The organic content in such soils
vary significantly. Peat is a somewhat fibrous aggregate of decayed and decaying
vegetation matter having a dark colour and odour of decay. Muck is formed when
peat deposits which have advanced in stage of decomposition to such extent that
the botanical character is no longer evident (NAVFAC, DM 7.1, 1982).
Organic soils have a high compressibility and vary depending on the amount of
decay. Such soils are weak in strength and hence need to be improved through
surcharging.

18
CVX4343 – Soil Mechanics

To sum up
This session introduces you to soils from an engineering perspective. Soils are
identified as sandy, clayey, lateritic/residual and organic/peat.
We intend to see how these soils respond to engineering situations.

Reference
NAVFAC DM7.1 (1982) Soil Mechanics Design Manual 7.1, Naval Facilities
Engineering Command, US Department of Navy.

19
Session 3
Soil and rock as a three-phase particulate system

What you will learn in this session


This is a 3-hour session. Section 3.1 explains how natural soils are made with solid
soil particles, water and air. The interaction between the three phases decides the
engineering behaviour of soils. Section 3.2 describes how soil parameters are
defined based on quantities representing volume and mass of the three phases.
These parameters quantify the amount of solids, voids, and water that fills voids.
Section 3.3 shows how these basic quantities express parameters commonly used in
the analysis and design of soil structures. Learners are expected to be able to
derive these parameters from basic definitions. Section 3.4 explains how total
vertical stress is computed. This is found to vary linearly with depth. In saturated
soils, water pressure reduces soil weight. Section 3.5 explains how the submerged
unit weight (or density) is determined.

3.1 Soil as a three phase system


Soils are made of three phases: solid, liquid and Units of measurement:
gaseous. A natural soil has solid particles (s), Mass g, kg
water (w) and air (a). The quantity of the three Volume m3
phases present in a soil mix and their interactions Density kg.m-3, g.cm-3
Unit weight kN.m-3
determine their behaviour.
We know that moist sand can be moulded while dry or saturated sands heap. A
potter knows the consistency of wet clay required to make pottery. The term
consistency indicates the amount of water in clay.
When soils are compacted water is added to achieve better compaction. We would
like to know the correct water content that is needed to lubricate solid grains.
When building a house on a soft compressible ground, we are concerned about
settlement. During ground settlement, we study how void ratio changes with time.
Flow of water through soils depends on how porous the soil is, hence flow depends
on its porosity.
In studying such behaviour, the three-phase soil matrix is represented by quantities
representing mass and volume. The parameters representing solid and water
phases are M s , M w , Vs and Vw. The respective total volume, V and total mass, M
are expressed as V = Vs + Vw + Va and M = Ms + Mw.
Water and solid phases are considered incompressible and hence the densities s
and w remain constant. Density of distilled water, w measured at 200C is
1.0g/cm3 (or 1000kg/m3).
The volume parameters are defined as Vs, Vw and Va (refer Figure 3.1). Here, we
neglect Ma since this is insignificant.

3.2 The definitions


The density of solids relative to water is termed the Specific Gravity. This is
denoted by Gs and is expressed by ratio s/w. The specific gravities of soils lie
between 2.65 (for sandy soils) to 2.72 (clayey soils). Silica minerals are less dense
than alumino-silicates.
CVX4343 – Soil Mechanics

Phase Volume Mass


Air VA 0
Water VW MW
Solid VS MS
Total V M

Figure 3.1: Quantities representing the three phases.


The solid soil particles being incompressible, hence their total volume, Vs is
unchanged. The total volume, however changes due to sliding and rotation. Water
is considered incompressible at relatively low soil stresses we consider; hence its
volume Vw. remains unchanged. In a fully saturated soil a continuous water phase
exists. This is because all void spaces are inter-connected. This continuous water
mass gives rise to a hydrostatic stress that varies linearly with depth.
The amount of void spaces in a soil decides how close the particles are placed.
This is expressed using volume parameters: void ratio, e and porosity, n. Void ratio
is expressed as Vv/Vs. This is defined with respect to solid volume. Porosity is
expressed as n = Vv/V, which is defined with respect to total volume. Table 3.1
lists typical void ratios of poorly graded and well-graded soils.
In-text question
e
Derive the relationship n  .
1 e
Table 3.1: Typical void ratios
Material emax - loose emin - dense
Equal spheres 0.92 0.35
Poorly graded 0.8 – 1.1 0.4 – 0.5
Well-graded 0.85 – 1.2 0.14 – 0.4

A partially saturated soil has its void spaces


partially filled with water and the amount of
water in soil is quantified using water content, w
and degree of saturation, S (refer Figure
3.2).Water content is based on mass and is
expressed as Mw/Ms. The degree of saturation is
defined in terms of volume and is expressed as Figure 3.2: Phase diagrammes
Vw/Vv. for a (a) fully saturated soil
and a partially saturated soil.
For a saturated soil, the two phases are quantified
in terms of w and Gs, while for a partially
saturated soil the three phases are expressed in terms of Gs, w and V.

21
Session 3: Soil and rock as a three-phase particulate system

In-text question
For a saturated soil which of the following parameters equal unity? S, w, e or n?
In a natural soil sample we can measure its water content. First we weigh the
container, then the soil sample; again after oven drying at 1050C.
Mass of container (g) = x
Mass of wet soil + container (g) = x + Mw + Ms
Mass of dry soil + container (g) = x + Ms
These equations yield Ms and Mw and hence the soil water content, w.

In-text question
n w n 1
Can we show that water content, w     . Starting from its
1  n s 1 - n G s
M  V  w Vw
definition w  w  w w  ; divide both numerator and
Ms  s Vs  s (V  Vv )
denominator by V and multiply the numerator by Vv/Vv (= 1) we get:
Vw V V V V
w w w v w w v
V V Vv Vv V Sn
w     .
(V  Vv )  s (1  n )  s (1  n ) (1  n )G s
s
V
To determine porosity, n we should know S, w and Gs. Gs can be measured in a
laboratory using a specific gravity bottle. If not we could assume a suitable value
between 2.65 – 2.72. We have also seen how w is measured. For saturated soils, S
n
= 1, hence w  .
(1 - n)Gs
Vw M w /  w M w / w
S is computed using the relationship: S    . You
Vv (V - Vs ) (V - M s /  s )
may note that we also need to know the total volume, V to compute S.
Void ratio is related to water content and is expressed as Se = wGs. Sometimes you
may be asked to show how such parameters are computed based on measured
quantities. Some learners attempt to memorise these relationships where as they
are expected to use first principles (i.e. starting from the definition) to prove these.
Let us see how this is done.
The term wGs is expressed as =(Mw/Ms)(s/w); and is simplified as
M M V V V V V
 W . s s  w . This gives us w . v  w .
M s M w Vw Vs Vv Vs Vs
Total density is the ratio of total mass to total volume. This is commonly known as
bulk density and is expressed as:
Ms  M w M
 bulk  . Dry density  d  s represents the ratio of dry mass to total
V V
volume. The saturated density is the total density when the soil is completely
saturated, i.e. when all voids are filled with water.

22
CVX4343 – Soil Mechanics

Ms  M w
This is expressed as  sat  . These densities are used to compute
Vs  Vw
vertical stresses with depth.

3.3 Indices used in engineering analysis and design


The indicesthat we’ve discussed in the above section represent engineering
properties, which are used in analysis and design. For clay soils Atterberg
consistency limits, namely, the Liquid Limit and the Plastic Limit characterises
`plasticity’. These limits are expressed as the water content at a particular
consistency.
Compressibility of a saturated soil is represented by its void ratio, e. The degree of
in-situ compaction is expressed in terms of dry density d.
We have seen how w, S, d,  bulk, and sat were defined using mass and volume
parameters.
The SI unit of density is kg/m3. The same in cgs system is g/cm3. Unit weight
(N.m-3) is the weight per unit volume. This is expressed as =g where  is the
density (kg.m-3); g is acceleration due to gravity (m.s-2).To obtain unit weight the
density should be multiplied by the acceleration due to gravity, g. The unit weight 
is expressed in kN/m3. The density of water is 1000kg/m3; its unit weight is
9.81kN/m3.The unit weight of water in fps system of units is 63.5lb/ft3. Pounds
(lb) is a measure of force and not a measure of mass!
These indices are found through measurements done on relatively undisturbed
M  M w Ms
bore-hole samples. For instance, bulk density  bulk  s  1  w  can
V V
be computed by measuring mass of solids, M s ; water content, w and the total
volume, V. We’ve seen in the previous section how water content is measured.
For a cylindrical shaped sample, total volume can be easily determined by
measuring average length and diameter. This sample can be oven dried to obtain
Ms .
Measured parameters are related to indices soil through certain identities. Learners
are expected to verify these using first principles. Following identities express
bulk density as:
1   s  Se   w  G s  Se 
 bulk     w
1  e  1 e 
(1  n ) s  Sn   w
 bulk   (1  n )G s  Sn    w
1
Let us see how we could prove this relationship. bulk is defined as the ratio of total
M  Mw
mass to total volume hence  bulk  s . Here the total volume is not
Vs  Vv
expressed as V=Vs+Vw+Va but as the sum of volumes of solids and voids. You
may also see that e is related to volume of solids and voids and not Va. If we
divide the numerator and denominator by Vs bulk density can be expressed as
M M
s  w s
Vs M s   w s
 bulk  . This is simplified as  bulk  s . Since Se = wGs, it
1 e 1 e

23
Session 3: Soil and rock as a three-phase particulate system

 s  Se  w
can be expressed as  bulk  . You would also see that w can be made a
1 e
 G  Se 
common factor in the numerator, hence  bulk   s  w .
 1 e 
When void spaces are saturated completely with water, the degree of saturation S
becomes 1. You may note that change in S does not affect the void ratio, e. Hence
we could use the expressions for bulk to obtain saturated densities as:
1  s  e   w  G s  e 
sat      w . This is simplified as:
1  e  1 e 
(1  n )s  n   w
sat   (1  n )G s  n    w
1
When void spaces are empty, S becomes 0. We could obtain an expression for dry
density d using the expression for bulk:
1 s  e   w  G s 
d 
1  e
    w and  d  (1  n)G s    w
1  e 
1   s G s   w  bulk
The dry density can also be expressed as  d    . Table 3.2
1 e 1 e 1 w
lists typical density ranges for the major soil groups.
Table 3.2: Typical density ranges for major soil groups.
Soil type sat (kN/m3) dry (kN/m3)
Gravel 20 – 22 15 - 17
Sand 18 – 20 13 - 16
Silt 18 – 20 14 - 18
Clay 16 – 22 14 - 21

The term Relative Density Dr is used to express how well a sandy soil is
compacted. It compares the in-situ void ratio with the densest and loosest packing
represented by emin and emax, respectively.
 e e   d max   d   d min 
D r   max   100%     100%
 e max  e min  d   d max   d min 

Actvity 3.1:
Can you explain how you would determine the densest and the loosest packing for
a pure sand?

The zero-air voids situation


The Standard Proctor Compaction Test requires you to plot the `zero air void line’.
Degree of saturation, S quantifies the amount of water in voids and zero air void
situation occurs when S = 1.
Since the relationship has to be plotted on a  d  w plot, d has to be expressed as
a function of w and S. If we are to start from the basic definition, d = Ms/V and

24
CVX4343 – Soil Mechanics

volume V = Vs+Vv, total volume V can be written as V = Vs(1+e). This gives us


G
 d  s . We saw earlier how void ratio is related to S, w and Gs and is
1 e
Gs
expressed as: Se = wGs. Hence d is written as:  d  . When S = 1,
1  wG s / s
Gs
this gives  d  . The `zero air-void line’ is in fact not a line, since w is
1  wG s
inversely proportional to d.

One-dimensional settlement
Figure 3.3 shows the three-phase diagram for a One Dimensional Consolidation
Test sample. The sample is placed in a solid brass ring, hence changes in volume
is proportional to changes in sample height.

Figure 3.3: Schematic showing the consolidation process.


The sample before consolidation has all three phases. During loading, air and
some water are squeezed out. When all air voids are compressed, the sample
becomes saturated. Hence at end of test,we have a completely saturated soil
specimen.
The height of water at end of consolidation, h wf can be computed knowing mass
of water and taking w=1.0g/cm3. The initial height of sample h0 is measured. The
dial gauge measures the change in height h. We know that height of solids
remains the same before and after, and can be expressed as h s  h 0  h  h wf .
h  hs
This enables us to determine initial void ratio e0, e 0  0 and final void ratio
hs
h  hs (h  h s )  (h f  h s ) (h 0  h f ) h
ef  f . e0  ef  0  . Since 1  e 0  0 ,
hs hs hs hs
e  e f h
these yield 0  .
1  e0 h0

3.4 Indeces representing rock properties


Index properties of rock are used to determine characteristics of rock. Properties of
rock mass however, depend on characteristics of rock and its discontinuities. Rock
porosities are found to reduce with increasing age and depth. Some typical
porosity values are given below: Granite 0.5 – 1.5; Sandstone 5 -25; Limestone 5 –
20; Quartzite 0.1 – 0.5; Gneiss 0.4 – 1.3.
Rock densities depend on both specific gravities of mineral constituents and
porosity. Dry densities in g/cm3: Granite 2.6 – 2.7; Basalt 2.8 – 3.0; Gneiss 2.6 –
2.9; Limestone 2.3 – 2.7; Marble 2.4 – 2.7; Coal 1.1 – 1.4; Sandstone 2.6 – 2.8;
Basalt 2.8 – 3.0

25
Session 3: Soil and rock as a three-phase particulate system

3.5 Computing soil pressures at a depth


The hydrostatic stress acting on an element of water at a 10m depth is
10m1000kg/m39.81m/s-2. This is 98,100N/m2 or 98.1kPa. Since water is an
isotropic material this pressure is the same in all directions. The stress is
proportional to depth and hence its variation with depth takes a triangular shape
(refer Figure 3.4).

Figure 3.4: Hydrostatic stress at element A is hg.


As we know soil layers extend several meters, hence the in-situ vertical stress of a
soil element located at a certain depth is due to soil weight above the soil element.
This is termed the weight of overburden. The vertical overburden stress caused by
a soil mass, at a given depth h is expressed as  = hg.
A soil with a unit weight  has a vertical stress equal to h. For instance, the
estimated vertical stress at 10m depth, for a soil with a bulk unit weight of
18kN/m3, is 180kPa.
Bulk unit weight of rock may vary significantly over a greater depth due to
variations in mineralogy and void spaces. This has to be considered in
underground construction work.
Soil being an anisotropic material, its horizontal stress is not the same as its vertical
stress. We will learn later how the horizontal stress resulting from the overburden
is computed.

3.6 Accounting for buoyancy in saturated soil and rock


We all weigh less in water and this is due to buoyancy. Buoyancy is due to the
upthrust caused by water pressure.
Seawater makes you weigh less, since its density is greater than that of fresh water.
Same is true for a saturated soil or blocks of rock with interconnected pore spaces.
High mineral content makes Dead Sea denser. This makes you float easily.
The net stress acting on a saturated soil element located a depth h (refer Figure 3.3)
would be the total stress, satgh, less the water pressure wgh. This net stress gh
is expressed as gh  sat gh   w gh ;  is defined as the submerged density.

What we `feel’ as our weight is the upward reaction given by surface we stand or
the chair we sit. When we are in water this reaction is reduced due to the up-thrust.
The above equation tells us that  can be expressed as:   sat   w .This is further
simplified to give:
 G  1 (1  n ) s  (1  n ) w
   s   w and    [1  n ]G s  1 w .
 1 e  1

26
CVX4343 – Soil Mechanics

To sum up
This session discussed how engineering parameters based on the three phase model
are derived. These parameters are obtained based on mass and volume
measurements made on laboratory samples of soil and rock. Learners are required
to be able to derive these parameters using first principles.
Learners are also expected to be familiar with typical values or ranges of values
observed for various soil and rock types.

27
Session 4
Towards classifying engineering soils

What you will learn in this session


This is a 2-hour learning session. Section 4.1 explains why engineering soils are
classified. Section 4.2 describes Sieve Analysis and Hydrometer Test and explains
how these tests are used to describe soils. Section 4.3 explains how plasticity is
expressed in terms of Liquid Limit and Plastic Limit. It further explains how
Casagrande’s Plasticity Chart distinguishes silts from clays and high plastic soils
from low plastic soils. This section also shows how Liquidity Index quantifies the
consistency of an in-situ clay soil. Section 4.4 explains the clay structure and what
makes clays to be `sticky’.

4.1 Introduction
Soils are classified to help us identify their suitability as an engineering material.
For instance, we may be interested to know whether a certain slope is stable; a
footing would not undergo significant settlement or whether an embankment dam
may breach due to piping.
Natural soil deposits aren’t homogeneous; however, they can be grouped spatially,
based on their material type and properties. More than 80% of Sri Lankan soils are
of residual origin. They display sandy and clayey properties.
Engineering soils are identified as `sticky’ (i.e. clayey) and `non-sticky’ (i.e.
sandy). Soils are granular since they are made of mineral grains. Earlier we learnt
that clayey soils interact with water as well as its own solid particles while sandy
soils do not.
Sandy soils are coarser than clayey soils. These are made of quartz with silica as
its constituent; particles are angular or spherical in shape; chemically inert hence
no interaction with water; drains rapidly through inter-connected voids; less
compressible; higher strength due to particle interlock.
Clayey soils are made of felspathic minerals such as kaolinite, illite or
montmorillonite. The particles are platy or needle like; display `plasticity’; adsorb
water and chemical ions in pore-water; attracts other particles; have less inter-
connected voids, less interlocking hence show a lower strength; high
compressibility.
Engineering soils are classified visually or based on laboratory test results. During
a sub-surface investigation, an experienced logger (a person who logs the sub-
surface profile) identifies soils through visual examination. These classifications
are verified using laboratory tests that determine particle sizes and size distribution,
and plasticity.

4.2 The distribution based on particle size


Particle size distribution of a soil gives us the distribution of soil grains expressed
as a percentage of mass, for varying particle sizes. You will learn in this section
that the distribution of `coarser’ particles are obtained using the Sieve Analysis
Test while `finer’ particles are estimated based on the Hydrometer Test.
CVX4343 – Soil Mechanics

Sieve Analysis Test


The Sieve Analysis Test determines per cent soil passing a particular particle size.
This is computed based on dry mass of soil retained on each sieve in a sieve stack
(refer Figure 4.1a). Sieves are arranged in the descending order of sieve sizes and
is shaken using a mechanical shaker (refer Figure 4.1b). After shaking, each sieve
will hold particles that pass through the previous sieve. Per cent soil for sizes
smaller than a particular sieve size is expressed as the per cent passing the
corresponding sieve size.

Figure 4.1 (a) A sieve stack, (b) the sieve shaker.


Soils are grouped as cobbles, gravel, sand, silt and clay based on particle size.
Table 4.1 gives the corresponding size ranges in millimetres. The size that divides
coarse-grained and fine-grained soils is 0.063mm (63-microns). 63 is the
smallest sieve opening used in a conventional stack of sieves and therefore this test
can only be used to determine the quantity of fines and not the per cent fraction of
clay or silt. The size distribution of sizes smaller than 63 is determined by the
Hydrometer Test. The size that differentiates clays from silts is taken as 2.
Table 4.1: Particle sizes dividing soil to main soil groups.
Clay Silt Sand Gravel Cobbles Boulders
< 0.002 0.002-0.063 0.063-2 2-60 60-200 > 200

The graph sheet that plots particle size distribution is shown in Figure 4.2. Observe
the scale and sieve sizes shown. You may note that silt, sand and gravel have sub-
groups identified as fine, medium and coarse.
Activity 4.1
Write the particle size ranges for the above soil groups and sub-groups. Do you see
a pattern in the identified size ranges?
Sieve Analysis Test is performed on soils with fines less than 12%. You may learn
later that this number is stipulated in the Unified Soil Classification System
(USCS).Particles of natural soils can sometimes be loosely cemented with
compounds of carbonates, and compounds of iron and other metals. Such soils are
first treated with dilute hydrochloric acid. Inorganic soil particles are sometimes
coated with organic matter, hence they are pre-treated prior to their use.

29
Session 4: Towards classifying engineering soils

The normal sieve analysis test is performed on a dry soil sample. The minimum
mass of dry soil to ensure proper sieving is specified as: 150g for clays, silts and
fine sand; 2.5kg for medium sand to fine gravel and 17g for medium gravel to
cobbles. This helps us represent respective proportions accurately.

Figure 4.2: Graph sheet used to plot the particle size distribution of a soil.
A triple beam balance (refer Figure 4.3) is used to
obtain mass of soil retained on each sieve.
Sometimes samples of naturally occurring soils
are pulverised lightly to separate particles. When
a partially weathered lateritic soil is tested a wet
sieve analysis is used. This is done by washing
the soil through a 63 sieve to separate particles
that are partially weathered and lightly cemented. Figure 4.3: Triple beam
The soil retained on the said sieve is then dried balance.
and sieved through the regular stack of sieves to
obtain its size distribution.
A soil is considered well graded when the sample contains a well distributed range
of particle sizes. A poorly graded soil may be uniformly graded or gap-graded. A
uniformly graded soil has only a few particle sizes represented while a gap-graded
soil is represented only by a certain range(s) of particle sizes.

Activity 4.2
Sketch a well graded soil, a uniformly graded soil and a gap graded soil on the
graph sheet shown in Figure 4.2.

30
CVX4343 – Soil Mechanics

Activity 4.3
Complete the table given below. The third column indicates the mass of soil
retained on each sieve and the pan. Then compute cumulative mass retained at
each sieve level. This would be the mass retained on the sieve considered plus the
mass retained on sieves stacked above it. Next, per cent of cumulative mass
retained at each sieve level can be computed and hence per cent passing.
Plot the per cent passing with particle size on the above graph sheet. Join the
points with a smooth curve. Determine the per cent fraction of the major soil
groups.
Sieve Size Mass Cumulative Percent Percent
(mm) retained (g) Mass Retained Retained Passing
(g) (%) (%)
63.0 0
20.0 10.2
6.3 30.0
2.0 48.7
0.6 84.2
0.212 59.9
0.063 37.4
pan 45.3
TOTAL

The gradation above is classified based on the


Coefficient of Uniformity, C u and Coefficient of
Curvature, C c . C u is defined as the ratio
D 60 / D10 . D 60 and D10 are particle sizes where
10% and 60% pass. A well-graded soil has a high
C u . C c is the ratio D 30  D10 D 60 . C c
2

between 0 and 1 indicates that the soil is well


graded. How would C u and C c values for the
three distribution curves compare?
The Hydrometer Test
The Hydrometer test is used to determine the
particle size distribution of fine-grained soils. A
hydrometer measures the density of soil in
suspension. At a given time if say 23g of solid
particles are in suspension, the weight contained
in 1litre is approximately 1023g. This is made of
~1000g of water plus 23g of solid, hence the
density of suspension is ~1.023 g.cm-3. Figure
4.4 shows the hydrometer stem. A calibrated
hydrometer will float with the free surface
reading 1.0 23. The hydrometer reading is
expressed as 23.
The hydrometer reading obtained at a given time Figure 4.4: Hydrometer stem
is used to compute the particle size and per cent with calibration.

31
Session 4: Towards classifying engineering soils

soil in suspension (i.e. per cent passing). The frictional drag acting on a settling
particle is expressed using Stokes Law.
Particles attain a terminal velocity, v when viscous drag equals the buoyant weight,
and the drag force on a `spherical’ soil grain with radius, a is expressed as
F  6  av.
When particles settle, clay particles tend to aggregate due to inter-particle
attraction, assisted by water molecules. This makes particles to form flocs making
them bigger and heavier. This makes these flocs to settle faster, resulting in an
incorrect size distribution. To overcome particle flocculation, a deflocculant is
added. Particles can also be cemented to each other due to carbonates, oxides and
organic matter and hence removing these prior to performing the hydrometer test is
required. The test apparatus is shown in Figure 4.5.

Figure 4.5: Hydrometer test apparatus


Performing a combined sieve-hydrometer test
A combined sieve-hydrometer test is recommended when a soil has finer particles
attached or cemented to coarser particles. This is seen in highly weathered
laterites. The hydrometer test is done for sizes finer than 63, which is later
combined with the wet-sieve analysis. If total soil mass = 155g; Mass retained on
63m sieve = 115g, hydrometer test is done on 45g. The combined test computes
the per cent passing based on 155g.

4.3 Atterberg Consistancy Limit Tests


The Atterberg Limit tests, the Liquid Limit and Plastic Limit tests quantify
plasticity of fine-grained soils.

Liquid Limit Test - Casagrande’s method


Clays when mixed with water can be moulded to different shapes. When a clay
soil is mixed with excess water it forms a clay-suspension. Suppose that we reduce
the water content by mixing with dry clay, it becomes thicker and more viscous;
when more clay is added it ceases to flow. The water content corresponding to the

32
CVX4343 – Soil Mechanics

lower consistency limit of viscous flow is termed the Liquid Limit. Can you
describe the consistency of clay when used for pottery or brick making?
When more dry soil is added, the soil becomes more plastic, which can be
moulded. Water content corresponding to the lower consistency limit, at which
plasticity is shown, is termed Plastic Limit. If its water content is reduced further,
soil comes to a consistency where further loss of moisture does not cause any
reduction in volume. This is termed the Shrinkage Limit (refer Figure 4.6). These
consistency limits were proposed by Atterberg to classify agricultural soils. Arthur
Casagrande showed that Liquid Limit and Plastic Limit have a special place in
characterising engineering soils.

Figure 4.6: Variation of total volume with water content


Atterberg consistency limits for different soils are established through standard
laboratory tests. Atterberg limit tests are performed on the soil fraction containing
fine sand, silt and clays, i.e. the fraction finer than 0.425mm size. Soil is mixed
with water and left for a period of 24 hours for it to attain chemical equilibrium.
Figure 4.7 shows the accessories used in Casagrande’s method. This method
identifies Liquid Limit as the consistency or water content when the standard
groove closes by 13mm at 25 blows. Figure 4.8 shows the plan view of the cup of
the Liquid Limit device, before and after groove closure.

Figure 4.7: Casagrande’s apparatus.


33
Session 4: Towards classifying engineering soils

Figure 4.8: Groove closure.


The water contents of a minimum of five specimens ranging 15 – 35 blow counts
are obtained. Figure 4.9 shows the plot of water content versus the blow count, and
the Liquid Limit, wL. The one point method approximately determines the Liquid
0.092
N
Limit, wLas w L  w   , where w is the water content corresponding to N
 25 
blows.

Figure 4.9: Relationship between water content and blow count.


Figure 4.10 shows the electronic balance with a
sensitivity of 0.001g. This is used to measure wet
and dry weights of soil specimens.

Figure 4.10: Electronic


balance

Liquid Limit Test - Cone Penetrometer Method


The Cone Penetrometer method uses a standard cone (refer Figure 4.11). The
Liquid Limit is identified as the consistency required to allow the cone to penetrate
20mm in to soil at a standard penetration rate. This point is established by plotting
penetration values for the water content range of 15 – 35% (refer Figure 4.12).

34
CVX4343 – Soil Mechanics

Figure 4.11: Cone penetrometer

Figure 4.12: Relationship between penetration and water content.

Plastic Limit Test


The Plastic Limit Test requires rolling a soil lump to a soil thread of 3mm
diameter. This is done on a glass plate while applying a uniform pressure. At
Plastic Limit consistency, the thread crumbles when folded. The test is repeated
until the calculated water contents differ by not more than 0.5%. Figure 5 shows
you how sample volume changes with water content.
The range over which a soil exhibits plasticity is termed the Plasticity Index. This
is the difference between Liquid Limit and the Plastic Limit.
Figure 4.13 shows Liquid Limit and Plasticity Index values when plotted on the
Plasticity Chart. This chart is named after Professor Arthur Casgrande. It is
evident that for the three main clay mineral types, the range of plasticity increases

35
Session 4: Towards classifying engineering soils

with the Liquid Limit. Soils with high plasticity retain more water within their clay
matrix. Montmorillonite has the highest Plasticity Index. What would be the range
of Liquid and Plastic Limit for Kaolinite?
The Liquidity Index (LI) compares the field moisture content of a soil sample with
w  wP
its plasticity range. This is expressed as LI  n . LI is a measure of the
PI
consistency of a saturated clay deposit. Over-consolidated clays have low in-situ
water contents and are closer to their Plastic Limit while normally consolidated
clays have water contents closer to their Liquid Limits.
The Flow Index (FI) is the difference in water contents corresponding to 10 blows
and 100 blows. It indicates loss in shear strength with increasing water content.
This is obtained from the semi-logarithmic plot shown in Figure 5.
The Toughness Index (TI) is the ratio PI/FI. Activity of a clay soil is defined as:
PI
A . Clays with high activity has A> 1.25. Such clays show a
% fraction 2
significant volume change during wetting and drying.

Figure 4.13: Casagrande’s Plasticity Chart.

4.4 Why clays differ from sands?


Clays are formed from inorganic crystalline grains of hydrated alumino-silicates
and sand constitutes mineral silica. Clays have sheets of silica tetrahedrons and
aluminium octahedrons (refer Figure 4.14).
These two types form the three main clay mineral types: Kaolinite, Illite and
Montmorillonite. Kaolinite has alternating layers of Silica and Alumina sheets
held together by hydrogen bonds (refer Figure 4.15). These sheets are stacked
together as layers to form Kaolinite grains. Illite minerals are formed with an
alumina sheet sandwiched between two silica sheets. Each unit is held together by
potassium ions. Montmorillonite minerals are formed with a silica-alumina
structure similar to that of Illite, held together by a layer of exchangeable ions,
easily penetrated by water.

36
CVX4343 – Soil Mechanics

Figure4.14: a) Silica tetrahedrons, b) silica sheets, c) single aluminium


octauhedrons, and d) aluminium sheets.

Figure 4.15: Structure of three main clay mineral types.

37
Session 4: Towards classifying engineering soils

We know that clay particles `hold’ water. Clay


particles have negatively charged surfaces and
positively charged edges. The surfaces also
attract water molecules (refer Figure 4.16).
A soil with a higher Liquid Limit can hold more
water. The attractive forces make layers of water
molecules to be adsorbed. This water layer is
highly viscous compared to `unbround’ water;
requires heating (up to 2000c) to break apart. The
presence of several such layers makes the soil
`plastic’. The term `plasticity’ traditionally refers
to mouldability. Clay particles have sizes of the
order of 2 or less; and its Specific Gravity
between 2.7-2.72. Figure 4.16: Charged clay
particles attracting water
Smaller the particle size, higher would be its molecules.
Specific Surface Area (SSA). SSA is the measure
of surface area per unit mass.
Table 4.1 compares physical properties of the three clay mineral types. We see that
Montmorillonite has the smallest particle size but has the greatest SSA. It also has
more sites for water molecules to adsorb. Adsorption refers to attraction of two or
more dissimilar material types.
Table 4.1: Properties of clay minerals
SSA Thickness Lateral
Type 2
(m /g) () dimensions ()
Kaolinite 15 0.1-1 1-20
Illite 80 0.05-0.5 1-5
Montmorillonite 800 0.01-0.05 1-5

This interaction between mineral type and pore-fluid type is termed the Physico-
Chemical Interaction. This makes some clay minerals highly compressible, some
expansive, less permeable and a weaker engineering material.

To sum up
This session explains the standard laboratory tests used to classify coarse-grained
and fine-grained soils. It describes the Sieve Analysis Test, the Hydrometer Test,
and Atterberg Consistency Limits. We have seen how coarse-grained soils are
identified based on particle size and its size distribution. The Liquid Limit and
Plastic Limit characterises the property `plasticity’ in fine-grained soils. This is
quantified using Casagrande’s Plasticity Chart.
We also learnt the reasons why clays absorb water and other dissolved ions, and
why sands do not show this behaviour.

38
Session 5
Laboratory classification of engineering soils

What you will learn in this session


This 2-hour session describes the soil classification systems. Section 5.1
introduces the Massachusetts Institute of Technology (MIT-BS) classification
system with the Unified Soil Classification System (USCS). Section 5.2 of this
session explains how soils are classified based on MIT system.

5.1 Introduction
MIT Soil Classification System and the Unified Soil Classification System (USCS)
classify soils based on Sieve Analysis, and Liquid and Plastic Limit tests. Both
methods use a `soil description’ and `a group symbol’.
The symbol contains a primary and a secondary character. The primary character
represents the main soil type and is identified as a Gravel (G), a Sand (S), a Silt
(M) or a Clay (C). A well-graded sand is identified as SW. If the soil is a mixture
of more than one type e.g. a well-graded silty sand, the dual symbol SW-SM is
used.
Soil description also describes other material types present in soil specimens. For
instance, SW-SM soil can have the description as well graded silty sand with some
clay.The MIT system differs from USCS, since size ranges of certain soil groups
differ. Table 5.1 and Figure 5.1 compare the two classification systems.
Table 5.1: Soil groups and their respective size ranges

Soil Group MIT USCS


Boulders > 200 mm > 300 mm
Cobbles 63 – 200 mm 75 – 300 mm
Gravel 2 – 63 mm 4.75 – 75 mm
Sand 63 - 2 mm 75 - 4.75 mm
Silt 2 – 63 2 – 75
Clay < 2 < 2

USCS identifies sand sizes to be between 75 to 4.75mm while MIT system uses
sand sizes between values 63 to 2.0mm. These differences sometimes create
confusion among geotechnical engineers when interpreting borehole logs. A
poorly graded gravel (GP) based on MIT system can be classified as a poorly
graded sand (SP) based on USCS. Many geotechnical testing laboratories in Sri
Lanka now prefer to use the USCS system and hence it is important that you are
conversant with both methods.

5.2 MIT Soil Classification System


Table 5.1 outlines the classification procedure. Column 1 explains to you how a
fine-grained soil is distinguished from a coarse-grained soil. A coarse grained soil
has less than 50% passing 63 sieve size. USCS uses 75 size instead. A coarse-
Session 6: Laboratory classification of engineering soils

grained soil is identified as sand if more than 50% of the coarse fraction is sand
(S). If more than 50% is gravel symbol G is used (refer Column 2).

Figure 5.1: Comparison of MIT system with USCS based on particle size ranges
Plasticity Chart distinguishes clays (C) from silts (M). Soils plotting above A –
line are identified as clays (refer Figure 5.2). Fine-grained soils are also identified
as low-plastic (L) or high-plastic soil (H) and are based on its Liquid Limit (wL).
When classifying coarse-grained soils, the quantity of fines determines the
secondary symbol. Column 4 of Table 5.2 shows how it is done. When per cent of
fines is less than 5, such soils are classified as well graded (SW, GW) or poorly
graded (SP, GP). The table also explains how Coefficient of Uniformity ( C u ) and
Coefficient of Curvature ( C c ) are used to describe gradation.
When per cent fines exceed 12 (implying that it is 12 – 50%) the secondary symbol
is obtained by referring to the Plasticity Chart. Such soils are classified as GC, SC
or GM, SM.
A soil is made of S-38%, G-17% and 45% fines. What would be its primary
symbol, G or S?
A fine-grained soil has wL=42 and wP=18. Is this soil a clay or a silt? Would you
classify it as high plastic (H) or low plastic? (Ans. CL)
Fine-grained soils are identified as clays or plastic fines (C), silts or non-plastic
fines (M) and organic soils (O). These soils are classified based on the Plasticity
Chart. Clays plot above A-line while silts and organic soils plot below. The
plasticity is expressed as low (L) and high (H) and is done based on a Liquid Limit
of 50.

A soil with per cent fines 5 – 12 falls between the two cases discussed above (i.e.
Fines < 5% and Fines > 12%) and hence is given a duel symbol. For instance,
well-graded sand with 8% silty fines will be classified as SW-SM. It should be
noted that the type of fines (i.e. C or M) is decided based on the Plasticity Chart.
The shaded region of the Plasticity Chart (refer Figure 5.2) represents soils with
symbol CL-ML. This region is bounded by the A-line and lines with PI=4 and
PI=7. The MIT Classification system identifies peat as Pt. Unlike peat, residual
soils are not identified by a separate group symbol.

40
CVX4343 – Soil Mechanics

Completely weathered soils are classified as fine-grained soils. They form silty
clays (CL) or clayey silts (ML). Partially weathered soils have `sand’ and `gravel’
size particles and hence are classified based on per cent coarse fraction.

Figure 5.2: Casagrande’s Plasticity Chart.

Example
A combined sieve-hydrometer test yields the Size % finer by
following results. The Atterberg limit tests yield (mm) weight
wL = 35 and wP = 25. Classify the soil using 2 100
USCS. 0.425 100
The learner should note that: i) the soil is a coarse 0.212 94.1
grained soil (65.9% is retained on 63 sieve) ii) 0.15 79.3
65.9% is sand (since 100% passes 2mm sieve) 0.063 34.1
iii) percent fines exceed 12%. These 0.04 28
observations should conclude that the Group 0.02 25.2
Symbol should be either SC or SM. 0.01 21.8
We also need to know the per cent fractions of 0.006 18.9
each soil group. iv) The soil has 60% fine sand 0.002 14
(63-212) and 5.9% medium sand, with 0%
coarse sand. This identifies the soil to be
predominantly fine sand. v) The Atterberg limits plot below the A-line hence the
fines are silty.
Therefore the soil is classified as a fine silty sand SM. This problem did not
require you to plot the size distribution.

41
Session 6: Laboratory classification of engineering soils

Table 5.2: MIT Soil Classification System


Laboratory criteria

symbol
Group
Description
Fines Grading Plasticity Notes

Well graded gravels, sandy


CU>4;
gravels, with little or no GW 0-5%

Dual symbols if above `A’ line and 4<PI<7


1<CC,<3
Gravels (More than 50% of coarse fraction of gravel size)

fines

Dual symbols if 5-12% fines.


GW requirement
Not satisfying
Poorly graded gravels,
sandy gravels, with little or GP 0-5%
no fines

Below `A’ line


or PI<4
Silty gravels, silty sandy
GM >12%
gravels
Coarse grained (More than 50% larger than 63m

line or PI>7
Above `A’
Clayey gravels, clayey
GC >12%
sandy gravels
BS sieve size)

Well graded sands,


CU>6;
gravelly sands, with little SW 0-5%
1<CC,<3
or no fines
Sands (More than 50% of coarse fraction of sand size)

Not satisfying SW
requirement

Poorly graded sands,


gravelly sands, with little SP 0-5%
or no fines
Below `A’ line or PI<4

Silty sands SM >12%

Above
Clayey sands SC >12% `A’ line
or PI>7

42
CVX4343 – Soil Mechanics

Table 5.2: MIT Soil Classification System (continued)


Inorganic silts, silty or

Fine grained (More than 50% smaller than 63m BS

Silts and clays (Liquid


clayey fine sands, with ML Use Plasticity Chart

limit less than 50)


slight plasticity
Inorganic clays, silty
clays, sandy clays of CL Use Plasticity Chart
low plasticity
Organic silts and
organic silty clays of OL Use Plasticity Chart
low plasticity
sieve size)

Inorganic silts of high


MH Use Plasticity Chart
Silts and clays (Liquid limit

plasticity
Inorganic clays of high
CH Use Plasticity Chart
plasticity
greater than 50)

Organic clays of high


OH Use Plasticity Chart
plasticity

Highly Peat and other highly Pt


organic soils organic soils

To sum up
This session explains how soils are classified based on MIT classification system.
This is based on British Standard Specification BS 5930 (1981).
Learners are expected to be familiar with the method of classification described in
this session.

43
Session 6
Classifying soils based on visual inspection and feel

What you will learn in this session


Section 6.1 of this session presents a soil classification technique based on visual
examination and feel. This system uses a qualitative approach to interpret
observations. Section 6.2 explains how fine-grained soils are classified based on
tests: dry strength, dilatancy, toughness at Plastic Limit and plasticity. These tests
are performed on a sample fraction finer than 0.425mm size. The particle size
distribution of coarse-grained soils is compared with size ranges defined as
cobbles, gravels and sand subgroups (9.3). This is a 2-hour learning session.

6.1 Introduction
An experienced borehole logger (i.e. a person who prepares the field borehole log)
or a soil laboratory technician can identify soil types through inspection. When
classifying soils through visual inspection one needs to differentiate fine-grained
soils from coarse-grained soils.
We have seen earlier that Atterberg Limit tests are performed on soils passing
0.425mm size. Such samples contain fine sand, silt and clay. Fine-grained soils
are classified based on plasticity. Soils display plasticity to varying degrees, from
non-plastic silt to high plastic clays. The visual method identifies fine-grained
soils based dry strength, dilatancy, plasticity and toughness. These are a set of
simple tests performed on moist soil.
Coarse-grained soils are classified based on particle size; comparing them with
cobbles, gravel and sand sizes. Physical characteristics are noted and are used to
describe such soils.

6.2 Classifying fine-grained soils


The Dry Strength Test
The dry strength test is used to express the strength of an air-dried specimen. The
test is performed on three 12mm-diameter test specimens. Soil is mixed with
water, slightly in excess; then is removed by kneading until it is mouldable to a
25mm-diameter ball.Air drying process facilitates the gradual removal of moisture
from the specimen. A laboratory oven with a temperature set at 60oC can also be
used. Dry strength is determined based on the pressure that the dry specimen could
withstand. Table 6.1 lists the criteria that describe Dry Strength.
Table 6.1: Criteria describing Dry Strength

Description Observation on dry specimen:


None Crumbles into powder with some pressure of handling.
Low Crumbles into powder with some finger pressure.
Medium Breaks into pieces or crumbles with considerable finger pressure.
High Cannot be broken with finger pressure. Specimen will break into
pieces between the thumb and hard surface.
Very high Cannot be broken between the thumb and hard surface.
CVX4343 – Soil Mechanics

The Dilatancy Test


When walking on fine beach sand you may have noticed that heel pressure makes
moisture surrounding your heal to disappear; becomes moist when the pressure is
released.
A specimen is moulded in to a 12mm-diameter ball by adding water to make it to a
soft but non-sticky consistency. It is then smoothened on the palm with fingers or
using a small spatula. The specimen is held on the palm; striking the palm several
times on the side with the other hand (refer Figure 6.1a).The rate at which water
appears on the specimen surface is observed. The specimen is then squeezed, by
closing the palm or pinching between fingers (refer Figure 6.1b). The rate at which
water disappears from surface is noted.

(a) Striking the palm.

(b) Squeezing the soil.


Figure 6.1: The Dilatancy Test.

45
Session 6: Classifying soils based on visual inspection and feel

Table 6.2 describes criteria to assess the reaction. Fine sands and silts respond well
to this test. Smaller the pore size, higher reaction is. Silt has the smallest pore size
and hence show a `rapid’ reaction.
Table 6.2: Criteria describing Dilatancy

Descripti Observation on dry specimen:


on
None No visible change in the specimen.
Slow Water appears slowly on the surface of the specimen during shaking
and disappears slowly upon squeezing.
Rapid Water appears quickly on the surface of the specimen during shaking
and disappears quickly upon squeezing.

Toughness of soil thread at Plastic Limit


Toughness measures the strength of a soil threat near its Plastic Limit. The soil
specimen is made to an elongated pat and rolled on a smooth surface or between
palms into a 3mm-diameter thread. When a uniform thread is formed at this
diameter, it is folded. If the thread does not crumble, it is wetter than Plastic Limit
and hence need to be re-rolled. If the specimen begins to crumble at the said
diameter, specimen has reached its Plastic Limit (refer Figure 6.2).

(a) The thread breaks when folded.

(b) It crumbles when rerolled.


Figure 6.2: Plasticity Test

46
CVX4343 – Soil Mechanics

The toughness of thread at Plastic Limit is expressed as low, medium and high,
based on its consistency and pressure required to roll it to a 3mm-diameter thread.
Table 6.3: Criteria describing Toughness

Description Observation on dry specimen:


Low Only slight pressure is required to roll the thread near Plastic
Limit. The thread and the lump are weak and soft.
Medium Medium pressure is required to roll the thread near Plastic Limit.
The thread and the lump have medium stiffness.
High Considerable pressure is required to roll the thread near Plastic
Limit The thread and the lump have very high stiffness.

Plasticity
Plasticity is reflected by the number of times kneading and re-rolling is done in
order to bring the 3mm-diameter thread to its plastic limit. Non-plastic soils cannot
be rolled to the required size. Longer it takes to bring the thread specimen to the
plastic limit consistency, greater would be its plasticity. Table 4 lists criteria that
describe plasticity.
Table 6.4: Criteria describing Plasticity
Description Observation on dry specimen:
Non-plastic A 3mm thread cannot be rolled at any water content.
Low Thread can barely be rolled; lump cannot be formed when drier
than Plastic Limit.
Medium Thread is easy to roll; not much time is required to reach the
Plastic Limit; cannot be re-rolled after reaching Plastic Limit;
lump crumbles when drier than Plastic Limit.
High Considerable time to roll and knead is required for the specimen
to reach the Plastic Limit. Thread can be re-rolled several times
after reaching Plastic Limit. The lump can be formed without
crumbling when drier than Plastic Limit.

Table 6.5 relates the four tests described above to four main fine-grained soil types:
ML, CL, MH and CH.
Table 6. 5: Relating soil test to soil type
Soil type Dry strength Dilatancy Toughness Plasticity
Low plastic None to low Slow to Low or thread Low or thread
silt, ML rapid cannot be cannot be
formed formed
Low plastic Medium to None to Medium Medium
clay, CL high slow
High plastic Low to None to Low to medium Low to
silt, MH medium slow medium
High plastic High to very None High High
clay, CH high

47
Session 6: Classifying soils based on visual inspection and feel

6.3 Classifying coarse-grained soils


The physical appearance of coarse-grained soils is observed in field specimens.
Table 6.6 lists adjectives that describe soils, the properties that they represent, and
criteria used to assess these properties.
Classification requires identifying proportions of soil groups and sub-groups. This
is done by estimating per cent fraction of cobbles, gravel (coarse, medium and
fine), and sand (coarse and medium). Soil grains are physically separated; a
standard set of jars representing size ranges are used for this purpose (refer Figure
6.3). These proportions identify predominant soil type.
In-text question
A well-graded sandy soil is made of 10% coarse gravel; 10% medium gravel; 20%
fine gravel; 40% coarse sand and 10% medium sand. How would you classify this
soil?
The per cent fractions indicate that fine sand and fines total to 10%. Total sand
percentage exceeds 50%. Fractions belonging to various soil groups and sub-
groups seem to be well represented hence it is a well-graded soil. Learner is
expected to use relevant descriptors (refer Table 6.6) to describe the soil. In
absence of any information the soil can be classified as Coarse to medium sand
with gravel, SW.
Table 6.6: Criteria that describes physical appearance of coarse grained soils.
Property Description Criteria
Water Dry Absence of moisture, dusty, dry to the touch.
content
Moist Damp but no visible water.
Wet Visible free water.
Consistency Very soft Thumb will penetrate soil more than 25mm.
Soft Thumb will penetrate soil about 25mm.
Firm Thumb will indent soil about 4mm.
Hard Thumb will not indent soil but readily indent with
thumb nail.
Very hard Thumb-nail will not indent soil.
Cementation Weak Crumbles or break with handling or little finger
pressure.
Moderate Crumbles or breaks with considerable finger pressure.
Strong Will not crumble or break with finger pressure.
Structure Stratified Alternating layers of varying material or color with
layers at least 6mm thickness.
Laminated Alternating layers of varying material or color with
layers less than 6mm thickness.
Fissured Breaks along definite planes of fracture; with little
resistance to fracturing.
Blocky Cohesive soil that can be broken down into small
angular lumps which resist further breakdown.
Lensed Inclusion of small pockets of different soils, such as
small lenses of sand scattered through a mass of clay.
Homogeneous Same colour and appearance throughout.

48
CVX4343 – Soil Mechanics

Figure 6.3: Sizes from left to right silt (2 – 63), fine sand (63-212), medium sand
(0.212-0.6mm), coarse sand (0.6-2mm), fine gravel (2-6.3mm), medium gravel
(6.3-20mm) and coarse gravel (20-63mm).

To sum up
The session describes how engineering soils are classified based on visual
inspection and feel. It described the methods used to classify both coarse-grained
and fine-grained soils.

49
Session 7
How soil layers carry load

What you will learn in this session


This is a 2-hour learning session. Section 7.1 explains how soil grains bear
external loads. We know that hydrostatic stress caused by a column of water, at a
certain depth is isotropic. This section explains why average soil stress at certain
depth in considered anisotropic. This makes the vertical and horizontal stresses
unequal. Section 7.2 explains the Principle of Effective Stress, put forward by
Professor Carl Terzarghi to explain why external loads cause internal stresses in
soil grains (termed as the effective stress)and in saturated pores (identified as pore-
water pressure). Section 7.3 discusses how you could determine the variation of
these stresses with depth. In the case of confined aquifers (refer Section 7.4)
resulting artesian pressures influence the effective stress. Section 5.0 introduces
you to stresses in a partially saturated soil where particles are held together through
capillary suction.

7.1 How soil grains carry load


Solid soil skeleton and pore water form the soil fabric. A soil grains rests on its
neighbouring soil grains touching them at points of contact. These grains form a
stable structure termed the soil fabric (refer Figure 7.1). Each soil grain has its
mass hence is weight is transmitted at points of contact. In a heap of sand, the
contact stress acting on a soil grain increases with depth. We also know that it
could still carry an imposed load without causing it to collapse; also transmitting
this load at points of contact. A well graded soil has a closer packing hence is
stronger than a poorly graded soil.
When a dry soil is loaded the average contact stress is dispersed. This dispersion
occurs at points of contact. We observe that the number of grains holding the load
increases with depth, hence the number of contact points will also increase.
Therefore it is reasonable to conclude that the contact stresses are dispersed within
a volume demarcated by the two dispersion lines (refer Figure 7.2); stresses beyond
this region is small and hence insignificant.

Figure 7.1: A granular fabric Figure 7.2: Granular fabric when


loaded.
The contact stresses tend to reduce with depth since more grains support the
external load and the number of contact points has increased. The contact stresses
decrease with depth hence you may identify a depth beyond which contact stresses
CVX4343 – Soil Mechanics

In an isotropic granular material the average


stresses in the horizontal and vertical directions
are unequal. This is a characteristic of granular
materials comprising of discrete particles.
Materials such as steel or water are a
homogeneous continuous mass and hence their
properties are isotropic. You may recall that
water pressure at 10m depth is 98.1kPa in all
directions. A sandy soil for instance, when
subject to a vertical stresses of 100kPa gives a
horizontal stress of 50kPa.
Perhaps you have learnt how Mohr’s circle Figure 7.3: Stresses on a soil
approach is used to determine normal and shear element.
stresses acting on a plane inclined at an angle, 
relative to the horizontal plane.
This concept can be used to predict stresses in an element of a granular material.
This requires us to establish two stress points on the circle. Here, we make an
assumption that shear stresses acting along vertical and horizontal planes are
negligible, i.e. vh = hv 0. This makes stresses v and h principal stresses, and
can be denoted by 1 and 2. For non-homogeneous and anisotropic granular
materials, different loading situations may give rise to stress changes v andh.
This makes the stresses in the selected two orthogonal directions equal to v +v
and h+h. Suppose that we make the assumption that these changes do not cause
any change in shear stresses acting on the horizontal and vertical planes, the
relationships vh = hv 0 would still hold. This also means that the principle stress
directions remain unchanged during loading.

7.2 The Principle of Effective Stress


A fully saturated soil has its voids completely
filled with water. The free water surface, known
as the ground water table is at the atmospheric
pressure (100kPa). The hydrostatic pressure at
the free water surface is zero. This is measured
relative to the atmospheric pressure. The
hydrostatic pressure increases with depth; we
know that it is proportional to water depth, hw,
and is equal to whw. This is true for a continuous
water body or for a saturated granular material.
In the case of a granular soil, void spaces inter-
connected hence form a water body. The term
that we use for this hydrostatic stress is pore-
water pressure.
We have learnt that hydrostatic pressure causes Figure 7.4: Stresses with water
an up-thrust on submerged bodies, which is also pressures.
the case for a saturated granular medium. The
buoyancy tends to reduce the grain-to-grain contact stresses by an amount equal to
its pore-water pressure. Carl Terzaghi presented this concept as the Principle of
Effective Stress.

51
Session 7: How soil layers carry load

He defined Effective Stress  as the parameter


that represents the average inter-granular stress.
A zero effective stress situation implies that
particles making the soil element are barely in
contact, and hence does not have any strength to
carry a load. When =100kPa soil particles
exert contact pressures on each other and hence
the soil skeleton is intact and can carry some
load.
Figure 7.5: Two particles in
Terzaghi visualised that a soil element when contact with pore water
loaded would bear it stresses by the soil skeleton pressures.
and the pore-water. This is expressed as:
    u
where is the total stress;  is the effective stress and u is the pore water pressure.
When a total force P is applied over an average area A (refer Figure 7.5):
P  P   u (A  A c ) .

P is the effective normal force transmitted at a grain contact. Dividing this by the
total area, A gives:
P P (A  A c )
 u .
A A A

By substituting  and , the above equation can be expressed as     u (1  a )


A
where a  c . When A c  0, a  0 hence we could conclude that     u .
A
Figure 7.6 represents three scenarios. For the dry granular assembly, u = 0, hence
   .When pores are completely saturated,     u .

Figure 7.6: Granular assembly i) dry ii) voids partially filled iii) fully saturated.

When pores are partially saturated the equation can be represented as:
    u a  (u a  u) . The terms are rearranged as:     u a (1  )  u .
When  = 1 we have the fully saturated condition;  = 0 gives the dry condition.
For partially saturated soils, parameter  represents the partial load carried by pore
water pressure. The term (1-) represents the partial load carried by pore air
pressure.

52
CVX4343 – Soil Mechanics

Tropical countries such as Sri Lanka have zones of partial saturation, located above
ground water level. Lateritic soils facilitate evaporation and evapo-transpiration
processes. Understanding how partially saturated soils behave is an important
aspect in assessing the stability of potentially unstable slopes.
Concepts related to the behaviour of partially saturated soils are beyond the scope
of this course. Hence we would limit our discussion to understand the behaviour of
dry and fully saturated granular soils.
Let us apply the principle of effective stress to determine the vertical stresses
acting on soil element A (refer Figure 7.4). The total stress in element A, in
vertical direction vis due to weight of soil and water on top. Hence  v  v  u
gives us  v  [ bulk (h  h w )   sat h w ] and u  [ w h w ] . This gives us the vertical
effective stress v   v  u .
If we to determine the vertical stresses on soil element A given bulk = 16.5 kN/m3;
sat = 21 kN/m3, h = 10 m, and hw= 9 m, computes u=9x9.81 = 88.2 kPa; v =
16.5x1+21x9 = 205.4kPa; v = 205.5-88.29 = 117.2 kPa.
For granular soils, the lateral stress h required to prevent lateral movement.
h  K 0 v , where K0 is termed the coefficient of earth pressure at rest.is
expressed as h  0.5v . If this is the case h=58.6 kPa. This also gives h =
h+u = 58.6+88.2 = 146.8 kPa.
For soil element A, K0= 146.8/205.4 = 0.71.
Assuming that these are principal stresses, Mohr’s circles of stress for total and
effective stresses are plotted (refer Figure 7.7).

Figure 7.7: Mohr’s circles of stress representing total


and effective stresses at point A.

7.3 Distribution of vertical stress with depth


Figure 7.8 shows a sheet pile wall embedded in a sandy soil. Soil on the right side
retains 9m of soil; a surcharge load of 50kN/m3; and 3m of water.
If depth is measured from the top of the sheet pile wall, at z = 0 the vertical stress is
equal to the applied surcharge load of 50kPa.
v   v  50kPa .

We know that vertical stress varies uniformly with depth similar to hydrostatic
stress, hence the variation is triangular. This makes the task easy since we only
need to compute stresses at points where conditions change. When z = -6m, v =

53
Session 7: How soil layers carry load

50+617.5 = 155kPa. Hence v   v  155kPa . At z = -9m, v =


[50+617.5]+320 = 215kPa, u = 39.81 = 29.4kPa.Hence
 v  215kPa, u  29.4kPa, v  215  29.4  185.6kPa .
At z = -11m, v = [50+617.5+320]+2x20 = 255kPa, u = [39.81]+2x9.81 =
49.1kPa.Hence  v  225kPa u  49.1kPa v  225  49.1  17.6kPa .

Figure 7.8: Variation of v and u with depth.


Can you plot the variations of v with depth?

7.4 Pore pressures in confined aquifers


We have learnt how static pore water pressures are computed at a given point. The
water table is a phreatic surface, since it is exposed to atmosphere pressure. Such
water bodies are identified as unconfined aquifers. The open dug-well at your
home is used to access the underlying shallow unconfined aquifer.
Figure 7.9 shows a situation where soil element A is in a confined or a closed water
body. This is known as a confined aquifer. You may observe that there is no
connectivity between the local water table level and the piezometric level of the
confined aquifer.
The effective vertical stress v may be lowered by high pore pressures present in
unconfined aquifers. The pore pressures in element A is due to the piezometric
height of the confined aquifer. In confined aquifers artesian conditions prevail.
This means that if you place a stand pipe (refer Figure 7.8), the water level would
rise to a height hw, which is above ground level.

54
CVX4343 – Soil Mechanics

Figure 7.9: Pore pressures in a confined aquifer


When computing the total stress on soil element A, total weight above point A due
to soil overburden and water due to the local water table need to be considered.
The pore pressure at A would be u = whw. You could still compute the vertical
effective stress as v   v  u .

7.5 Stresses in partially saturated soils


Soils above water table are partially saturated. Fine grained soils form a capillary
fringe above water table. Within this zone water molecules are held to soil
particles through surface tension forces.
For partially saturated soils some interconnected pores are exposed to atmospheric
pressure while those forming capillary tubes form negative pressures (refer Figure
7.10). The height of water rise in a capillary tube is expressed as:
4T cos 
Zc  , where T is the coefficient of surface tension; d is the diameter of
d w
the tube; w the unit weight of water;  the angle at fluid-solid interface. For water
4T 1
 = 00, Z c   . T at 200C is 73 dynes/cm; w = 9.81 kN/m3. This gives:
d w d
 0.03m
Zc  .
d (in mm)
The depth of capillary fringe depends on pore size. Smaller the particle size
smaller would be the pore diameter. This gives a higher capillary rise. Table 7.1
lists heights of capillary rise in different soils.
When a soil is subject to negative pore pressures its effective stress would be
higher. This keeps the particles intact and hence would increase its shear strength.
Perhaps this explains why we could mould fine sand to make sand castles. When
pores become fully saturated or completely dried, the negative pressures seize to
exist thus reducing its effective stress.

55
Session 7: How soil layers carry load

Figure 7.10: The capillary tube.


Table 7.1: Heights of capillary rise in different soils (Hansbo, 1975)
Loose Dense
Coarse sand 0.03 – 0.12m 0.04 – 0.15m
Medium sand 0.12 – 0.50m 0.35 – 1.10m
Fine sand 0.30 – 2.0m 0.40 – 3.5m
Silt 1.5 – 10m 2.5 – 12m
Clay  10m

Activity
When you walk on moist beach sand, you may have noticed that `wetness’ of sand
surrounding your foot disappears due to foot pressure. Did sand beneath your foot
become softer or stiffer? Did the sand volume increase, decrease or did not
change?

To sum up
This session explained the Principle of Effective Stress and is use. We also learnt
that the effective stress represents the stresses between soil grains, and hence a high
effective stress represents a tighter packing, hence a stronger soil. This principle is
used to determine vertical stress distribution with depth, in a saturated soil mass, a
soil mass subjected to artesian conditions, and stresses prevailing in a partially
saturated soil.

56
Session 8
Stresses in soils due to imposed loads

What you will learn in this session


Learners are expected spend around six hours on this study session. Stresses due to
imposed loads are transmitted get dispersed in soil strata. Section 8.2 discusses a
simple model that represents dispersion of imposed vertical loads for granular soils.
Section 8.3 explains why granular soils behave as an elastic material at low strain
values. This allows us to use theory of elasticity to predict stresses and strains.
Both methods enable us to define the zone of influence due to imposed loads.
Section 8.4 presents several numerical and graphical techniques that are used to
estimate changes in stress in the soil mass.

8.1 Introduction
Soil strata are formed during in-situ weathering or when weathered products are
transported and deposited elsewhere. During deposition, strata get compressed
under its own weight; when depth overburden increases, soil beneath it is subjected
to higher stresses, and becomes stronger.
Natural soil formations are subjected to impose loads during land filling, reservoir
filling, and construction of embankment dams, levees and highway embankments.
When high loads are applied over a short time period, loss of bearing capacity or
excessive settlements may cause in weaker soil strata.
Therefore it is required to estimate changes in effective stresses with depth due to
imposed loads. This enables us to check whether soil strata have sufficient strength
to resist the imposed loads, and to estimate possible settlements that may occur.
This session discusses how one could estimate the vertical effective stress with
depth due to an imposed load.
In-situ penetration tests or probing tests help identify locations and depths of such
weak layers while samples are tested to identify their engineering properties.

8.2 How stresses disperse in a soil continuum


When a strip footing transmits the building load
to ground, its zone of influence is limited to a few
meters below its founding level. This also
disperses laterally beyond the width of the
footing. Figure 8.1 shows a strip footing that
exerts a force of 20kN per meter run. This results
in a uniform average contact stress of
20kN / m
 44kPa . Figure 8.1: A strip footing.
0.45m  1m
In the case of granular soils, we know why the contact stress disperses with depth.
If the material beneath is homogeneous rock or concrete, dispersion of stress would
not occur.
When a granular material is poured or shovelled, it tends to form a heap. If you
repeat this experiment with a few marbles you would not be able to form the heap.
Session 8: Stresses in soils due to imposed loads

In granular soils surface friction of grains and their ability to interlock with each
other makes it to heap. The maximum angle that the sloping surface makes with
the horizontal,  (refer Figure 8.2a) is equal to the angle of repose, . The angle of
repose depends on the frictional characteristics of the material. We will learn in a
later session that  is in fact the angle of internal friction of the granular material.
Therefore when  = , the heap could sustain the weight of its solid grains. For
sandy soils this angle is around 30-350.
Suppose that we intentionally main the side slope at an angle that is less than the
angle of repose, this would allow us to place a load at the top (refer Figure 8.2b).
Figure 8.2c shows the same condition as 2b; the shaded area identifies the zone of
stress dispersion. This also is the zone of influence.
Figure 8.2d shows when the footing is located at a depth, d. The surcharge, d
provides a lateral thrust which contracts the zone of stress further. A dispersion
angle of 2-vertical:1-horizontal is considered for most soils.

Figure 8.2: (a) Carrying self weight  =  (b) carrying 50kPa additional load <
(c) 50kPa load applied on a footing at ground level (d) carrying 50kPa footing load
at depth, d;>.

Activity 8.1
If a strip footing with B = 0.45m carries a load of 44kPa what would be the stress
at 1m below the founding level? The force at founding level = 44kPa  1m x
1
0.45m = 19.8kN. The area at 1m below has a width of 0.45  0.45x x 2  0.9m
2
19.8kN
.Therefore the stress 0.5m below the footing level is   22kPa .
1mx0.9m
Dispersion of vertical stress limits stresses to within the zone of influence. If this is
the case we would also like to know the depth at which the vertical stress becomes
insignificant. In engineering designs, this depth is taken to be the depth at which
the vertical stress becomes 10% of the footing pressure (i.e. 4.4kPa as computed in

58
CVX4343 – Soil Mechanics

this example).In soft compressible soils, this depth is taken to be at 5% of footing


pressure.
When a large extent of land is pre-loaded with a surcharge fill the zone of influence
is taken to be the total depth up to sound bed rock. For instance, the stress due to a
1m soil layer with bulk = 18kN/m3causes a uniform stress of 18kPa.
The above analysis is a simplified approach to determine a vertical stress increase
beneath a footing. It assumes that vertical stress increase at a particular depth is
uniform within the zone of influence.
This however is not true. If we compare two soils elements, one located along the
loading axis, and one further away from the loading axis, the latter is stressed to a
lesser extent.
You would also note that two soil elements located on either side of the zone
boundary, and located at the same depth, have incompatible stress values.
The above method yields stresses in acting in the vertical direction, where as a
precise analysis may require stresses in other directions as well.
Advances in continuum mechanics made applied mathematicians to analysis of
stresses and strains as an elastic continuum, assuming soil to be a homogeneous
and isotropic material with linear-elastic behaviour.

8.3 Stress-strain behaviour of soils


Figure 8.3 shows typical results obtained from standard tensile tests performed on
low-carbon steel and aluminium. During this test a standard metal rod is subjected
to a tensile force until rupture. The Engineering Stress  is plotted against
Engineering Strain,.
The axial extension test of a steel wire gives us the variation of engineering stress
with engineering strain. Engineering Stress is defined as the ratio of force to
original area. Engineering Strain is the ratio of extension to the original length.

Figure 8.3: Stress-strain behaviour of (a) Low-carbon steel (b) Aluminium alloy.
Both metal types show a linear elastic region. Steel has distinct yield points (i.e.
the upper and the lower) between which yielding takes place. It then undergoes
strain hardening, followed by necking and rupture. Aluminium alloy does not
show a distinct yield point. It undergoes work hardening and then fails. Both
metals display elastic (recoverable) and plastic (permanent) strains. When we test

59
Session 8: Stresses in soils due to imposed loads

soils to obtain its stress-strain behaviour a cylindrical sample is compressed until it


fails (refer Figure 8.4).
Failure of a soil specimen is defined when it reaches a maximum stress beyond
which excess deformations occur. This we identify as the `Strength’. When we
tested the steel specimen we determined the Yield Strength, Y and the Ultimate
Strength, U. For soils it is rather difficult to determine the Yield Strength, since
plastic deformations occur even at very low strain values.
Deviatoric stress, q is the net axial stress that
compresses the sample. This is equal to v  h .
Axial strain is ratio between change in height L
and initial height, L0. Volumetric strain is the
ratio between change in volume, V and V0. We
have considered compressive strains to be
positive.
Figure 8.4: Effective stresses
on a soil element.
Activity 8.2
Can you prove that volumetric strain V can be expressed as  V   v  2 h . v and
h are strains in the vertical and horizontal directions, respectively.
Figure 8.5(a) shows the stress-strain behaviour of a dense-sand obtained during a
Consolidated Drained triaxial loading test. You may also notice that for small
strains we could define the Young’s modulus, Es. Figure 8.5(b) shows that the
volumetric strain changes with a change in vertical strain. This means that when a
sample is compressed, axially, it strains in the lateral direction. Can you explain
the behaviour of soil due to a negative volumetric strain?

Figure 8.5: (a) Variation of deviatoric stress versus axial strain (b) volumetric
strain versus axial strain, for a dense sand

60
CVX4343 – Soil Mechanics

Soil is not an elastic material. It is neither homogeneous nor isotropic. We also


observe that stress-strain behaviour of a soil is non-linear. However, at very low
strains, its behaviour can be considered to be linear-elastic. This enables us to
determine its Young’s modulus, Es and Poisson’s ratio, . Table 8.1 shows
experimentally determined values for common soil types.
Researchers have confirmed that elastic strains computed based on experimentally
determined Es and  values agree well with corresponding in-situ measurements.
Present day designs use numerical techniques such as the finite element method to
analyse complex field situations.
Table 8.1: Young’s moduli and Poisson’s ratios

Young’s modulus, Es Poisson’s ratio, 


Type of soil
MN/m2
Loose sand 10.35 – 24.15 0.20 – 0.40
Medium dense sand 17.25 – 27.60 0.25 – 0.40
Dense sand 34.50 – 55.20 0.30 – 0.45
Silty sand 10.35 – 17.25 0.20 – 0.40
Sand and gravel 69.00 – 172.50
0.15 – 0.35
Soft clay 2.07 – 5.18
Medium clay 5.18 – 10.35
0.20 – 0.50
Stiff clay 10.35 – 24.15

8.4 Computing `elastic’ stresses

Stresses due to a point load


The theory of elasticity is used to represent the
stress at a point due to a point load Q (refer
Figure 8.6). Here we assume an elastic medium.
The equilibrium and compatibility conditions of a
soil element located at coordinates (r, , z)is
subjected to an axi-symmetric point load, Q. The
respective stress increments of this element in the
three orthogonal directions are expressed as:
5/ 2
3Q  1 
 z   
2z 2  1  ( r / z) 2 
  Figure 8.6: Stress increase due
to a point load Q
Q 3r 2 z 1  2 
 r   2  2 

2  ( r  z )
2 5/ 2
r  z  z ( r 2  z 2 )1 / 2
2 

Q  z 1 
    (1  2) 2  2 

2  (r  z )
2 3/ 2
r  z  z ( r 2  z 2 )1 / 2
2

3Q  rz 2 
 rz   2 2 5/ 2 
2  ( r  z ) 

61
Session 8: Stresses in soils due to imposed loads

Activity 8.3
Compute the stresses acting on an element with z = 2.5m; r = 1m for a point load of
40kN. Compare the magnitude of normal stresses with shear stress rz. Do you
find  rz negligible compared to  z and   ?

Activity 8.4
Make a spread sheet on MS-EXCEL to compute the above stresses, when r and z
are varied. Plot the variation of  z with z along the centre of loading. Plot the
variation of  z with r for a certain depth. Comment on your observations.

Stress due to a line load


The stresses due to a line load (refer Figure 8.7)
are expressed in terms of the rectangular
coordinate system. These are expressed as:
2Qz 3 2Qx 2 z
 z  ;  x  ;

 x2  z2 
2

 x2  z2 
2

2Qxz 2
 zx  .

 x2  z2  2

Figure 8.7: Stress due to a line


load.
Activity 8.5
Make a spreadsheet on MS-EXCEL to compute the above stresses, when x and z
are varied. Plot the variation of  z with z along the line of loading. Plot the
variation of  z with x for a certain depth. Comment on your observations.

Stress due to a strip load


We come across situations where stresses at
points due to a strip load need to be determined
(refer Figure 8.8). This can be the foundation
pressure exerted by a strip footing. The
respective stress increments are expressed as:

 z 
q
  sin  cos(  2)

 x    sin  cos(  2)
q

q
 zx  [sin  sin(  2)]

Figure 8.8: Stress due to strip
Activity 8.6 load q.

Make a spreadsheet on MS-EXCEL to compute  z below the centreline of a strip


footing of 450mm width, carrying a load of 60kPa. Plot the variation of  z with
z. Determine the depth at which  z becomes 10% of q.

62
CVX4343 – Soil Mechanics

Area transmitting a triangular strip load


When determining stresses due to an earth embankment, we come across triangular
loads (refer Figure 8.9). The stresses due to a triangular load with a maximum
intensity q are expressed as:
qx 1 
 z     sin 2 
B 2 

q  x z R 12 1 
 x     ln 2  sin 2 
B B R2 2 
q z 
 xz  1  cos 2  2  
 B 

During stress analysis, the simplified assumptions


permit us to use the principal of superposition.
This allows us to add the effect due to individual
loads to give the total stress at a given point. Let
us see how this is done.
Figure 8.9: Stress due to a
triangular strip load
Activity 8.7
Principle of superposition allows you to add the effect due to several loading
conditions, to give the net effect.
Figure 8.10 shows a road embankment fill. Determine the vertical stress at point A
of a soil element along the centre line.

Figure 8.10: A road embankment fill.


Hint: First determine the stress at A due to the rectangular load; then consider the
stress at A due to the two triangular loads.

Stress due to a uniformly loaded rectangular area


Many foundations are rectangular in shape. The stress increases below a corner of
a rectangular loaded area of width B and length L are expressed as:

q  1 LB LBz  1 1 
 z   tan    
2  zR 3 R 3  R 12 R 22 

63
Session 8: Stresses in soils due to imposed loads

q  1 LB LBz 
 x   tan  2 ;
2  zR 3 R 1 R 3 
q  1 LB LBz 
 y   tan  2 ;
2  zR 3 R 2 R 3 
q  B z2B 
 xz    2  , where R 1  L2  z 2 ;
2  R 2 R 1 R 3 

R 2  B 2  z 2 ; R 3  L2  B 2  z 2 . The stress
increment in the vertical direction,  z , acting below
a corner of a rectangular loaded area is expressed as
 z  qI z , where

1  2mn m 2  n 2  1  m 2  n 2  2  
1  2mn m  n  1 
2 2 
Iz   2    tan 
2  m  n 2  m 2 n 2  1  m 2  n 2  1   m 2  n 2  m 2 n 2  1 
  
m = B/z and n = L/z. Figure 8.11 shows the chart that
represents the influence factor, Iz used to compute the
vertical stress increase below a corner of a rectangular
area LxB.

Figure 8.11: Influence factor Iz below a corner of a rectangular area LxB

Activity 8.8
Figure 8.12 shows a rigid loaded area with a footing pressure of 100kPa.
Determine the stress at point A, located at a 2m depth.
Determine Izfor : m=1.5m/2m = 0.75; n=0.5m/2m = 0.25; Iz = 0.065. Izfor : m
= n = 0.5m/2m = 0.25; Iz = 0.03.
Iz = 0.065+0.065+0.03+0.03 = 0.19;  z  qI z  100x0.19 = 19kPa.

64
CVX4343 – Soil Mechanics

Figure 8.12: A rigid loaded area.

The pressure bulb


Figure 8.13 shows stress contours that determine soil stresses with depth. The left
hand side shows the stress distribution for a square footing. The right hand side
depicts stress distribution for a continuous footing. The depth is normalised with
respect to B while the vertical stress increment is expressed is normalised with
respect to the contact stress, q0.
For a continuous footing, 10% of contact stress occurs at a depth of 4B while for a
square footing this occurs at 2.2B.

Figure 8.13: The pressure bulb, the variation of  z with depth z, along the centre
of footing.

Vertical stress increase below an arbitrarily shaped area


The influence of vertical stress at a certain depth due to a flexible loaded area is
found using Newmark’s chart (refer Figure 8.14).
This is a graphical approach where the footing length and width is scaled down so
that length AB (refer Figure 8.14) represents z, the depth from footing to the point
where the stress is evaluated.The figure is divided in to 200 units. Draw the scaled
down footing on a transparency sheet; place it on the chart so that the projection of

65
Session 8: Stresses in soils due to imposed loads

the point where the vertical stress,  z is to be determined at the centre point.
Then count the number of squares, N enclosed within the footing.  z is
N
expressed as  z  q0
200

Activity 8.9: In text question


Determine the vertical stress at point A for the configuration shown in Figure 8.14.

Figure 8.14: Newmark’s chart

To sum up
This session discussed how the zone of influence is identified due to an increase in
vertical stress. We have seen how elasticity theory is used to express stresses at
various points. The variations of deviatoric stress versus axial strain and
volumetric strain versus axial strain, observed for a dense-sand were presented.
The stress-strain behaviour observed for steel and aluminium was presented for
comparison.
We have discussed several numerical and graphical techniques that are used to
compute horizontal, vertical and shear stresses, due to an imposed stress increment.
The use of the Principle of Super-position when assessing the cumulative effect
due to several loads is demonstrated.

66
Session 9
Computing stresses acting on a soil element

What you will learn in this session


This is a 2-hour study session. Section 9.1 explains how you could compute
normal and shear stresses acting on any predetermined plane. This is analysed for
a two dimensional stress system. This section may help you recall how you could
do the same using Mohr’s circle of stress.
Section 9.2 discusses stress changes a soil element may undergo when subject to
loading and unloading. It also draws your attention to assumptions made in order
to simplify the analysis.
We have seen in Session 13 how soils behaves elastically, undergoing both linear
and non-linear elastic strains. Section 9.3 discusses how you could determine
strains based on `elastic’ soil parameters.

9.1 Introduction
We have learnt in the Strength of Materials
course that the normal stress, n and shear stress,
 can be represented on    n space. The two
stresses are plotted as a point on the Mohr’s circle
of stress. Figure 9.1 shows the standard
convention used in the two dimensional Cartesian
space. The sign convention considers
compressive stresses and clockwise shear stresses
as positive.
Figure 9.2 shows how stresses (,) are expressed
in terms of stresses  xx ,  yy ,  xy and . The
Mohr’s circle has its centre located at
  xx   yy  Figure 9.1: Stress at a point
 ,0  . Its radius

 2 
2 xy
2
  xx   yy 
R      2xy . The angle  is expressed as T an 
 .
 2  ( xx   yy )

 xx   y y
n   R cos(2  ) and   R sin(2  ) .
2
 xx  yy
n   cos 2 [R cos ]  sin 2 [R sin ] is simplified as:
2
 xx  yy  xx   y y
n   cos 2   xy sin 2 .
2 2
  sin 2[R cos ]  cos 2[R sin ] is simplified as:
 xx   y y
 sin 2   xy cos 2 .
2
Session 9: Computing stresses acting on a soil element

Figure 9.2: Expressing stresses (,) in terms of  xx ,  yy ,  xy and .

Activity 9.1
Figure 9.3 shows an element with the given
stresses. We need to determine a) the major and
minor principle stresses and their respective
directions; b) the stresses on a plane inclined at
350 from the horizontal; c) The maximum shear
stress and the inclination of the plane on which it
acts.
The coordinates at the centre of sphere is (17.5,0)
and R=60.5.
Step 1: 1 = 17.5+60.5 = 78 kPa; 2 = 17.5-60.5
= -43 kPa; the angle  = 29.740. Figure below
shows the angles that the principal stresses make
with  xx . The angle 1 makes with  xx is
sin(180  2)  30 / 60.5  2  119.7 0  59.9 0. Figure 9.3: Representation of
This is measured clockwise. The angle 2 makes stresses on a soil element.
with  xx is 59.9 + 90 = 149.9 . The figure below
0

shows how the principal stresses are oriented with respect to  xx .

Step 2: Refer figure to your right. Here we


would locate the point (n,) on the Mohr’s circle,
which makes 600 from  xx . Sin = 30/60.5,
hence  = 29.70. The angle 2 + = 70+29.7 =
99.70. Since this is greater than 900, point A lies
to the right of the centre of circle. The angle
point A makes with the positive  n direction is
180-99.7 = 80.30. Hence at A,
 n  17.5  60.5 cos80.3  27.7 kPa.   60.5 sin 80.3  59.6 kPa.

68
CVX4343 – Soil Mechanics

Step 3: The maximum shear stress 60.5 kPa, and is equal to the radius of the circle.
The angle is (90-29.7)/2 = 30.10 clockwise from  xx .

9.2 Examples of loading and unloading situations


Let us examine some common loading situations.
Figure 9.4 shows a loading where  1 is increased
while keeping  2   c a constant. 1 is
increased by 1 while maintaining 2 = 0. The
x-coordinate of the centre of Mohr’s Circle, C
and its radius, R can be expressed as
  2   2
C 1 and R  1 . When expressed
2 2
in the increment form:
   2 1    2 1
C  1  and R  1 
2 2 2 2

This tells us that the centre and the radius Figure 9.4: A conventional
1 triaxial loading test.
increases by the same amount .
2
A conventional triaxial loading test simulates
such a condition. Here, we assume that principal
stresses remain vertical and horizontal through
out the test.
K0 loading condition considers that the horizontal
stress is always proportional to the vertical stress.
This condition prevails during one-dimensional Figure 9.5: K0 condition.
consolidation, which is when a large extent of
land is being compressed by placing a uniform Lateral strain 2 can be made
zero by having a fixed vertical
surcharge load. K0 condition yields 1  K 0  2
wall.
and  2  0 .
Figure 9.5 shows how Mohr’s circle of stress changes with loading. Here, the
centre and radius of the Mohr’s circle can be expressed as:
  K 0 1 1   K 0 1 1
C 1  1  K 0  and R  1  1  K 0  . This gives us
2 2 2 2
1
C  1  K 0  and  R  1 1  K 0  .
2 2
Figure 9.6 shows how Mohr’s circles of stress
change during a constant average stress test, i.e.
1   2
  c .We also call this a constant-p test.
2
To satisfy this condition, an increase in vertical
stress, 1 should ensure a decrease in horizontal
stress, 2 by the same amount, i.e. 1 = -2. Figure 9.6: Constant average
This gives us C=0 and R=1. stress condition.
When a retaining wall is pushed out by a
deforming soil mass, its lateral stress, 2 is reduced while the vertical stress, 1

69
Session 9: Computing stresses acting on a soil element

remains unchanged. When a retaining wall pushes the soil in, the lateral stress is
increased beyond its constant vertical stress.

Activity 9.2
Can you sketch the Mohr’s Circles of stress for the situation described above?
Can you also sketch the Mohr’s Circles of stress when the lateral stress is increased
while keeping the vertical stress a constant? Note that all stresses are considered to
be principal stresses.

9.3 Plane stress and plane strain conditions


Three dimensional strains can be expressed in terms of stresses  xx ,  yy and  zz as
shown below:

 xx 
1
E
  1
 
 xx  ( yy   zz ) ;  yy   yy  ( xx   zz ) ;
E
 zz
1

  zz  ( xx   yy )
E

Where E and  are Young’s modulus and Poisson’s ratio, respectively.
Plane stress situation satisfies  zz  0 , i.e. normal
stresses act only along x–y plane. This gives us
1

strains  xx   xx   yy ,
E

1
  

 yy   yy   xx and  zz    xx   yy .
E E

Plane strain situation (refer Figure 9.7) satisfies
 zz  0 . This gives  zz  ( xx   yy ) This
gives the expression for  xx and  yy as:
1 Figure 9.7: Plane strain
 xx  (1  )( xx   yy ) and
E situation.
1
 yy  (1  )( yy   xx ) .
E

To sum up
This session showed how learners could represent a 2-dimensional stress situation
on a Mohr’s Circle plot. We saw how Mohr’s circle concept is used to determine
principal stresses and stresses acting on a plane inclined at an angle .
Geotechnical applications result in changes in in-situ stresses. These changes are
represented by Mohr’s Circles of stress. We have discussed a few such examples
and learnt how these changes are represented using Mohr’s Circles of stress.
Learners may have noted that the cases that we’ve considered assume that principal
stress directions are the vertical and horizontal directions. This assumption
simplifies computations since we do not have to deal with shear stresses.
We have expressed elastic strains in x, y, and z directions and saw how stresses for
plane-strain and plane-stress conditions are derived.

70
Session 10
Compacting soils

What you will learn in this session


This 3-hour session explains compaction process by which soil grains are brought
closer to attain a dense packing. Section 10.2 explains the Standard Proctor
Compaction Test that is used to obtain the compaction curve for a particular soil
type. This test is used to assess suitability of natural or artificially mixed soils for
engineering applications. Section 10.3 explains how densities are measured during
field compaction; compared with the maximum dry density obtained during the
Proctor Compaction Test.

10.1 Introduction
Soil compaction is the process by which we bring soils to a closer packing.
Methods used for this vary from using a wooden tamper to compact a gravely soil
prior to rendering for your floor or using a steel roller to compact a sub grade
material. We compact soils to enhance its engineering properties. Soil beneath
floors is compacted to reduce cracking. Embankment dams are compacted to make
them water tight and to give strength to retain the water body. A sub grade is
compacted to be able to carry repetitive wheel loads. A well compacted fill
reduces building settlement.
Compaction reduces the amount of void spaces and is quantified in terms of dry
density d. We know that it is easy to compact a moist soil than a dry soil. Higher
densities can be achieved through a high energy input; i.e. heavier hammer, higher
drop, more number of blows and number of layers.
Dry density d = Ms/V. This does not mean that the soil is always `dry’.
We also know that various soils have their characteristic maximum dry densities
dma. A well-graded sand can be compacted to a higher dry density than a
uniformly or poorly graded sand. A fat-clay with high plasticity is difficult to
compact.

10.2 The Standard Proctor Compaction Test

Figure 10.1: Standard Proctor Compaction Apparatus.


Session 10: Compacting soils

Standard Proctor Compaction test uses a 2.5kg hammer dropped by 300mm. This
imparts an energy of 2.5x9.81x0.3 Nm =7.36 Nm. The total energy imparted to the
sample during test is equal to 7.36/blow x 25 blows/layer x 3 layers = 552Nm.
When a 1000cc mould is used, the compacting effort works out to be 552kJ/m3.
The test procedure establishes the changes in dry density d with the amount of
moisture in soil. The variation d-w when plotted gives us the characteristic
compaction curve for a soil. Figure 10.2 shows the
Figure 10.2 shows the compaction curve obtained for a sandy-clay. You may
observe that d increases with w till it attains a maximum dry density. The
corresponding moisture content is termed the optimum moisture content, wopt.
Further increase in moisture makes the dry density to drop.

Figure 10.2: Compaction curve and the zero air-voids line for a sandy-clay.
Figure 10.3 shows the changes in the three phases. For a given mould volume, V it
is observed that dry density increases with an increase in moisture content; when
moisture content increases further, the dry density drops.
An increase in dry density shows that particles achieve a closer packing. Therefore
`dry’ of optimum condition signifies both an increase in Ms and Mw. An increase
in water in pore spaces facilitates particles to come closer, acting as a lubricant to
reduce contact friction.

Figure 10.3: Conditions `dry’ of optimum, at optimum and `wet’ of optimum.

72
CVX4343 – Soil Mechanics

Compaction beyond optimum moisture content results in a decrease in dry density.


This shows a reduction in Ms and an increase in Mw. The degree of saturation,
Vw 1
S  . As we know that saturation increases with increasing
Vw  Va 1  Va / Vw
water content; the air volume Va should hence decrease. When water is further
added to soil, its saturation nears 100%. It is also a fact that one cannot `compact’
a soil when all its voids are completely filled with water. This tells us that the
segment of compaction curve `wet’ of optimum cannot `intersect’ the S=100% or
zero air void region. This is in fact a curve closer to a line (refer Figure 10.2). You
would also observe that all compaction curves lie to the left of the said line, some
being nearly asymptotic.
Figure 10.3 shows compaction curves obtained for different soils types. First of all
we should recall that all observations are comparable since a standardised test
procedure was adopted.
A well-graded sand gives the highest compaction while a uniformly graded sand
does not compact even as much as a high plasticity clay. This is due to achieving a
denser packing in the case of the former soil type.
A uniform-sand has the loosest packing similar to a box full of marbles.

Figure 10.3: Standard Proctor Compaction Test curves obtained for different soil
types.
The zero air void curve relates d to w, s, S=1, w = 1g/cm3. For soils variation
in s, i.e. 2.65-2.72 gives approximately the same curve.
Activity 10.1: In-text question:
The curve that represents the relationship between ρd and w for a given degree of
wS
saturation S is expressed as  d  . For a given S you may find that

w w S
s
ρd1/w. Using first principles and the three phase diagram can you derive this
expression?
Clay soils can hold water within their pores and hence results in a reduction in its
dry density. Perhaps this explains why the particular sand clay had a lower
compaction compared with the well-graded sand.

73
Session 10: Compacting soils

A 1ml (1cm3) space of clay soil contains 2.72g while the same volume of water
contains just 1g.
Plasticity enhances the amount of water held in pores. The observed drop in dmax
is also associated with an increase in wopt. When high plasticity soils are
compacted, the compacted volume contains more water thus replacing solids with
water.
Activity 10.2
The results of a Standard Proctor Compaction Test performed on a Clayey Sand are
as follows:
Sample No. 1 2 3 4 5
Moisture can no. 55 56
Wt. of can + wet soil 97.65 89.43
Wt. of can + dry soil 88.98 80.85
wt. of water
wt. of can 17.83 17.89
wt. of dry soil
Water Content, w

Assumed w 12%
Measured w
Wt. of soil + mould 4108.1
Wt. of mould 1932.2
Wt. of soil in mould 2175.9
Wet density, kN/m3
Dry density kN/m3

The Specific Gravity of the soil is found to be 2.68. The measured average
diameter of the mould is 10.32cm and the average height is 12.04 cm.
a) Determine the dry unit weight (in kN/m3) for Test No. 3.
b) Determine the respective masses of the three soil phases.
c) Determine the respective volumes of the three soil phases.
d) Determine the degree of saturation of the soil.
e) During Test No. 4 it was found that the bulk unit weight decreased with a 2%
increase in water content. Indicate the changes of the three phases with respect
to Test No. 4.

74
CVX4343 – Soil Mechanics

10.3 Measuring field densities


Field compaction is carried out at moisture contents closer to its optimum. Even
though we add the correct amount of moisture, achieving a density closer to d max
depends on how well we do the compaction. We now know that this requires the
correct layer thickness and tamping. The term Relative Compaction compares field
dry density achieved with d max obtained through a standardised laboratory test.
 d field
The Relative Compaction is expressed as .
 d max

Sand Cone Test is the traditional method that we


use to measure field densities. The sand cone
(refer Figure 10.4) is used to determine the
volume of the dug hole by weighing the volume
occupied by a uniformly graded sand. A uniform
sand gives you the same particle packing, hence a
constant density. The Standard Ottawa Sand is
used in the standardised procedure.
During this test we also weigh the moist weight
of soil removed and also take samples to obtain
the moisture content.
In-text question:
Explain how d is computed knowing bulk and w?
Computing Relative Compaction is important to
compare against laboratory compacting effort. Figure 10.4: Sand Cone Test
Earlier we saw that this is around 552kJ/m3. Apparatus.

We should also note that certain engineering works such as compaction beneath
airfield runways may require establishing Relative Compaction with respect to a
higher laboratory maximum dry density value. In such instances we increase the
compactive effort by increasing hammer weight and number of blows per layer.
This test is termed the `Modified Proctor Compaction Test’.

To sum up
The indices Dry Density and Moisture Content are used to quantify compactness of
a soil. For a given compaction effort a particular soil type gives a characteristic
compaction curve. This session has explained to you why the dry density increases
when samples are compacted dry of optimum and why it decreases wet of
optimum. We have seen how Maximum Dry Density and the Optimum Moisture
Content is used to define the state of compactness.
The Sand Cone test is used to determine in-situ dry densities and their respective
average moisture contents. This information is used to determine Relative
Compaction, a measure that indicates how well we’ve achieved good compaction

75
Session 11
Water movement through soil and rock

What you will learn in this session


This is a 3-hour session. Understanding fluid flow through porous media has many
applications in Civil Technology, such as in groundwater extractions and
controlling seepage under an embankment dam. Section 01 explains how loss in
hydraulic energy (or hydraulic head) over a length determines the average velocity
of flow. This gives us Hydraulic Conductivity, k, a parameter that represents fluid
and material properties that makes fluid flow.
Section 02 discusses two simple laboratory methods that determine Hydraulic
Conductivities of soils ranging from clays to fine sands. It further demonstrates
how the concept of hydraulic head is used to establish the hydraulic gradient across
the sample.
Section 03 explains some common empirical correlations that estimates hydraulic
conductivities of engineering soils.

11.1 Fluid flow through porous media


Water flows through soil and rock. It finds its way through interconnected soil
pores or rock fractures taking a tortuous path. Surface water flows under
atmospheric conditions while sub-surface water sustains pressures greater than
atmospheric.
Water flow in a conduit is similar to water flow in soil or rock. We know that
water is incompressible and satisfies the equation of continuity. The energy
contained in a water body is transferable.
If there is no accumulation, the equation of continuity states that what goes in
should go out.
Bernoulli’s Theorem is about conservation of energy. Bernoulli’s Theorem
p v2
expresses the total energy per unit weight   z to be a constant if there is
g 2g
no loss of head. In soils we consider that laminar flow conditions exist. The
velocities are so low that we could ignore the velocity head. This makes the total
p
head H equal to the Hydraulic Head z.
g
Unlike water flow in a smooth pipe, we cannot
ignore the loss of energy in friction. This results
the head to drop over a certain travel length and is
defined as Hydraulic Gradient, i (refer Figure
11.1). Darcy’s experimental findings on flow
through porous media shows that under laminar
flow conditions, the average velocity v  i , hence
v  ki . The term k is defined as the hydraulic
conductivity. It is also termed coefficient of Figure 11.1: A stream line.
permeability.
CVX4343 – Soil Mechanics

Hydraulic conductivity quantifies the ease of fluid flow through soil or rock. This
is expressed in cm/s. The hydraulic conductivities of fine sand is in the order of
10-2cm/s while for clay soils it is 10-6.

11.2 Laboratory measurement of k

Constant Head Permeameter:


Laboratory measurement of k involves measuring the total head drop between two
points of a soil column with a uniform cross sectional area. We measure the
hydraulic gradient (i.e. H / L ) and the average discharge (in litres per second)
First we will discuss the constant head permeability test that measures k of high
permeable granular soils. Constant head test applies a constant head difference,
H to make water flow through soils (refer Figure 11.2). This is done by locating
the inlet and outlet levels at H1 and H2, respectively. I have also plotted the
variation of total head, pressure head and elevation head, with sample length L.

Figure 11.2: Constant head permeameter and variation of head with sample length.

We assume that the total head drops linearly between inlet and outlet points. This
is a reasonable assumption since soil friction influences the head in a uniform
manner.
The elevation head changes linearly with sample length, hence this makes the
pressure head to vary linearly.
If we take the datum at the bottom of sample, the elevation head, z = 0; the height
of the water column, H1 is equal to the pressure head, thus the total head is equal to
the pressure head.
At the top of the sample, the elevation head = L; the pressure head is the height of
water column above the outlet level. The total head is the summation these and is
equal to H2.
When water flows through soil the head difference H, over a length L gives a
steady flow. Knowing H and L enable us to compute the hydraulic gradient,
i  H L .

77
Session 11: Water movement through soil and rock

Alternatively, we could make use of the two stand pipes to establish the drop in
hydraulic head, and hence determine the hydraulic gradient, i  H L .

Q
The average velocity, v is determined by  Av , where we collect a certain
t
volume of water, Q over a time period, t. A is the average area of cross-section of
the permeameter. This computes the average velocity determined over a few trial
measurements. Then we make use of Darcy’s equation v = ki to determine k.
Falling head permeameter:
Falling head permeameter (refer Figure 11.3) is more suited for fine-grained soils
with low permeability. Its stem with a small cross-section allows us to measure the
volume change by measuring the amount of fluid displaced during a time period t.
The average velocity through the sample during a
change in height of water column, h is expressed
h
as v  k . The amount of water displaced
L
through the annular tube over time dt is a.dh
hence, the incremental discharge can be
h h
expressed as a  Ak By rearranging terms
t L
h2
dh Ak
this is expressed as: 
h1
h

a
t . This gives

a h 
the solution as k  ln 1  .
At  h 2 
Activity 11.1: In-text question
This requires us to measure areas a and A
accurately. Explain how you could do this.
The hydraulic conductivity is found by measuring
the average time period t, for a water level drop Figure 11.3: Falling head
h1 to h2. permeameter.

11.3 Hydraulic Conductivity, k


Hydraulic conductivity, k depends on both material and fluid properties. When the
porous media has more voids it makes the fluid to flow faster. When its path is
tortuous flow, occurs with a high energy (or head) loss.
Hydraulic Conductivity is proportional to a higher power of in-situ void ratio, e.
For e.g. fine sands k  e3. Hazen’s correlation empirically relates k to effective
size D10 and is given as k  CD 10
2
. This gives good agreement for soils with D10
between 0.1 – 3mm and with a coefficient of uniformity CU < 5. Hazen’s
coefficient C is between 0.8 – 1.2, hence a value of 1 is generally used.
Hydraulic Conductivity depends on the density,  and dynamic viscosity,  of the
fluid. Both these parameters change with temperature, T. This gives us
 
k T T  k 20 20 .
T  20

78
CVX4343 – Soil Mechanics

The kinematic viscosity  is defined as =/


Therefore, it is important for us to express k at a standard temperature, e.g. at 200c.
Table 11.1 lists typical range of k for different soil types.
Table 11.1: Range of k values for saturated
unconsolidated soils.

Material k (cm/s)
Clay 10-10 – 10-7
Silt, sandy silts, clayey sands 10-6 – 10-4
Silty sands, fine sands 10-3– 10-1
Coarse sand 10-2 – 1
Gravel 1 – 100

11.4 Methods to measure in-situ Hydraulic Conductivities


Hydraulic conductivities of insitu soils may defer significantly from laboratory
determined values. We know that in-situ test represents `average’ conditions than
a few test samples. It is also difficult to obtained relatively undisturbed samples.
Well pumping Tests
Well Pumping Tests are used to determine hydraulic conductivities of
homogeneous aquifers with high permeability, i.e. sands and gravels. These can
be confined or unconfined. A confined aquifer is sandwiched between two
pervious layers, while an unconfined aquifer has a free water surface. Figures 11.4
and 11.5 show the two aquifer types, which is a simplified scenario.

Figure 11.4: Pumping test performed on a confined aquifer.

79
Session 11: Water movement through soil and rock

The extraction well penetrates the full depth of


acquifer and is pumped at a constant rate, Q0.
The two monitoring wells measure the depth to
piezometric surface. It is important that the
learner is able to differentiate between the
piezometric surface and the water level. During
pumping draw down causes a cone of influence
(refer Figure 11.5). When drawdown does not
change with time, pumping has reached a steady
state.
During pumping water seeps to the well steady
and its flow is considered laminar. This enables
us to use Darcy’s Law to determine its radial
velocity. Let us quantify the flow that occurs
between two cylindrical sections A and B spaced
h
at r. It’s velocity is given as: v  k .
r
Considering the fact that water is incompressible,
Figure 11.5: Piezometric
equation of continuity yields that surface.
 h 
Q 0  (2rL)  k  0.
 r 
Q0 dh
This gives us r in the differential form. The integral form of the
2kL dr
r h
Q0 2 1 2

above differential equation is


2kL r r  h

dr  dh , when solved
1 1

Q0 r Q0 r
ln 2  h 2  h 1 . K is expressed as: k  ln 2 .
2kL r1 2L(h 2  h 1 ) r1
Figure 11.6 shows well pumping in an
unconfined aquifer. If two cylindrical elements A
and B are compared, height of flow, L becomes
the variable height h. This alters the differential
Q0 dh
equation to read as r ; the integral
2kh dr
r h
Q 21 2

form being 0
2k r r  h

dr  h dh (refer Figure
1 1

11.7). The solution takes the form


Q0 r h2 h2
ln 2  2  1 . The expression for k is
2k r1 2 2
Q0 r
k ln 2 .
 [h 2  h 1 ] r1
2 2

Figure 11.7: Water table.

80
CVX4343 – Soil Mechanics

Figure 11.6: Pumping test performed on an unconfined aquifer.


Borehole Tests
In-situ permeability of a particular stratum can be determined using a borehole test.
The two types of tests are a) constant head test, and b) variable head test.
The constant head test (refer Figure 11.8) uses a
pipe which is cased on the sides. Soil and debris
in the pipe are removed. Water is discharged to
the stratum at the bottom of pipe, while a constant
head h is maintained throughout the test. The
water flow rate q, is measured under steady
conditions.
Hydraulic conductivity, k is expressed as
q
k ; d is the diameter of the pipe. In a
2.75dh
variable head tests water is allowed to infiltrate
and the time, t taken for the head to drop from
height h1 toh 2 is measured. Two methods of
variable head tests are discussed below (refer
Figure 11.9). The first has a cased pipe
penetrating a shallow depth in a stratum of
infinite depth, length D being between 0.15m to Figure 11.8: Constant head
test.
1.5m. Water is discharged at the bottom of the
hole (refer Figure 11.9 Case I).
d  h 1 
Hydraulic conductivity, k is expressed as k  ln  
11t  h 2 

81
Session 11: Water movement through soil and rock

Case II (refer Figure 11.9) uses a perforated pipe, with perforated length, L in a
stratum of infinite depth. This method is suited for fine grained formations located
d2  2L   h  L
at greater depths. K is expressed as k  ln   ln  1 ;  4.
8Lt  d   h 2  d

Figure 11.9 Case I: shallow cased pipe, Case II: perforated pipe

11.5 Water movement through rock


Measuring rock permeability is important when water is pumped out from a
fractured rock formation. When making a reservoir water-tight permeability
measurements decide whether we need to grout in rock to form an impervious
curtain. When limestone is encountered, seepage can cause weathering through
solution.
Permeability of rock mass depends largely on rock type and presence of
discontinuities. In most formations water flow occurs through fault planes and
fractures, and fracture aperture controls water movement. When rock is subjected
to a heavy overburden stress, fractures tend to close. They can be clogged with
fine-grained soils and chemical compounds can cement these together. Presence of
air in fractures hinders flow.
Hydro-geologists use permeability, K to identify a water bearing formation. This is
measured in cm2/s.
Can we estimate in-situ permeability in a laboratory environment? A laboratory
model should simulate inward radial flow towards the extraction well. Inward
flows under high water pressures, cause the stratum to compress, thus increasing
resistance to flow. Outward flows during injection can cause to open-up the
formation, reducing resistance to flow. Do we know how the rock formation looks
like?

82
CVX4343 – Soil Mechanics

When an ill-defined problem is encountered,


engineering scientists look for ideal
configurations that can be represented
mathematically. Figure 11.10 shows a formation
with mutually orthogonal fractures with spacing,
s and aperture d. Assuming that laminar flow
conditions exist, k can be expressed as
 d3 
k  w  .
6  s 
Figure 11.10: Visualising how
rock fractures look like.

Activity 11.2: In-text question


Can you identify the terms that represent a) the fluid and b) the rock fabric?
The hydraulic conductivity is computed based on a pumping test conducted in an
Q0 r
unconfined aquifer, k  ln 2 . Then aperture thickness, d is
2L(h 2  h 1 ) r1
established for a known fracture distance, s. These hypothetical parameters serve
as quantitative indices that describe rock mass quality.

To sum up
This session presented to you various models that are used to estimate hydraulic
conductivities of soil and rock. These are determined through laboratory or field
measurements or by using empirical methods.
We have learnt to apply the concepts of hydraulic head and Darcy’s Law, to
compute hydraulic conductivity of soil and rock.
Hydraulic conductivity, k is a property of the soil fabric as well has the permeating
fluid. Permeability, K is defined as an intrinsic property of the soil fabric.

83
Session 12
Computing settlements in soils

What you will learn in this session


This is a 3-hour learning session. Section 12.1 explains how changes in effective
stress with time give rise to consolidation settlement. Changes in effective stress
occur with the dissipation of pore water pressure. Section 12.2 explains how
consolidation takes place in-situ. This behaviour is seen as a One-Dimensional
Consolidation problem (refer Section 12.3) and hence is simulated in a laboratory
test. The test results predict total settlements due to imposed loads.

12.1 Introduction
We have seen in Session 10 how moist soils are compacted. In the case of high
plasticity clays when compacted `wet of optimum’ may bring the soil to near
saturation. Clay soils deposited in marine or lake environments form layers of
saturated deposits. Such deposits also result from in-situ weathering of residual
soils and rock.
Thickness of saturated clay deposits can vary from a few meters to several hundred
meters. Clay minerals with their bound water and free water molecules form a
porous structure. When a soil is compressed, the void volume is reduced. When a
soil is loaded at the top, load is first taken by pore water and later gradually
transferred to the soil matrix. Figure 12.1 shows an analogy which models this
behaviour.
When a load P is applied to the piston it would
first compress the fluid. This is true since the
fluid is incompressible and hence a slight
displacement of the piston would pressurise the
fluid to a pressure equal to P/A.
The fluid migrates through the tiny hole, thus
reducing the generated excess fluid pressure,
there by compressing the spring by an amount x.
This spring with a stiffness k would carry a load Figure 12.1: A spring loaded
equal to kx while the fluid would carry a piston carrying a load.
pressure of (P- kx)/A. The piston would
displace until the total displacement kx = P.
Water is an incompressible fluid. Water in clay pores are either bound to clay
particles or stays free. Pores, when inter-connected form flow channels. Their
pore diameters are much smaller than those of sands, hence make clay soils less
permeable. A saturated soil element at a certain depth may have its vertical
stresses satisfying the principle of effective stress,  v  v  u 0 . When a soil
element is subjected to an increase in total stress , it raises its pore water
pressure, thus:  v    v  [u 0  u] where   u .
With time, a certain amount of excess pore water pressure u dissipates due to
migration of water. This reduction in water pressure u is now transferred to the
solid soil skeleton, which increases the effective stress by the same amount.
CVX4343 – Soil Mechanics

The change in effective stress from  v to  v  u causes consolidation


settlement. Dissipation of excess pore water pressures will take place until they
finally become equal to u0. The effective stress at this stage would be equal to
v  u . This transformation of stress is expressed as:

 v    [v  u]  u 0

where [v  u] is the new vertical effective stress. Compression process as we
have seen takes place with dissipation of excess pore water pressures. Change in
volume caused by expulsion of excess pore water from its saturated pores is termed
Consolidation.

12.2 How settlements occur


Major construction projects such as the Colombo-Matara highway project,
Keravalapitiya-Hendala industrial estate, involved improving underlying soils.
Such improvements require us to reduce settlements and to improve their load
carrying capacity. Figure 12.2 shows how an embankment, 50m wide is
constructed over a compressible saturated clay layer.

Figure 12.2: Layout of road embankment


When a surcharge load is placed on a clay layer,
and the area of the surcharge is large, the
compressible stratum is subjected to a uniform
vertical stress. In this instance, the effective
vertical stress, v is increased by v.v for
elements A and B prior to loading (refer Figure
12.3(i)) would be 3(19.2-9.81) = 28.2kPa; the
change in stress v=27.8kPa.
This gives rise to a vertical strain, hence a
settlement. The stress increment v gives rise
to a horizontal stress incrementh= Kv. An
increase in horizontal stress may give rise to
compressive strains in the horizontal direction.
When two adjacent elements are considered the
two compressive strains in the horizontal Figure 12.3: Stresses in soil
direction violates strain compatibility requirement elements A and B (i) before
at the common boundary of the two elements. loading and (ii) at end of
This argument concludes that lateral strains loading.

85
Session 12: Computing settlements in soils

cannot exist, except in regions closer to the edges of the embankment.


When an element is subjected to zero lateral strain we say that `at rest’ condition or
`K0’ condition prevails. K0 is expressed empirically as K0 = (1-sin).
For most soils the angle of internal friction varies between 280 to 360. When  =
300 gives us K0 = 0.5.Can you compute h, and h at end of loading?
A pad foundation sitting on a compressible layer (refer Figure 12.4) subjects soil
elements A and B to varying vertical and lateral stresses. The stress values can be
calculated using the pressure bulb (refer Figure 8.13), which is based on the theory
of elasticity. When two soil elements located at the same depth are subjected to
varying vertical stresses, both elements would experience lateral strains. Hence
they no longer satisfy the `at rest’ condition.

Figure 12.4: Pad-footing sitting on a compressible layer; before and after loading
situations for elements A and B.
We have seen earlier that stress dispersion with depth is limited to a small region
on either side of the loading plane. This region approximately falls within the
dispersion of 2v:1h.

Simulating field changes in a laboratory environment


The one-dimensional consolidation test, which is also known as the oedometer test
estimates in-situ settlements.Figure 12.5 shows three test devices used for this
purpose.
Figure 12.6 gives a detailed view of the oedometer cell. A relatively undisturbed
saturated soil specimen that represents in-situ conditions is used for this purpose.
When a saturated soil specimen is extruded in to a tube sampler, its in-situ vertical
effective overburden stress vo is removed. This also changes its lateral
component ho , causing a slight disturbance. The specimen is then placed on the
consolidation test ring, after trimming it to the required size.

86
CVX4343 – Soil Mechanics

Figure 12.5: Table top oedometer devices (Courtesy ELE International,


SOILTEST)
A 1kg mass on the loading arm (refer Figure 12.5) gives 9.81N. A lever ratio of
1:10 gives 98.1N force on the sample. A 2.5 (63.5mm) diameter gives a cross
sectional area of 3166.92 cm2.
This gives a uniform vertical pressure v, of 31kPa.
The sample consolidates under this effective stress and the settlement, H is
recorded using a dial gauge that measures to 0.002mm accuracy.
The specimen consolidates under a particular load for 24 hours, or until settlement
stops.
At this stage the load is doubled; i.e. a conventional test is loaded from 1kg to 2, 4,
8, 16, 32, 64kg. The corresponding pressures are 31kPa to 62, 124, 248, 512, 991,
1984kPa.

The specimen is then loaded in the vertical direction. Testing requires subjecting
the sample to similar average field conditions. We first ensure that the prepared
soil specimen has the same orientation as in the field. Then a vertical load is
applied so that a uniform stress acts upon the total cross section and length of the
sample.
Figure 12.7 shows the stresses and strains in the soil specimen during loading. The
oedometer ring has a rigid circular wall made of brass or steel. This wall prevents
lateral straining, i.e. h = 0, thus gives `at rest’ or `K0’ loading condition. The test
enables us to simulate the `at rest’ condition in the laboratory, hence we are able to
represent the true in-situ loading situation.

Figure 12.7: Stress /strain representation.

87
Session 12: Computing settlements in soils

The right hand side column explains how settlement is measured for a given load.
The sampling depth is obtained from borehole record is used to estimate the
effective overburden stress, v 0 .
We also need to estimate the vertical stress increment  that the soil element may
experience due to the proposed construction. Suppose that v0  85kPa and
  42kPa the vertical stress at end of loading would be 85+42 = 127kPa. This
may require you to stress your laboratory sample to an effective vertical
consolidation stress of 248kPa. This is done by loading the oedometer to a final
load of 8kg, which would take around 4-6 days to carry out the test. The example
given in Session 3 discusses how the in-situ void ratio of the specimen, e0 and the
initial degree of saturation, S is determined.
The void ratio e1 corresponding to a total settlement of H (= H0-H1) is expressed
H e 0  e1 H
as:  . This enables us to plot void ratio, e or vertical strain  v 
H0 1  e0 H
versus v .

Figure 12.6: Oedometer cell

Activity 12.1: In-text question


Table 12.1 gives the results of an oedometer test performed on an organic peat
sample tested at Open University’s geotechnical testing laboratory. The test
sample was taken at 8.5-9.0m depth.
Initial height = 20mm; total settlement at end of test = 4.538mm; mass of water,
Mw = 41.535g; the equivalent volume Vw = 41.535 cm3; height of water.
41.535
Hw   1.0432cm, Hs = 20 - 4.538 - 10.432 = 5.03mm. The initial void
 2
7.12
4
H  H s 20.0  5.03
ratio, e 0    2.976 . Table12. 2 summarises the results for
Hs 5.03

88
CVX4343 – Soil Mechanics

H e 0  e1 e
each load.  gives the relationship   . This, with the initial
H0 1  e0 1  e0
void ratio e0 is used to determine the void ratios at end of primary consolidation,
under a given load. Table 12.2 list the variation of void ratio with effective vertical
stress.
Table 12.1: Oedometer test results for the organic peat.
Loading Loading Loading Loading Loading Loading
P(kg) = 0.39 P(kg) = 0.78 P(kg) = 1.565 P(kg) = 3.126 P(kg) = 6.26 P(kg) = 12.5
 = 10.75kPa  = 20.36kPa  = 39.7kPa  = 78.16kPa =155.38kPa =308.83kPa
Time H Time H Time H Time H Time H Time H
(min) (mm) (min) (mm) (min) (mm) (min) (mm) (min) (mm) (min) (mm)
0 0.000 0 -0.496 0 -0.760 0 -1.132 0 -1.656 0 -2.532
0.1 -0.054 0.1 -0.532 0.1 -0.804 0.1 -1.192 0.1 -1.764 0.1 -2.630
0.25 -0.064 0.25 -0.536 0.25 -0.818 0.25 -1.210 0.25 -1.784 0.25 -2.664
0.5 -0.075 0.5 -0.540 0.5 -0.830 0.5 -1.240 0.5 -1.806 0.5 -2.696
1 -0.088 1 -0.548 1 -0.844 1 -1.272 1 -1.836 1 -2.740
2 -0.102 2 -0.556 2 -0.862 2 -1.302 2 -1.884 2 -2.800
4 -0.126 4 -0.569 4 -0.887 4 -1.342 4 -1.940 4 -2.892
8 -0.154 8 -0.588 8 -0.919 8 -1.388 8 -2.028 8 -2.986
16 -0.194 15 -0.606 15 -0.953 15 -1.445 15 -2.118 15 -3.156
30 -0.236 30 -0.630 30 -0.996 30 -1.498 30 -2.226 30 -3.350
60 -0.287 60 -0.653 60 -1.032 60 -1.542 60 -2.316 60 -3.538
120 -0.348 120 -0.675 120 -1.058 120 -1.579 120 -2.378 120 -3.658
240 -0.400 242 -0.706 256 -1.084 260 -1.613 240 -2.432 240 -3.739
478 -0.450 418 -0.726 473 -1.106 415 -1.631 415 -2.466 399 -3.786
1460 -0.496 1458 -0.760 1423 -1.132 1425 -1.656 1437 -2.532 1464 -3.904

Table 12.1: continued


Loading Loading Unloading Unloading Unloading
P(kg) = 25 P(kg) = 50 P(kg) = 25 P(kg) = 1 P(kg) = 0
=616.74kPa  =1233.1kPa = 617.11kPa  = 25.78kPa  = 1.14kPa
Time H Time H Time H Time H Time H
(min) (mm) (min) (mm) (min) (mm) (min) (mm) (min) (mm)
0 -3.904 0 -5.869 0 -7.758 0 -7.518 0 -5.496
0.1 -4.000 0.1 -5.935
0.25 -4.040 0.25 -6.000
0.5 -4.100 0.5 -6.055
1 -4.156 1 -6.128
2 -4.243 2 -6.226
4 -4.370 4 -6.366
8 -4.555 8 -6.560
16 -4.774 15 -6.786
30 -5.069 30 -7.079
60 -5.357 63 -7.339
120 -5.541 120 -7.485
240 -5.653 240 -7.598
412 -5.761 466 -7.682
1420 -5.869 1460 -7.758 1428 -7.518 1437 -5.496 1423 -4.538

89
Session 12: Computing settlements in soils

Table12.2: Variation of e with v


Load  v H @ ~24 Void ratio,
(kg) (kPa) hrs. (mm) e
0 1.14 2.976
0.39 10.75 0.496 2.877
0.78 20.36 0.760 2.825
1.565 39.70 1.132 2.751
3.126 78.16 1.656 2.647
6.26 155.38 2.532 2.472
12.5 308.83 3.904 2.200
25 616.74 5.869 1.809
50. 1233.10 7.758 1.434
25 617.11 7.518 1.481
1 25.78 5.496 1.883
0 1.14 4.538 2.074

Figure 12.7 shows the variation of void ratio with v when the specimen settles
under the stated loads. You may have observed that the gradient of this plot
reduces with v.

Activity 12.2
Compute the change in void ratio e corresponding to a change in v from 85–127
kPa during loading?
Did you observe that the gradient of the loading curve change depending on the
stress range v?

Figure 12.7: e – v plot.


Figure 12.8 shows the same variation when plotted on a semi-logarithmic scale.
Note the loading and unloading segments of the compression curve. The latter
portion of the loading curve falls on a straight line.

90
CVX4343 – Soil Mechanics

Figure 12.8: e – logv plot.

e – v plot
The gradient of e  v curve is termed coefficient of compressibility, a v and is
 de
expressed as a v  . a v when computed for the stress range of 30 – 125kPa is
dv
 (2.53  2.8)
equal to  0.00284kPa1 .
(125  30)

The change in void ratio from 85 – 127 kPa is e = -0.00284  (127-85) = 0.1194.
Here we make the assumption that the in-situ soil strata and the laboratory
experience the same change in void ratio when subjected to the same stress change.
This assumption is true since the in situ void ratio e0 corresponds to the in-situ
effective overburden stress v 0 , and e0 is estimated from the laboratory specimen.
 e
This leads us to the relationship  where  is the settlement of soil layer
H 1  e0
with thickness H.
We have computed e0 to be 2.976. For a layer thickness of say 4.5m, the above
e H 0.1194x 4.5
equation will estimate the settlement,  as     0.135m .
1  e0 1  2.976
Activity 12.3: In-text question:
Compute a v for stress ranges 0 – 125 kPa, 125 – 250 kPa, and 500 – 625 kPa.
e – log10v plot
Figure 12.8 shows the same data set when plotted on semi-logarithmic scale. The
gradient of the straight line portion is defined as the compression index, Cc. This is
 (e 2  e1 )
expressed as C c  .
log10 v 2  log10 v1

91
Session 12: Computing settlements in soils

Compression index can be easily computed by considering e for a one log cycle.
One cycle is the distance between 100 -1000 kPa. Could you also take it between
110 – 1100?
The denominator when simplified works out to be log1010, which is equal to 1.
This makes Cc = e!
Activity 12.4: In-text question:
Let us compute the Cc. The void ratio, e for 1000kPa is 1.54; e for 100kPa is 2.85.
Hence Cc = 2.85-1.54 = 1.31.
If we look closely at e  log v plot we see that its shape changes around 180kPa.
This point has the maximum curvature (or the minimum radius of curvature). This
change in compression signifies the past maximum vertical effective stress the soil
specimen has undergone during its stress history. This stress is identified as the
pre-consolidation pressure, p .

Arthur Casagrande proposed a method to determine p . Figure 12.9 shows a


close-up view of e  log v curve. We determine point A that has the maximum
curvature.

Figure 12.9: A close up view of e  log v curve (a) and (b).

The point of maximum curvature can be obtained by moving a straight edge


inwards from its tangent point. The point that has the smallest radius needs to be
identified. At A we draw the tangent and the horizontal to the curve. We draw a
bisector to the angle the tangent and the horizontal makes at A. This line is
extended until it meets the straight line representing the compression index, at
point B. This point represents the pre-consolidation pressure, p  210kPa .

Constructing the field curve


The e  log v curve obtained above need to be corrected to account for sample
disturbance. Figure 12.9 shows how the in-situ compression curve is generated
from the laboratory compression curve.

92
CVX4343 – Soil Mechanics

This in situ effective overburden stress, v 0 has a void ratio equal to e0 . The soil
element recalls stresses v 0 to p that it had experienced before. This was during
a previous loading /unloading cycle. This re-compression is quantified using the
re-compression index, C r and is obtained by measuring the gradient of the
loading-unloading curve. Therefore line segment AB has a gradient C r .
Line segment BC represents the in-situ compression index. C is located on the
laboratory compression line with e  0.42e 0 .
Activity 12.5: In-text question
Can you draw the idealised in-situ compression curve for the example we’ve
looked at?
The in-situ settlement  is estimated based on the idealised in-situ compression
curve. To use this we need to know the thickness of the compressible layer, its in-
situ void ratio e0, in-situ overburden stress v 0 , pre-consolidation pressure,  p
compression index, C c and recompression index, C r .

Figure 12.10: Idealised in-situ compression curve.


A compressible state will settle when its stress is raised from its in-situ overburden
stress v 0 . The stress increment v decides by how much settlement it would
undergo. Suppose that v  p  v0 then re-compression of soil takes place
and hence C r is used to determine the settlement. If v  p  v0 then one has
to consider C c and C r .

A soil that experiences stresses for the first time since its formation is said to
undergo normal consolidation. Such a soil is a normally consolidated (NC) soil.
A soil that recalls stresses as stresses it had experienced during its stress history is
termed an over-consolidated (OC) soil.
Activity 12.6: In-text question
A compressible soil layer, 4.5m thick has the following properties:
e 0  2.8, C r  0.2, C c  0.8, v0  85kPa, p  210kPa

93
Session 12: Computing settlements in soils

Determine: the settlement for a stress increment a) v  75kPa and for b)
v  175kPa .
e0  ef
f) v  p  v0 therefore C r  0.2  .  e 0  e f  0.05494.
85  75
log10
85
 e0  ef 0.05494
 gives   x 4.5m = 65mm.
H 1  e0 1  2.8
e 0  e1 e1  e f
g) v  p  v0 therefore C r  0.2  and C c  0.8 
210 210  50
log10 log10
85 210
.
Hence e 0  e1  0.07856 and e1  e f  0.07420 therefore, e 0  e f  0.15276.
 e0  ef 0.15276
 gives   x 4.5m = 181mm.
H 1  e0 1  2.8

To sum up
The One-dimensional Consolidation Test, when performed on an undisturbed
representative soil sample gives us the necessary information to predict the
consolidation settlement. We have seen how the in-situ compression curve is
obtained having made necessary adjustments to account for sample disturbance.
The test yields parameters e0, p , v 0 , Cc and Cr. The compression curve is used to
determine the change in void ratio e when the vertical effective stress is raised
from v0 to v0  v . The consolidation settlement of the compressible layer, 
is determined using parameters e, e0 and H.

94
Session 13
Computing time for settlement

What you will learn in this session


This is a 2-hour session. Section 13.1 of this session explains how compressible
soils such as clayey soils and organic peats settle with time. The One-dimensional
consolidation test discussed in the previous session gives us the variation of
settlement with time, for a given load. Section 13.2 explains the mathematical
model put forward by Professor Carl Terzarghi to explain the sample response.
Section 13.3 shows the procedure to compute the model parameters that represents
rate of settlement.

13.1 Introduction
Settlement of compressible soils does not happen instantaneously but takes place
over several years. Significant settlements cause structural damage to buildings
making them unusable. When building on a soft compressible soil, the usual
practice is to place a fill material of sufficient thickness. This has to be well
compacted so that it would provide a good bearing stratum for shallow footings. It
also raises the ground elevation, which may keep the zone of influence within the
fill, and thus reduce settlements after construction. Such fills cause settlements in
underlying layers, hence we would like to estimate the time duration for a
significant settlement to occur.
The reason for delayed settlement is slow dissipation of excess pore water
pressures (refer Figure 13.1). A water molecule located at the centre of the layer
has a tortuous journey to reach the free surface, which is at atmospheric pressure.
The movement is governed by permeability of the layer. The travel time depends
on the length of flow path; a two way draining layer would settle faster than a layer
with single drainage.

Figure 13.1: The model which explains settlement due to dissipation of pore
pressures.
This session explains how the time taken for a certain degree of consolidation to
occur is computed based on settlement-time data obtained from an oedometer test.
Session 13: Computing time for settlement

The settlement at time t when expressed relative to the total settlement is termed
Degree of Consolidation, U. If settlement after 3 years is 387mm and the total
settlement is estimated to be 548m U = 70.6%.
U is also expressed in terms of excess pore water pressure generated, u and the
u u
initial excess pore water pressure, uias U  i .
ui
Figure 13.2 shows settlement changed with time when the effective stress was
increased from 39.7 to 78.2kPa. This graph is drawn to a linear scale and hence
does not show enough detail to comment on initial settlement behaviour. This is
overcome by plotting either   t (refer Figure 13.3) or   log(t ) (refer Figure
13.4).

Figure 13.2:  - t curve for stress increment, v from 39.7 to 78.2kPa.

Figure 13.3:   t curve for stress increment,  v from 39.7 to 78.2kPa.

96
CVX4343 – Soil Mechanics

Figure 13.4:   log(t ) curve for stress increment, v from 39.7 to 78.2kPa.
It shows three distinct regions. The settlement during the first few minutes occurs
at high rate. During the next few hours settlement rate reduces with time. Beyond
8 hours settlement takes place at a slower rate.
When the sample is loaded, it undergoes rapid settlement. This initial compression
occurs when the soil fabric deforms like an elastic material. Following this,
settlement becomes gradual with the rate reducing with time. This is termed
primary consolidation. Primary consolidation causes a significant reduction in
void space.
Settlement rate is correlated to the increase in effective stress and excess pore
water pressure dissipated during the process. Suppose that we continue with the
test for several months settlement may continue at a much slower rate. This is
termed secondary compression. Secondary compression is high in organic clays
and peat and hence when designing structures to suit such conditions, we need to
account for secondary compression.
Activity 13.1: In-text question
Initial compression is immediate. Primary consolidation may follow initial
compression or do both occur simultaneously?
Activity 13.2: In-text question
Can secondary compression and primary consolidation occur simultaneously?
The following section discusses the development of a mathematical model which
models primary consolidation. The model enables us to determine settlements with
time when subjected to a particular load. It can also be used to compute the degree
of consolidation for a particular time period.

13.2 Theorising settlement () –time (t) behaviour


Carl Terzaghi proposed a mathematical formulation to explain -t behaviour. His
analysis was based on the assumptions listed in the right-hand side column. He
computed the net upward flow of water across the soil element with volume
(dx.dy.dz) due to vertical compression. This is simplified to give

97
Session 13: Computing time for settlement

k 2u
dQ out  dQ in    dx .dy.dz.dt (refer Figure 13.5). The change in height
 w g z 2
a d
due to compression of void space, (dz) is expressed as dz   v dz.dx .dy ,
1  e1
where the initial void ratio is e1. The change in excess pore water pressure du,
causes the effective stress to increase by d, hence du = d.

Figure 13.5: (a) saturated clay soil element dx, dy, dz subject to vertical
compression, (dz), (b) the variation of stresses with depth.
Terzarghi’s1-d consolidation theory makes the following assumptions:
h) Clay is homogeneous and completely saturated.
i) Drainage at top and bottom of layer.
j) Water movement through soil governed by Darcy’s law.
k) Solid soil grains and water are incompressible.
l) Compression and flow occur in one direction.
k  2u a du
These two equations give us the partial differential equation  v .
 w g z 2
1  e1 dt
du 2u k 1  e1
This is expressed as  c v 2 where c v  . cv is found to be a
dt z w g a v
constant for a given small load increment. It is termed the coefficient of
consolidation.
For a small increment it is assumed that layer height, hydraulic conductivity, k
compressibility, av remain constant.
The relationship de  a v v remains constant over applied stress increment. No
secondary compression.

du 2u
The partial differential equation  c v 2 is solved using Fourier
dt z
Transformations. The theoretical solution is expressed in terms of two
dimensionless quantities namely, the degree of consolidation, U and time factor T.
This relationship is found to correlate well with following expressions:

98
CVX4343 – Soil Mechanics

 2
T U for U  60% ; T  1.781 0.933log10 (100  U%) for U  60% .
4
ui  u
Earlier we defined degree of consolidation, U as U  . Time factor T is
ui
cvt
defined as T  , elapsed time, t length of drainage path, d.
d2
Figure 13.6 and 13.7 express this relationship graphically. Suppose that observed
in-situ settlements resemble the theoretical solution shown in Figures 13.6 and
13.7, we could for instance, use the theoretical approach to determine the time for a
particular degree of consolidation to occur. This, however requires us to perform a
field test to establish coefficient of consolidation, cv, which would be time
consuming and costly.
The oedometer test can be used effectively to determine a laboratory cv value. We
have discussed that the laboratory test simulates field conditions, and hence
accurate measurements of laboratory settlement would estimate cv fairly accurately
for inorganic clays. cv is determined by matching a point on the theoretical curve
with the corresponding point on the laboratory curve.

Figure 13.6: Square root of time plot of Degree of Consolidation U versus Time
Factor,Tv

13.3 Computing cv using t and log-t methods


When matching the experimental curve, -t with the theoretical curve, U-T we are
seldom able to find an identical match.
Figure 13.8 shows the method used to determine t 90 . t 90 is the time for 90%
primary consolidation, i.e. U=90%. This graphical method requires you to
construct a tangent to the straight line portion beyond initial compression. This
line is extended till it meets point A. This point is taken to be  0 , settlement when
primary consolidation commenced. We then draw line AC located at 1.15 time the
distance to line AB. The point where line AC intersects with the curve corresponds
to  90 . This gives t 90  5.62 and hence t 90  31.58 min .

99
Session 13: Computing time for settlement

This method tells us that T90 matches well with the corresponding point on the
experimental curve. This point would have t = t90, which is 31.58 min.

Figure 13.7: Semi logarithmic plot of Degree of Consolidation U versus the Time
Factor T.
Time factor T90 is computed as T90  1.781  0.933log10 (100  90%)  0.848 .
Now we need to determine the average length of drainage path. Table 12.2 gives
the settlements corresponding to 78.12 kPa stress. At t = 0,  = -1.132; at t = 1425
min., = -1.656; therefore the sample height varies from (20-1.132) mm to (20-
1.656) mm. The drainage path d is the maximum distance a water molecule may
travel to relieve its excess pore water pressure. For the two-way draining layer of
thickness H, this is equal to H/2. This gives the average sample height to be
(1.132  1.656)
20   18.606 mm . Hence length of drainage path d = 9.303 mm.
2
The time factor at U = 90% is expressed as:
c t 0.848x 9.3032 mm2
T90  v 290  c v   0.039 mm2/s or 0.003 m2/day.
d 31.58 x 60 sec
Figure 13.9 shows the method used to determine t 50 . t 50 is the time for 50%
primary consolidation, or U=50%. This graphical method requires you to establish
 0 and  100 to determine  50 . To find  0 we must first identify a time t 1 near the
first segment of the curve. Here, I have picked 0.1 min. point. Then we need to
determine the settlement corresponding to t 2  4t 1 , in this case it is 0.4 min. point.
The settlement, x between these two time values are then determined.  0 is
determined by measuring a distance x from the first point.
100 is established by drawing tangents as indicated in Figure 13.9. This signifies
end of primary consolidation. t 50 corresponding to  50 is found to be 9 min.

T is computed as T  (0.5) 2  0.196 . Using the same drainage path, d = 9.303
4
c v t 50 0.196 x 9.3032 mm2
mm, T50   cv   0.031 mm2/s or 0.0027 m2/day.
d2 9 x 60 s

100
CVX4343 – Soil Mechanics

Figure 13.8: t method to determine t 90 .

You may have already noticed that we have not plotted the settlement for t=0. This
is because we cannot define t=0 on a logarithmic scale.

Figure 13.9: Log(t) method to determine t 50 .

Modern day Rowe cells are instrumented to measure the generation and dissipation
of excess pore water pressure with time. This enables us to determine settlement
when primary consolidation commenced,  0 and when it ended 100 .

101
Session 13: Computing time for settlement

Activity 13.2: In-text question


Figure 13.10 shows a recently placed residual fill overlying a 8m thick organic
clay. Determine the time taken for 80% of the total settlement to occur.
For U=80%, T80  1.781  0.933log10 (100  80%)  0.567 . The drainage length
would be 8m since the bottom layer is impermeable. Hence
T d 2 0.567x[8x10 3 ]2
t  80   28.8 years . This is a long wait. You might want
cv 0.04
to find a way to accelerate the settlement rate.

Figure 13.10: Accelerating settlement using sand drains.

To sum up
This session have shown you how in-situ settlement rate is estimated based on
laboratory observations.
Professor Terzarghi’s settlement model based on One-dimensional Consolidation
Theory was used to explain the observed settlement behaviour. We have also seen
how coefficient of consolidation, cv is computed. This is found to vary with
vertical effective stress.
This session also have shown how you could estimate the time for a certain degree
(e.g. 80%) of primary consolidation. The same analysis could also be used to
determine the degree of consolidation that is achieved during a particular time
period (i.e. within the first 10 years after completing construction).

102
Session 14
Soil strength

What you will learn in this session


This is a 1-hour session. Section 14.1 explains how inter-particle friction and
particle interlocking contribute to soil strength. A soil loses its strength when
applied forces overcome soil resistance. Soil strength is expressed using Mohr-
Coulomb failure criteria. Section 14.2 explains how soil failure is determined
based on shear strength parameters  and c.

14.1 What makes a soil mass loose its strength?


Granular soils gain strength due to inter-granular
friction occurring at particle contact points, and
particle interlocking. A high effective stress also
enhances soil strength.
When an external force is applied to a granular
assembly (refer Figure 14.1) at points of contact,
normal force, N produces a resistance of N to
overcome the tangential force, T. In the case of
mineral grains, the coefficient of static friction 
is around 0.3. When T  N we find that particle
contacts slide. When the number of sliding
contacts increases, particles undergo rotation and
translation causing the assembly to lose its Figure 14.1: A granular soil
carrying load.
`strength’. The above mechanism seems a simple
and a logical explanation.
Granular soils, unlike spherical material do not simply roll or slide. Their irregular
shaped rough surfaces make particles to inter-lock giving it additional strength.
Some grains can also form bonds due to cementation.
Therefore, at particle level a normal force may produce a resistance Tmax  N  c ,
where c is a constant independent of N.
What makes a soil mass loose its strength? When a rubber band is stretched, the
tensile stress increases and suddenly it snaps. A balloon bursts when membrane
stresses become considerable. A beam may deflect more when tensile or
compressive stresses become excessive. A soil that is well confined in all sides
cannot loose its strength unless applied stresses cause particle contacts to crush.
This requires external pressures of several mega-pascals.
Figure 14.2 shows how a soil element holds together with effective stresses v and
h . This is equivalent to a deviatoric stress of v  h , and a hydrostatic stress,
h . The hydrostatic stress component compresses the soil mass together, thus it
does not cause any loss of strength. Here, we assume that the compressive strength
is not too high to crush soil grains. The deviatoric stress component v  h is an
unbalanced stress that may make the soil sample to lose its strength.
Session 14: Soil strength

Figure14.2: Representing stresses in a soil


element.
Let us compare two elements – element A made
of solid aluminium and element B being a
granular material. When both elements are
subjected to a deviatoric stress of v  h ,
element A would either crush or buckle due to
compression; element B would shear along grain
boundaries. Figure 14.3 shows two inclined
planes along which failure occur.
Figure 14.3: Shearing planes
14.2 Failure criterion that defines soil
failure
Figure 14.4 shows the stress-strain behaviour of a cylindrical dense sand sample
when subjected to a deviatoric stress.

Figure 14.4: Stress-strain behaviour of a dense sand.


In order to represent these stresses on a Mohr’s circle of stress plot, we make two
important assumptions. They are: a) initial stresses in the horizontal and vertical
stresses are principle stresses and b) during deformation principal stress directions
remain unchanged. These assumptions allow us to use principal stresses instead
(refer Figure 14.5). Suppose that the Mohr’s circle in Figure 14.6 represents the
failure condition.

Figure 14.5: Stresses approximated to principal stresses.

104
CVX4343 – Soil Mechanics

During a tensile or compressive failure, failure stress is normal to its plane of


failure. When a soil element fails, failure takes place on an inclined plane (refer
Figure 14.6). Therefore point on Mohr’s circle corresponding to failure has
coordinates (n ,  f ) . When samples are tested at varying hydrostatic stresses,
points of failure lies on an envelope termed Mohr-Coulomb failure envelope (refer
Figure 14.7). This approximately represents a straight line having the equation
 f  n tan   c . Shear strength parameters  and c represent the angle of
internal friction and apparent cohesion.

Figure 14.6: Mohr’s circle representing failure.

Figure 14.7: Soil failure at varying deviatoric stresses.


Activity 14.1: In-text question
A drained triaxial loading test is performed on a loose-sand with
2  2f  500kPa. Deviatoric stress at failure = 820kPa. Determine shear
strength parameters. Determine the `theoretical’ angle that the failure plane makes
with the horizontal.
Figure 14.8 shows the Mohr’s circle of stress at failure. The soil being a loose
sand, it has zero apparent cohesion, c  0kPa. The length OA = (1320+500)/2 =
910kPa. The radius of the Mohr’s circle = 820/2 = 410kPa.
Sin  410 / 910;   26.8 0 . 2  900      58.4 0 .

105
Session 14: Soil strength

Figure 14.8: Mohr’s circle of stress at failure.


Figure 14.8 can be used to compute the angle between the failure plane and the
horizontal direction. The angle between the two principal stresses is 900 since
principal stresses are orthogonal. The Mohr’s circle of stress shows this angle as
1800. You may recall that any angle measured between two stress points on the
Mohr’s circle of stress, e.g. 2, represent an angle  between the two planes under
consideration.
The stress point B (refer Figure 14.8) lies on the failure plane while the stress point
corresponding to major principal stress represents the horizontal plane. Therefore,
angle  would be the angle that the failure plane makes with the horizontal
direction.
The triangle ABO gives us the relationship 2 = 900+ , and hence
  450   / 2

To sum up
This session explained the representation of 2-dimensional stress in a soil element
using Mohr’s circle of stress concept. When a soil specimen is subjected to a
deviatoric loading, the specimen fails along a plane where its shear stress  reaches
the failure shear stress f. The failure shear stress f is explained by Mohr-
Coulomb failure criterion,  f  n tan   c . The failure plane that corresponds
to this failure shear stress makes an angle 450   / 2 with the horizontal direction.

106
Session 15
Laboratory tests to assess soil shear strength

What you will learn in this session


Section 15.1 explains the conventional laboratory tests performed on saturated soil
samples. Representative undisturbed samples are tested to determine soil strength
parameters.
The Unconfined Compression (UC) Test (Section 15.2)is carried out while
maintaining undrained conditions in the specimen; this gives us total strength
parameters c u with  u  0. The UC test is used to perform a total stress analysis.
Section 15.3 discusses three conventional traixial loading tests. The test performed
under consolidated drained (CD) condition gives us  and c. These parameters
represent inter-granular or effective strength, hence can be used to perform an
effective stress analysis.
The consolidated undrained (CU) triaxial loading test allows us to measure excess
pore water pressures. This gives us strength parameters with respect to effective
stresses  and c and those with respect to total stresses  and c . CU trixial test
results are useful in performing an effective stress analysis and a total stress
analysis.
The unconsolidated undrained (UU) triaxial loading test is done under `undrained’
conditions. This gives us strength parameter c u with  u  0. These strength
parameters are used to perform a total stress analysis.

15.1 Introduction
In-situ undisturbed soil samples are collected
during a geotechnical investigation. These
samples are preserved in sampling tubes and later
transported to a geotechnical laboratory for
testing. These samples need to be handled with
care in order to prevent any disturbance. Such
disturbances may cause changes in the soil fabric
of soft compressible soils. In the case of granular
soils disturbances cause samples to compact.
Loss of moisture from the sample may require
back pressure saturation, prior to testing. If such
errors go undetected the strength parameters
would overestimate the available strength.
Laboratory samples are subjected to a deviatoric
stress until the applied axial force starts to drop.
The loading apparatus (ELE International,
SOILTEST) shown in Figure 15.1 is used to
perform the Unconfined Compression test. Figure 15.1: Unconfined
compression testing apparatus
The base moves upwards, thereby pressing the (Courtesy ELE International,
sample against the proving ring. The test SOILTEST)
apparatus maintains a constant strain rate. When
Session 15: Laboratory tests to assess soil shear strength

the proving ring compresses, its load dial records its displacement. The proving
ring is calibrated to give the force, P. The calibration constant represents the
stiffness of the ring.
The displacement dial reads the displacement, L. The vertical strain is given by
L
1  , where L 0 is the initial height of sample. The area of cross section at a
L0
given strain  1 depends whether the sample volume is allowed to change. We saw
that the volumetric strain  V is can be expressed as 1  2 2   V , where
V d
volumetric strain  V  ; the lateral strain,  2  . During undrained loading
V0 d0
sample volume remains constant. This is so since no drainage is allowed from a
completely saturated sample. Hence during undrained loading the condition
1  2 2  0 .
During drained loading volume change occurs with the dissipation of excess pore
water pressures. Sometimes you may find that the sample volume increases during
loading. The change in average cross-sectional area is expressed as
2
A  / 4(d  d) 2   / 4d 2 d  d 
 2   2 2 . This expression gives the
A0  / 4d 2
d  d 
new area A  A 0  A  A 0 [1  2 2 ] . When loading takes place under undrained
conditions A  A 0 [1  1 ] . During drained loading A  A 0 [1   V  1 ] .

15.2 Unconfined compression (UC) test


The Unconfined Compression Test as the name implies, is performed without
confining or providing an all round pressure, thus  2  0 . This makes the
deviatoric stress 1   2  1 .
The soil is tested under undrained conditions. This means that excess pore
pressures dissipation is prevented. This is achieved by applying a high axial strain
rate to the specimen, thereby not giving it enough time to dissipate its excess pore
water pressures. This assumption would be valid for clayey soils which has
significantly low permeabilities. The UC test is also called the Quick Test. The
UC test is a simple test and hence we cannot measure excess pore pressures during
testing. Therefore, what we measure is the total stress. Total stresses give a
Mohr’s circle with respect to total stress and hence at failure, this computes shear
strength parameters c and.
During undrained loading we will learn later why
the angle of internal friction with respect to total
stress,  u is always zero. When  u  0 , Mohr-
Coulomb failure envelope becomes a horizontal
surface touching the Mohr’s circle at its
maximum shear stress point (refer Figure 15.2).
This gives the undrained shear strength c u .
Figure 15.2: Mohr-Coulomb
The undrained shear strength parameter c u is failure envelope based on
used to analyse the stability of soil structures undrained  u  0 condition.
under undrained situations. Usually this
corresponds to immediate condition where excess

108
CVX4343 – Soil Mechanics

pore water pressures are yet to dissipate. c u obtained from an undisturbed sample
would directly determine the soil’s in-situ strength.

15.3 Conventional triaxial loading tests


The conventional triaxial loading test represents
in-situ conditions better than the UC test. It
allows us to apply a confining stress to the soil
sample. This gives a lateral total stress  2 .
Unlike the UC test, the sample is isolated from its
surrounding with a flexible latex membrane. This
enables us to maintain a saturated specimen. It
also enables us to control sample drainage,
independent of the applied strain rate. Two
drainage lines are connected to the top and
bottom platens. This enables us to de-air the
sample and also to allow water to seep through.
Figure 15.3 shows a standard triaxial loading cell.
A completely saturated soil sample can be tested
under undrained conditions by closing all
drainage values during the testing stage. The
Figure 15.3: Triaxial loading
drainage lines enable us to connect the sample to cell (Courtesy ELE
a gauge that measures the excess pore water International, SOILTEST).
pressure. These are sometimes used to saturate
slightly unsaturated samples by sending water to
the soil specimen under a given pressure. This is
termed back pressure saturation.
The total stress in the vertical direction consists of cell
pressure  2 and the deviatoric stress 1  2 . This
gives a net stress of 1 .

The conventional triaxial loading test consists of


two stages of loading. Initially the sample is
allowed to consolidate under a confined stress.
This is done by increasing the cell pressure to the
desired value while keeping the drainage lines
open. This results in an isotropic (i.e. all round)
consolidation. Constant pressure devices (refer
Figure 15.5) is used to maintain a constant cell
pressure and a constant back pressure. Figure
15.6 shows the standard triaxial loading Figure 15.5: Constant pressure
apparatus. device used to apply cell
pressure and back pressure
Isotropic Consolidation
(Courtesy ELE International,
During isotropic consolidation we measure the SOILTEST).
volume change V and the height change L.
These give us  V and 1 . The new area of cross
section is expressed as A1  A 0 [1   V  1 ] .
The consolidation phase is carried out to bring the
sample to a pre-determined average stress.

109
Session 15: Laboratory tests to assess soil shear strength

Figure 15.4: The conventional triaxial loading apparatus


Applying a deviatoric stress
The deviatoric stress 1   2 is applied to the
sample through the loading device. At a given
strain  1 the deviatoric stress is computed as
P
1   2  , where P is computed based on
A1
the load dial reading; A 1 is obtained as described
above.
During drained loading, a slower strain rate  1 is
maintained to give sufficient time for the excess
pore water pressure to dissipate. The drain line is
connected to the volume measuring device. This
device is either kept opened to the atmosphere or
connected to constant pressure device that
supplies a back pressure. Undrained loading Figure 15.6: Triaxial pressure
requires us to maintain zero volumetric strain. panel (Courtesy ELE
International, SOILTEST).

When a pressure panel (refer Figure 15.6) is used to measure the excess pore water
pressures during an undrained loading test, it is connected to the drainage outlet via
a null indicator, which is used to ensure that pore water neither goes in nor comes

110
CVX4343 – Soil Mechanics

out of the soil specimen. Modern triaxial loading devices are equipped with pore-
pressure transducers that are directly connected to a drainage outlet. Figure 15.7
shows the twin burette volume measuring device used during a consolidated
drained triaxial loading test.

Consolidated drained (CD) triaxial loading test


CD test determines drained shear strength
parameters c  and   . We are required to perform
at least three CD tests (at three cell pressures) to
establish the Mohr-Coulomb failure envelope.
The three circles (refer Figure 15.8) is then used
to establish the Mohr-Coulomb failure
envelope.This test allows drainage, therefore we
could directly determine the effective stresses
1  1 and  2  2 .
Measuring the inclination of failure plane is not
done due to inaccuracies.
Consolidated undrained (CU) triaxial
loading test
CU test determines shear strength parameters c
and  , with respect to total stress and c  and  ,
with respect to effective stress.
This test is an undrained test, with excess pore
water pressure measurements. The test enables Figure 15.7: The twin burette
volume measuring device
us to compute the effective stresses 1  1  u
and  2  2  u , and hence obtain the two sets of
circles (refer Figure 15.9).

Figure 15.8: Mohr-Coulomb failure envelope based on effective stress.


Unconsolidated undrained (UU) triaxial loading test
UU test is performed at three different cell pressures without allowing the
specimens to consolidate. In a fully saturated soil, when the confining stress is
increased while keeping the drains closed, the excess stress is taken by pore water;
hence the soil structure remains unaltered. This also means that its water content
remains unchanged during the test period. Figure 15.10 shows the Mohr-Coulomb

111
Session 15: Laboratory tests to assess soil shear strength

failure envelope drawn based on Mohr’s circles of stress with respect to total and
effective stresses. You may find that both envelopes overlap.

Figure 15.9: Mohr-Coulomb failure envelope based on (a) total stress and (b)
effective stress.
The Mohr’s circles of stress at failure have the same radius. Further, the three
Mohr’s circles based on total stress give us the same circle based on effective
stress.
 u  0 is the outcome of the way the test is performed. Here we have not allowed
the sample to change its soil structure.

Figure 15.10: Mohr-Coulomb failure envelope for undrained  u  0 condition.


Activity 15.1: In-text question
Explain what would happen if you unknowingly allow the specimen to consolidate.
Will it cause the specimen to consolidate? Can you explain what effect it would
have on Mohr’s circle of stress at failure?

112
CVX4343 – Soil Mechanics

To sum up
This session explained the conventional tests performed on in-situ soils to
determine strength parameters. It discussed four tests namely, the UC test, CD test,
CU test and the UU test. These tests yield strength parameters with respect to total
and effective stresses.
The above parameters are used to perform stability analyses.

113

You might also like