Você está na página 1de 120

University of Khartoum

Building and Road Research Institute (BRRI)


Geotechnical Engineering Department

GeotechnicalPropertiesofClayeySand
(SC)soils

A thesis submitted in fulfillment for the degree of Master of Science


in Building Technology at University of Khartoum

Presented By:
Haitham Shammat Al-Amin Saeed

December 2009

ABSTRACT

Montmorilloinic clayey sand (SC) soils tend to have relatively high values of
L.L & P.I and noticeable variations of volume change as their moisture
content changes. The expansive clay in (SC) soil tends to increase with the
increase of its colloidal properties.

In this research, it is intended to examine the behaviour of an artificial and


natural soils made by combining different ratios of sand with clay of known
characteristics. The number of soil types tested in this program can give good
idea about the geotechnical properties of (SC) soil. The effect of increasing
sand fraction on soil was determined at different types of clay minerals. The
use of artificial and natural soil samples is intended to clarify the effect of
sand fraction on the behaviour of sand-clay mixture. Increasing sand fraction
content will affect the volume change characteristics whereas it changes the
properties of expansive clay and then changes the whole soil structure.

Series of index tests, compaction, consolidation and swelling tests were


performed to understand the geotechnical properties of (SC) soil. Package
SPSSWIN was used for the regression analysis of laboratory tests data. The
test results were analyzed and presented.

The research concluded that artificial soils with 5% bentonite content and not
more than 15% are classified as having low swelling potential and
recommended for landfill liners of low hydraulic conductivity. Increasing the
sand fraction from 7% to 10% will give low volume change characteristics
and high strength properties for (SC) soils for using under foundations or fill
materials.



. .


.
) ( .

.
.
.


.
SPSS .

5 15

. 7 10

.

XI

Table of Contents

Table of Contents..I

List of Figures..V
List of Tables...VIII

AbstractX


 ...XI

CHAPTER 1 INTRODUCTION
1.1 Introduction.1
1.2 Thesis Organization2

CHAPTER 2 LITERATURE REVIEW


2.1 Introduction.........4

2.2 Soil composition.4

2.2.1 Sand and silt fraction.....4


2.2.2 Clay fraction..5
2.2.2.1 Mineralogy of clay...........6
2.2.2.2 Bentonite10

2.2.2.3 Influence of the clay fraction.11

2.3 Mechanism of swelling.12

2.4 Expansive soil.......14

2.5 Identification of the expansive soils.15

2.6 Volume change of the sand-bentonite mixtures16

2.7 Stabilization of the clayey sand (SC) soil.........18

2.7.1 Lime stabilization (General review)............19

I
2.7.2 Stabilization of the clayey sand (SC) soil using fly ash mixed with
small amount..........20
2.7.3 Strength behavior of Stabilized clayey sand (SC) soil using lime-
polyamide fiber..........22

2.8 Clayey sand (SC) soils and their properties..26

2.8.1 Compression behavior of Clayey Sand (SC) soils..26


2.8.2 Compaction of clayey sand (SC) soils.27
2.8.3 Clayey sand (SC) soils as fill and embankment material28

2.9 Conclusion....30

CHAPTER 3 EXPERIMENTAL WORKS AND RESULTS


3.1 Introduction.......34

3.2 Test program.34


3.3 Preparing of soil samples..35

3.3.1 Artificial soil sample...........35


3.3.2 Natural soil samples............36

3.4 Artificial soil samples.......37

3.4.1 Classification tests...........37


3.4.2 Moisture-density relationship test (Proctor compaction)...38

3.4.3 Consolidation tests...........39

3.5 Natural soil samples..40


3.5.1 Classification tests..........40

3.5.2 Moisture-density relationship test (Proctor compaction)............42

3.5.3 Consolidation tests...........43

3.6 Conclusion44

CHAPTER 4 DATA ANALYSIS AND DISCUSSIONS


4.1 Introduction...47

II
4.2 Artificial soil samples...47

4.2.1 Classification tests results............47


4.2.1.1 Atterbergs limits...........47

4.2.1.2 Classification and identification of expansiveness................49

4.2.1.3 Prediction of Liquid limit (Schreiners method).........50

4.2.1.4 Prediction of specific gravity (Estebans method)..........51

4.2.1.5 Predicting of compression index (Cc) (Skemptons


method).......53

4.2.2 Compaction test results..........54

4.2.3 Consolidation test results56

4.2.3.1 Permeability coefficient (k) and consolidation coefficient


(Cv)62
4.3 Natural soil samples.66

4.3.1 Classification tests results............66

4.3.1.1 Effect of increase of sand content on Atterbergs limit...66

4.3.1.2 Effect of increase of sand content addition on swelling


pressure (S.P)....66

4.3.1.3 Classification and identification of expansiveness..67

4.3.1.4 Prediction of Liquid limit (Schreiners method)......69

4.3.1.5 Prediction of compression index (Cc) (Skemptons


method)........70

4.3.2 Compaction test results72

4.3.3 Consolidation test results.74


4.3.3.1 Permeability coefficient (k) and consolidation coefficient
(Cv)..78
4.4 Conclusion....83

III
CHAPTER 5 SUMMARIES AND RECOMMENDATIONS
5.1 Introduction.......86

5.2 Summaries........86
5.2.1 Literature review................86
5.2.2 Experimental work.....88
5.2.3 Analysis and discussions89
5.3 Recommendations91
References......92
Appendix (A): Test procedures for Specific gravity (S.G) of artificial soil
samples .............102
Appendix (B)...104

IV
List of Figures
Page
Figure No. Name of figure
No.
2.1 Clay minerals: Basic units. 7

2.2 Diagrammatic sketch of the Kaolinite structure. 7

2.3 Diagrammatic sketch of the Montmorillonite structure. 8

2.4 The Kaolinite minerals. 9

2.5 The Illite clay minerals. 9

2.6 The Montmorillonite clay minerals. 10

2.7 Phase relations of clay - enriched granular soil. 12

2.8 Internal electrochemical system of soil. 13

2.9 Optical micrograph of a cut-polyamide strip 23

Comparison between compressive strength of reinforced and


2.10 24
non-reinforced samples vs. clay content for 3% lime content.
Comparison between compressive strength of reinforced and
2.11 25
non-reinforced samples vs. clay content for 6% lime content.
Comparison between compressive strength of reinforced and
2.12 25
non-reinforced samples vs. clay content for 9% lime content.
Particle size distribution for the fine sand used in the
3.1 36
artificial soils samples.

3.2 Moisture-density relationships for the artificial soil samples. 38

3.3 Consolidation curve results (artificial soil samples). 39

3.4 Grain size distribution curves for the natural soils samples. 41

V
3.5 Moisture-density relationships for the natural soil samples. 42

3.6 Consolidation curve results (natural soil samples). 43

Relationship between bentonite content and Atterbergs


4.1 48
limits.
Relationship between sand content and Atterbergs limits
4.2 48
(artificial soil samples).
Relationship between Schreiners method and L.L from
4.3 51
laboratory testing (artificial soil samples).
Relationship between Estebans method and S.G from
4.4 53
laboratory testing (artificial soil samples).
Relationship between M.D.D and bentonite content (artificial
4.5 55
soil samples).

4.6 Relationship between M.D.D and P.I (artificial soil samples). 56

Relationship between bentonite content and swelling


4.7 59
pressure.
Relationship between swelling pressure and P.I (artificial soil
4.8 59
samples).
Relationship between mv and Atterbergs limits (artificial
4.9 60
soil samples).

4.10 Relationship between Cc and P.I (artificial soil samples). 60

Relationship between swelling pressure and Cc (artificial soil


4.11 61
samples).

4.12 Relationship between Cc and M.D.D (artificial soil samples). 61

Relationship between k and sand content (artificial soil


4.13 64
samples).
Relationship between k and bentonite content (artificial soil
4.14 65
samples).

4.15 Relationship between k and P.I (artificial soil samples). 65

Relationship between sand content and Atterbergs limits


4.16 67
(natural soil samples).
VI
Relationship between Schreiners method and L.L from
4.17 70
laboratory testing (natural soil samples).
Relationship between Skemptons method and Cc from
4.18 72
laboratory testing (natural soil samples).
Relationship between M.D.D and swelling pressure (natural
4.19 73
soil samples).
Relationship between swelling pressure and L.L (natural soil
4.20 75
samples).
Relationship between swelling pressure and P.I (natural soil
4.21 76
samples).

4.22 Relationship between Cc and mV (natural soil samples). 76

Relationship between Cc and swelling pressure (natural soil


4.23 77
samples).

4.24 Relationship between Cc and M.D.D (natural soil samples). 77

Relationship between k and sand content (natural soil


4.25 80
samples).
Relationship between k and clay content< 2m (natural soil
4.26 80
samples).

4.27 Relationship between k and P.I (natural soil samples). 81

4.28 Relationship between k and M.D.D (natural soil samples). 81

Relationship between k and applied stress (natural soil


4.29 82
samples).
Plot of clay minerals on Casagrandes chart (natural soil
4.30 85
samples)

VII
List of Tables

Table Page
Name of Table
No. No.

2.1 Compaction characteristics of clayey sand soils. 29

2.2 Specification of clayey sand soil as embankment or fill material. 29

3.1 Physical Properties of research materials. 35

3.2 Index properties of the artificial soil samples. 37

Standard Proctor compaction tests results for artificial soil


3.3 38
samples.
Consolidation and free swelling tests results for artificial soil
3.4 40
samples.

3.5 Index properties for the natural soil samples. 41

Standard Proctor compaction tests results for natural soil


3.6 42
samples
Consolidation and free swelling tests results for natural soil
3.7 44
samples.
Classification and identification of expansiveness for the
4.1 49
artificial soil samples.
Predicted L.L values of artificial soils samples (according to
4.2 50
Schreiners method).
Predicted specific gravity values of artificial soil samples
4.3 52
(according to Estebans method).
Predicted (Cc) values of artificial soil samples (according to
4.4 54
Skemptons method).

4.5 Estimated permeability coefficients of artificial soil samples. 62

Classification and identification of expansiveness for the natural


4.6 68
soil samples.

VIII
Predicted L.L values of natural soils samples (according to
4.7 69
Schreiners method)
The compression index (Cc) of natural soil samples (according to
4.8 71
Skemptons method).

4.9 Estimated permeability coefficients of natural soil samples. 78

IX
CHAPTER ONE
INTRODUCTION

1.1 Introduction

Clayey sand (SC) soils are known to exist in different places in Sudan, especially
in the eastern and centre of Sudan and along the river Nile valley.

The Unified Soil Classification System (USCS) describes the (SC) soils varying to
great extent in the percentages of clay fraction and the consistency limit L.L, P.L
and P.I. Soils of sand content greater than 50% and percentage finer less than sieve
#200 less than 50% can have very high L.L and P.I.

Although soils are identified as clayey sand (SC) they can have an appreciable
volume change characteristics that will affect the safety of the foundations and the
floors of the buildings. Such type of soil is available in different areas in Sudan.
The specifications for fill material under the foundation and floor of building must
comply with the requirement of strength and settlement limits.

Expansive clay soil swells when wetted and shrinks when dry and usually results in
ground movement. Depending on its severity, swelling can cause structural
damage to low-rise buildings. Upon expansion, the soil exerts an upward pressure
on foundation and structures founded on it. If this pressure is greater than the
foundation pressure, then uplift or differential uplift can cause walls, beams, and
columns to crack (Basma et al., 1995). Assessing the swelling pressure and amount
of swelling is an important stage in designing foundations on expansive soils.

Several factors can influence the swelling potential of the clay soils; these factors
include the amount and the type of clay minerals, the cation exchange capacities of

1
clay minerals, availability of moisture, and the initial water content and other
factors related to the clay deposition history such as fabric and overburden
pressures. Among the three major structural groups of clay minerals, the smectite
group, consisting mainly of montmorillonites, constitutes the most highly
expansive soils.

Artificial and natural soil samples were used to clarify the effect of sand fraction
on the expansive behaviour of sand-clay mixture. The effect of increasing sand
fraction on soil was determined on different types of clay minerals, i.e. sodium
bentonite (pure montmorillonite) and highly expansive natural soil, and various
mixing ratios.

1.2 Thesis Organization


Chapter two of this thesis includes brief literature review of the expansive soils, its
mineralogy composition, identifications and mechanism of swelling. It also
contains the volume change behaviour of sand-bentonite mixtures. This chapter
also demonstrates the properties of clayey sand (SC) soils to be used as fill or
under foundation material. The historical background of the theories and the
assumptions used to stabilize the (SC) soils are also reviewed in this chapter.

Chapter three presents the results of experimental tests for the artificial and natural
soil samples. The results were presented in tabular forms or graphs drawn to show
the relationship between different parameters.

The analysis of the experimental and discussion of results obtained from the
experimental work for artificial and natural soil samples are included in Chapter
four.

Chapter five is a concluding chapter, presenting the conclusions drawn from the
different chapters, and also presents some recommendations for future research.

2
CHAPTER TWO
LITERATURE REVIEW

2.1 Introduction
This chapter aims to summarize the literature review pertaining to the subject of
this research. The literature review will cover the general properties of the
expansive soils, their properties and mineralogy composition of clay and the
mechanism of swelling.

The literature review shall also cover the bentonite properties and the volume
changes of sand-bentonite mixtures. A literature review of clayey sand (SC)
soils and their properties will be presented in this chapter.

2.2 Soil composition


2.2.1 Sand and silt fraction
Particle shapes of sand and silt particles can be described in terms of angularity
and roundness. Elongated and platy particles develop a preferred orientation
within the soil mass. This can result in anisotropic behavior of the soil
(Mitchell, 1993).

Sand and silt particles contain mainly primary minerals (most of them being
quartz and feldspars). Quartz has a very stable structure and high relative
hardness, which accounts for the persistence of quartz in soils (Mitchell, 1993).
Feldspars have an open structure with low bond strength between units.
Feldspars are therefore easily broken down. Due to the small amount of surface
area exposed per unit weight, their presence has little influence on the physical
properties of soils that are associated with surface phenomena (Mitchell, 1993).
Mineralogical analyses have shown that feldspars are not found in abundance in
fractions of soil particles smaller than 2 m (Baver et al, 1972).

4
2.2.2 Clay fraction
Clay is the common name for a number of fine-grained, earthy materials that
become plastic when wet. Chemically, clays are hydrous aluminium silicates,
usually containing minor amounts of impurities such as potassium, sodium,
calcium, magnesium, or iron (Mitchell, 1993).

Crystalline clay minerals may be divided into three main groups:


1. Kaolin group with a 1:1 lattice type. Characterized by having one silica and
one alumina sheet (see Fig. 2.2). Members are kaolinite, dicktite, nacrite and
halloysite. There is little isomorphous substitution in the crystal lattice of
these minerals. The unit layers are bonded together so tightly (through
hydrogen bonding) that ions or water molecules cannot permeate the
interlayer positions between adjacent kaolinite layers. This means that their
colloidal properties are determined by external surfaces only. There is little
swelling or shrinkage and low plasticity. The Cations Exchange Capacity
(CEC) is less than 10 meq/100 g of clay (Baver et al, 1972).

2. Hydrous mica group with a 2:1 lattice type. Consists of two silica sheets and
one of alumina. Two main members are illite and vermiculite. The major
substitutions of illite occur in the silica sheets (Al for Si). The negative
charges produced are balanced by potassium ions. The CEC varies between
20 and 40 meq per 100 g of clay. However this is one mineral in which the
CEC does not reflect the degree of isomorphous substitution in the crystal.
This is because the unit layers are held together by the potassium ions and its
exchange capacity resides on the external surfaces.

3. Smectite group: the structure of the smectite mineral group consists of an


octahedral sheet sandwiched between two silica sheets (see Fig. 2.3). The
crystal lattice expands or contracts depending on the amount of water and
interlayer cations present. The main members of the group are

5
montmorillonite, beidellite and nontronite. In montmorillonite there is some
substitution of Al for Si in the silica sheet and Fe and Mg for Al in the
alumina sheet. The negative charges generated are balanced by exchangeable
cations that are found between the unit layers. Bonding between successive
layers is weak and they are easily separated by cleavage or adsorption of
water or other polar liquids. Montmorillonite is the main component of
bentonite clayey materials.

2.2.2.1 Mineralogy of clay


Most clay is chemically and structurally analogous to other phyllosilicates but
contains varying amounts of water and allows more substitution of their cations.
Physical characteristics that define this group:
1) They can absorb water or lose water from simple humidity changes.

2) When water is absorbed, clays will often expand as the water fills the
spaces between the stacked silicate layers.

3) Due to the absorption of water, the specific gravity of clays is highly


variable and is lowered with increased water content.

4) Clays tend to be formed from weathering and secondary sedimentary


processes with only a few examples of clays forming in primary igneous
or metamorphic environments.

The silicon-oxygen tetrahedron (silica tetrahedron) consists of four oxygen


atoms nested around a silicon atom. (Fig. 2.1.a) Alumina octahedron consists of
an aluminium atom surrounded octahedrally by six oxygen ions. (Fig. 2.1.b)

6
Fig. 2.1: Clay Minerals: Basic Units (Craig, 1994).

The basic units combine to form sheet structures. Silica sheet is the combination
of silicon-oxygen tetrahedrons. Alumina sheet is formed by combination of
alumina octahedrons. Diagrammatic sketches of the kaolinite and
montmorillonite structures are shown on Figures 2.2 and 2.3.

Fig. 2.2: Diagrammatic Sketch of the Kaolinite Structure (Craig, 1994).

7
Fig. 2.3: Diagrammatic Sketch of the Montmorillonite Structure (Craig, 1994).

The various clay minerals are formed by the stacking of combinations of the
basic sheet structures with different forms of bonding between the combined
sheets.

Kaolinite consists of a structure based on a single sheet of silica tetrahedrons


combined with a single sheet of alumina octahedrons. There is very limited
isomorphism substitution. The combined silica-alumina sheets are held together
fairly tightly by hydrogen bonding: a kaolinite particle may consist of over one
hundred stacks (see Fig. 2.4).

8
Fig. 2.4: The Kaolinite Mineral (Cernica, 1995).

Illite has a basic structure consisting of a sheet of alumina octahedrons between


and combined with two sheets of silica tetrahedrons. The combined sheets are
linked together by fairly weak bonding due to potassium ions held between
them (see Fig. 2.5).

Fig. 2.5: The Illite Clay Mineral (Cernica, 1995).


Montmorillonite has the same basic structure as illite. In the octahedral sheet,
there is a partial substitution of aluminium by magnesium. Water molecules
occupy the space between the combined sheets and exchangeable metallic ions

9
(see Fig. 2.6). There is a weak bond between the combined sheets due to these
ions. Considerable swelling of montmorillonite can occur due to additional
water adsorbed between the combined sheets (Craig, 1994). The large swelling
capacity of montmorillonites, particularly sodium montmorillonites, marks these
minerals as the most troublesome ones with respect to engineering design and
construction (Popescu, 1986).

Fig. 2.6: The Montmorillonite Clay Minerals (Cernica, 1995).


2.2.2.2 Bentonite
Bentonite is part of the montmorillonite clay family, usually formed from the
weathering of volcanic ash. It is noted for its expansive properties in the
presence of water. The term bentonite is now well established, and used to
describe a clay material whose major mineralogical component is formed by the
smectite group and whose physical properties are dictated by the smectite
minerals (Grim and Guven, 1978).

As a result, bentonite is a highly expansive material. By mixing bentonite with


soils of high hydraulic conductivity, the coefficient of hydraulic conductivity
can be reduced to acceptable values for containment purposes.

The use of bentonite as a hydraulic containment can be put into these categories

10
(Gleason et al, 1997) as a soil additive for compacted soil barriers in landfill and
cover systems, as a low-hydraulic conductivity component in geosynthetic clay
liners, and in vertical cut off walls that are backfilled with mixtures of soil-
bentonite or cement-bentonite.

Most bentonites are either sodium or calcium bentonite. They are characterized
by the type of external cation (i.e. calcium or sodium) that is absorbed into the
surface of the clay particle during the mineral formation, or, for treated
bentonite, during the processing of the bentonite.

Sodium bentonite is used more extensively than calcium bentonite because of its
superior swelling capacity and its very low hydraulic conductivity to water.
Although calcium bentonite has a smaller swelling capacity and a higher
hydraulic conductivity to water than sodium bentonite, some researchers have
suggested that calcium bentonite may be more stable than sodium bentonite
when exposed to chemical constituents in permeating fluids (Gleason et
al,1997). Cernica (1995) reported that sodium bentonite has a liquid limit of
500% or more.

As such, it was found to have beneficial uses as a general grout in preventing


leakage from reservoirs, for plugging leaks in tunnel construction, and as a
drilling mud in connection with soil borings and oil and gas wells. It prevents
flocculation and facilitates the removal of the drill cuttings of the rotary drill. In
practice, the proportion of bentonite used in landfill liners varies from 5% to
13% (Chapuis, 1990), with the lower limit corresponding to well graded
material and the upper limit corresponding to uniform granular material.

2.2.2.3 Influence of the clay fraction


In a granular soil with bentonite, the greater the bentonite content, the greater in
volumetric changes with respect to changes in the water content (i.e. swelling
and shrinkage). In fully saturated conditions, as the bentonite content increases,

11
the soil compressibility increases and its hydraulic conductivity decreases.
There is also a reduction in the angle of shearing resistance.

Assuming that all the water in a soil is associated with the bentonite, the amount
of bentonite required to fill the voids of the granular soil to prevent contact
between the granular particles can be estimated as a function of the water
content. The mass and volume relationships of a fully saturated clayey soil are
shown in Fig. 2.7 as a function of the clay content. The clay content is defined
as a percentage of the solids mass.

Volume Mass

wWs/w WATER wWs

V
cWs/Gcw CLAY PHASE cWs

Ws
(1-c)Ws/Ggw
GRANULAR
(1-c)Ws
PHASE

Fig. 2.7: Phase relations of clay - enriched granular soil.

Where Ws = the mass of solids, Ww = the mass of water


Gc = specific gravity of clay, Gg = specific gravity of granular phase
w= density of water , c = clay content , w = water content

2.3 Mechanism of swelling


There are two basic mechanisms involved in swelling phenomena:
1. The swelling is chiefly a characteristic of the montmorillonite group of

minerals. The layers that make up the individual single crystals of


montmorillonite are weakly bonded, mainly by water in combination with

12
exchangeable cations. On wetting, water enters not only between the single
crystals, but also between the individual layers that make up the crystals
(Popescu, 1986).
2. In montmorillonites the interlayer cations become hydrated, and the large
hydration energy involved is able to overcome the attractive forces between
the unit layers. Since in the prototype minerals interlayer cations are absent,
there is no cation hydration energy available to separate the layers. (Cernica,
1995).

There can be two reasons of swelling: Clay particles are generally platelets
having negative charges on their surfaces and positively charged edges (Fig.
2.8). Cations in the soil water attach to the surfaces of the platelets and the
negative charges on the surfaces of clay particles. The unbalanced electrostatic
charges on clay-particle surfaces draw water molecules into the area between
silicate sheets and force the sheets apart.

Fig. 2.8: Internal Electrochemical System of Soil (Mitchell, 1976).

The other factor is provided by cations attracted to the clay surfaces. Due to the
attraction of negatively charged clay surfaces for the cations, the concentration
of cations between the clay-particle surfaces is higher than the concentration of
cations in the pore fluid. This creates an osmotic potential difference between

13
the pore fluid and clay-mineral surfaces. In the actual case cations should
migrate from the (higher potential) to the (lower potential) to equalize the cation
concentration. But due to the attraction of clay surfaces, cations cant move and
water moves into the area between clay-mineral surfaces. Due to this condition
a repulsive force is exerted on the clay-mineral surfaces and the volume of clay
soil increases.

Generally, the swelling of expansive clays is affected by a large number of


factors due to its complex mechanism. The behavior of geological medium
depends on number of factors such as density, stresses, stress history,
experience of discontinuities, temperature, time, and existence of liquids and
gases in the pores etc. The value of density affected the initial value of heave
(elastic heave) as well as the rate of heave of density (Elarabi, 2004).

2.4 Expansive soil


Expansive clays are confined to semi - arid regions of the tropical and temperate
climate zones. Naturally occurring expansive soils are found to exist in partially
saturated state. Seasonal rainfall especially in areas where evapo-transpiration
exceeds precipitation, fluctuations of the ground water table leakage of sewers
and water supply systems are all contributing to the problem associated with
expansive clays (El turabi, 1985).

Expansive soils prevail over large areas of Sudan and have caused significant
damages to irrigation systems, water lines, sewer lines and other structures
located on these soils. Damages caused by expansive soils before 1984 exceeded
6 million USD per year (Osman, 1984).

Expansive soil is a term generally applied to any soil or rock material that has a
potential for shrinking or swelling under changing moisture conditions .The term
expansive soil will generally be used in a universal sense to refer to soils that both
shrink and swell on drying and wetting (Nelson 1992). Expansive soil is a soil

14
that under some condition is capable of increasing its volume when wetted,
normally, a soil containing expansive clay mineral.

Expansive clays have the ability to generate tremendous pressure on structures


such as concrete foundations. These high pressures are the key to the destructive
power of expansive clay in creating foundation problems. Some reports and
investigations noted that these pressures can reach over 700kN per meter
square.

An over consolidated soil is more expansive than the same soil at the same void
ratio, but amount of swell under light loading has been shown to be unaffected
by aging (Mitchell, (1976); Kassiff and Baker, (1971)).

Expansive soil will also exert pressure on the vertical face of foundation,
basement or retaining wall resulting in lateral movement. Shrink-swell soils
which have expanded due to high ground moisture experience a loss of soil
strength or capacity and the resulting instability can result in various forms of
foundation problems and slope failure. Expansive soil should always be a
suspect when there is evidence of active foundation movement. Plastic clay
termed as expansive soils or active soils exhibit volume change when subjected
to moisture variations. In addition, the engineering properties of expansive soil
are affected due to the activity of the clay minerals present.

In order for expansive soil to cause foundation problems, there must be


fluctuations in the amount of moisture contained in the foundation soils. If the
moisture content of the foundation soils can be stabilized, foundation problems
can often be avoided.

2.5 Identification of Expansive Soils


Identification of potentially swelling or shrinking soils is important tool for
selection of appropriate foundation (Hamilton, 1977). Despite the lack of standard

15
definition of swell potential (Nelson and Miller, 1992), there are various
geotechnical techniques to identify the swelling potential of soils. Surface
examination, geological and geomorphologic description can give indicators of
expansive soils. The morphological description includes a list of many things
such as ground water table level, color of the soil, soil groups... etc.

The indicators of swell potential are obtained by performing geotechnical index


property tests such as Atterberg limits and grain size analysis. Other tests (direct
tests) to determine the swell potential include volume change tests (free swell and
swell in oedometer tests). The direct methods consist essentially of laboratory
swelling tests while indirect methods are based on the correlation of certain
physicochemical properties and mineralogical composition of the soils with
swelling or shrinkage parameter. Empirically correlating soil parameters and
placement conditions to its expansiveness is indirect method while tests like
oedometer swell tests are direct methods.

2.6 Volume change of sand-bentonite mixtures


Sand-bentonite mixtures are commonly used as engineering barriers which are
used in the containment of hazardous materials. The addition of relatively small
amount of bentonite (5-15%) can improve the performance of granular material
providing both low permeability and enhanced mechanical stable material (Julio
Esteban, 2002). Changes in water content after compaction can cause swelling
or collapse depending on the stress and suction conditions before the water
content change occurs.

One material, which can meet the hydraulic conductivity criteria without
suffering from volumetric changes following the addition of water, is sand-
bentonite mixture (Mollins et al, 1996). The sand limits the shrinkage when
drying and below limiting clay content, the sand particles are in contact
providing good mechanical stability. With respect to the clay content there are

16
two general states. The first state is, when the clay content is too high and the
sand particles are not in contact. This material is very likely to change its
volume with changes in water content (i.e. swelling when the water content
increases and shrinking when the water content reduces). The second general
state is when the clay content is below certain limiting value and the sand
particles are in contact, thereby providing mechanical stability and preventing
shrinkage. When wet, the clay fills the sand voids producing a very low
hydraulic conductivity for the compacted barrier.

The behaviour of compacted bentonite can be explained by the following simple


idealized mechanism. At the commencement of hydration the bentonite starts to
expand into the existing voids as it tries to overcome the confining pressure. As
the clay continues to swell the voids become filled and overall expansion takes
place at a reducing rate with time.

Conversely, at low initial dry densities, the volume of voids is relatively high
and the permeability of the clay is also relatively high. For high density
compacted bentonite it takes longer time to absorb water and therefore the
initial rate of swelling strain is smaller. Komine and Ogata (1994) reported that
at higher confining pressures this difference in rate between low and high
density samples is not so apparent.

When compacted at higher water contents, the early rate of swelling


deformation for the high initial dry density is faster than for the low initial dry
density. For low vertical stresses the rate of swelling deformation is higher than
that of the samples compacted at low initial water content.

For the low initial water content, at low dry densities the void volume of the
compacted bentonite is large and it takes some time for the swelling clay
particles to fill up the voids. Munthohar and Hashim (2002) reported that the

17
clay fraction in sand-bentonite mixtures increased almost linearly with
increasing bentonite content.

From the observation of the swelling behaviour of compacted mixtures of sand-


bentonite, 50% bentonite content, for use in radioactive waste containment,
Komine and Ogata (1999) have proposed the following mechanism for the
development of swelling pressure and swelling deformation of sand-bentonite
mixtures:

1. The volume of the bentonite in the mixture increases by absorbing water,


and the voids in the mixture are filled with this additional volume.

2. If the mixture can swell as in the oedometer test at constant vertical stress,
the volume of the mixture increases as the volume of the bentonite increases,
until the swelling pressure of the bentonite is equal to the vertical stress.

3. If the volume of the mixture is restricted as in a constant volume test, the


bentonite will swell filling up the voids (at constant total volume of the
mixture). After all the voids are filled, the bentonite cannot change volume
anymore and the pressure caused by the bentonite will correspond to the
swelling pressure of the mixture.

Similar behaviour is observed with dense samples of compacted sand-bentonite


mixtures which show a small amount of swelling after flooding following
oedometer testing under constant vertical stress. Complete or partial void filling
will depend on the tortuosity and continuity of the pore structure of the soil.

2.7 Stabilization of the Clayey Sand (SC) soil


Clayey Sand soils are which have low plasticity, low compressibility, and high
strength under loads, are suitable for use as base material for any engineering
construction projects as well as for roads and building construction. Decrease of
plasticity and compressibility and increase in strength of these (SC) soils can be

18
obtained by many different methods.

Increasing the sand fraction can improve the (SC) soil properties. It fills the
voids, increases the strength, and decreases the plasticity of soil.

One of various methods of increasing the strength of (SC) soils is lime


stabilization. It is a common, applicable and easy-to-use approach that can
improve geotechnical properties of clayey sand soils. Lime stabilization usually
used in highway pavements. Some literature review will be presented to show
the previous approach and studies in this type of stabilizer with clayey sand
soil.

2.7.1 Lime Stabilization (General Review)


It is an age-old practice to use lime in one form or the other to improve the
engineering behavior of clayey soils. Because of the proven success of lime
stabilization in the field of highways and air-field pavements, this technique is
now being extended for deep in-situ treatment of clayey soils to improve their
strength and reduce compressibility. The improvements in the properties of soil
are attributed to the soil-lime reactions (Clare and Cruchley, 1957; Ormsby and
Kinter, 1973; Locat et al. 1990).

Lime stabilization is covered extensively in the literature (Rogers and


Glendinning, 2000; Little et al. 1987; National Lime Association (NLA), 2004;
Thompson, 1966). Lime will primarily react with medium, moderately fine, and
fine-grained soils to produce decreased elasticity, increased workability,
reduced swell, and increased strength. Such improved soil properties are the
result of three basic chemical reactions, Cation exchange and flocculation-
agglomeration, Cementation (pozzolanic reaction) and Carbonation.

The cation exchange process involves an agglomeration of the fine clay


particles into coarse particles. The cementation process develops from the
reaction between calcium present in lime and silica and alumina in the soil,

19
forming calcium-silicate and calcium-aluminate or calcium-aluminate-silicates.
The cementitious compounds produced are characterized by their high strength
and low-volume change. Previous researchers reported that small lime additions
(from 2% to 8%) significantly decrease the liquid limit, plasticity index,
maximum dry density, and swell, and increase plastic limit, the optimum
moisture content, and strength of expansive soils (Abduljauwad, 1995; Basma
et al., 1995). It was reported by Sivapullaiah et al., (1995) that lime added in
excess of the amount required for cation exchange could only produce
cementitious compounds, which blind the flocculated particles and develop
extra strength (Al-Rawas et al., 2002).The most commonly used products are
hydrated high calcium lime Ca (OH) 2, MgO, calcitic quick lime CaO, and
dolomitic quick lime CaO.MgO. Quick lime is used widely for soil
stabilization.

Hydrated lime is a fine powder, whereas quicklime is a more granular


substance. Quick lime is more caustic than hydrated lime, so additional safety
procedures are required with this material. The type of the lime used as a
stabilizing agent varies from country to country. Although using quick lime is
more popular in Europe, hydrated lime is used mainly for stabilization but
proportion of quick lime that is used increased to about 25% in 1987 from about
15% in 1976 (Rollings, 1996).

2.7.2 Stabilization of clayey sand soil using fly ash mixed with small amount
of lime (by Thaweesak J., Supakij N. and Korchoke C. (2003))
In studying to use fly ash mixed with small amount of lime to improve
properties of soil. The soil represents the lowest permissible clay fraction in this
category. Results indicate that (SC) soil has satisfied required physical
properties for geotechnical construction such as bottom liner for landfill, while
some engineering properties must be improved such as shear strength and
hydraulic conductivity.

20
A proportion of lime to fly ash of (1:20) was selected based on compaction
results giving the highest maximum dry density (M.D.D). (SC) soil specimens
were then mixed with this proportion at stabilizer contents of 3, 6 and 10
percent by dry weight.

The samples were prepared by standard compaction with varying water content
from optimum moisture content (O.M.C) to 2.0% wet of optimum
(O.M.C+2%). Hydraulic conductivity tests were performed by constant head
method after curing time of 14 and 28 days. Strength test samples were cured
for 7, 14, 28 and 90 days and divided into 2 groups (unsoaked and soaked
samples) before performing by unconfined compression tests.

The conclusions are that, the proportion 1:20 of lime to fly ash was selected as a
mixed proportion giving the highest (M.D.D). In addition, if different sizes of
mold were used with equivalent compaction energy applied, similar relationship
of compaction characteristics was also observed.

Secondly, Hydraulic conductivity of improved (SC) soil markedly decreased


with an increase of stabilizer content while curing period had a little effect to
reduce their hydraulic conductivity. At the same content of stabilizers, results
also show that soil compacted at OMC gave lower hydraulic conductivity when
compared to that of compacted on 2 percent wet of optimum (OMC+2%).

Basically, for both soaked and unsoaked samples, strength significantly


increases as stabilizer content increases. Rapid development of strength of
improved (SC) soil was observed at early curing period perhaps due to
hydration reaction among fly ash, lime and water. For long-term strength, it was
attributed that pozzolanic reaction would perform to develop their strength
continuously.

21
2.7.3 Strength behavior of stabilized clayey sands using lime-polyamide
fiber (by Mahyar Arabani, Akbar K. Haghi and Mehdi Veis Karami
(2005))
This study presents experimental investigations for strength improvement of
clayey sands obtained from northern Iran and reinforced with lime-polyamide
fiber. This study presents the effect of polyamide strips on engineering
properties of lime stabilized clayey sand soil. In this method, Polyamide 66
strips were placed between compacted layers. The results of reinforced and
unreinforced samples are compared with several lime-stabilized soils among the
middle-east countries. The comparison shows the superiority of this new
composite in improving the strength.

The study was done in Guilan in north of Iran. Guilan has a humid and rainy
climate placed near Caspian Sea in the north of Iran. Most of soil layers were
formed by the sediments of Quaternary, transported by Sefid-Roud River,
which is the biggest river in north of Iran. In Guilan, the groundwater table is
high and the natural moisture content ranged between 11-56%. The general soil
texture in this area consists of sand and clay. Kaolinite and Illite clay are the
general clay minerals. Because of this, lime-stabilization and cement-
stabilization are the most common techniques for soil improvement in this
province. However, the benefits of polyamide additions did not reported
previously and can be included as the primary ingredient necessary for
stabilizing the clayey sand. In this study, a good adhesion between the
component of cements and polyamide obtained.

The fine content of mixes, provided from dominant clay of the area, were 5%,
15%, 22%, 30% and 36% in order to stabilize the samples. Water content of
specimens had been determined by modified proctor compaction test according
to AASHTO guideline. The samples were prepared for further testing
procedures reinforced with polyamide 66.

22
All samples were then placed compacted in 5 layers in standard uniaxial test and
indirect tensile test cylindrical molds and compacted to reach to relative density
of 90 to 95 %. Polyamide 66 strips were placed between any these compacted
layers.

Fig. 2.9: Optical micrograph of a cut-polyamide strip.

After 24 hours curing in laboratory open condition, at 23C, the samples were
taken out from molds and placed in an oven in constant temperature at 45C, in
order to increase the rate of curing, in humid condition to allow chemical
reactions to take place for a 7-day strength (Clough, G.W., N. Sitar, R.C.
Bachus and N. S. Rad; 1981). While a curing of 48 hours was taken, samples
were tested. Three samples were prepared and cured for each test in this
manner. The tests were performed on specimens just 15 minutes after they were
taken out from the oven. A capping procedure with chemical materials were
performed on upper and lower surfaces of specimens to make these surfaces
slick and obtain a uniform stress distribution on surfaces and body of the
samples. The tests were performed according to I. R. Iran Code for highway
design based on AASHTO T 220.

23
Brazilian tests were performed to obtain the tensile strength of specimens, based
on ASTM C496 for indirect tensile test. The size and curing time of both tensile
and compressive samples were similar.

In Figures 2.10 to 2.12 the results compare the compressive strength of


polyamide 66 reinforced and non-reinforced lime-stabilized materials versus
different clay contents stabilized with 3%, 6% and 9% lime content respectively.
Based on the variation of compressive strength it is clear that the higher the clay
content, the higher the compressive strength. Meanwhile, the compressive
strength of specimens reinforced with polyamide 66 is higher than the others.
Compressive strength can even be raised up to 25% when polyamide 66 strips as
a reinforcing agent is used.

Fig. 2.10: Comparison between compressive strength of reinforced and


non-reinforced samples vs. clay content for 3% lime content.

24
Fig. 2.11: Comparison between compressive strength of reinforced and
non-reinforced samples vs. clay content for 6% lime content.

Fig. 2.12: Comparison between compressive strength of reinforced and


non-reinforced samples vs. clay content for 9% lime content.

The results of this study lead to the following conclusions, stripping of


polyamide can improve the cohesion properties, and so it is possible to increase
the bearing capacity of foundations.

The cracking patterns observed were minimal, and then the stiffness of the
mixture is significantly improved to use these materials in pavement layers (i.e.
decrease the deflection and increase the strengths under traffic loads).

25
The results indicate that the lime/polyamide can be a good candidate for the
stabilization of clayey sands as base, sub-base, stabilized sub-grade for highways
and airport pavement as well as for any construction practices.

The variation of stress-strain in the mix design indicates that the lime/polymeric
stabilizer can be used in heavy-duty pavements.

Stability of slopes can be achieved using such materials as clayey sand in natural
and/or artificial soils.

2.8 Clayey sand (SC) soils and their properties


2.8.1 Compression behavior of Clayey Sand (SC) soils
Clayey sand soils can be considered as a composite matrix of coarse and fine
grains. The interaction between coarser and finer grain matrices affects the
overall stressstrain behavior of these soils (Monkula, Ozdenb, 2006). Soils that
contain platy particles are more compressible than those composed entirely of
bulky grains. A fine sand or silt that contains mica flakes maybe quite
compressible. There are other factors that can affect in soil density like the
temperature, humidity and depositional characteristics (United States (US)
Army 1992 - Field Manual FM 5-410). Fine content may affect the compression
characteristics of coarse grained soils as well. In some models proposed for
compression behavior of cohesionless soils such as those by Hardin (1987) and
Pestana and Whittle (1995); effects of initial void ratio, relative density, particle
shape, mineralogy, structure and applied stress conditions were mentioned.
These factors were also prominent in the experimental researches related to the
compression of sands (Yamamuro et al. 1996, Chuhan et al. 2003).

Skempton (1985) indicated for the clayey soils that if the clay fraction is less
than about 25%, the soil behaves much like a sand or silt, whereas residual
strength is controlled almost entirely by sliding friction of the clay minerals
when the fraction is above 50%. Georgiannou et al. (1990) concluded that up to

26
a fraction of 20%, clay does not significantly reduce the angle of shearing
resistance of the granular component.

The compressibility of a loose sand deposit is much greater than that of the
same sand in a relatively dense condition. Generally, structures should not be
located on loose sand deposits. Some cohesionless soils, including certain very
fine sands and silts, have loose structures with medium settlement
characteristics.

2.8.2 Compaction of clayey sand (SC) soils


The achievement of proper soil is related by durability and stability of a
structure. Structural failure of roads and airfields and the damage caused by
foundation settlement can often be traced back to the failure to achieve proper
soil compaction.

Large scale experiments have done by the (US Army Corp of Engineers) and
indicated that the unconfined compressive strength of a clayey sand (SC) soil
could be doubled by compaction, within the range of practical field compaction
procedures. Some cohesionless soils including certain very fine sands and silts
have loose structures with medium settlement characteristics.

Sorochan (1991) had carried out several experiments on disturbed and


undisturbed soil with different values of densities and conclude that the swelling
increases when the density of soil is increased. This happened due to the
increase in number of particles and consequently the total surface of the solid
phase.

However, when some silts and very fine sands {predominantly (ML) and (SC)
soils} are compacted in the presence of a high water table, they will pump water
to the surface and become quick, resulting in a loss of shearing strength.
These soils cannot be properly compacted unless they are dried.

27
(US Army Corp of Engineers) in their high career, showed that if these soils can
be compacted at the proper moisture content, their shearing resistance is
reasonably high.

Every effort should be made to lower the water table to reduce the potential of
having too much water present. If trouble occurs with these soils in localized
areas, the soils can be removed and replaced with more suitable ones. If
removal, or drainage and later drying, cannot be accomplished, these soils
should not be disturbed by attempting to compact them. Instead, they should be
left in their natural state and additional cover material used to prevent the layers
from overstressed.

2.8.3 Clayey sand (SC) soils as fill and embankment materials


Some backfill materials (e.g. clean sands or gravels and crushed rock)
compaction can increase their internal friction angel. In practice, the coarse
granular soils or oversize materials (i.e. less than 100mm) can be used for the
construction of embankment bases for their ability to facilitate drainage and
prevent saturation.

The manual of US Army Corp of Engineers suggested that the selected fill
material of non-expansive sandy clay or clayey sand with P.I of 5% to 15% and
L.L of 35% or less. Generally, soils with L.L < 35% and P.I less than 12% have
no potential.

Table 2.1 presents compaction requirements and values of unit dry weight of the
(SC) soils and Table 2.2 shows the characteristics of (SC) soil as fill or
embankment material.

When a granular soil which contains little amount of fine content and poorly
drained, such clayey sand soil, it can be used for backfill behind retaining walls,
but they must be protected against frost action and may require elaborate

28
drainage provisions. As known, ideal backfill materials are purely granular soils
containing < 5% of fines. (See Table 2.2)

The characteristics of (SC) soils are so similar to (CL) and (ML) soils in
compaction equipment and values of maximum unit dry weight, permeability,
and foundation which make the (SC) soil undesirable as under foundation
material or fill / embankment material.

Table 2.1 Compaction characteristics of clayey sand soils


Standard
Sub-
Compaction Unit AASHTO
CBR Value grade
characteristics dry Max. Unit Drainage
design as sub- modulus
and weight dry characteristic
values 3 grade (k)
equipment (g/cm ) weight
(g/cm3)
(g/cm3)
Fair, Rubber-
Poor to
tired roller, Poor to
5-20 1.6-2.16 1.68-2.0 1.6-4.81 practically
sheep foots fair
impervious
roller

Table 2.2 Specification of clayey sand soil as embankment or fill material


Frost Requirements
Value for Value for Compressibility Permeability
action for seepage
foundations embankment and expansion (cm/sec)
potential control
Fairly stable,
Good to may be used
poor for Slight to Slight to
10-3 10-8 None
bearing impervious medium high
value core for flood
structures

29
2.9 Conclusions
Soil is composite material of sand, silt and clay fractions. Sand and silt are
containing in most quartz and some feldspars. Clay is hydrous aluminium
silicates and containing minor amounts of potassium, sodium, calcium,
magnesium, or iron.
Three main groups were divided for crystalline clay minerals, namely, kaolin
group, hydrous mica group, and smectite group. Montmorillonite is one of the
main members of smectite group. In smectite group, its so easy to separate the
bonding between layers by water or any polar liquids.

Bentonite is one of montmorillonite clay family and classified as highly


expansive material when water presence. Sodium bentonite is commonly
famous than calcium bentonite for its low hydraulic conductivity and the high
tendency for swelling. Some benefits can get from use of bentonite such as
preventing leakage from reservoirs, as drilling material in boreholes, and as
barrier material in waste deposits facilities.

When the bentonite content increased in granular soil, the volumetric changes
will increase. There is a reduction in angle of friction in fully saturated soil
when bentonite content increased. Besides that, increasing in compressibility
and decreasing in hydraulic conductivity will occurred. Generally, when a given
amount of clay tend to expand (increase in volume) as it absorbs water and
shrink (lessen in volume) as water is drawn away, it is called (expansive clay).

There are many methods to predict the swell potential of clays therefore to
select the tolerable type of foundation under expansive soils. Swell potential
classification can be obtained by Index tests (grain size analysis and Atterbergs
limits) beside the direct methods (free swell and swell in odeometer tests) and
indirect methods (which depended on physiochemical properties and
mineralogy composition of soil).

30
The sand-bentonite mixture is a suitable material for facing the volume changes
due to moisture contents variation and preventing shrinkage when bentonite
content providing certain percentage (5% to 15 % bentonite). If the content of
bentonite is high and sand particles are not in contact, the mixture will have
tendency to change its volume at water content variation.

In sand-bentonite mixtures the bentonite, at moisture increasing, begins to


expand into the existing voids as it tries to overcome the confining pressure and
continues to heave until the voids filled. The expansion rate of the mixture
reduces with time.

Sand-bentonite mixtures have high volume of voids and permeability at low


initial dry densities. At high densities, the absorption is taking time and
therefore the swelling rate is longer and the swelling strain is smaller.

In sand-bentonite mixtures, at higher water contents, fast swelling rate


deformation occurring to mixtures for high initial dry density and its slow in
low initial dry density.

The common stabilizer used for (SC) soils is lime. This is because its easiness
in use and its improvement for soil properties. Little amount of lime (2% to 8%)
can improve the soil properties. The using of fly ash with small amount of lime
giving high value of (M.D.D), increasing the strength and decreasing the
hydraulic conductivity of the (SC) soils. The use of lime-polyamide fiber
showed that it is improving the cohesion properties of the (SC) soil which make
it suitable as construction material as under foundation material or highway
material as base, sub-base, and stabilized sub-grade under heavy load
pavements.

Clayey sand soils can be considered as composite matrix of coarse and fine
grains and the interaction between these matrices can affect the overall stress
strain behavior of these soils. Fine content may affect the compression

31
characteristics of coarse grained soils as well. Many experimental researches
and models were proposed for the compression behavior of cohesionless soils
and some factors such as initial void ratio, relative density, particle shape,
mineralogy, structure and applied stress conditions were noted to illumine this
behavior.

Many authors reported that when the clay fraction is less than 25-30%, the soil
tends to be sand or silt soils; other wise the soil behaves much like clay.

Clayey sand soil is poor material in pavements if not compacted at the proper
moisture content, in this case its shear resistance can be low. The unconfined
compressive strength of clayey sand soil can be doubled by compaction. The
compressibility of the cohesionless soil is affected by grading and grain shape.
After large experiments done by Sorochan (1991), it was found that the swelling
increases when the density of soil is increased.

Clayey sand soil can be used behind retaining walls, when the protection against
frost action takes place. However, the ideal backfill material should contain less
than 5% of fines. When the L.L of soil is less than 35% and P.I is less than 12%,
no soil swelling occurs.

Clayey sand soils (SC) are similar to (CL) and (ML) soils in its physical
properties (i.e. compaction and permeability). Therefore (SC) soil is undesirable
as fill or under foundation material, but it is suitable in roads and highways as a
base material when frost action protection is considered.

32
CHAPTER THREE
EXPERIMENTAL WORK AND RESULTS

3.1 Introduction
This chapter presents the results and describes the procedures followed for the
soil sample preparation and experimental techniques adopted. Several tests were
carried out to classify and evaluate the geotechnical properties of the clayey
sand (SC) soil. The tests were carried out according to the British standards (BS
1337 - 1990) and the classification of the soil samples results were established
in accordance to Unified Soil Classification System (USCS).

Natural soil samples selected for research was obtained from (Soba) district on
the south east of Khartoum town. Two test pits were excavated about 1.5m
depth. The disturbed soil samples were obtained at 0.5m depth intervals and
taken in plastic bags. The field moisture content and field density were recorded
every 0.5m and established using the standard requirements methods.

3.2 Testing program

The laboratory testing programs were realized in this research. The first testing
program was comprised of the artificial soil samples which were made by
combining sodium bentonite with different sand contents. The second testing
program was designed for testing natural clayey soils taken from (Soba) region.
The natural soil samples were subjected to series of tests to cover the range of
sand percentages added to the relatively high plastic clay soil.

The following laboratory tests were carried out including:

1. Grain size analysis.

2. Atterbergs limits.

3. Specific gravity test.

34
4. Proctor compaction tests.

5. Consolidation and Swelling pressure tests.

All the tests were performed in the laboratories of the Building and Road
Research Institute (BRRI), University of Khartoum in accordance with the
British standards.

3.3 Preparing of soil samples


3.3.1 Artificial soil samples

The artificial soils which used in this research consist of clay of sodium
bentonite type and fine sand. These soils were mixed to produce soil samples
with different ratios. Percentage of bentonite by weight was varied as follows
10%, 15%, 20%, 25%, and 30%. The physical properties of sand-bentonite
mixtures are shown in Table 3.1 and Fig. 3.1.

The sand is composed of 4.0% coarse, 62.0 %medium, and 33.0% fine size.
Uniformly coefficient (Cu) and coefficient of curvature (Cc) were 2.5 and 0.9
respectively. The density of sand was obtained by placing known weight of
sand in a mould having a known volume.

The sand was classified as poor graded (see Fig. 3.1). The fine sand was washed
and passed through #40 sieve (0.425 mm) and retained in #200 sieve (0.075
mm).

Table 3.1 Physical Properties of research materials

L.L P.L P.I


Soil description S.G
(%) (%) (%)
Fine sand 2.59 Non plastic
Sodium bentonite 2.86 395 45 350

35
Fig. 3.1: Particle size distribution for the fine sand used in the artificial soils
samples.

3.3.2 Natural soil samples

The natural soils were taken from (Soba) district in the vicinity of Medical
Science College in Khartoum. The collection of samples was made during
November 2007. The samples were taken from trial pits of 1.5m depths.
Disturbed samples were obtained at 0.5m depth intervals. The first layer (i.e.
from 0.0m to 0.5m depth) classified as sandy silt (ML) soil and as silty clay of
high plasticity (CH) for the second and third layers.

The soil samples have relatively high field water contents (field moisture
contents ranged from 17% to 23%). The field water content was noticed to
increase with depth. The samples for testing program were taken from 1.0m to
1.5m depth (which gave the highest value of L.L and P.I).

36
The natural soils were mixed with fine sand in percentage of the total weight of
mixture. Sand percentages varied from 17 % (as origin sample), 25%, 40%,
55%, 60%, and 65% from the total weight. The description of all tests and their
results will be presented in this chapter.

3.4 Artificial soil samples


3.4.1 Classification tests

The wet sieving method (BS 1377: Part 2: 1990: 9.2) was used for the sand.
This was adopted to make sure that the sand is clean from any fine material.
British standard procedures were followed to determine the liquid and plastic
limits. The liquid limit was established by the cone penetrometer method (BS
1377: Part 2: 1990:4.3) and the determination of the plastic limit followed the
standard method (BS 1377: Part 2: 1990:5.3). (See Table 3.2)

Standard procedures were used to obtain the specific gravity of different ratios
of the sand-bentonite mixtures. The procedures were in accordance with the
(BS 1377: Part 2: 1990: 8.3)

Table 3.2 Index properties of the artificial soil samples


Soil description
Soil L.L P.L P.I L.S
(Bentonite content) Activity S.G
Code (%) (%) (%) (%)
(%)
A1 30 115 21 94 19.3 3.13 2.99
A2 25 95 17 78 15.7 3.12 2.95
A3 20 78 20 58 12.4 3.0 2.83
A4 15 64 19 45 8.6 2.9 2.77
A5 10 46 21 25 5.0 2.5 2.63

Note:
L.L: Liquid Limit, P.L: Plastic Limit, P.I: Plasticity Index,
L.S: Linear Shrinkage, S.G: Specific Gravity

37
3.4.2 Moisture-density relationship test (Proctor compaction)
Standard proctor compaction test was carried out on five samples of the sand-
bentonite mixture in order to obtain their moisture-density relationship (i.e. their
optimum moisture contents (O.M.C) and maximum dry densities (M.D.D)). The
compaction tests were carried out according to British standard (BS 1377: Part
4: 1990). (See Table 3.3 and Fig. 3.2)

Fig. 3.2: Moisture-density relationships for the artificial soil samples.

Table 3.3 Standard Proctor compaction tests results for artificial soil samples
Soil description
O.M.C M.D.D
Soil Code (Bentonite content)
(%) (g/cm3)
(%)
A1 30 18.4 1.670
A2 25 10.3 1.691
A3 20 11.8 1.692
A4 15 7.0 1.748
A5 10 10.5 1.762

38
3.4.3 Consolidation tests
The test was made according to British standard (BS 1377: Part 4: 1990). The
soil specimens were prepared at O.M.C which obtained by the compaction test
(see Table 3.3), cured for 24 hours, placed in standard ring and carefully placed
in the oedometer with porous stones on each face. The oedometer was placed
into the loading device and the dial gage was attached. The consolidation results
are showed in Table 3.4 and Fig. 3.3.
The coefficient of volume compressibility (mv) and the compression index (Cc)
were calculated for the artificial and natural samples by the same stresses which
appointed in the primary consolidation curve (see Table 3.4 & 3.7, Fig. 3.3 and
Fig. 3.6).

Fig. 3.3: Consolidation curve results (artificial soil samples)


39
Table 3.4 Consolidation and swelling tests results for artificial soil samples
Initial
Free Swelling Swelling
Soil I.M.C F.M.C void mv Pc
Swelling pressure potential Cc Cr
Code (%) (%) ratio (m2/kN) (kN/m2)
(%) (kN/m2) (%)
(eo)

A1 20.4 36.1 300 140 0.48 0.81 0.296 0.079 1.39E-4 235

A2 16.0 26.7 210 120 0.33 0.72 0.158 0.069 7.47E-5 177

A3 11.8 21.6 190 113.8 0.24 0.70 0.188 0.023 5.07E-5 225

A4 9.3 19.0 130 105 0.14 0.62 0.110 0.012 1.83E-5 148

A5 23.9 18.1 130 40 0.06 0.54 0.066 0.018 2.03E-5 53

3.5 Natural soil samples


3.5.1 Classification tests

The wet sieving method (BS 1377: Part 2: 1990: 9.2) and the sedimentation by
the hydrometer method (BS 1377: Part 2: 1990: 9.5) were done for the natural
soil samples. (See Fig. 3.4)

The liquid limit was established by the cone penetrometer method (BS 1377:
Part 2: 1990:4.3) and the plastic limit following the standard method (BS 1377:
Part 2: 1990:5.3). (See Table 3.5)

Specific gravity test performed according to the (BS 1377: Part 2: 1990: 8.3).
(See Table 3.5)

40
Fig. 3.4: Grain size distribution curves for the natural soils samples

Table 3.5 Index properties for the natural soil samples


Clay
Soil Sand content L.L P.L P.I L.S fraction
Activity S.G
Code (%) (%) (%) (%) (%) (< 2m)
(%)
N1 17 78 23 54 14.7 53 1.02 2.84
N2 25 77 25 52 13.6 40 1.30 2.81
N3 40 66 19 40 10.6 28 1.43 2.77
N4 55 52 18 34 7.9 16 2.13 2.75
N5 60 37 14 23 7.1 13 1.77 2.72
N6 65 35 14 21 6.2 11 1.91 2.68

41
3.5.2 Moisture-density relationship test (Proctor compaction)
Standard proctor compaction test was carried out on the six samples in
accordance with the British standard (BS 1377: Part 4: 1990). The results are
shown in Table 3.6 and Fig. 3.5.

Fig. 3.5: Moisture-density relationships for the natural soil samples

Table 3.6 Standard Proctor compaction tests results for natural soil samples

O.M.C M.D.D
Soil Code
(%) (g/cm3)
N1 29.3 1.471
N2 21.9 1.650
N3 18.0 1.728
N4 16.7 1.790
N5 14.0 1.820
N6 13.0 1.810

42
3.5.3 Consolidation tests
The disturbed specimens were prepared at the O.M.C. The specimens were
cured for 24 hours, placed in standard ring, carefully loaded in the oedometer
with saturated porous stones on each face. The procedure of the test is according
to British standard (BS 1377: Part 4: 1990). The results and curves of
consolidation test are shown in Table 3.7 and Fig. 3.6.

Fig. 3.6: Consolidation curve results (natural soil samples)

43
Table 3.7 Consolidation and free swelling tests results for natural soil samples.

Initial
Free Swelling Swelling
Soil I.M.C F.M.C void mv Pc
Swelling pressure potential Cc Cr
Code (%) (%) ratio (m2/kN) (kN/m2)
(%) (kN/m2) (%)
(eo)
N1 27.6 33.3 520 320 0.35 0.89 0.517 0.154 1.28E-4 180
N2 21.4 29.8 490 315 0.32 0.67 0.309 0.079 4.97E-5 390
N3 13.7 22.5 460 160 0.27 0.63 0.282 0.033 8.13E-5 149
N4 16.5 19.9 340 80 0.18 0.55 0.210 0.052 6.36E-5 144.5
N5 10.9 17.7 250 65 0.16 0.55 0.199 0.024 6.06E-5 106
N6 11.8 14.7 200 30 0.14 0.47 0.154 0.025 4.93E-5 60

3.6 Conclusions
Relatively high plastic clay soils were obtained from (Soba) district on the south
east of Khartoum town. Tests pits were excavated about 1.5m depth. The
disturbed soil samples were obtained at 0.5m depth intervals and taken in plastic
bags. The field moisture content and field density were recorded every 0.5m and
established using the standard requirements methods.

Two testing programs were accomplished in this research. The first testing
program was comprised of the artificial soil samples which were made by
combining sodium bentonite with different sand contents. The second testing
program was designed for testing natural soil samples selected from (Soba)
region. The natural soil samples were subjected to series of tests to cover the
range of sand percentages added to the relatively high plastic clay soil.

The artificial soils were containing of clay of sodium bentonite type and fine
sand. Percentage of bentonite by weight was varied from 10%, 15%, 20%, 25%,
and 30%.

The natural soils were obtained from the expansive clay from (Soba) mixed
with fine sand in percentage of the total weight of mixture. Sand percentages

44
varied from 17 % (as origin sample), 25%, 40%, 55%, 60%, and 65% from the
total weight.

Classification tests and Proctor compaction tests were performed. Consolidation


tests and swelling pressure tests were all performed. The tests were carried out
according to the British standards (BS 1337 - 1990) and the classification of the
soil samples results were established in accordance to Unified Soil
Classification System (USCS).

All the tests were performed in the laboratories of Building and Road Research
Institute (BRRI), University of Khartoum.

45
CHAPTER FOUR
DATA ANALYSIS AND DISCUSSIONS

4.1 Introduction

This chapter describes the analysis and discussion of the experimental results
obtained in chapter four of this research program. Data of sand-bentonite mixtures
will be analysed first while the analysis of natural samples data will follow.

The use of artificial and natural soil samples is intended to clarify the effect of sand
fraction on the behaviour of sand-clay mixture. The number of soil types tested in
this program can give good idea about the geotechnical properties of (SC) soils.

There are many differences between the engineering properties of artificial and
natural soil samples. This is attributed to the structural arrangement of particles
(i.e. bonding, fabric and combination). The regression analysis was performed to
estimate the significant and best relationships between the parameters of index
tests results, compaction, swelling and consolidation test results. Package
SPSSWIN was used as program to analyze the data, make the degree of association
and give correlation relationships between the variables.

4.2 Artificial samples


4.2.1 Classification tests results
4.2.1.1 Atterbergs limits

Atterbergs limits attain for the artificial samples show high values of L.L and P.I;
this is attributed to the mineralogy of bentonite. Increasing bentonite content from
10% to 30% made L.L and P.I to increase from 46% to 115% and from 25% to
94% respectively (see Table 3.2). The correlation coefficient (R2) for this
relationship equals 0.98 (see Fig 4.1 and 4.2).

47
Fig. 4.1: Relationship between bentonite content and Atterbergs limits

Fig. 4.2: Relationship between sand content and Atterbergs limits for artificial
soil samples

48
4.2.1.2 Classification and identification of expansiveness

Identifications and classifications of soil expansiveness were carried out for the
artificial soil samples (see Table 4.1). Four samples were classified as (CH)
according to (USCS) and one sample classified as (CL) (which contains 10%
bentonite). Table (4.1) indicates that all sand-bentonite mixtures were classified as
having very high swelling potential, except samples (A4) and (A5) which showed
high and medium swelling behaviour, respectively. The classification method
advanced by Garcia (1980) indicated that the artificial soil samples have low
swelling potential according to their swelling pressure values.

Table 4.1 Classification and identification of expansiveness for the artificial soil
samples
Artificial soil
A1 A2 A3 A4 A5
Classification
USCS classification CH CH CH CH CL
Skemptons method Active Active Active Active Active
Swelling potential Very Very
High Medium Medium
(Chen, 1988) high high
Swelling potential
Low Low Low Low Low
(Garcia, 1980)
Potential Extra Extra Very
High Medium
expansiveness high high high
swelling swelling
(Dakshanamurthy swelling swelling swelling
(CH) (CI)
and Raman, 1973) (CE) (CE) (CV)
Potential
Extra Extra Very
expansiveness High Medium
high high high
(Van der Merwe, swelling swelling
swelling swelling swelling
1964)
Swelling potential Very Very Very Very
High
(after Snethen, 1980) high high high high
Swelling potential Very Very
High High Medium
(Seed,1962) high high

49
The results presented in Table (4.1) showed very high expansive behaviour. This is
due to the mineralogy of the sodium bentonite having physical properties almost
dictated by the smectite minerals (Grim and Guven, 1978). From these
classifications and identifications, it is recommended to use 5% bentonite and not
more 15% to obtain low hydraulic conductivity and safe landfill liners or barriers
from heaving and cracks.

4.2.1.3 Prediction of Liquid limit (Schreiners method)


Schreiner, H.D (1999) developed an empirical equation to predict the L.L of the
soil using the percentage passing #40 sieve (see equation 4.1). Table (4.2) shows
the predicted values from Schreiners equation.

L.LShc = L.Lorigin sample - (%age passing #40 sieve * L.Lorigin sample /100) (4.1)

Table 4.2 Predicted L.L values of artificial soils samples (according to Schreiners
method)
L.L values Predicted L.L values
Soil (Bentonite content)
(from laboratory test (Schreiners
Code (%)
results) method)
A1 30 115 119
A2 25 95 99
A3 20 78 79
A4 15 64 59
A5 10 46 40

Where L.LShc = Predicted L.L from Schreiners method.


L.Lorigin sample = L.L of original sample before increasing of sand content.

From Table (4.2), it is noted that the L.L values from Schreiners method is similar
to the L.L values from laboratory testing. Very good relationship was observed

50
from the correlation between the predicted L.L values by Schreiners method and
the L.L values obtained from laboratory testing results (R2 = 0.99) (see Fig. 4.3).

This method is suitable to give very good prediction in field or laboratory when the
percentage passing #40 sieve is known.

Fig. 4.3: Relationship between Schreiners method and L.L from laboratory testing
(artificial soil samples)

Where L.Ltest = L.L values from laboratory testing.

4.2.1.4 Prediction of specific gravity (Estebans method)

Julio Esteban (2002) established an equation to predict the specific gravity of sand-
bentonite mixtures (see equation 4.2). Table 4.3 showed the predicted specific
gravity values of artificial soil samples by equation (4.2).

100
S.Gmi x=
(
%Sand
S .Gsand
) (
+ Bentonite
%
S .Gbetonite
) (4.2)

51
Where S.Gmix = Specific gravity of mixture.
%Sand = Percentage of sand, %Bentonite = Percentage of bentonite.
S.Gsand = Specific gravity of sand, S.Gbentonite = Specific gravity of bentonite

Table 4.3 Predicted specific gravity values of artificial soil samples (according to
Estebans method)
Soil (Bentonite content) S.G values from Predicted S.G values
Code (%) laboratory testing (Estebans method)
A1 30 2.99 2.68
A2 25 2.95 2.67
A3 20 2.83 2.66
A4 15 2.77 2.64
A5 10 2.63 2.63

The predicted S.G values are generally lower than the experimental values by a
factor between 4.7 to 10.4%. The relationship between the two variables gave
value of correlation coefficient (R2) equals 0.95. (See Fig. 4.4)

52
Fig. 4.4: Relationship between Estebans method and S.G from laboratory testing
for artificial soil samples.

4.2.1.5 Prediction the compression index (Cc) (Skemptons method)

Skempton and his associates established a relationship between the compressibility


of normally consolidated clay, indicated by its compression index (Cc), and L.L
and expressed in equation (4.3). Skemptons relationship was established based on
test results conducted on clay soils from different parts of the world. Table (4.4)
shows the predicted values from Skemptons equation.

Cc = 0.007(L.L-10) (4.3)

53
Table 4.4 Predicted (Cc) values of artificial soil samples (according to Skemptons
method)
L.L values (from Cc values (from Cc values
Soil (Bentonite content)
laboratory tests laboratory tests (Skemptons
Code (%)
result) result) equation)
A1 30 115 0.296 0.735
A2 25 95 0.158 0.595
A3 20 78 0.188 0.476
A4 15 64 0.110 0.378
A5 10 46 0.066 0.252

The values of the approved compression indices are not similar to the predicted
values (see Table 4.4). From Table (4.4), It is noted that Skemptons method is not
tolerable method to predict the compression index for sand-bentonite mixtures (R2
= 0.88). This method was established to predict the compression index for the soft
clay (normally consolidated clay of low to medium sensitivity).

There are many parameters used to predict (Cc) such as the specific gravity
(Nagaraj & Murthy, 1985), natural moisture content and initial void ratio (Azzouz
et al, 1976), but Skemptons method is the world wide one.

4.2.2 Compaction test results

Standard compaction tests were performed to study the behavior of density of the
sand-bentonite mixtures if subjected to mechanical effort. The results of the
standard compaction tests are shown in Table 3.3 and Fig. 3.2.

The artificial soil samples showed increasing in M.D.D with increase of sand
fraction (see Table 3.3). Santucci de Magistris (1998) was working with relatively
low bentonite contents and reported that a general tendency for O.M.C to increase
with high contents of bentonite.

54
The O.M.C results are inversely proportion with M.D.D results. The results of
M.D.D are relatively reduced with the increasing of bentonite content and P.I (see
Fig. 4.5 and 4.6). Good relationship was obtained between M.D.D and increasing
of bentonite content with correlation coefficient (R2) equals 0.92. The best derived
correlation coefficient (R2) for M.D.D and P.I relationship was equals 0.88.

Extensive laboratory testing is recommended to attain more correlations for


prediction of M.D.D for the sand-bentonite mixtures.

Fig. 4.5: Relationship between M.D.D and bentonite content (artificial soil
samples)

55
Fig. 4.6: Relationship between M.D.D and P.I (artificial soil samples)

4.2.3 Consolidation test results


The consolidation parameters of sand-bentonite mixtures such as initial moisture
content (wi), coefficient of volume compressibility (mv), compression index (Cc),
pre-consolidated pressure (pc) and swelling pressure (S.P) are noted to increase as
the bentonite content is increased (see Table 3.4 and Fig. 4.7 to 4.12).

The relationship between swelling pressure (S.P) and bentonite content gave
correlation coefficient (R2) value of 0.89 (see Fig. 4.7). The volume
compressibility coefficient (mv) also gave very good correlation with L.L and P.I
of 0.99 and 0.98 respectively (see Fig. 4.9). Good relationship was presented for
P.I results with compression index (Cc) with correlation coefficient (R2) of 0.88
(see Fig. 4.10).

56
The values of M.D.D were noticed to decrease with increasing bentonite content
and Cc (see Fig. 4.12) with correlation coefficient (R2) equals 0.91. Conversely,
when the dry density of a clayey soil is increased, the Cc is noted to decrease.

The relationship between swelling pressure (S.P); P.I and Cc gave correlation
coefficient (R2) value of 1.0 (see Fig. 4.8 and 4.11). This correlation is an
indication that a new equation can be derived to predict S.P of the sand-bentonite
mixtures from P.I and Cc. More experiments are needed to make a general equation
to predict S.P from these sample parameters.

In expansive soils, larger change in moisture implies higher degree of volume


change (swelling and settling) in the soil structure (Mesfin Kassa, 2005). But the
influence of the volume change on the consolidation characteristics of expansive
soil is not similar to non-expansive clay soils. In non-expansive clay soils, it shows
flatter e-log P curves (Mesfin Kassa, 2005). Whereas, the laboratory test results on
sand-bentonite mixtures have shown that the soil exhibit a steeper e-log P plot (see
Fig. 3.3).

At low stresses, there is a slight variation in void ratio (see Fig. 3.3). It is noted that
with increase of bentonite content the void ratio increased and indicated that the
bentonite is very sensitive to the applied stresses below 200 kpa. L.H. Mollins et al
(1996) and Studds et al (1998) carried out extensive studies to predict the
properties of sand-bentonite mixtures. They reported that, at low effective stress
levels the behaviour of sand-bentonite mixtures was indistinguishable from the
bentonite alone, and thus it can be inferred that the externally applied stress is
supported by the bentonite. When the externally applied stress exceeds a threshold
value (the maximum surcharge that can be supported solely by the bentonite) there
is little variation in the bentonite void ratio with increasing effective stress, and it

57
can be inferred that the difference between the externally applied stress and the
threshold value is supported largely by the sand, whose particles must now be in
contact. The threshold value at which this change in the mixture's behaviour occurs
appears to be a function of the bentonite content of the mixture (L.H. Mollins et al
(1996); Studds et al, 1998).

In literature review, the bentonite was defined as a powder consisting primarily of


randomly orientated clusters of aligned unit particles and it was reported that this
cluster-based fabric is preserved during swelling if the bentonite is confined
(Studds et al, 1998). The mechanical action of sand particles supposed to partially
disrupting the cluster-based fabric of the wet bentonite within the mixture by the
shearing by compaction when the mixture was compacted at O.M.C. Thus, this
disrupting will increase if the sand content increased and then reduce the swell
potential of the mixture.

Graham et al. (1986) proposed a similar explanation of the mechanical behaviour


of sand-bentonite mixtures assuming that the internal stress distribution is a
combination of inter-particle forces between sand grains and osmotic pressure
between clay particles. In such a model, the osmotic pressure between the clay
particles decreases as the clay void ratio increases. Therefore, at low confining
stress, the clay in the mixture is able to swell against the surcharge and separate the
sand particles to reach the same void ratio, for a given surcharge, as 100%
bentonite powder. Whereas at high confining stress, the clay void ratio upon filling
the sand pores is greater than the clay void ratio in equilibrium with surcharge
stress, and the sand particles remain either partially or fully in contact.

58
Fig. 4.7: Relationship between bentonite content and swelling pressure

Fig. 4.8: Relationship between swelling pressure and P.I (artificial soil samples)

59
Fig. 4.9: Relationship between mv and Atterbergs limits (artificial soil samples)

Fig. 4.10: Relationship between Cc and P.I (artificial soil samples)

60
Fig. 4.11: Relationship between swelling pressure and Cc (artificial soil samples)

Fig. 4.12: Relationship between Cc and M.D.D (artificial soil samples)

61
4.2.3.1 Permeability coefficient (k) and consolidation coefficient (Cv)
The permeability of the sand-bentonite mixtures was calculated using equation
(4.4). The permeability results of artificial soil samples were shown in Table 4.5.
k = mv * w * Cv (4.4)

Where k = Coefficient of permeability


Cv = Coefficient of consolidation
w = Unit weight of water

Table 4.5 Estimated permeability coefficients of artificial soil samples

Cv values Estimated
(Bentonite void Stress
Soil O.M.C M.D.D (by root time permeability
content) ratio level
Code (%) (g/cm3) method) coefficient
(%) (e) (kPa)
(cm2/min) (k) (cm/sec)
A1 30 18.4 1.670 0.743 320 0.579 1.18E-07
A2 25 10.3 1.691 0.711 160 0.632 1.44E-07
A3 20 11.8 1.692 0.651 280 2.126 3.13E-07
A4 15 7.0 1.748 0.618 122.5 2.201 4.29E-07
A5 10 10.5 1.762 0.527 80 2.275 5.44E-07

The estimated permeability coefficient (k) of the artificial soil samples was
calculated at the start of primary consolidation. The Coefficient of consolidation
(Cv) for the sand-bentonite mixtures was calculated by Taylors method (root time
method) (see Appendix B). Olson R.E (1987) has shown that the estimated
permeability coefficient values are always less than the measured values, but the
permeability values estimated from Taylors method are more close to the
measured values than the values obtained from Casagrandes method.

62
Evaluation of permeability from consolidation tests has the advantage of being
more rapid than that derived from standard permeability tests.

Terzaghis theory, used in the interpretation of consolidation rates in terms of


permeability, is based on some assumptions that do not necessarily fit exactly with
the actual soil behavior. But assuming the requirement being fulfilled, the
permeability may be derived from the one-dimensional consolidation equation.

The permeability coefficient values exhibits that the soil samples are of sand and
clay mixtures according to many authors (Terzaghi, 1940; Ali Aysen, 2002; Alam
Singh, 1987; Casagrande and Fadam, 1940). The range of values for the
permeability coefficients indicates behaviour of impermeable soils according to
EM 1110-2-1901 - Manual of US Army Corp of Engineers.

The coefficients of permeability for sand-bentonite mixtures are increasing as the


sand fraction content increased (see Table 4.5) (see Fig. 4.13 and 4.14). It gave
correlation coefficient (R2) of 0.97.

Well relationships were obtained between permeability coefficients and P.I with
correlation coefficient (R2) of 1.0 (see Fig. 4.15). More laboratory tests are
recommended to make a general equation used for the prediction of permeability
coefficient from P.I.

Many authors (Julio Esteban, 2002; Studds et al, 1998; L.H. Mollins et al, 1996;
Graham et al., 1986) recommended that the required hydraulic conductivity is
frequently from 10-6 to 10-8 cm/sec for various projects.

There are many factors affect in soil permeability such as particles shape and
texture, particles size distribution, void ratio and type of flow. Mineralogy
composition also is an important factor that may affect the value of soil

63
permeability (Arvind V. Shroff and Dhananjay L. Shah, 2003). L. Lovas and D.
Se. (1963) made large scale experiments on different types of rock forming clay
minerals, such as aplite (K and Na-feldspar), kaolin (kaolinitc), illite and bentonite
(montmorillonite), were mixed with washed sand. They reported that Na-clay
minerals do not display higher water adsorbing capacities proportionate to their
larger specific surface, and to their higher degree of dispersion. That is also
because Na-clay minerals, consequently their adsorption equilibrium sets in later.

Fig. 4.13: Relationship between k and sand content (artificial soil samples)

64
Fig. 4.14: Relationship between k and bentonite content (artificial soil samples)

Fig. 4.15: Relationship between k and P.I (artificial soil samples)

65
4.3 Natural samples
4.3.1 Classification test results
4.3.1.1 Effect of increase of sand content on Atterbergs limits

The samples obtained from the field were classified according to (USCS) as (CH)
and after addition of sand it was classified as (SC) as shown in Table 4.6. The
increase of sand content reduced the values of Atterbergs limits (see Fig. 4.16).
The original soil sample has sand content of 17% and the sand content was
gradually increased till it reached 65%. Atterbergs limits results showed that L.L
decreased from 78% to 35% and P.I decreased from 54% to 21% (see Table 3.5
and Fig. 4.16).

Soil sample designated (N4) showed high values of L.L and P.I despite the sand
fraction is more than 50% (see Table 3.6 and Fig. 4.16). The P.I varied against
swelling pressure showed correlation coefficient (R2) of 0.98 (see Fig. 4.21).

4.3.1.2 Effect of increase sand content on swelling pressure (S.P)

For the natural soil samples, swelling pressure approached 50 kPa when the sand
fraction reached 62.5%. At this percentage of sand content, the value of L.L & P.I
of the mixture is about 29.5 and 10.7% respectively (see Fig. 4.16 and 4.21). These
results are applicable to soil samples obtained from (Soba) zone.

The results indicated that the amount and type of the clay mineral
(montmorillonite) needs 62.5% of sand content to decrease the effect of swelling
pressure to near zero value. It is postulated that the sand will fill the voids and that
will result in decreasing swelling pressure (S.P).

Fig. 4.30 presents the clay minerals of the natural soil samples on Casagrandes
chart (Chleborad A.F., et al., 2005).

66
The correlations have been derived for the variation of sand content with L.L, P.I
and swelling pressure. All relationships gave correlation coefficients (R2) of 0.98,
indicating very good relationship.

Fig. 4.16: Relationship between sand content and Atterbergs limits (natural soil
samples)

4.3.1.3 Classification and identification of expansiveness


The swelling potential of the natural soil samples was classified according to
different methods obtained from literature. The classification and identification of
expansiveness for the natural samples results were shown in Table (4.6). Soil
samples (N1) and (N2) were classified as having very high swelling potential soils
and sample (N3) was classified as having high to very high swelling soil. Soil
samples (N4), (N5) and (N6) with sand contents greater than 50%, represent (SC)
soils. These were noted to have medium to high swelling behaviour. The natural

67
soil samples, classified according Garcias method (1980), have medium to low
swelling potential.

Table 4.6 Classification and identification of expansiveness for the natural soil
samples

Natural soil
N1 N2 N3 N4 N5 N6
Classification
USCS classification CH CH CH SC SC SC
Skemptons method
Normal Normal Active Active Active Active
(1953)
Swelling potential Very
Very high Very high Medium Medium Medium
(Chen, 1975) high
Potential
Very high Very high High High Medium Medium
expansiveness
swelling swelling swelling swelling swelling swelling
(Dakshanamurthy
(CV) (CV) (CH) (CH) (CI) (CI)
and Raman, 1973)
Swelling potential Very
Very high Very high High Moderate Moderate
(Snethen, 1980) high
Swelling potential
High High Medium Low Low Low
(Aronld, 1984)
Swelling potential
Medium Medium Low Low Low Low
(Garcia, 1980)
Potential
Very
expansiveness Very high Very high Medium Low Low
high
(Van der swelling swelling swelling swelling swelling
swelling
Merwe,1964)
Degree of expansion
Very high Very high High Medium Medium Medium
(Holtz & Gibbs,1956)
Swelling potential
Very high Very high High Medium Low Low
(Seed,1962)

From Table (4.6), sample (N4) exhibits expansive soil with medium to high
swelling potential from several classifications. Although the soil sample (N4)
contains sand content more than 50%, but it has expansive behaviour. The

68
behaviour is attributed to mineralogy composition, particles shape and texture and
particles size distribution of the sand-clay mixture.

4.3.1.4 Prediction of Liquid limit (Schreiners method)

The predicted L.L values from Schreiners method are similar to the L.L values
from laboratory test (see Table 4.7). This method gives very good prediction for
both natural and artificial soils (see Tables 4.2 and 4.7). The relationship between
the predicted L.L values by Schreiners method and L.L values from laboratory
testing is presented in Fig. 4.17.

Table 4.7 Predicted L.L values of natural soils samples (according to Schreiners
method)

Soil (Sand content) Predicted L.L values L.L values


Code (%) (Schreiners method) (from testing)
N1 17 - 78
N2 25 72 77
N3 40 60 66
N4 55 48 52
N5 60 45 37
N6 65 41 35

The results in Table (4.7) gave very good correlation (R2 = 0.98). From the
mentioned correlations (see clause 4.2.1.3 and 4.3.1.4), it is indicated that
Schreiners method is applicable for natural and artificial soil samples. Schreiners
method is a good method to predict L.L of soil when the percentage passing #40
sieve is known in laboratory or field. It is suggested to perform additional research
program to cover soils from different areas in the country to get a more general
equation.

69
Fig. 4.17: Relationship between Schreiners method and L.L from laboratory
testing (natural soil samples)

4.3.1.5 Prediction of compression index (Cc) (Skemptons method)

Series of consolidation tests were conducted in the laboratory on different sand-


clay mixtures. The compression index (Cc) was calculated and compared with
empirical equations obtained from literature such as Skemptons method. The
results are shown in Table (4.8). Skemptons method, in this type of soil samples,
shows similar values to laboratory test values. Its a tolerable method to predict the
compression index of the natural clayey soils.

70
Table 4.8 The compression index (Cc) of natural soil samples (according to
Skemptons method)

Predicted Cc
L.L values Cc values (from
Soil (Sand content) values
(from laboratory laboratory
Code (%) (Skemptons
tests result) tests result)
equation)
N1 17 78 0.476 0.517
N2 25 77 0.469 0.309
N3 40 66 0.392 0.282
N4 55 52 0.294 0.210
N5 60 37 0.189 0.199
N6 65 35 0.175 0.154

The relationship between the laboratory tests result and the predicted values by
Skempton gave correlation coefficient of (R2) 0.93 with 2nd degree polynomial
relationship (see Fig. 4.18). Skemptons method, in this type of soil samples, can
be used to predict the compression index of clayey soils. Skemptons method was
derived for the case of normally consolidation clay and not applicable for the sand-
bentonite mixtures (see Table 4.4 and 4.8).

71
Fig. 4.18: Relationship between Skemptons method and Cc from laboratory testing
(natural soil samples)

4.3.2 Compaction test results


The natural soil samples were tested in the laboratory to get the moisture-density
relationships. The results of the standard compaction tests showed in Fig. 3.5 and
Table 3.6. The aim of test is to understand the moisture and density behaviour of
such soils when used as backfill material under foundation or material floors of
building. This is necessary to save cost of fill material by using portion of the
excavated material from the foundation pits.

The natural soil samples mixed with sand gave increasing in M.D.D with the
increasing of sand fraction. Its postulated that the sand will fill the voids and that
will result increasing dry density. The swelling pressure results are inversely
proportion with M.D.D and gave correlation relationship with (R2) = 0.81 (see Fig.

72
4.19). This equation will make it easy to predict S.P from M.D.D results without
performing swelling pressure tests.

Many authors (Nelson and Miller, 1992; Elarabi, 2004) and conducted experiments
and concluded that the swelling rate increases when the density of soil is increased.

The sample (N4) has relatively high value of (P.I) and (M.D.D) which allows it to
heave when there is no control on moisture (see Table 3.6 and Fig. 3.5).
Furthermore, the increase of water content may cause the soil to lose some of its
strength.

Although samples (N4) and (N5) have different value of P.I due to the increasing
of sand fraction, they approximately have the same M.D.D (see Table 3.6 and Fig.
3.5).

Fig. 4.19: Relationship between M.D.D and swelling pressure for natural soil
samples.

73
4.3.3 Consolidation test results
Oedometer tests were performed to understand the swelling and compressibility
behaviour of clayey sand (SC) soils which have volume change characteristics.
When this type of soils used as fill material under foundation, its very important to
know the volume change characteristics of these soils when subjected to loads
from structures and how uncontrolled moisture affect the structure of clayey sand
(SC) soil. The results of the consolidation test were showed in Table 3.7 and Fig.
3.6.

Generally, the natural soil samples showed reduction in swelling and


compressibility characteristics as the percentage of sand fraction was increased
(see Table 3.7 and Fig. 4.20 to 4.21). Very good correlation coefficient was
obtained between L.L, P.I with swelling pressure (R2= 0.98) (see Fig. 4.20 and
4.21).

The initial void ratio (eo) decreased from 0.89 for sample (N1) to 0.47 for (N6)
which indicated that the sand grains fill the void of the clay (see Table 3.7). The
coefficient of volume compressibility (mv) is proportionally increased with (Cc) (R2
= 0.77) (see Fig. 4.22).

The results of Cc increased with S.P and gave correlation relationship with R2=
0.83(see Fig. 4.23). The increasing of M.D.D values reduce the values of Cc with
correlation coefficient (R2) of 0.96 (see Fig. 4.24).

Samples (N4) and (N5) have the same initial void ratio, swell potential and
compression index (Cc) (see Table 3.7 and Fig. 3.6). As can be seen from the test
results, the factors that affect the swelling characteristics of expansive soils (i.e.
moisture content variation and density) have also affected the consolidation
characteristics of the expansive soil.

74
Mesfin Kassa (2005) performed a study on the effect of swelling on consolidation
characteristics of expansive soils. The study showed that, swelling is a result of
disturbance in the internal stress equilibrium of the soil particle due to changing in
moisture content. To attain new internal stress equilibrium, the soil particles start
to swell and a new particle arrangement will take place in the soil mass. Therefore,
swelling brings out a soil with different particle arrangement and different
consolidation characteristics.

Consolidation is the property of the soil mass that is highly dependent on


permeability which depends on the structural arrangement of soil particles. On the
other hand, swelling is the property of the soil particle, which depends on the
mineralogy of the soil particle. In effect, both phenomena bring about volume
change in the soil mass (Mesfin Kassa, 2005).

Fig. 4.20: Relationship between swelling pressure and L.L (natural soil samples)

75
Fig. 4.21: Relationship between swelling pressure and P.I (natural soil samples)

Fig. 4.22: Relationship between Cc and mV (natural soil samples)

76
Fig. 4.23: Relationship between Cc and swelling pressure for natural soil samples

Fig. 4.24: Relationship between Cc and M.D.D for natural soil samples

77
4.3.3.1 Permeability coefficient (k) and consolidation coefficient (Cv)
The estimated permeability coefficients (k) of the natural soil samples were
calculated using the same equation and methods applied to the artificial soil
samples (clause 4.2.3.1 page No.62) (see Appendix B). The permeability
coefficient results of natural soil samples were shown in Table 4.9.

The estimated Permeability coefficients of the natural soil samples were calculated
at the start of primary consolidation. The results of the calculated permeability are
increasing as the sand fraction content increased (see Table 4.9). The estimated
permeability coefficients of the natural soil samples didnt correlate well with
M.D.D results. The consolidation coefficient (Cv) was plotted against increasing
sand content and found out to correlate with each other.

Table 4.9 Estimated permeability coefficients of natural soil samples


Clay Cv values Estimated
Applied
Soil (Sand content) fraction O.M.C M.D.D void (by root time permeability
Stress
Code (%) (< 2m) (%) (g/cm3) ratio (e)
(Kpa)
method) coefficient (k)
(%) (cm2/min) (cm/sec)
N1 17 53 29.3 1.471 0.815 320 0.021 5.79E-09
N2 25 40 21.9 1.650 0.640 560 0.498 1.78E-08
N3 40 28 18.0 1.728 0.560 320 2.125 4.54E-07
N4 55 16 16.7 1.790 0.523 160 0.531 1.57E-07
N5 60 13 14.0 1.820 0.512 160 1.991 5.10E-07
N6 65 11 13.0 1.810 0.436 40 2.126 1.76E-06

Some parameters in Table 4.9 (i.e. % sand, clay content < 2m and the applied
stress) gave good relationships with permeability coefficients with correlation
coefficient value (R2) of 0.87, 0.84 and 0.92 respectively (see Fig. 4.25, 4.26 and
4.29).

The estimated permeability coefficients gave good relationship with P.I results
with correlation coefficient R2 = 0.84 (see Fig. 4.27). More experimental works are

78
recommended to obtain more data to examine the relationship between
permeability; L.L and P.I. The correlation coefficient (R2) between M.D.D and
permeability coefficient (k) is 0.80 (see Fig. 4.28).

In accordance with manual of US Army Corp of Engineers, the range of values of


the permeability coefficients shows a nature of poorly drained soils. The values of
estimated permeability presented in Table (4.9) shows that the natural samples are
mixtures of fine sand, silt and clay (Terzaghi, 1940; Ali Aysen, 2002; Alam Singh,
1987; Casagrande and Fadam, 1940). These classifications indicate that the
existing clay in the (SC) soil is highly impermeable clay with high montmorillonite
content.

Arvind V. Shroff and Dhananjay L. Shah (2003) reported that mineralogy of soil is
important factor that affects the of soil permeability besides other factors such as
void ratio, particles size distribution, particles shape and texture.

It has been accepted by many authors who studied the relationship between
compacted moisture content and permeability for granular soils (Cedergren, 1977;
Lambe & Whiteman, 1979). It obtained that permeability for granular soils is a
function of density. In cohesive soils increasing the percentages of silt and clay in
soil cause it to form clods and create different flow conditions after compaction
(S.P. Wright, P.J. Walden, C.M. Sangha & N.J. Langdon; 1996).

In saturated soil, permeability is function of void ratio (Lambe & Whitman, 1979)
and significantly is affected by combined changes in (e) and saturation degree of
unsaturated soils. As a result, permeability with respect to the water phase
decreases rapidly as the space available for water flow reduces (Fredland, 1981).

79
Fig. 4.25: Relationship between k and sand content (natural soil samples)

Fig. 4.26: Relationship between k and clay content< 2m (natural soil samples)

80
Fig. 4.27: Relationship between k and P.I (natural soil samples)

Fig. 4.28: Relationship between k and M.D.D (natural soil samples)

81
k Vs Applied stress

2.00E-6

1.50E-6 k = -6.7E-08 + 7.26E-05/Appstress

k (cm/sec)
1.00E-6

5.00E-7

0.00E0

0.00 100.00 200.00 300.00 400.00 500.00 600.00

Applied stress (kpa)

Fig. 4.29: Relationship between k and applied stress (natural soil samples)

82
4.4 Conclusions

The artificial soil samples gave relatively high values of Atterbergs limits. All
sand-bentonite mixtures containing more than 15% bentonite content were
classified as having very high swelling potential. Its better to construct landfill or
barriers with low bentonite contents (5% to 15%).

Schreiners equation is a suitable method to predict the L.L of soils in the


laboratory and field when the percentage passing #40 sieve is known. It gave
correlation coefficient near to 1.0 for. Its applicable to sand-bentonite mixtures
and soils from Soba district. Probably, its applicable to wide range of soil types.

Skemptons method was derived to predict the compression index (Cc) for the case
of normally consolidation clay and not applicable for the sand-bentonite mixtures

All soil samples showed decreasing in volume change characteristics with increase
of sand fraction while the M.D.D of soils is noted to increase with increasing sand
fraction.

The results of M.D.D for natural soils gave good correlation coefficients with
compression index (Cc) (R2 = 0.96) and permeability coefficient (k) (R2 = 0.80).
This makes the density as one of the parameters to predict the behaviour of (SC)
soil and to classify its expansiveness concept.

Increasing the sand fraction from 7% to 10% will make the (SC) soils of low
plasticity, low compressibility, and high strength under loads and suitable as a base
material the for any engineering construction projects as well as for roads.

Well correlation coefficients were obtained between sand content, Atterbergs


limits and swelling pressure and gave correlation coefficient near to 1.0. New
prediction methods were established for natural soils in clause (4.3.1.1) and (4.3.3)

83
[page (66) and (74)] and clause (4.2.1.1) [page (47)] for natural and artificial soils
to characterize the volume change of the clay-sand mixtures, but more experiments
needed to design more general equation.

Swelling pressure (S.P) results present good correlation coefficients with Cc results
(R2 = 1.0). Swelling pressure can be used to predict the value of compression index
besides the natural moisture content and void ratio. It is recommended to conduct
more laboratory tests to predict the compression index.

The estimated permeability coefficients for the artificial and natural soil samples
were gave well relationships with good correlation coefficients with sand fraction,
Atterbergs limits and M.D.D. But more experimental works are recommended to
obtain more data to examine the relationship especially between permeability and
Atterbergs limits.

The mineralogy and permeability of expansive clay are the essential parameters in
clayey sand (SC) soils. It suggested performing more researches in the mineralogy
and permeability of (SC) soils.

Actually, the artificial soils are easier to control than the natural soils (i.e.
controlled in clay content and moisture content). The different properties
associated with the artificial and natural soils are due to different structure.

The natural soils contain clay + sand + silt, but the artificial soils contain clay
(sodium bentonite type) + sand & no silt fraction. This difference between the two
types makes difference in bonding, combination between soil particles and fabric.

84
CHAPTER FIVE
SUMMARIES AND RECOMMENDATIONS

5.1 Introduction

This is a concluding chapter. It will summarize the finding of the research carried
out on natural expansive clayey soils, taken from trial pits in (Soba) district and the
artificial soils (sand-bentonite mixtures) prepared in laboratory. Bentonite and
natural clayey sand (SC) soil were mixed in various ratios with fine sand fraction.
Series of tests performed in laboratory to cover the geotechnical properties of (SC)
soils by increasing range of sand percentages.

The data, which was taken from the experimental works, is used in SPSSWIN
computer program and correlation coefficients were calculated to choose the best
relationship between variables. Good correlation coefficients were obtained for
artificial and natural soil samples but more experiments are recommended to
confirm the findings obtained in this research.

5.2 Summaries
5.2.1 Literature review

From the chapter of literature review the following conclusions have been reached:

1. Sand-bentonite mixture, at moisture increasing, starts to heave into the existing


voids as it tries to overcome the confining pressure and continues to heave until
the voids filled and the rate of expansion reduces with time.

2. The common stabilizer used for (SC) soils is lime because of its easiness in use
and its improvement for soil properties. The improvement of soil properties can
be obtained for 2% to 8% lime content.

86
3. A proportion of lime to fly ash of (1:20) is giving high value of (M.D.D),
increasing the strength and decreasing the hydraulic conductivity of the (SC)
soils.

4. The use of lime-polyamide fiber showed that it is improving the cohesion


properties of the (SC) soil which make it suitable as construction material as
under foundation material or highway material as base, sub-base, and stabilized
sub-grade under heavy load pavements.

5. Fine content may affect the compression characteristics of coarse grained soils
as well. Many experimental researches and models were proposed for the
compression behavior of cohesionless soils and some factors such as initial void
ratio, relative density, particle shape, mineralogy, structure and applied stress
conditions were noted to illumine this behavior.

6. Many authors reported that when the clay fraction is less than 25-30%, the soil
tends to be sand or silt soils; other wise the soil behaves much like clay.

7. Clayey sand (SC) soil is poor material in pavements if not compacted at the
proper moisture content, in this case its shear resistance can be low. The
unconfined compressive strength of clayey sand soil can be doubled by
compaction. Many authors (e.g. Sorochan, 1991; Elarabi, 2004) confirmed that
the swelling of (SC) soil increases when the density of soil is increased.

8. Clayey sand (SC) soil can be used behind retaining walls provided that
precautions are taken against frost action. Generally, the ideal backfill material
should be containing less than 5% of fines and the L.L of soil is less than 35%
and P.I is less than 12%. Soils having such properties will experience a
negligible amount of heave.

87
9. Clayey sand (SC) soils are generally undesirable as fill or under foundation
material, due to the low compaction and permeability properties but they can be
modified to be suitable in roads, highways as sub-base material and other civil
engineering applications.

5.2.2 Experimental works

The following conclusions have been summarized from the chapter of


experimental works:

1. Two testing programs were accomplished in this research. The first testing
program was comprised of the artificial soil samples which were made by
combining sodium bentonite with different sand contents. The second testing
program was designed for testing natural soil samples selected from (Soba)
region.

2. Relatively high plastic clay soils were obtained from (Soba) district. Tests pits
were excavated about 1.5m depth. The disturbed soil samples were obtained at
0.5m depth intervals and taken in plastic bags. The field moisture content and
field density were recorded every 0.5m and established using the standard
procedures.

3. The artificial soils were made up of clay of sodium bentonite type and fine
sand. Percentage of bentonite by weight was varied from 10%, 15%, 20%, 25%,
and 30%.
4. The natural soils were mixed with fine sand in percentage of the total weight of
mixture. Sand percentages varied from 17 % (as origin sample), 25%, 40%,
55%, 60%, and 65% from the total weight.

5. Classification tests and Proctor compaction tests were performed. Consolidation


tests and swelling pressure tests were also performed. The tests were carried out

88
according to the British standards (BS 1337 - 1990) and the classification of the
soil samples results were established in accordance to Unified Soil
Classification System (USCS). All the tests were performed in the laboratories
of Building and Road Research Institute (BRRI), University of Khartoum.

5.2.3 Analysis and discussions

The conclusions were obtained from the chapter of analysis and discussions are
that:

1. Sand-bentonite mixtures containing more than 15% bentonite content were


classified as having very high swelling potential. Therefore low bentonite
contents from 5% to 15% are recommended to design safe landfill liners or
barriers from volume changes hazard.

2. Increasing the sand fraction from 7% to 10% for (SC) soil will give relatively
high strength and low volume change properties and makes it suitable for use
under foundations and floors and also in the construction of roads.

3. The density is one of the parameters which clarify and classify the expansive
behaviour of (SC) soils. The mineralogy and permeability of expansive clay are
the essential parameters in expansive (SC) soils.

4. It is recommended to take precautions in using (SC) soil as highway material,


fill or under foundation materials. Hazards of volume change of (SC) soils can
be avoided by moisture controlling in field. Moisture controlling is
recommended also to get good density for pavements and under superstructure.

5. According to the difference in soil structure between the artificial and natural
soils, therefore its easy to control the artificial soils more than the natural soils
because the natural soils contain different ratios of clay, sand and silt fractions,

89
but the artificial soils just contain sand, pure montmorillonite clay fraction and
no silt fraction.

6. Schreiners equation is a suitable method to predict the L.L of soils in the


laboratory and field when the percentage passing #40 sieve is known. It gave
correlation coefficient near to 1.0. Its applicable to sand-bentonite mixtures
and soils from Soba district. Probably, its applicable to wide range of soil
types.

7. Skemptons equation was derived to predict the compression index (Cc) for the
case of normally consolidation clay and not applicable for the sand-bentonite
mixtures.

8. New prediction methods with good correlation coefficients, near to 1.0, were
obtained between sand content, Atterbergs limits and (S.P) for natural and
artificial soils to characterize the volume change of the clay-sand mixtures.

9. Swelling pressure (S.P) results showed good correlation coefficients with (Cc)
results (R2 = 1.0). The relationships indicating that the (S.P) can be used to
predict the value of compression index with high correlation coefficient beside
the natural moisture content and void ratio tests.

10. Well relationships were obtained for the estimated permeability coefficients
with sand-bentonite mixtures and the natural soil samples. Good correlation
coefficients were obtained with sand fraction, Atterbergs limits and M.D.D.
But more laboratory tests are recommended to make a general equation used for
the prediction of permeability coefficient from Atterbergs limits.

90
5.3 Recommendations for future works

1. More experiments on wide range of natural soils around the country are needed
to obtain more general equations for the characterization and prediction of
clayey sand (SC) soils.

2. More experiments with intermediate percentage of sand are recommended to


verify.

3. It is suggested to perform more research on the mineralogy and permeability of


(SC) behaviour.

4. Permeability coefficients of soil can be predicted from the Atterbergs limits,


but more studies and extensive experiments on more local soils are
recommended to confirm this relationship.

5. Compression index (Cc) can be predicted from of swelling pressure (S.P),


natural moisture content and void ratio tests with high correlation coefficient.
More studies are recommended to confirm this relationship and to get more
general equation.

91
References
1. A. Lasledj and M. Al-Mukhtar, Effect of Hydrated Lime on the
Engineering Behaviour and the Microstructure of Highly Expansive Clay,
Proceedings of the 12th International Conference of International
Association for Computer Methods and Advances in Geomechanics
(IACMAG), October 2008, Goa, India

2. Abduljauwad, S.N. (1995), Improvement of Plasticity and Swelling


Potential of Calcareous Expansive Clays, Geotechnical Engineering
Journal, June 1995, pp. 3-16.

3. Alam Singh (2009), Soil engineering in theory & practice fundamentals &
general principles, Vol. 1, India.

4. Ali Aysen (2002), Soil mechanics: Basic concepts and Engineering


Applications, 1st edition, Swets & Zeitlinger B.V, Lisse, The Netherlands.

5. Al-Rawas A.A., Hago A.W., Al-Sarmi H. (2005), Effect of lime, cement


and Sarooj (artificial pozzolan) on the swelling potential of an expansive
soil from Oman, Building and Environment, Vol.40, pp.681687.

6. Arabani, M. and M. Veis Karami (2005), Geomechanical properties of


lime stabilized clayey sands, A paper accepted for Publication in the
Arabian Journal for Science and Engineering (AJSE), King Fahd University
of Petroleum, Dhahran, Saudi Arabia.

7. Aronld, M. (1984), The Genesis, Mineralogy and Identification of


Expansive soils, Proceeding the 5th International Expansive soil
Conference, Adelaide, South Australia, pp. 32-36.

8. Azzouz, A.S., Kriezek, R.J., and Corotis, R.B. (1976), Regression


Analysis of soil compressibility, Soils Found., Tokyo, Vol. 16, No. 2.

9. BS 1377 (1990), Soils for civil engineering purposes: part 2: classification


tests.

10. Basma A.A., Al-Homoud A.S. and Hussein A. (1995), Laboratory


assessment of swelling pressure of expansive soils, Department of Civil
Engineering, Jordan University of Science and Technology, Applied Clay
Science 9 (1995), pp. 355368.
92
11. Basma, A.A., and Tuncer, E.R. (1991), Effect of lime on volume change
and compressibility of expansive clays, Transportation, Research Record
1295, pp. 5261.

12. Baver, L.D, Grander, W.L and Grander, W.R (1972), Soil physics,
John Wiley & Sons, Inc. New York.

13. Casagrande, A., and Fadum, R.E. (1940), Notes on Soil Testing for
Engineering Purposes, Soil Mechanics Series, Graduate School of
Engineering, Harvard University, Vol. 8, pp. 268.

14. Casagrande A (1936), Characteristics of cohesionless soils affecting the


stability of slopes and earth fills, J. Boston Soc. Civil Engineering, Vol.
23(1), pp.1332.

15. Cedergren, Harry R. (1977), Seepage, Drainage, and Flow Nets, Wiley
-Interscience.

16. Chapuis, R.P. (1990), Sand-Bentonite Liners: Predicting Permeability


from Laboratory Tests, Canadian Geotechnical Journal, Vol. 27, pp. 47-57.

17. Chleborad, A.F., Diehl, S.F., Cannon, S.H. (1996), Geotechnical


properties of selected materials from the Slumgullion landslide, In: The
Slumgullion earth flow: a large-scale natural laboratory, D.J. Varnes, W.Z.
Savage, (eds), U.S. Geological Survey Bulletin 2130, pp. 6771.

18. Cerina J.N. (1995), Geotechnical engineering soil mechanics, Canada,


John Wiley & Sons, Inc. New York.

19. Chalermyanont T. and Arrykul S. (2005), Compacted sand-bentonite


mixtures for hydraulic containment liners, Songklanakarin Journal of
Science and Technology, Thailand, Vol. 27(2): 313-323.

20. Chen, F.H. (1975), Foundations on Expansive soils, American Elsevier


publ., New York, pp. 280.

21. Clare K.E., Cruchley A.E. (1957), Laboratory experiments in the


stabilization of clays with hydrate lime, Gotechnique 7, pp. 97-111.

93
22. Clough, G.W., Sitar, N., Bachus, R.C., Rad, N.S., Cemented sands
under static loading, Geotechnical Engineering Division, ASCE 107.
(1981), No. GT6, pp. 799817.

23. Craig J.R., & Vaughan D.J. (1994), Ore Microscopy and Ore
Petrography, 2nd edition, John Wiley & Sons Inc., New York.

24. Dakshanamurthy,V. and Rahman,V. (1973), A simple method of


identifying an Expansive soil, Jananese Society of Soil Mech. and
Foundation Engineering, Vol. 13, pp. 97-104.

25. Elarabi, H (2004), Factors influencing swelling behavior of expansive


soils, BRRI journal, BRRI, University of Khartoum, Vol. 6.

26. El turabi, M.A.D. (1985), A study on expansive clay soil in Sudan, thesis
submitted for the degree of M.Sc in civil engineering, Building and Road
Research institute (BRRI), University of Khartoum.

27. Fredlund, D.G. and Rahardjo, H. (1993), Soil mechanics for unsaturated
soils, John Wiley and Sons, Inc. New York.

28. Garacia, L., Martinez, J. and Paolini, G., Design and construction of
short piles for expansive soils in Coro, Venezuela, Proceeding 4th
International Conference on Expansive Soils, Denver, Colorado, USA, Vol.
11, pp. 850865.

29. Georgiannou, V.N., Burland, J.B., Hight, D.W. (1990), The undrained
behaviour of clayey sands in triaxial compression and extension,
Geotechnique, Vol. 40 (3), pp. 431449.

30. Gleason, M.H., Daniel, D.E. and Eykholt (1997), Calcium and Sodium
bentonite for hydraulic containment applications, Journal of Geotechnical
and Geo-environmental Engineering, Vol. 123, No. 5, pp. 438-445.

31. Graham, J., Gray, M.N., Sun, B.C. & Dixon, D.A (1986), Strength and
volume change characteristics of a sand-bentonite buffer, Proceeding 2nd
International Conference on Radioactive Waste Management, Winnipeg,
Man., pp.188-194.

94
32. Grim, R. E. and Guven N. (1978), Bentonites: Geology, Mineralogy,
Properties and Uses, Elsevier Scientific Publishing Company, New York,
pp. 256.

33. G. E. Abdelrahman, M.M. Shahien (2004), Swelling treatment by using


sand for Tamia swelling soil, Proceeding 4th International Engineering
Conference (4th IEC), Sharm El Sheikh, Egypt.

34. Hamilton, J. J. (1977), Foundations on swelling or shrinking subsoils,


Canadian Building Digest, National Research Council, Canada, Vol. CBD-
184.

35. Hardin, B.O. (1987), 1-D Strain in Normally Consolidated Cohesionless


Soils, Journal of Geotechnical Engineering, ASCE. Vol. 113, No. 12. pg
1449 1467.

36. Holtz, W. and Gibbs, J.J (1956), Engineering properties of Expansive


Clays, Journal of the Soil Mech. and Foundation Div., American Society for
Civil Engineers (ASCE), Transactions paper No. 2814, Vol. 121, pp. 641- 663.

37. Julio Esteban (2002), Suction and volume changes of compacted sand-
bentonite mixtures, thesis submitted for the degree of Doctor of
philosophy (Ph.D) in Faculty of engineering, Imperial College of Science
and Medicine, London University.

38. Kassiff and R. Baker (1971), Aging Effects on Swell Potential of


Compacted Clay, Journal of the Soil Mechanics and Foundation Division,
American Society for Civil Engineers (ASCE), Vol. 97, pp. 529-540.

39. Komine, H., and Ogata, N. (1994), Experimental study on swelling


characteristics of compacted bentonite, Canadian Geotechnical Journal,
Vol. 31, pp. 478490.

40. Komine, H., and Ogata, N. (1996), Predicting of on swelling


characteristics of compacted bentonite, Canadian Geotechnical Journal,
Vol. 33, pp. 1112.

41. Komine, H., and Ogata, N. (1999), Experimental study on swelling


characteristics of sand-bentonite for waste disposal, Soil and foundations,
Vol. 39, pp. 8397.

95
42. Lambe, W.T. and Whitman, R.V. (1979), Soil mechanics, John Wiley
& Sons, Inc. New York.

43. Little, D.L. (1987), Fundamentals of the Stabilization of Soil with Lime,
(Bulletin 332), National Lime Association (NLA), Arlington, VA.

44. L. Lovas & D. Se. (1963), The effect of clay minerals on the permeability
of sand soils, International Association of Scientific Hydrology, Berkeley
Assembly, Publication 64, pp. 274-289.

45. L.H. Mollins, D.I. Stewart AND T.W. Cousens (1996), Predicting the
properties of bentonite-sand mixtures, Clay Minerals, Vol. 31, pp. 243-
252.

46. Locat, J., Berube, M.A. and Choquette, M. (1990), Laboratory


investigations on the lime stabilization of sensitive clays: shear strength
development, Canadian Geotechnical Journal, Vol.27, pp. 294-304.

47. Mahyar Arabani, Akbar K. Haghi and Mehdi Veis Karami (2005), A
study on the strength behavior of stabilized clayey sands using lime-
polyamide fiber, Electronic Journal of Geotechnical Engineering (EJGE),
Vol. 10, Bundle D.

48. Mesfin Kassa (2005), Relationship between consolidation and swelling


characteristics of expansive soils of Addis Ababa, thesis submitted to
Addis Ababa University for the degree of M.Sc in Civil Engineering.

49. Mitchell J.K. (1993), Fundamentals of soil behavior, John Wiley &
Sons, Inc. New York.

50. Mollins, L.H., Stewart, D.I. and Cousens, T.W. (1996), Predicting the
properties of bentonite-sand mixtures, Clay Minerals. Vol. 31, pp. 243-
252.

51. Monkul, M.M., Ozden, G. (2005), Effect of intergranular void ratio on


one-dimensional compression behavior, Proceedings of International
Conference on Problematic Soils, International Society of Soil Mechanics
and Geotechnical Engineering, Famagusta, Turkish Republic of Northern
Cyprus, 3, pp. 12031209.

96
52. Muni Budhu (2000), Soil mechanics and foundations, John Wiley &
Sons, Inc. New York.

53. Muntohar, A.S. and Hashim, R. (2002), Properties of engineered


expansive soils, Proceeding of the 1st Technical Postgraduate Symposium,
University of Malaya, October 2002, pp. 272276.

54. Nagaraj, T.S., and Murthy, B.R.S. (1985), Prediction of the Pre-
consolidation Pressure and Recompression Index of Soils, Geotechnical
Testing Journal, American Society for Testing and Materials, Vol. 8, No. 4,
pp. 199-202.

55. National Lime Association (NLA) (2004), Lime Treated Soil


Construction Manual: Lime Stabilization and Lime Modification.

56. N. K. Ameta, Abhay Shivaji Wayal (2008), Effect of bentonite on


permeability of dune sand, Electronic Journal of Geotechnical Engineering
(EJGE) Vol. 13, Bundle A.

57. Nelson, J.D. and Miller, D.J. (1992), Expansive soils: Problem and
practice in foundation and pavement engineering, John Wiley and Sons,
New York.

58. Olson, K.R. (1985), Characterization of pore size distribution within soils
by mercury intrusion and water release methods, Soil science, Vol.139,
No.5, pp. 400-404.

59. Olson, K.R. (1987), Method to measure soil pores outside the range of
mercury intrusion Porosimeter, Soil science, Journal Vol. 51, pp. 132-135.

60. Ormsby W.C., Kinter E.B (1973), Strength development and reaction
products in lime-montmorillonite-water systems, Public Roads, Vol. 37 (4),
pp.136-148.

61. Osman, M.A. and Ali, E.M. (1984), Construction expansive soils in
Sudan, Journal of construction Engineering and Management, Vol. 110,
pp.359374.

97
62. Osman, M.A. and Hamadto, M.E.M (1987), Identification &
classification of expansive soils in arid and semi arid regions, Proceeding
of the 9th Regional conference for Africa on Soil Mechanics and
Foundation Engineering, Lagos, Vol. 9, pp. 113-116.

63. Pestana, J.M. and Whittle, A.J. (1995), "Compression Model for
Cohessionless Soils", Gotechnique, Vol. 45 (4), pp. 611-631.

64. Popescu, M. E., 1986, A Comparison between the Behavior of Swelling


and Collapsing Soils, Engineering Geology, Vol. 23, No. 2, pp. 145-163.

65. Roslan Hashim and Agus Setyo Muntohar (2006), Swelling rate of
expansive clay soils, Expansive Soils - Recent advances in characterization
and treatment, edited by Amer Ali Al-Rawas & Mattheus F.A. Goosen.

66. Santosh Kumar Garg (2005), Soil mechanics and foundation


engineering: for Civil Engineering degree students, 6th edition, Khanna,
Indian.

67. Santucci de Magistris (1998), Physical and mechanical properties of


compacted silty sand with low bentonite fraction, Canadian Geotechnical
Journal, Vol. 35, pp. 909-925.

68. Schreiner, H.D,(1999), A new approach to the determination of the


expansiveness of soils, Proceeding of the 12th African Regional conference
on Soil Mechanics and Foundation Engineering, Durban, South Africa.

69. Seed, H.B., Woodword, R.J., and Lurdgren, R. (1962), Prediction of


swelling potential for compacted clays, Journal of the Soil Mech. and
Foundation Division, American Society for Civil Engineers (ASCE), Vol.
202, pp. 337-344.

70. Shroff, Arvind V.; Shah, Dhananjay L. (2003), Soil Mechanics and
Geotechnical Engineering, Publisher: Taylor & Francis.

71. Sivapullaiah, P.V., Sridharan, A., and Ramesh H.N. (1995),


Mechanisms controlling the index properties of lime treated black cotton
soil in the presence of sulphate, Indian Geotechnical Journal, Vol.25,
pp.379-394.

98
72. Skempton AW (1944), Notes on the compressibility of clays, Q. J.
Geological Society, London, Vol. 100, pp. 119-135.

73. Skempton, A.W. (1953), The colloidal "activity" of clays, Proceeding


3rd international conference on Soil Mechanics and Foundation
Engineering, Zurich, Vol. 1, pp. 57-61.

74. Snethen, D. R. (1980), Characterization of expansive soils using soil


suction data, Proceeding 4th International Conference on Expansive Soils,
American Society for Civil Engineers (ASCE), Vol. 1, pp. 54-75.

75. Sorochan, E.A. (1991), Construction of Buildings on Expansive Soils,


Publisher: Taylor and Francis.

76. Studds, P.G., Stewart, D.I. and Cousens, T.W. (1998), The Effects of
Salt Solutions on the Properties of Bentonite-Sand Mixtures, Clay
Minerals, Vol. 33, pp. 651-660.

77. Studds, Stewart, D.I., P.G. and Cousens, T.W. (2003), The factors
controlling the engineering properties of bentonite-enhanced sand,
Applied Clay Science, Vol. 23, pp. 97-110.

78. S.P. Wright, P.J. Walden, C.M. Sangha & N.J. Langdon (1996),
Observations on soil permeability, moulding moisture content and dry
density relationships, Quarterly Journal of Engineering Geology and
Hydrogeology, Vol. 31, pp. 75-76.

79. Terzaghi K. T. (1943), Soil mechanics in engineering practice, 3rd


edition, 1996, John Wiley & Sons, Inc. New York.

80. Thaweesak Jirathanathaworn, Supakij Nontananandh and Korchoke


Chantawarangul (2003), Stabilization of clayey sand using fly ash mixed
with small amount of lime, Proceedings of the 9th National Convention on
Civil Engineering (NCCE), Petchaburi, Thailand, pp. 6.

81. Thompson, M.R. (1966), Shear Strength and Elastic Properties of Lime-
Soil Mixtures, Transportation Research Record No. 139, Transportation
Research Board, Washington, D.C.

99
82. Rogers, C. D. F., Glendinning, S., and Roff, T.E.F. (1997), Lime
modification of clay soils for construction expediency, Proceedings of the
Institution of Civil Engineers, Geotechnical Engineering, Vol. 125, pp. 242-
249.

83. Rollings, M.P. and R.S. Rollings. 91996), Geotechnical Materials in


Construction, New York: McGraw-Hill.

84. United States (US) Army Corp of Engineers Manual - EM 1110-2-1911,


Earth-Fill and Rock-Fill Construction, Chapter five 30 Sep 95.

85. United States (US) Army - Engineering school (1992), Military soils
engineering, Field Manual FM 5-410.

86. Van Der Merwe, D .H. (1964), Prediction of heave from the plasticity
index and percentage clay fraction of soils, South African Institute of civil
Engineers, Vol. 6, pp. 103-107.

87. V.N.S. Murthy (2003), Principles and practices of soil mechanics and
foundation engineering, Marcel Dekker Inc., New York.

88. Yamamuro, J.A., Bopp, P.A., Lade, P.V. (1996), One dimensional
compression of sands at high pressures, Journal of Geotechnical
Engineering, ASCE, Vol. 122 (2), pp. 147154.

100
A-1 Test procedures for Specific gravity (S.G) of artificial soil samples

The dry sample of fine sand was added in the density bottle at first and then the
bentonite. The weight of bottle and dry sample were measured is (W2).

Distilled water was added to the dry sample in the density bottle. The distilled
water approximately filling of volume of the bottle and cured for 24 hours to
sure that the bentonite particles were saturated. Then the bottles were filled by
water and shaking well. The weight of the bottle with the sample and water
equals to (W3).

The bottles were setting in big dish contained of water of 26 C 1 until the
mixture in bottles reached the degree of the dish. The bottle was washed to
remove soil, filled with distilled water and the measured weight is (W4).

The above procedure was repeated for ten specimens and the difference between
the tolerable values in variation of 0.025 was taken (i.e. difference should be
less than 0.03).

The average of values was calculated and it represented the actual value of the
specific gravity (S.G). (See Table 3.2)

S.G =
(W2 W1 )
(W4 W1 ) (W3 W2 ) (1)

For the pure bentonite, the above procedures were repeated in addition of
preparing of five small dishes their weights were known. The dishes were
placed down the bottles when the solution of bentonite and distilled water filled
in the bottle. The flowing out of the solution was got by the dish and the
measured weight was (W3). The bottles and dishes were placed in the oven for
24 hours and the measured weight of bottle, dishes, and the dry sample were
measured in the balance was (W2). The results were presented in Table 3.1.

102
H Vs t (10% bentonite content)
7.325
7.32
7.315
7.31
7.305
7.3
7.295
7.29
7.285
7.28
7.275
7.27
7.265
7.26
7.255
Deformation (H) / Div

7.25
7.245
7.24
7.235
7.23
7.225
7.22
7.215
7.21
7.205
7.2
7.195
7.19
7.185
7.18
7.175
7.17
7.165
7.16
7.155
7.15
7.145
7.14
7.135
7.13
7.125
7.12
7.115
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39

t / min

Fig. B.1: Consolidation coefficient (Cv) by Taylor's method (10% bentonite content) 100
H Vs t (15% bentonite content)
6.05

5.95

5.9

5.85

5.8

5.75

5.7

5.65

5.6
Deformation (H) / Div

5.55

5.5

5.45

5.4

5.35

5.3

5.25

5.2

5.15

5.1

5.05

4.95
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39

t / min

Fig. B.2: Consolidation coefficient (Cv) by Taylor's method (15% bentonite content) 101
H Vs t (20% bentonite content)
3.595
3.575

3.555
3.535

3.515

3.495
3.475

3.455
3.435

3.415
3.395
Deformation (H) / Div

3.375
3.355

3.335

3.315
3.295

3.275
3.255

3.235
3.215

3.195
3.175

3.155

3.135
3.115

3.095
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B-3: Consolidation coefficient (Cv) by Taylor's method (20% bentonite content) 102
H Vs t (25% bentonite content)
6.575

6.57
6.565

6.56
6.555

6.55
6.545

6.54
6.535

6.53
Deformation (H) / Div

6.525

6.52

6.515
6.51

6.505
6.5

6.495
6.49

6.485
6.48

6.475
6.47

6.465
6.46

6.455
6.45
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B-4: Consolidation coefficient (Cv) by Taylor's method (25% bentonite content) 103
H Vs t (30% bentonite content)
6.65

6.6

6.55

6.5

6.45

6.4
Deformation (H) / Div

6.35

6.3

6.25

6.2

6.15

6.1

6.05

5.95
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.5: Consolidation coefficient (Cv) by Taylor's method (30% bentonite content) 104
H Vs t (17% sand content)
7
6.95
6.9
6.85
6.8
6.75
6.7
6.65
6.6
6.55
6.5
6.45
6.4
6.35
6.3
Deformation (H) / Div

6.25
6.2
6.15
6.1
6.05
6
5.95
5.9
5.85
5.8
5.75
5.7
5.65
5.6
5.55
5.5
5.45
5.4
5.35
5.3
5.25
5.2
5.15
5.1
5.05
5
4.95
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.6: Consolidation coefficient (Cv) by Taylor's method (17% sand content) 105
H Vs t (25% sand content)
6.05

5.95

5.9

5.85

5.8

5.75

5.7
Deformation (H) / Div

5.65

5.6

5.55

5.5

5.45

5.4

5.35

5.3

5.25

5.2

5.15

5.1

5.05

4.95
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.7: Consolidation coefficient (Cv) by Taylor's method (25% sand content) 106
H Vs t (40% sand content)
7.25

7.2

7.15

7.1

7.05

6.95
Deformation (H) / Div

6.9

6.85

6.8

6.75

6.7

6.65

6.6

6.55

6.5

6.45

6.4

6.35

6.3
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.8: Consolidation coefficient (Cv) by Taylor's method (40% sand content) 107
H Vs t (55% sand content)
7.135

7.12

7.105

7.09

7.075

7.06

7.045
Deformation (H) / Div

7.03

7.015

6.985

6.97

6.955

6.94

6.925

6.91

6.895

6.88

6.865

6.85
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.9: Consolidation coefficient (Cv) by Taylor's method (55% sand content) 108
H Vs t (60% sand content)
7.825

7.8

7.775

7.75

7.725

7.7

7.675

7.65
Deformation (H) / Div

7.625

7.6

7.575

7.55

7.525

7.5

7.475

7.45

7.425

7.4

7.375

7.35

7.325

7.3
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.10: Consolidation coefficient (Cv) by Taylor's method (60% sand content) 109
H Vs t (65% sand content)
8.5

8.475

8.45

8.425

8.4

8.375
Deformation (H) / Div

8.35

8.325

8.3

8.275

8.25

8.225

8.2

8.175

8.15

8.125

8.1
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39
t / min

Fig. B.11: Consolidation coefficient (Cv) by Taylor's method (65% sand content) 110

Você também pode gostar