You are on page 1of 241

1

SOIL MECHANICS

INTRODUCTION

Geotechnical Engineering is that part of engineering which is


concerned with the behaviour of soil and rock. Soil Mechanics
is the part concerned solely with soils. From an engineering
perspective soils generally refer to sedimentary materials that
have not been cemented and have not been subjected to high
compressive stresses.

As the name Soil Mechanics implies the subject is concerned


with the deformation and strength of bodies of soil. It deals with
the mechanical properties of the soil materials and with the
application of the knowledge of these properties to engineering
problems. In particular it is concerned with the interaction of
structures with their foundation material. This includes both
conventional structures and also structures such as earth dams,
embankments and roads which are they made of soil.

As for other branches of engineering the major issues are


stability and serviceability. When a structure is built it will
apply a load to the underlying soil; if the load is too great the
strength of the soil will be exceeded and failure may ensue. It
is important to realise that not only buildings are of concern, the
failure of an earth dam can have catastrophic consequences, as
can failures of natural and man made slopes and excavations.
Buildings or earth structures may be rendered unserviceable by
excessive deformation of the ground, although it is usually
differential settlement, where one side of a building settles more
than the other, that is most damaging. Criteria for allowable
settlement vary from case to case; for example the settlement
allowed in a factory that contains sensitive equipment is likely
to be far more stringent than that for a warehouse. Another
important aspect to be considered during design is the effect of
any construction on adjacent structures, for example the
excavation of a basement and then the construction of a large
building will cause deformations in the surrounding ground and
2

may have a detrimental effect on adjacent buildings or other


structures such as railway tunnels.

Many of the problems arising in Geotechnical Engineering stem


from the interaction of soil and water. For example, when a
basement is excavated water will tend to flow into the
excavation. The question of how much water flows in needs to
be answered so that suitable pumps can be obtained to keep the
excavation dry. The flow of water can have detrimental effects
on the stability of the excavation, and is often the initiator of
landslides in natural and man made slopes. Some of the effects
associated with the interaction of soil and water are quite subtle,
for example if an earthquake occurs, then a loose soil deposit
will tend to compress causing the water pressures to rise. If the
water pressures should increase so that they become greater than
the stress due to the weight of the overlying soil then a
quicksand condition will develop and buildings founded on this
soil may fail.

Soil mechanics differs from other branches of engineering in


that generally there is little control over the material properties,
we have to make do with the soil at the site and this is often
highly variable. By taking samples at a few scattered locations
we have to determine the soil properties and their variability. At
this stage in a project knowledge of the site geology and
geological processes is essential to successful geotechnical
engineering.

Soil mechanics is a relatively new branch of engineering


science, the first major conference occurred in 1936 and the
mechanical properties of soils are still incompletely understood.
The first complete mechanical model for soil was published as
recently as 1968. Over the last 40 years there has been rapid
development in our understanding of soil behaviour and the
application of this knowledge in engineering practice. The
subject has now reached a phase of development similar to that
of structural mechanics a century ago and the words of William
Anderson in 1893 about structural engineers are relevant today
for geotechnical engineering, "There is a tendency among the
3

young and inexperienced to put blind faith in formulas


(computer programs), forgetting that most of them are based
upon premises which are not accurately reproduced in practice,
and which, in any case, are frequently unable to take into
account collateral disturbances which only observation and
experience can foresee, and common sense provide against."
4

1. SOILS AND THEIR CLASSIFICATION

1.1 Introduction

Soil mechanics is concerned with particulate materials (soils)


found in the ground that are not cemented and not greatly
compressed. These soils usually have a sedimentary origin,
however, they can also occur as the result of rock weathering
without any transport of the particles. The soil particles can
have varying sizes, shapes and mineralogies, although these
properties are usually interrelated. For instance the larger sized
particles are generally composed of quartz and feldspars,
minerals that have high strengths and the particles are fairly
round. The smaller sized particles are generally composed of
the clay minerals kaolin, illite and montmorillonite, minerals
that have low strengths and form plate like particles. One of the
most important aspects of particulate materials is that there are
gaps or voids between the particles. The amount of voids is
also influenced by the size, shape and mineralogy of the
particles.

Because of the wide range of particle sizes, shapes and


mineralogies in a typical soil a detailed classification of each
soil would be very expensive and inappropriate for most
geotechnical engineering purposes. However, some form of
simple classification system giving information about the
engineering properties is required on all sites. Why is this
necessary?

• Usually the soil on site has to be used. Soils differ from other
engineering materials in that one has very little, if any,
control over their properties.

• The extent and properties of the soil at the site have to be


determined.

• Cheap and simple tests are required to give an indication of


the engineering properties such as stiffness and strength for
preliminary design.
5

To achieve this continuous samples are recovered from


boreholes, drilled to a depth that will depend on the scale of the
project. Observation of the core enables the different soil layers
to be determined and then classification tests are performed for
these different strata. The extent of the different soil layers can
be determined by correlating the results from different
boreholes and this information is used to build a picture of the
sub-surface profile.

An indication of the engineering properties is determined on


the basis of particle size. This crude approach is used because
the engineering behaviour of soils with very small particles,
usually containing clay minerals, is significantly different from
the behaviour of soils with larger particles. Clays can cause
problems because they are relatively compressible, drain
poorly, have low strengths and can swell in the presence of
water.

1.2 Particle Size Definitions

The precise boundaries between different soil types are


somewhat arbitrary, but the following scale is now in use
worldwide.

Gravel Sand Silt Clay


C M F C M F C M F C M F
60 20 6 2 0.6 0.2 0.06 0.02 .006 .
002 .0006 .0002

where C, M, F stand for coarse, medium and fine respectively,


and the particle sizes are in millimetres.

Note
• the logarithmic scale. Most soils contain mixtures of sand,
silt and clay particles, so the range of particle sizes can be
very large.
6

• not all particles less than 2 µm are comprised of clay


minerals, and some clay mineral particles can be greater than
2 µm. (A micron, µm, is 10-6m).

1.3 Broad Classification

1.3.1 Coarse-grained soils

These include sands, gravels and larger particles. For these


soils the grains are well defined and may be seen by the naked
eye. The individual particles may vary from perfectly round to
highly angular reflecting their geological origins.

1.3.2 Fine-grained soils

These include the silts and clays and have particles smaller than
60 µm.

• Silts These can be visually differentiated from clays


because they exhibit the property of dilatancy. If a moist
sample is shaken in the hand water will appear on the
surface. If the sample is then squeezed in the fingers the
water will disappear. Their gritty feel can also identify silts.

• Clays Clays exhibit plasticity, they may be readily


remoulded when moist, and if left to dry can attain
high strengths

• Organic These may be of either clay or silt sized particles.


They contain significant amounts of vegetable
matter. The soils as a result are usually dark
grey or black and have a noticeable odour from decaying
matter. Generally only a surface phenonomen but layers
of peat may be found at depth. These are very poor
soils for most engineering purposes.

1.4 Procedure for grain size determination


7

Different procedures are required for fine and coarse-grained


material. Detailed procedures are described in the Australian
Standard AS 1289.A1, Methods of testing soil for engineering
purposes. These will be demonstrated in a laboratory session.

• Coarse Sieve analysis is used to determine the distribution


of the larger grain sizes. The soil is passed through a
series of sieves with the mesh size reducing
progressively, and the proportions by weight of the
soil retained on each sieve are measured. There are a
range of sieve sizes that can be used, and the finest
is usually a 75 µm sieve. Sieving can be performed
either wet or dry. Because of the tendency for fine
particles to clump together, wet sieving is often
required with fine-grained soils.

• Fine To determine the grain size distribution of material


passing the 75µm sieve the hydrometer method is
commonly used. The soil is mixed with water and a
dispersing agent, stirred vigorously, and allowed to
settle to the bottom of a measuring cylinder. As the
soil particles settle out of suspension the specific
gravity of the mixture reduces. An hydrometer is
used to record the variation of specific gravity with
time. By making use of Stoke’s Law, which relates
the velocity of a free falling sphere to its diameter,
the test data is reduced to provide particle diameters
and the % by weight of the sample finer than a
particular particle size.
8

Figure 1 A schematic view of the hydrometer test

1.5 Grading curves

The results from the particle size determination tests are plotted
as grading curves. These show the particle size plotted against
the percentage of the sample by weight that is finer than that
size. The results are presented on a semi-logarithmic plot as
shown in Figure 2 below. The shape and position of the grading
curve are used to identify some characteristics of the soil.
9

100

80
% Finer

60

40

20

0
0.0001 0.001 0.01 0.1 1 10 100
Particle size (mm)

Figure 2 Typical grading curves

Some typical grading curves are shown on the figure. The


following descriptions are applied to these curves

W Well graded material


U Uniform material
P Poorly graded material
C Well graded with some clay
F Well graded with an excess of fines

The use of names to describe typical grading curve shapes and


positions has developed as the suitability of different gradings
for different purposes has become apparent. For example, well
graded sands and gravels can be easily compacted to relatively
high densities which result in higher strengths and stiffnesses.
For this reason soils of this type are preferred for road bases.
The suitability of different gradings is discussed in some detail
by Terzaghi and Peck (1967).
10

From the typical grading curves it can be seen that soils are
rarely all sand or all clay, and in general will contain particles
with a wide range of sizes. Many organisations have produced
charts to classify soils giving names for the various
combinations of particle sizes. One such example is given in
Figure 3 below.

0
100
10
90
20
80
30
70
40 Clay
)

Cla
(%

60

y
es

50

Siz
Siz

50

e s (%
d

60
Sa n

Sandy Clay Silty Clay 40

)
70
30
80 Clay-Sand Clay-Silt
20
90
Silty Sand Sandy Silt 10
100 Sand
0
0 10 20 30 40 50 60 70 80 90 100
Silt Sizes (%)
LOWER MISSISSIPPI VALLEY DIVISION,
U. S. ENGINEER DEPT.

Figure 3 Classification Chart

Important observations from figure 3 are that any soil


containing more than 50% of clay sized particles would be
classified as a clay, whereas sand and silt require 80% of the
particles to be in that size range. Also any soil having more
than 20% clay would have some clay like properties.

The hydrometer test is usually terminated when the percentage


of clay sized particles has been determined. However, there are
significant differences between the behaviour of the different
clay minerals. To provide additional information on the soil
behaviour further classification tests are performed. One such
11

set of tests, the Atterberg Limit Tests, involve measuring the


moisture contents of the soil at which changes in the soil
properties occur.

1.6 Atterberg Limits

These tests are only used for the fine-grained, silt and clay,
fraction of a soil (actually the % passing a 425 µm sieve). If we
take a very soft (high moisture content) clay specimen and
allow it to dry we would obtain a relation similar to that shown
in Figure 4.

As the soil dries its strength and stiffness will increase. Three
limits are indicated, the definitions of which are given below.
The liquid and plastic limits appear to be fairly arbitrary, but
recent research has suggested they are related to the strength of
the soil.

Volume

SL PL LL
Moisture Content (%)

Figure 4. Moisture content versus volume relation

• (SL) The Shrinkage Limit - This is the moisture content the


soil would have had if it were fully saturated at the point at
which no further shrinkage occurs on drying.
12

weightof water w
moisturecontent = = w (1)
weightof solids ws

In the shrinkage test the soil is left to dry and the soil is
therefore not saturated when the shrinkage limit is reached. To
estimate SL it is necessary to measure the total volume, V, and
the weight of the solids, ws. Then

γ wV 1
SL = m = − (2)
ws Gs

where γ w is the unit weight of water, and


Gs is the specific gravity

• (PL) The Plastic Limit - This is the minimum water content


at which the soil will deform plastically

• (LL) The Liquid Limit - This is the minimum water content


at which the soil will flow under a small disturbing force

• (PI or Ip) The Plasticity Index. This is derived simply from


the LL and PL

IP = LL - PL
(3)

• (LI) The Liquidity Index - This is defined as


m − PL m − PL
LI = = (4)
LL − PL Ip
The Atterberg Limits and relationships derived from them are
simple measures of the water absorbing ability of soils
containing clay minerals. For example, if a clay has a very high
LI and LL it is capable of absorbing large amounts of water,
and for instance would be unsuitable for the base of a
pavement. The LL and PL are also related to the soil strength.

Remember that only the fraction finer than 425 µm is tested in


the Atterberg Tests. If this fraction is only small (that is, the
13

soil contains significant amounts of sand or gravel) it might be


expected that the soil would have better properties. While this
is true to some extent it is important to realise that the soil
behaviour is controlled by the finest 10 - 25 % of the particles

1.7 Classification Systems for Soils

Several systems are used for classifying soil. This is because


these systems have two main purposes

1. To determine the suitability of different soils for various


purposes (see p8 Data Sheets)

2. To develop correlations with useful soil properties, for


example, compressibility and strength

The reason for the large number of such systems is the use of
particular systems for certain types of construction, and the
development of localised systems.

1.7.1 PRA (AASHO) system

An example is the PRA system of AASHO (American


Association of State Highway Officials), which ranks soils
from 1 to 8 to indicate their suitability as a subgrade for
pavements. The detailed classification is given in the Data
Sheets p9.

1. Well graded gravel or sand; may include fines


2. Sands and Gravels with excess fines
3. Fine sands
4. Low compressibility silts
5. High compressibility silts
6. Low to medium compressibility clays
7. High compressibility clays
8. Peat, organic soils

1.7.2 Unified Soil Classification


14

The standard system used worldwide for most major


construction projects is known as the Unified Soil
Classification System (USCS). This is based on an original
system devised by Cassagrande. Soils are identified by symbols
determined from sieve analysis and Atterberg Limit tests.

• Coarse Grained Materials

If more than half of the material is coarser than the 75 µm


sieve, the soil is classified as coarse. The following steps are
then followed to determine the appropriate 2 letter symbol

1. Determine the prefix (1st letter of the symbol)

If more than half of the coarse fraction is sand then use prefix S

If more than half of the coarse fraction is gravel then use prefix
G

2. Determine the suffix (2nd letter of symbol)

This depends on the uniformity coefficient Cu and the


coefficient of curvature Cc obtained from the grading curve, on
the percentage of fines, and the type of fines.

First determine the percentage of fines, that is the % of material


passing the 75 µm sieve.

Then if % fines is < 5% use W or P as suffix


> 12% use M or C as suffix
between 5% and 12% use dual symbols. Use
the prefix from above with first one of W or P and
then with one of M or C.

If W or P are required for the suffix then Cu and Cc must be


evaluated
15

D60
Cu =
D10
D302
Cc =
( D60 × D10 )

If prefix is G then suffix is W if Cu > 4 and Cc is


between 1 and 3
otherwise use P

If prefix is S then suffix is W if Cu > 6 and Cc is


between 1 and 3
otherwise use P

If M or C are required they have to be determined from the


procedure used for fine grained materials discussed below.
Note that M stands for Silt and C for Clay. This is determined
from whether the soil lies above or below the A-line in the
plasticity chart shown in Figure 5.

For a coarse grained soil which is predominantly sand the


following symbols are possible

SW, SP, SM, SC


SW-SM, SW-SC, SP-SM, SP-SC
• Fine grained materials

These are classified solely according to the results from the


Atterberg Limit Tests. Values of the Plasticity Index and
Liquid Limit are used to determine a point in the plasticity
chart shown in Figure 5. The classification symbol is
determined from the region of the chart in which the point lies.

Examples CH High plasticity clay


CL Low plasticity clay
MH High plasticity silt
ML Low plasticity silt
OH High plasticity organic soil (Rare)
Pt Peat
16

60
Comparing soils at equal liquid limit
50 Toughness and dry strength increase
l ine
with increasing plasticity index "
"A
40
Plasticity index

CH
30

20 OH
CL or
10 CL OL
or MH
ML
0 ML
0 10 20 30 40 50 60 70 80 90 100
Liquid limit

Figure 5 Plasticity chart forPlasticity chart classification of fine


laboratory
for laboratory classification of fine grained soils
grained soils

The final stage of the classification is to give a description of


the soil to go with the 2-symbol class. For a coarse grained soil
this should include:

• the percentages of sand and gravel


• maximum particle size
• angularity
• surface condition
• hardness of the coarse grains
• local or geological name
• any other relevant information

If the soil is undisturbed mention is also required of

• stratification
• degree of compactness
• cementation
• moisture conditions
• drainage characteristics
17

The information required, along with all the details of the


Unified Classification Procedure is given in Figure 6. Note that
slightly different information is required for fine-grained soils.

Unified soil classification (including identification and description)

Field identification procedures Group Information required for Laboratory classification


(Excluding particles larger than 75mm and basing fractions on symbols Typical names describing soils criteria
estimated weights) 1

Depending on percentages of fines (fraction smaller than .075mm


Wide range of grain size and substantial GW Well graded gravels, gravel- D
Give typical names: indicate ap- C U = --- 60 Greater than 4

Bordeline case requiring use of dual symbols


Gravels with Clean gravels

Determine percentages of gravel and sand from grain size curve


(little or no

amounts of all intermediate particle sand mixtures, little or no proximate percentages of sand D 10
2
fines)

sizes fines and gravel: maximum size: (D30 )


More than half of coarse

C c = --------------
------- Between 1 and 3
fraction is larger than

Predominantly one size or a range of GP Poorly graded gravels, gravel- angularity, surface condition, D10 x D60

sieve size) coarse grained soils are classified as follows


sizes with some intermediate sizes sand mixtures, little or no and hardness of the coarse
More than half of material is larger than

missing fines grains: local or geological name Not meeting all gradation requirements for GW
2.36mm
Gravels

and other pertinent descriptive


amount of fines)

Use grain size curve in identifying the fractions as given under field identification
Non-plastic fines (for identification GM Silty gravels, poorly information and symbol in
(apreciable

procedures see ML below) graded gravel-sand-silt mixtures parentheses. Atterberg limits below Above "A" line with
fines
The .075mm sieve size is about the smallest particle visible to the naked eye

"A" line or PI less than 4 PI between 4 and 7


Coarse grained soils
.075mm sieve size

Plastic fines (for identification pro- GC Clayey gravels, poorly graded For undisturbed soils add infor- are borderline cases
cedures see CL below) gravel-sand-clay mixtures mation on stratification, degree Atterberg limits above "A" requiring use of dual

More than 12% GM, GC, SM, SC


line with PI greater than 7 symbols

GW, GP, SW, SP


of compactness, cementation,
Wide range in grain sizes and sub- moisture conditions and drain- D
SW Well graded sands, gravelly
stantial amounts of all intermediate age characteristics. C U=--- 60 Greater than 6
Clean sands
(little or no

particle sizes sands, little or no fines D 10


fines)
More than half of coarse

(D 30 )2
fraction is smaller than

Example:
Predominantely one size or a range of SP Poorly graded sands, gravelly C c = --------------
-------- Between 1 and 3
Silty sand, gravelly; about 20% D 10 x D60
sizes with some intermediate sizes missing sands, little or no fines hard angular gravel particles
2.36mm

12.5mm maximum size; rounded Not meeting all gradation requirements for SW
Sands

amount of fines)

Non-plastic fines (for identification pro- SM Silty sands, poorly graded and subangular sand grains
(appreciable

Less than 5%
Sands with

cedures, see ML below) sand-silt mixtures coarse to fine, about 15% non- Atterberg limits below Above "A" line with

5% to 12%
plastic lines with low dry PI between 4 and 7
fines

"A" line or PI less than 4


Plastic fines (for identification pro- SC Clayey sands, poorly graded strength; well compacted and are borderline cases
cedures, see CL below) sand-clay mixtures moist in places; alluvial sand; Atterberg limits above "A" requiring use of dual
(SM) line with PI greater than 7 symbols
Identification procedure on fraction smaller than .425mm
sieve size
Dry strength Dilatency Toughness
More than half of material is smaller than

crushing (consistency
Silts and clays

(reaction
less than 50
liquid limit

character- to shaking) near plastic


istics limit)
Inorganic silts and very fine sands, Give typical name; indicate degree
None to Quick to None rock flour, silty or clayey
slight slow ML and character of plasticity, 60
Fine grained soils
.075mm sieve size

fine sands with slight plasticity


amount and maximum size of Comparing soils at equal liquid limit
Medium to None to very Inorganic clays of low to medium coarse grains: colour in wet con-
Medium CL,CI plasticity, gravelly clays, sandy 50 Toughness and dry strength increase
in
e
high slow clays, silty clays, lean clays dition, odour if any, local or "l
with increasing plasticity index
geological name, and other pert- "A
Slight to Organic silts and organic silt- inent descriptive information, and 40
Plasticity index

medium Slow Slight OL clays of low plasticity CH


symbol in parentheses
30 CI
inorganic silts, micaceous or
Silts and clays

Slight to Slow to Slight to For undisturbed soils add infor-


greater than

dictomaceous fine sandy or


liquid limit

medium none medium MH mation on structure, stratif- OH


silty soils, elastic silts 20
ication, consistency and undis-
50

High to very Inorganic clays of high or


None High turbed and remoulded states, CL OL
high CH plasticity, fat clays moisture and drainage conditions 10
MH
CL-ML or
ML
Medium to None to very Slight to Organic clays of medium to Example 0
high high medium OH high plasticity Clayey silt, brown: slightly plastic: 0 10 20 30 40 50 60 70 80 90 100
Liquid limit
Readily identified by colour, odour small percentage of fine sand:
Highly organic soils spongy feel and frequently by fibrous Pt Peat and other highly organic soils numerous vertical root holes: firm Plasticity chart
texture and dry in places; loess; (ML) for laboratory classification of fine grained soils

Figure 6 Unified Soil Classification Chart


18

Example - Classification using USCS

Classification tests have been performed on a soil sample and


the following grading curve and Atterberg limits obtained.

100

80
% Finer

60

40

20

0
0.0001 0.001 0.01 0.1 1 10 100
Particle size (mm)
Determine the USCS classification.

Atterberg limits:Liquid limit LL = 32, Plastic Limit, PL =26

Step 1: Determine the % fines from the grading curve

%fines (% finer than 75 µm) = 11% - Coarse grained,


Dual symbols required

Step 2: Determine % of different particle size fractions (to


determine G or S), and D10, D30, D60 from grading curve
(to determine W or P)

D10 = 0.06 mm, D30 = 0.25 mm, D60 = 0.75 mm

Cu = 12.5, Cc = 1.38, and hence Suffix1 = W

Particle size fractions: Gravel 17%


Sand 73%
Silt and Clay 10%
19

Of the coarse fraction about 80% is sand, hence Prefix is


S

Step 3: From the Atterberg Test results determine its Plasticity


chart location

LL = 32, PL = 26. Hence Plasticity Index Ip = 32 - 26 = 6

From Plasticity Chart point lies below A-line, and hence


Suffix2 = M

Step 4: Dual Symbols are SW-SM

Step 5: Complete classification by including a description of


the soil

2. BASIC DEFINITIONS AND TERMINOLOGY

Soil is a three phase material which consists of


solid particles which make up the soil skeleton and
voids which may be full of water if the soil is
saturated, may be full of air if the soil is dry, or
may be partially saturated as shown in Figure 1.

Solid

Water

Air

Figure 1: Air, Water and Solid phases in a typical


soil

It is useful to consider each phase individually as


shown in Table 1.
20

Phase Volume Mass Weight


Air VA 0 0
Water VW MW WW
Solid VS MS WS

Table 1 Distribution by Volume, Mass, and Weight

2.1 Units

For most engineering applications the following


units are used:

Length metres
Mass tonnes (1 tonne = 103
kg)
Density (mass/unit volume) t/m3
Weight kilonewtons (kN)
Stress kilopascals (kPa) 1 kPa
2
= 1 kN/m
Unit Weight kN/m3

To sufficient accuracy the density of water ρ w is


given by

ρw = 1 tonne/m3
= 1 g/cm3

In most applications it is not the mass that is


important, but the force due to the mass, and the
weight, W, is related to the mass, M, by the
relation

W = Mg

where g is the acceleration due to gravity. If M is


measured in tonnes and W in kN, g = 9.8 m/s2
21

Because the force is usually required it is often


convenient in calculations to use the unit weight, γ
(weight per unit volume).

W
γ =
V

Mg
γ =
V

= ρg

Hence the unit weight of water, γ w = 9.8 kN/m3

2.2 Specific Gravity

Another frequently used quantity is the Specific


Gravity, G, which is defined by

Densityof Material ρ
G = =
Densityof Water ρw

Unit Weight of Material γ


G = =
Unit Weight of Water γw

It is often found that the specific gravity of the


materials making up the soil particles are close to
the value for quartz, that is

Gs ≈ 2.65

For all the common soil forming minerals 2.5 <


Gs < 2.8

We can use Gs to calculate the density or unit


weight of the solid particles
22

ρs = Gs ρw
γ s = Gs γ w

and hence the volume of the solid particles if the


mass or weight is known

Ms Ws
Vs = =
Gs ρ w Gs γ w

2.3 Voids Ratio and Porosity

Using volumes is not very convenient in most


calculations. An alternative measure that is used is
the voids ratio, e. This is defined as the ratio of the
volume of voids, Vv to the volume of solids, Vs, that
is

Vv
e =
Vs
where Vv = Vw +Va

V = Va + V w + V s

A related quantity is the porosity, n, which is


defined as ratio of the volume of voids to the total
volume.

Vv
n =
V

The relation between e and n can be determined


by noting that

Vs = V - Vv = (1 - n) V
23

Now

Vv Vv n
e = = =
Vs (1 − n) V 1− n

and hence
e
n =
1+e

2.4 Degree of Saturation

The degree of saturation, S, has an important


influence on the soil behaviour. It is defined as the
ratio of the volume of water to the volume of voids

Vw
S =
Va + Vw

The distribution of the volume phases may be


expressed in terms of e and S, and by knowing the
unit weight of water and the specific gravity of the
particles the distributions by weight may also be
determined as indicated in Table 3.

Vw V
S = = w
Vv eVs

V w = e S Vs

Va = Vv - Vw = e Vs (1 - S)

Phase Volume Mass Weight


Air e (1 - S) 0 0
Water eS e S ρw e S γw
24

Solid 1 Gs ρw Gs γ w

Table 2 Distribution by Volume, Mass and Weight


in Soil

Note that Table 2 assumes a solid volume Vs = 1


m3, All terms in the table should be multiplied by
Vs if this is not the case.

2.5 Unit Weights

Several unit weights are used in Soil Mechanics.


These are the bulk, saturated, dry, and submerged
unit weights.

The bulk unit weight is simply defined as the


weight per unit volume

W
γ bulk =
V

When all the voids are filled with water the bulk
unit weight is identical to the saturated unit
weight, γ sat, and when all the voids are filled with
air the bulk unit weight is identical with the dry
unit weight, γ dry. From Table 2 it follows that

W γ G + γ w eS γ (G + e S )
γ bulk = = w s = w s
V 1+ e 1+ e

γ w ( G s + e)
γ sat = S=1
1+ e

γ w Gs
γ dry = S=0
1+ e

Note that in discussing soils that are saturated it is


common to discuss their dry unit weight. This is
25

done because the dry unit weight is simply related


to the voids ratio, it is a way of describing the
amount of voids.

The submerged unit weight, γ ´, is sometimes


useful when the soil is saturated, and is given by

γ ´ = γ sat - γ w

2.6 Moisture content

The moisture content, m, is a very useful quantity


because it is simple to measure. It is defined as the
ratio of the weight of water to the weight of solid
material

Ww
m =
Ws

If we express the weights in terms of e, S, Gs and


γ w as before we obtain

W w = γ w V w = γ w e S Vs

Ws = γ s V s = γ w Gs V s

and hence

eS
m =
Gs

Note that if the soil is saturated (S=1) the voids


ratio can be simply determined from the moisture
content.

Example – Mass and Volume fractions


26

A sample of soil is taken using a thin walled


sampling tube into a soil deposit. After the soil is
extruded from the sampling tube a sample of
diameter 50 mm and length 80 mm is cut and is
found to have a mass of 290 g. Soil trimmings
created during the cutting process are weighed
and found to have a mass of 55 g. These
trimmings are then oven dried and found to have a
mass of 45 g. Determine the phase distributions,
void ratio, degree of saturation and relevant unit
weights.

1. Distribution by mass and weight

Phase Trimmin Sample Sample


gs Mass Mass, M Weight, Mg
(g) (g) (kN)
Total 55 290 2845 × 10-6
Solid 45 237.3 2327.9 × 10-6
Water 10 52.7 517 × 10-6

2. Distribution by Volume

Sample Volume, V = π (0.025)2 (0.08) = 157.1 ×


10-6 m3

Ww 517 × 10 −6
Water Volume , Vw = = = 52.7 × 10 −6 m 3
γw 9.81
Ws 2327.9 × 10 −6
Solid Volume , Vs = = = 89.5 × 10 −6 m 3
Gs γ w 2.65 × 9.81

Air Volume, Va = V - Vs - Vw = 14.9 × 10-6 m3

3. Moisture content

Ww 10 52.7 × 10 −6
m = = = = 0.222
Ws 45 237.3 × 10 −6
27

4. Voids ratio

Vv 14.9 × 10 −6 + 52.7 × 10 −6
e = = = 0.755
Vs 89.5 × 10 −6

5. Degree of Saturation

Vw 52.7 × 10 −6
S = = = 0.780
Vv 52.7 × 10 −6 + 14.9 × 10 −6

6. Unit weights

W 2845 × 10 −6 kN
γ bulk = = = 18.1 kN / m 3
V 157 .1 × 10 −6 m 3

Ws 2327.9 × 10 −6
γ dry = = = 14.8 kN / m 3
V 157.1 × 10 −6

If the sample were saturated there would need to


be an additional 14.9 × 10-6 m3 of water. This
would weigh 146.2 × 10-6 kN and thus the
saturated unit weight of the soil would be

(2845 × 10 −6 + 146.2 × 10 −6 )
γ sat = = 19.04 kN / m 3
157.1 × 10 −6

Example – Calculation of Unit Weights

A soil has a voids ratio of 0.7. Calculate the dry


and saturated unit weight of the material. Assume
that the solid material occupies 1 m3, then
assuming Gs = 2.65 the distribution by volume and
weight is as follows.

Phase Volume Dry Weight Saturated


(m3) (kN) Weight
(kN)
Voids 0.7 0 0.7 × 9.81
28

= 6.87
Solids 1.0 2.65 × 9.81 26.0
= 26.0

26.0 kN
• Dry unit weight γ dry = = 15.3 kN / m 3
. m3
17

(26.0 + 6.87)
• Saturated unit weight γ sat =
17
.
= 19.3 kN / m 3

If the soil were fully saturated the moisture content


would be

6.87
• Moisture content m =
26.0
= 0.264 = 26.4%

Alternatively the unit weights may be calculated


from the expressions given earlier which are on p.
5 of the Data Sheets

Gs γ w
γ dry =
1+ e
( G s + e) γ
γ sat = w

1+ e
29

3. COMPACTION

Compaction is the application of mechanical


energy to a soil to rearrange the particles and
reduce the void ratio.

3.1 Purpose of Compaction

• The principal reason for compacting soil is to


reduce subsequent settlement under working
loads.

• Compaction increases the shear strength of the


soil.

• Compaction reduces the voids ratio making it


more difficult for water to flow through soil. This
is important if the soil is being used to retain
30

water such as would be required for an earth


dam.

• Compaction can prevent the build up of large


water pressures that cause soil to liquefy during
earthquakes.

3.2 Factors affecting Compaction

• Water content of the soil

• The type of soil being compacted

• The amount of compactive energy used

3.3 Laboratory Compaction tests

There are several types of test which can be used


to study the compactive properties of soils.
Because of the importance of compaction in most
earth works standard procedures have been
developed. These generally involve compacting
soil into a mould at various moisture contents.

• Standard Compaction Test AS 1289-E1.1

Soil is compacted into a mould in 3-5 equal layers,


each layer receiving 25 blows of a hammer of
standard weight. The apparatus is shown in Figure
1 below. The energy (compactive effort) supplied
in this test is 595 kJ/m3. The important dimensions
are

Volume of Hammer mass Drop of


mould hammer
1000 cm3 2.5 kg 300 mm
31

Because of the benefits from compaction,


contractors have built larger and heavier machines
to increase the amount of compaction of the soil. It
was found that the Standard Compaction test
could not reproduce the densities measured in the
field and this led to the development of the
Modified Compaction test.

• Modified Compaction Test AS 1289-E2.1

The procedure and equipment is essentially the


same as that used for the Standard test except
that 5 layers of soil must be used. To provide the
increased compactive effort (energy supplied =
2072 kJ/m3) a heavier hammer and a greater drop
height for the hammer are used. The key
dimensions for the Modified test are

Volume of Hammer mass Drop of


mould hammer
1000 cm3 4.9 kg 450 mm

Handle

collar
(mould
extension)

Metal guide to
control drop of
hammer

Cylindrical
soil mould

Hammer for
compacting soil
32

Base plate

Figure 1 Apparatus for laboratory compaction


tests

3.4 Presentation of Results

To assess the degree of compaction it is important


to use the dry unit weight, γ dry, because we are
interested in the weight of solid soil particles in a
given volume, not the amount of solid, air and
water in a given volume (which is the bulk unit
weight). From the relationships derived previously
we have

Gs γ w
γ dry =
1+ e
which can be rearranged to give
Gs γ w
e = −1
γ dry

Because Gs and γ w are constants it can be seen


that increasing dry density means decreasing
voids ratio and a more compact soil.

In the test the dry density cannot be measured


directly, what are measured are the bulk density
and the moisture content. From the definitions we
have

W to f S o lid s Ws W to fW a te r Ww
γ dry = = m= =
T o taVl o lu m e V W to f S o lid s Ws

W Wt of Solids + Wt of Water W + Ww
γ bulk = = = s
V TotalVolume V
33

(1 + m) Ws
=
V

= (1 + m) γ dry

This allows us to plot the variation of dry unit


weight with moisture content, giving the typical
reponse shown in Figure 2 below. From this graph
we can determine the optimum moisture content,
mopt, for the maximum dry unit weight, (γ dry)max.

(γdry)
Dry unit weight

max

mopt

Moisture content

Figure 2 A typical compaction test result

If the soil were to contain a constant percentage,


A, of voids containing air where

Va
A (% ) = × 100
V

writing Va as V - Vw - Vs we obtain
34

A V + Vs
1− = w
100 V

then a theoretical relationship between γ dry and m


for a given value of A can be derived as follows

A
(Ws + Ww ) (1 − )
γ W + Ww 100
γ dry = bulk = s =
1+ m V (1 + m) (Vs + Vw ) (1 + m)

Ws Ww mWs
Now Vs = Vw = =
Gs γ w γw γw

A  Gs γ w 
Hence γ dry = (1 − ) 
100  Gs m + 1

If the percentage of air voids is zero, that is, the


soil is totally saturated, then this equation
becomes

 G γ 
γ dry =  s w 
 G s m + 1

From this equation we see that there is a limiting


dry unit weight for any moisture content and this
occurs when the voids are full of water. Increasing
the water content for a saturated soil results in a
reduction in dry unit weight. The relation between
the moisture content and dry unit weight for
saturated soil is shown on the graph in Figure 3.
This line is known as the zero air voids line.
35

ze
ro
Dry unit weight -a
ir-
vo
id
s
lin
e

Moisture content
Figure 3 Typical compaction curve showing no-
air-voids line

3.5 Effects of water content during


compaction

As water is added to a soil ( at low moisture


content) it becomes easier for the particles to
move past one another during the application of
the compacting forces. As the soil compacts the
voids are reduced and this causes the dry unit
weight ( or dry density) to increase. Initially then,
as the moisture content increases so does the dry
unit weight. However, the increase cannot occur
indefinitely because the soil state approaches the
zero air voids line which gives the maximum dry
unit weight for a given moisture content. Thus as
the state approaches the no air voidsline further
moisture content increases must result in a
reduction in dry unit weight. As the state
approaches the no air voids line a maximum dry
unit weight is reached and the moisture content at
this maximum is called the optimum moisture
content.
36

3.6 Effects of increasing compactive effort

Increased compactive effort enables greater dry


unit weights to be achieved which because of the
shape of the no air voids line must occur at lower
optimum moisture contents. The effect of
increasing compactive energy can be seen in
Figure 4. It should be noted that for moisture
contents greater than the optimum the use of
heavier compaction machinery will have only a
small effect on increasing dry unit weights. For this
reason it is important to have good control over
moisture content during compaction of soil layers
in the field.

ze
increasing compactive ro
-a
energy ir-
Dry unit weight

vo
id
s
lin
e

Moisture content

Figure 4 Effects of compactive effort on


compaction curves

It can be seen from this figure that the compaction


curve is not a unique soil characteristic. It depends
on the compaction energy. For this reason it is
important when giving values of (γ dry)max and mopt
to also specify the compaction procedure (for
example, standard or modified).
37

3.7 Effects of soil type

The table below contains typical values for the


different soil types obtained from the Standard
Compaction Test.

Typical Values
(γ dry )max mopt (%)
3
(kN/m )
Well graded 22 7
sand SW
Sandy clay 19 12
SC
Poorly graded 18 15
sand SP
Low plasticity 18 15
clay CL
Non plastic silt 17 17
ML
High plasticity 15 25
clay CH

Note that these are typical values. Because of the


variability of soils it is not appropriate to use
typical values in design, tests are always required.

3.8 Field specifications

To control the soil properties of earth constructions


(e.g. dams, roads) it is usual to specify that the soil
must be compacted to some pre-determined dry
unit weight. This specification is usually that a
certain percentage of the maximum dry density, as
38

found from a laboratory test (Standard or Modified)


must be achieved.

For example we could specify that field densities


must be greater than 98% of the maximum dry
unit weight as determined from the Standard
Compaction Test. It is then up to the Contractor to
select machinery, the thickness of each lift (layer
of soil added) and to control moisture contents in
order to achieve the specified amount of
compaction.

Acce Accept
Reject
pt
Dry unit weight

Reje
ct

Moisture content
(a)
(b)
Figure 5 Possible field specifications for
compaction

There is a wide range of compaction equipment.


For pavements some kind of wheeled roller or
vibrating plate is usually used. These only affect a
small depth of soil, and to achieve larger depths
vibrating piles and drop weights can be used. The
applicability of the equipment depends on the soil
type as indicated in the table below
39

Equipment Most Typical Least


suitable application suitable
soils soils
Smooth Well Running Uniform
wheeled graded surface, sands
rollers, sand- base
static or gravel, courses,
vibrating crushed subgrades
rock,
asphalt
Rubber Coarse Pavement Coarse
tired rollers grained subgrade uniform
soils with soils and
some fines rocks
Grid rollers Weathered Subgrade, Clays, silty
rock, well subbase clays,
graded uniform
coarse soils materials
Sheepsfoot Fine Dams, Coarse
rollers, grained embankme soils, soils
static soils with > nts, with
20% fines subgrades cobbles,
stones
Sheepsfoot as above, subgrade
rollers, but also layers
vibratory sand-gravel
mixes
Vibrating Coarse Small clays and
plates soils, 4 to patches silts
8% fines
Tampers, All types Difficult
rammers access
areas
Impact Most Dry, sands
rollers saturated and gravels
and moist
soils
40

3.9 Sands and gravels

For soils without any fines (sometimes referred to


as cohesionless) the standard compaction test is
difficult to perform. For these soil types it is normal
to specify a relative density, Id, that must be
achieved. The relative density is defined by

emax − e
Id =
emax − em in

where e is the current voids ratio,


emax, emin are the maximum and minimum
voids ratios measured in the laboratory from
Standard Tests (AS 1289-5.1)

Note that if e = emin, Id = 1 and the soil is in its


densest state
e = emax, Id = 0 and the soil is in its loosest
state

The expression for relative density can also be


written in terms of the dry unit weights associated
with the various voids ratios. From the definitions
we have

Gs γ w
e = − 1
γ dry

and hence

1 1

γ drymin γ dry γ drym ax (γ dry − γ drym in )
Id = =
1 1 γ dry (γ drym ax − γ drym in )

γ drymin γ drym ax
41

The description of the soil will include a description


of the relative density. Generally the terms loose,
medium and dense are used where

Loose 0 < Id < 1/3


Medium 1/3 < Id < 2/3
Dense 2/3 < Id < 1

Note that you cannot determine the unit weight


from knowing Id. This is because the values of the
maximum and minimum dry unit weights (void
ratios) can vary significantly. They depend on soil
type (mineralogy), the particle grading, and the
angularity.
42

4. EFFECTIVE STRESS

4.1 Saturated Soil

A saturated soil is a two phase material consisting


of a soil skeleton and voids which are saturated
with water. It is reasonable to expect that the
behaviour of an element of such a material will be
influenced not only by the forces applied to its
surface but also by the water pressure of the fluid
in the pores.

Suppose that a soil sample having a uniform cross


sectional area A is subjected to an applied load W,
as shown in Fig la, then it is found that the soil will
deform. If however, the sample is loaded by
increasing the height of water in the containing
vessel, as shown in Fig lb, then no deformation
occurs.

W
W

Soil Soil

Fig(1a) Soil loaded by an applied weight W Fig(1b) Soil loaded by water weighing W
43

In examining the reasons for this observed


behaviour it is helpful to use the following
quantities:
Vertical Force
σ v = Vertical Stress =
Cross Sectional Area (1)

and to define an additional quantity the vertical


effective stress, by the relation

σv′ = σ v − u w
(2)

Let us examine the changes the vertical stress,


pore water pressure and vertical effective stress
for the two load cases considered above.

∆σv ∆uw ∆σv´


Case (a) W 0 W
A A
Case (b) W W 0
A A
These experiments indicate that if there is no
change in effective stress there is no change in
deformation, or alternatively that deformation only
occurs when there is a change in effective stress.

Another situation in which effective stresses are


important is the case of two rough blocks sliding
over one another, with water pressure in between
them as shown in Fig 2.
44

Fig 2 Two pieces of Rock in contact


The effective normal thrust transmitted through
the points of contact will be

N′ = N − U
(3a)

where U is the force provided by the water


pressure

The frictional force will then be given by T = µ N ′


where µ is the coefficient of friction. For soils and
rocks the actual contact area is very small
compared to the cross-sectional area so that U/A is
approximately equal to uw the pore water pressure.
Hence dividing through by the cross sectional area
A this becomes:

τ = µσv′
(3b)

where τ is the average shear stress and σ′ v is the


vertical effective stress.

Of course it is not possible to draw a general


conclusion from a few simple experiments, but
there is now a large body of experimental
evidence to suggest that both deformation and
45

strength of soils depend upon the effective stress.


This was originally suggested by Terzaghi in the
1920’s, and equation 2 and similar relations are
referred to as the Principle of Effective Stress.

4.2 Calculation of Effective Stress

It is clear from the definition of effective stress


that in order to calculate its value it is necessary to
know both the total stress and the pore water
pressure. The values of these quantities are not
always easy to calculate but there are certain
simple situations in which the calculation is quite
straightforward. The most important is when the
ground surface is flat as is often the case with
sedimentary (soil) deposits.
4.2.1 Calculation of Vertical (Total) Stress

Consider the horizontally "layered" soil deposit


shown schematically in Fig.3,
Surcharge q

Layer 1 γ bulk = γ 1 d1

Layer 2 γ = γ z d2
bulk 2

Layer 3
γ bulk = γ 3 d3
σ v

Fig 3 Soil Profile

If we consider the equilibrium of a column of soil of


cross sectional area A it is found that
46

Force on base = Force on Top + Weight of Soil


Aσ v = Aq + Aγ 1d 1 + Aγ 2 d 2 + Aγ 3 ( z − d 1 − d 2 ) (4)
σv = q+ γ 1d 1 + γ 2d 2 + γ 3 ( z − d1 − d 2 )

 Calculation of Pore Water Pressure

Water table

Fig 4 Soil with a static water table


Suppose the soil deposit shown in Fig. 4 has a
static water table as indicated. The water table is
the water level in a borehole, and at the water
table uw = 0. The water pressure at a point P is
given by

u w (P) = γ wH
(5)

Example

A uniform layer of sand 10 m deep overlays


bedrock. The water table is located 2 m below the
surface of the sand which is found to have a voids
ratio e = 0.7. Assuming that the soil particles have
a specific gravity Gs = 2.7 calculate the effective
stress at a depth 5 m below the surface.

Step one: Draw ground profile showing soil


stratigraphy and water table
47

Layer 1 γ bulk = γ 1 2m
Water Table
Layer 2 3m
γ bulk = γ 2

Fig 5 Soil Stratigraphy


Step two: Calculation of Dry and Saturated Unit
Weights

Ww = Vv * γ w kN
Voids Vv=e Vs Voids Ww=0 Voids = 0.7 * 9.8 kN
= 0.7m3 = 6.86 kN

Ws = Vs * G s * γ w Ws = Vs * G s * γ w
Solid Vs= 1m3 Solid Solid
= 1 * 2.7 * 9.8 kN = 1 * 2.7 * 9.8 kN
= 26.46 kN = 26.46 kN

Distribution by Volume Distribution by Weight Distribution by weight


for the dry soil for the saturated soil

Fig 6 Calculation of dry and saturated unit weight

26.46 kN
γ dry = = 15.56 kN / m3
. m3
170

(26.46 + 6.86) kN
γ sat = = 19.60 kN / m3
. m3
170
(6)

Step three: Calculation of (Total) Vertical Stress


48

σv = 15 .56 × 2 +19 .60 ×3 = 89 .92 kPa (kN / m 2 )


(7)

Step four: Calculation of Pore Water Pressure

u w = 3 × 9.8 = 29 .40 kPa


(8)

Step five: Calculation of Effective Vertical Stress

σ v′ = σ v − u w = 89 .92 − 29 .40 = 60 .52 kPa


(9)

Note that in practice if the void ratio is


known the unit weights are not normally
calculated from first principles considering
the volume fractions of the different phases.
This is often the case for saturated soils
because the void ratio can be simply
determined from

e = m Gs

The unit weights are calculated directly from


the formulae given in the data sheets, that is

Gs γ w
γ dry =
1+e

( Gs + e) γ w
γ sat =
1 + e
49

Effective Stress under general conditions

In general the state of stress in a soil cannot be


described by a single quantity, the vertical stress.
To fully describe the state of stress the nine stress
components (6 of which are independent), as
illustrated in Fig. 7 need to be determined. Note
that in soil mechanics a compression positive sign
convention is used.
σzz

σyz
z
σzy
σxz

σyy
σzx
y
σxy
σyx

σ xx x

Figure 7 Definition of Stress Components

Fig 7 Definition of Stress Components

The effective stress state is then defined by the


relations

σ′xx = σ xx − u w ; σ′yz = σ yz
σ′yy = σ yy − u w ; σ′zx = σ zx
σ′zz = σ zz − u w ; σ′xy = σ xy
(10)
50

Example – Effects of groundwater level


changes

Initially a 50 m thick deposit of a clayey soil has a


groundwater level 1 m below the surface. Due to
groundwater extraction from an underlying aquifer
the regional groundwater level is lowered by 2 m.
By considering the changes in effective stress at a
depth, z, in the clay investigate what will happen
to the ground surface.

Due to decreasing demands for water the


groundwater rises (possible reasons include de-
industrialisation and greenhouse effects) back to
the initial level. What problems may arise?

Assume

• γ bulk is constant with depth


• γ bulk is the same above and below the
water table (clays may remain saturated for
many metres above the groundwater table
due to capillary suctions)

The vertical total and effective stresses at depth z


are given in the Table below.

Initial GWL Lowered GWL


σv γbulk × z γbulk × z
u γw ×( z −1) γw × ( z − 3)
σv´ z × (γbulk − γ w ) − γ w z × (γbulk −γw ) + 3×γw

At all depths the effective stress increases and as


a result the soil compresses. The cumulative
effect throughout the clay layer can produce a
significant settlement of the soil surface.
51

When the groundwater rises the effective stress


will return to its initial value, and the soil will swell
and the ground surface heave (up). However, due
to the inelastic nature of soil, the ground surface
will not in general return to its initial position. This
may result in:

• surface flooding

• flooding of basements built when GWL


was lowered

• uplift of buildings

• failure of retaining structures

• failure due to reductions in bearing


capacity

5. STEADY STATE FLOW

5.1 Introduction

The flow of water in soils can be very significant,


for example:

1. It is important to know the amount of water that


will enter a pit during construction, or the
amount of stored water that may be lost by
percolation through or beneath a dam.

2. The behaviour of soil is governed by the


effective stress, which is the difference between
52

total stress and pore water pressure. When


water flows the pore water pressures in the
ground change. A knowledge of how the pore
water pressure changes can be important in
considering the stability of earth dams, retaining
walls, etc.

5.2 Darcy’s law

Because the pores in soils are so small the flow


through most soils is laminar. This laminar flow is
governed by Darcy's Law which will be discussed
below.

5.2.1 Definition of Head

IMPORTANT
z(P)
z is measured vertically
UP from the DATUM

Datum

Fig 1 Definition of Head at a Point

Referring to Fig. (1) the head h at a point P is


defined by the equation

u w ( P)
h ( P) = + z( P)
γw
(1)
53

In this equation γ w (9.8 kN/m3) is the unit weight of

water, and uw(P) is the pore water pressure .


54

Note

1. The quantity u(P)/γ w is usually called the


pressure head.

2. The quantity z(P) is called the elevation head (its


value depends upon the choice of a datum).

3. The velocity head (not shown in Equation 1) is


generally neglected. The only circumstances
where it may be significant is in flow through
rock-fill, but in this circumstance, the flow will
generally be turbulent and so Darcy's law is not
valid.

Example - Calculation of Head

Static water table 2m


1m X

5m
P
1m
Impermeable stratum

Fig 2 Calculation of head using different datums

Fig 2 Calculation of head using different datum


1. Calculation of Head at P

Datum at the top of the impermeable layer

uw ( P ) = 4γ w
55

z (P) = 1 m
4γ w
h( P ) = + 1 = 5m
γw

2. Calculation of Head at X

Datum at the top of the impermeable layer

u w ( X ) = 1γ w
z (X) = 4 m
γw
h( X ) = + 4 = 5m
γw

It appears that when there is a static water table


the head is constant throughout the saturated
zone.
3. Calculation of Head at P (Datum at the
water table)

uw ( P ) = 4γ w
4γ w
z (P) = - 4 m h( P ) = − 4 = 0m
γw

4. Calculation of Head at X (Datum at the water


table)

u w ( X ) = 1γ w
γw
z (X) = - 1 m h( X ) = − 1 = 0m
γw

When there is a static water table the head is


constant throughout the saturated zone, but its
numerical value depends on the choice of datum.

It is very important to carefully define the datum.


The use of imaginary standpipes can be helpful in
visualising head. The head is then given by the
56

height of the water in the standpipe above the


datum

Note also that it is differences in head (not


pressure) that cause flow

5.2.2 Darcy’s Experiment

∆h

Soil Sample

∆L

Fig 3 Darcy’s Experiment

During his fundamental studies of the flow of water


in soil Darcy found that the flow Q was:

1. Proportional to the head difference ∆h

2. Proportional to the cross sectional are A

3. Inversely proportional to the length ∆L of the soil


sample.

Thus Darcy concluded that:


57

∆h
Q = k A
∆l

(2a)
where k is the coefficient of permeability or
hydraulic conductivity.

Equation (2a) may be rewritten:


Q = k Ai
v = ki
(2b)

where
i = ∆h/∆L is the hydraulic gradient
v = Q/A is the Darcy or superficial velocity.

Note that the actual average velocity of the water


v
in the pores (the groundwater velocity) is n where
n is the porosity. The groundwater velocity is
always greater than the Darcy velocity.

5.3 Measurement of Permeability

5.3.1 Constant Head Permeameter


58

inlet
constant head
device
load

H
Manometers
outlet

device for flow Sampl


sample L
measurement e
porous disk

Fig. 4 Constant Head Permeameter


This is similar to Darcy's experiment. The sample
of soil is placed in a graduated cylinder of cross
sectional area A and water is allowed to flow
through. The discharge X during a suitable time
interval T is collected. The difference in head H
over a length L is measured by means of
manometers
.
From Darcy’s law we obtain
X H
= kA
T L

X L
k =
AH T

(3)

The piston is used to compact the soil because the


permeability depends upon the void ratio
59

5.3.2 Falling Head Permeameter


Standpipe of
cross-sectional
area a

porous disk
H1
H

Sample of H2
L area A

Fig. 5 Falling Head Permeameter

During a time interval δ t

δH
The flow in the standpipe = −a
δt

H
The flow in the sample = k A
L
and thus
d H H
−a = k A
dt L

(4a)

Equation (4a) has the solution:

kA
a ln ( H ) = t + cons tan t
L
(4b)
60

Now initially at time t = t1 the height of water in


the permeameter is H = H1 while at the end of the
test, t = t2 and H = H2 and thus:

 H1 
ln 
H  
aL  2
k =
A (t 2 − t1 )
(4c)

5.4 Typical permeability ranges

Soils exhibit a very wide range of permeabilities


and while particle size may vary by about 3-4
orders of magnitude, permeability may vary by
about 10 orders of magnitude.

10-1 10-2 10-3 10-4 10-5 10-6 10-7 10-8 10-9 10-10 10-11 10

Gravels
Grovels Sands Silts Homogeneous Clays

Fissured & Weathered Clays

Fig 6 Typical Permeability Ranges

Permeability is often estimated from correlations with


particle size. For example

k = ( d 10 ) 2

This expression was first proposed by Hazen in


1893. It is satisfactory for sandy soils but is less
reliable for well graded soils and soils with a large
fines fraction.
61

5.5 Mathematical form of Darcy's law

Because of their geological history soils tend to be


deposited in layers and hence have different flow
properties along the layering and transverse to the
layering.
z(Vertical)

∆z
B C
∆x
O x (Horizontal)

Fig 7 Definition of Hydraulic Gradients

Suppose that the permeability in the horizontal


plane is kH, then the velocity vx in the x direction is
approximately given by:

v x = −k H i x
h ( C ) − h( B )
ix ≈
∆x
∂h
vx = −k H
∂x

(5a)

The negative signs in these equations have been


introduced because flow occurs in the direction of
decreasing head.

Similarly if the permeability in the vertical


direction is kv then the velocity vz is given by

v z = −k v i z
h ( A ) − h( B )
iz ≈
∆z
62

∂h
vz = −k v
∂z

(5b)

Should there also be flow in the y direction this is


similarly governed by
∂h
vy = −k H
∂y
63

5.5.1 Plane Flow

In many situations, as in the dam shown below,


there will be no flow in one direction (usually taken
as the y direction).

Cross section of a long dam


Dam

Soil Flow

Impermeable bedrock

Fig. 8 Plane Flow under a Dam


5.5.2 Continuity Equation

In order to be able to analyse the complex flows


that occur in practice it is necessary to examine
the water entering and leaving an element of soil.
Consider plane flow into the small rectangular box
of soil shown below:
64

vz

Soil
vx D
Element B ∆z

∆ x

Fig 9 Flow into a soil element


Fig. 9 Flow into a soil element

Net flow into element =


( v x ( B ) − v x ( D )) ∆y ∆z + ( v z ( C ) − v z ( A )) ∆x ∆y (6a)

For steady state seepage the flow into the box will
just equal the flow out so the net flow in will be
zero, thus dividing by ∆x∆y∆z and taking the limit
for an infinitesimal element, it is found:

∂v x ∂v z
+ = 0
∂x ∂z
(6b)

When equation (6) is combined with Darcy’s law it


is found that:

∂  ∂h  ∂  ∂h 

k H ∂ x 
 + ∂z 
k v ∂ z 
 = 0
∂x    
(7a)

For a homogeneous material in which the


permeability does not vary with position this
becomes:
65

∂2 h ∂2 h
kH + k v = 0
∂x2 ∂z2
(7b)

and for an isotropic material in which the


permeability is the same in all directions (kH = kv):

∂2 h ∂2 h
+ = 0
∂x2 ∂z2
(7c)

For the more general situation in which there is


flow in three dimensions the continuity equation
becomes:

∂v x ∂v y ∂v z
+ + = 0
∂x ∂y ∂z
(8)

The equation governing seepage then becomes:

∂  ∂h  ∂  ∂h  ∂  ∂h 

kH 
 + 
k H ∂ y 
 + ∂z 
k v ∂z 
 =0
∂x  ∂x  ∂y    
(9a)

For a homogeneous material in which the


permeability does not vary with position (x, y, z)
this becomes:

∂2 h ∂2 h ∂2 h
kH + k H + k v = 0
∂x2 ∂ y2 ∂z2
(9b)

and for an isotropic material in which the


permeability is the same in all directions:

∂2 h ∂2 h ∂2 h
+ + = 0
∂x2 ∂ y2 ∂z2
(9c)
66
67

6. FLOW NETS

6.1 Introduction

Let us consider a state of plane seepage as for


example in the dam shown in Figure 1.

Phreatic line

Unsaturated
Soil Drainage
blanket

z Flow of water

Fig. 1 Flow through a dam

For an isotropic material the head satisfies

∂ ∂
Laplace's equations, thus analysis involves the

∂ ∂
solution of:
2 2
h h
+ =
0
x2 z2
subject to certain boundary conditions.

6.2 Representation of Solution

At every, point (x,z) where there is flow there will


be a value of head h(x,z). In order to represent
these values we draw contours of equal head as
shown on Figure 2.
68

Flow line (FL)

Equipotential (EP)

Fig.2 Flow lines and equipotentials

These lines are called equipotentials. On an


equipotential (EP). by definition:

h( x, z ) = constant
(1a)

it thus follows

∂h ∂h
dx + dz = 0
∂x ∂z
(1b)

and hence the slope of an equipotential is given by

 dz  ∂h / ∂x
 dx  = − ∂h / ∂z
EP

(1c)

It is also useful in visualising the flow in a soil to


plot the flow lines (FL), these are lines that are
tangential to the flow at a given point and are
illustrated in Figure 2.

It can be seen from Fig. (2) that the flow lines and
equipotentials are orthogonal. To show this notice
that on a flow line the tangent at any point is
parallel to the flow at that point so that:
69

[ dx: dz] ∝ [ v x : v z ]
(2a)

it follows immediately that:

 dx  vx
 dz  =
FL vz
now from Darcy ' s law
∂h
vx = −k
∂x
∂h
vz = −k
∂z
thus
 dx  ∂h / ∂x
 dz  =
FL ∂h / ∂z
(2b)

and so

 dx   dx 
 dz  •  dz  = −1
FL EP

(3)

and thus the flow lines and equipotentials are


orthogonal in an isotropic material.
70

6.3 Some Geometric Properties of Flow Nets

Consider a pair of flow lines, clearly the flow


through this flow tube must be constant and so as
the tube narrows the velocity must increase.
Suppose now we have a pair of flow lines as shown
in Figure 3.

h+ ∆h Y
h

Z FL
h+2 ∆h
T
EP

t X ∆Q
y
X x FL
z

∆Q
Fig. 3 Equipotentials intersecting a pair of
Flow Lines

Suppose that the flow per unit width (in the y


direction) is, ∆Q, then the velocity v in the tube is
given by
∆Q
v=
yx

(4a)
Also let us assume that the potential drop between
any adjacent pair of equipotentials is ∆h then it
follows from Darcy’s law:
71

∆h
v =k
zt

(4b)

It thus follows that:

∆Q yx
=
k∆h zt

(4c)
using an identical argument to that used in
developing equation(4c) it can be shown that:

∆Q YX
=
k∆h ZT

(4d)

and hence that:


yx YX
=
zt ZT

(5)
Thus each of the elemental rectangles bounded by
the given pair of flow lines and a pair of
equipotentials (having an equal head drop) have
the same length to breadth ratio.

Next consider a pair of equipotentials cut by flow


tubes each carrying the same flow ∆Q, as shown in
Fig. (4)
72

∆Q B

∆Q C EP( h )
FL D

d A
b
a EP ( h + ∆h )
c

Fig. 4 Flow lines intersecting a pair of


Equipotentials

Then we see that if it is assumed that each of the


tubes is of unit width (in the y direction) then the
velocity in the tube is:
∆Q
v=
cd

(6a)

and using Darcy's law:


∆h
v =k
ab

(6b)
It thus follows that:

∆Q cd
=
k∆h ab

(6c)
It can be similarly shown that:

∆Q CD
=
k∆h AB

(6d)
Hence again a pair of flow tubes carrying equal
flows will intersect a given pair of equipotentials in
73

elemental rectangles which have the same length


to breadth ratio.
In drawing flow nets by hand it is usual to draw
them so that each flow tube carries the same flow
and so that the head drop between adjacent
equipotentials is equal. In such cases all elemental
rectangles will be similar. It is usually most
convenient to draw the net so that these
rectangles are 'square' (it is possible to draw an
inscribed circle). This is illustrated in Fig.(5).

Fig. 5 Inscribing circles in a Flow Net


Fig. 5 Inscribing Circles in a Flow Net
To calculate quantities of interest, that is the flow
and pore pressures, a flow net must be drawn.
The flow net must consist of two families of
orthogonal lines that ideally define a square mesh,
and that also satisfy the boundary conditions. The
three most common boundary conditions are
discussed below.

6.4 Common boundary conditions

6.4.1 Submerged soil boundary - Equipotential

Consider the submerged soil boundary shown in


Figure 6
74

Water
H-z
H

Datum

Fig. 6 Equipotential boundary

The head at the indicated position is calculated as


follows:
uw
h = +z
γw
now
uw = γ w (H − z)
so
( H − z )γ w
h = +z=H
γw
(7)
That is, the head is constant for any value of z,
which is by definition an equipotential.
Alternatively, this could have been determined by
considering imaginary standpipes placed at the
soil boundary, as for every point the water level in
the standpipe would be the same as the water
level. The upstream face of the dam shown in
Figures 1 and 2 is an example of this situation.

6.4.2 Flow Line

At a boundary between permeable and


impermeable material the velocity normal to the
boundary must be zero since otherwise there
75

would be water flowing into or out of the


impermeable material, this is illustrated in Figure
7.

Permeable Soil
vn=0
Flow Line
vt

Impermeable Material

Fig. 7 Flow line boundary

The phreatic surface shown in Figures 2 and 8 is


also a flow line marking the boundary of the flow
net. A phreatic surface is also a line of constant
(zero) pore pressure as discussed below.

6.4.3 Line of Constant Pore Pressure

Sometimes a portion of saturated soil is in contact


with air and so the pore pressure of the water just
beneath that surface is atmospheric. The phreatic
surface shown in Figure 8 below is an example of
such a condition. We can show from the expression
for head in terms of pore pressure that
equipotentials intersecting a line of constant pore
pressure do so at equal vertical intervals as
follows:
76

uw
h= +z
γw
thus
∆u w
∆h = + ∆z
γw
now ∆u w = 0

and so
∆h = ∆z
(8)

Fig. 8 Constant pore pressure boundary

6.5 Procedure for Drawing Flow Nets

1. Mark all boundary conditions

2. Draw a coarse net which is consistent with the


boundary conditions
and which has orthogonal equipotential and
flow lines. ( It is usually
77

easier to visualise the pattern of flow so start


by drawing the flow lines).

3. Modify the mesh so that it meets the


conditions outlined above and so
that the rectangles between adjacent flow
lines and equipotentials are
square.

4. Refine the flow net by repeating step 3.


78

6.6 Calculation of Quantities of Interest from


Flow Nets

6.6.1 Calculation of Increment of Head

In most problems we know the head difference (H)


between inlet and outlet and thus:
H
∆h =
Number of potential drops .
(9)

15 m

h = 15m h=0
5m
P

h = 3m
h = 12m h = 9m h = 6m

Fig. 9 Value of Head on Equipotentials

For example let us assume that the depth of water


retained by the dam is 15 m, and that downstream
of the dam the water table is level with the ground
surface. For this case it can be seen that the total
head drop is 15 m. Inspection of Fig. 2 or Fig. 9
shows that the are 5 potential drops and hence the
head drop between each pair of potentials is ∆h =
15/5 = 3 m.

6.6.2 Calculation of flow

The flow net has been drawn so that the elemental


rectangles are approximately square thus referring
79

to Fig (3) and equation(4) it can be seen that


between each pair of flow tubes the flow is:
∆Q = k ∆h
(10a)

It should be noted in the development of this


formula it was assumed that each flow tube was of
unit width and so equation (10a) gives the flow per
unit width (into the page).

Suppose that the permeability of the underlying


soil is k=10-5 m/sec (typical of a fine sand or silt)
then the flow between each pair of flow tubes is:

∆Q =10 −5 ×3 m 3 /sec/(m width)


(10b)

there are 5 flow tubes and so the total flow per


unit width of dam is:

Q = 5 ×10 −5 ×3 m 3 /sec/(m width)


(10c)

and if the dam is 25m wide the total flow under the
dam:

Q = 25 ×5 ×10 −5 ×3 m 3 /sec
(10d)

The flow per unit width can alternatively be


calculated from the formula

N f
Q =k H
Nh
(10e)
80

This equation (10e) is given in the Data Sheets to


calculate the flow per unit width. In this equation
Nf is the number of flowtubes (The number of
flowlines – 1), and Nh is the number of
equipotential drops (The number of equipotential
lines – 1).

Note that there are occasions where this formula


(10e) cannot simply be applied, but equation (10a)
will always be applicable for individual flow tubes.
It is often necessary to determine ∆h from
consideration of a single flow tube. If a square
flow net has been constructed that value of ∆h will
apply to all flow tubes.

6.6.3 Calculation of Pore Pressure

The pore pressure at any point can be found using


the expression

uw
h= +z
γw
(11a)

Now referring to Fig. 9 suppose that we wish to


calculate the pore pressure at the point P. Taking
the datum to be at the base of the dam it can be
seen that z = - 5m and so:

uw = 2 −
[1 5)] γ
(− w =
17 γ w

(11b)
81

Example – Calculating pore pressures

The figure below shows a long vessel, 20 metres


wide, stranded on a sand bank. It is proposed to
pump water into a well point, 10 metres down,
under the centre of the vessel to assist in towing
the vessel off. The water depth is 1 metre.

The sand has a permeability of 3 × 10-4 m/sec.


Assuming that a head of 50 m can be applied at
the well point calculate:

1. The pore pressure distribution across the


base of the vessel
2. The total upthrust due to this increase in
pore pressure
3. The rate at which water must be pumped
into the well point.
82

Stranded Vessel

Water Supply

Soft Sea Bottom


Reaction Pile
Well Point

Figure 10 Schematic diagram of vessel on sandbank

Step 1: Choose a convenient datum. In this example the sea


floor has been chosen

Then relative to this datum the head at the well


point, H1 = 40 m
And the head at the sea floor, H2 = 1 m.

The increment of head, ∆h = 39/9 = 4.333 m

Figure 11 Flow net for situation in Figure10


83

5m 2.5 m 1.8

Figure 12 Enlarged view of flownet in the vicinity of


the base of the vessel

Step 2: Calculate the head at points along the base


of the vessel. For convenience these are
chosen to be where the EPs meet the vessel (B
to E) and at the vessel centerline (A). Hence
calculate the pore water pressures.

A B C D E
Head H1 – H1 –5 H1 – 6 H1 – 7 H1 – 8
(1) 4.5 ∆h ∆h ∆h ∆h ∆h
Head H2 + H2 + 4 H2 + 3 H2 + 2 H2 + ∆h
(2) 4.5 ∆h ∆h ∆h ∆h
Head 20.5 18.33 14.0 9.67 5.3
(1 and
2) (m)
Pressur 201.1 179.8 137.3 94.9 52.3
e (kPa)
γ w(h –
z)
84

Step 3: Measure the lengths off the flow net (Note


that diagram must be drawn to scale) and hence
calculate force from pressure distribution. For
simplicity assume linear variation in pressure
between points. Then the TOTAL UPTHRUST (per
unit length of the vessel) is

 201 .1 + 179 .8  179 .8 +137 .3  137 .3 +94 .9   94 .9 +52 .3 


2 × ×5 +  ×2.5 +  ×1.8 +  ×0
 2   2   2   2 
= 3218 kN/m

Without pumping Upthrust = 20 × 1 × 9.81 =


196 kN/m

Upthrust due to Pumping = 3218 – 196 = 3022


kN/m

Nf
Flow required, Q = kH
Nh
=
14
3 ×10 −4 × 39 ×
9
= 1.8 ×10 −2 m3/m/sec
85

7. FLOW NETS FOR ANISOTROPIC MATERIALS

7.1 Introduction

Many soils are formed in horizontal layers as a


result of sedimentation through water. Because of
seasonal variations such deposits tend to be
horizontally layered and this results in different
permeabilities in the horizontal and vertical
directions.
86

7.2 Permeability of Layered Deposits

Consider the horizontally layered deposit, shown in


Figure 1, which consists of pairs of layers the first
of which has a permeability of k1 and a thickness of
d1 overlaying a second which has permeability k2
and thickness d2.

k=k 1 d1

k=k 2 d2

Fig. 1 Layered Soil


Fig. 1Layeredsoil deposit
First consider horizontal flow in the system and
suppose that a head difference of ∆h exists
between the left and right hand sides as indicated
in Fig. 2. It then follows from Darcy’s law that:

∆h ∆h
v1 = k1 ; Q1 = k1 d1 (1a)
L L
and
∆h ∆h
v2 = k2 ; Q2 = k2 d2 (1b)
L L

h =h0 h =h 0 −∆h

v=v 1 d1

v=v 2 d2

Fig.
Fig. 2 Horizontal flowina layered soil 2 Horizontal
deposit flow through
layered soil
87

It therefore follows:
Q1 + Q 2 ∆h (2a)
v = = kH
d1 + d 2 L
where
k 1d 1 + k 2 d 2 (2b)
kH =
d1 + d 2

Next consider vertical flow through the system,


shown in Fig.3. Suppose that the superficial
velocity in each of the layers is v and that the head
loss in layer 1 is ∆h1, while the head loss in layer 2
is ∆h2

h = h0

v d1
h = h 0 − ∆ h1

v d2

h = h 0 − ∆ h1 − ∆ h 2
L
Fig. 3 Vertical flow through
Fig. 3Vertical flow in a layered soil deposit
layered soil

∆h 1
In layer 1: v = k1
d1
v d1
so ∆h 1 =
k1
(3a)

∆h 2 vd2
Similarly in layer 2 v = k2 and ∆h 2 =
d2 k2
(3b)

The total head loss across the system will be


∆h=∆h1+∆h2 and the hydraulic gradient will be
given by:
88

v d1 v d 2
+
∆h ∆ h1 + ∆ h 2 k1 k2
i = = =
d d1 + d 2 d1 + d 2
(3c)

For vertical flow Darcy’s Law gives

∆h
v = kV
d
(3d)
and hence
d d d
= 1 + 2
kV k1 k 2
(3e)

Example

Suppose that that the layers are of equal thickness


d1 = d2 = d0 and that k1 = 10-8 m/sec and that
k2 = 10-10 m/sec, then:
− −
d ×
10 8
+do ×
10 10

kH = o = 5 ×
5 .0 10 9
m/
do +do
and
do +do −
kV = = 8 ×
1.9 10 10
m/s
ec
do do

+
1
0 8
0 −
1 10

Showing that, as is generally the case, the vertical


permeability is much less than the horizontal.

7.3 Flow nets for soil with anisotropic


permeability

Plane flow in an anisotropic material having a


horizontal permeability kH and a vertical
permeability kv is governed by the equation:
89

∂2 h ∂2 h
kH + k V =0
∂x 2 ∂z 2
(4)

The solution of this equation can be reduced to


that of flow in an isotropic material by the
following simple device. Introduce new variables
defined as follows:
x = αx
and
z = z
(5a)

the seepage equation then becomes

kH ∂2h ∂2h
+ =0
α 2 k V ∂x 2 ∂z 2
(5b)

Thus by choosing:

kH
α=
kV
(5c)

It is found that the equation governing flow in an


anisotropic soil reduces to that for an isotropic soil,
viz.:
∂2 h ∂2 h
+ =0
∂x 2 ∂z 2
(5d)

and so the flow in anisotropic soil can be analysed


using the same methods (including sketching flow
nets) that are used for analysing isotropic soils.

Example - Seepage in an anisotropic soil


90

Suppose we wish to calculate the flow under the


dam shown in Figure 4;

H1 Impermeable
dam
H2
z L
Z
Soil layer
x

Impermeable bedrock

Fig. 4 Dam on a permeable soil layer over


impermeable rock (natural scale)

For the soil shown in Fig. (4) it is found that kH = 4× kV


and therefore

4 × kV
α = =2
kV
so
x
x = 2x or x=
2
z =z
(6)
In terms of transformed co-ordinates this becomes
as shown in Figure 5
91

H1
H2
z L/2
Z
x Soil layer

Impermeable bedrock

Fig. 5 Dam on a permeable layer over


impermeable rock (transformed scale)

The flow net can now be drawn in the transformed


co-ordinates and this is shown in Fig.6

5m

Impermeable bedrock

Fig. 6 Flow-net transformed coordinates


Fig. 6 Flow net for the transformed
geometry
92

It is possible to use the flow net in the transformed


space to calculate the flow underneath the dam by
introducing an equivalent permeability

k eq = kH kV

(7)

A rigorous proof of this result will not be given


here, but it can be demonstrated to work for purely
horizontal flow as follows:
h h-∆
h h h-∆
h


Q
transformed
Natural scale t scale

x x
Fig. 7 Horizontal flow through anisotropic
soil
∆h
∆Q = kH t
x

For the natural scale


(7a)
∆h ∆h k H
For the transformed scale ∆ Q = k eq t = k eq t
x x kV
(7b)

From Equations 7a and b it can be seen that


k eq = kH kV

Example

Suppose that in Figure 6 H1 = 13m and H2 =


2.5m, and that kv = 10-6 m/sec and kH =4 × 10-6
m/sec The equivalent permeability is:

k eq = ( 4 × 10 −6 ) × (10 −6 ) = 2 × 10 −6 m / sec

(8a)
93

The total head drop is 10.5 m and there are 14


head drops and thus:

(13 − 2.5)
∆h = = 0.75 m
14
(8b)

The flow through each flow tube, ∆Q = keq ∆h =


(2× 10-6 )× (0.75) = 1.5 × 10-6 m3/s/m

There are 6 flow tubes and so the total flow , Q


= 6 × 1.5 × 10 -6

= 9.0× 10-6
m3/sec/(m width of dam)

For a dam with a width of 50 m Q =


450 × 10-6 m3/sec = 41.47 m3/day

7.4 Piping

Many dams on soil foundations have failed


because of the sudden formation of a piped
shaped discharge channel. As the store water
rushes out the channel widens and catastrophic
failure results. This results from erosion of fine
particles due to water flow. Another situation
where flow can cause failure is in producing
‘quicksand’ conditions. This is also often referred
to as piping failure.

In order to analyse this situation consider water


flowing upwards through the element shown in
Figure 8.
94

u2

(z=z 2 , h=h 2 , u=u 2,)

Elevation

(z=z 1 , h=h 1 ,u=u 1,)

u1

Plan Area =A

Fig. 7 Analysis ofFig. 8 Analysis of Piping


Piping

Uplift Force = A ( u1 − u 2 )

Force due to weight = Aγ sat ( z 2 − z1 )

(9a)
The pore pressure can be calculated from the head
and so:

u 2 = γ w ( h 2 − z2 )
and
u1 = γ w ( h1 − z1 )
(9b)

For piping to occur the Uplift must be greater than


the self-weight of the soil
A( u 2 − u 1 ) > Aγ sat ( z 2 − z 1 )
(9c)

γ w ( h 1 − h 2 ) − γ w (z 1 −z 2 ) > γ sat ( z 2 − z 1 )

γw (h1 −h 2 ) > γ sat ( z 2 − z 1 ) − γ w ( z 2 − z 1 )


(h1 −h 2 ) γ sat − γ w
>
(z 2 −z 1 ) γw
(9d)
or alternatively
95

i > i crit

where
h1 −h 2
i = hydraulic gradient =
z 2 − z1
and
γ sat − γ w
i crit = critical hydraulic gradient =
γw

Example

Suppose the dam shown in Figure 6 is 39 metres


wide (this may be determined from the scale
drawing), the water levels are the same as in the
previous example (H1 = 13 m, H2 = 2.5 m), and the
saturated unit weight of the soil is 18 kN/m3. Piping
is most likely to occur at the toe of the dam, the
hydraulic gradient there can be obtained from the
flow net:

h1 - h2 = ∆h = 0.75 m (calculated
from Fig. 6)

z2 - z1 = 1.125 m (scaled
from Fig. 6)

thus
0.75
i = = 0.67
1125
.
Now
18 − 9.81
i crit = = 0.83
9.81
(10)

The safety factor against piping failure is thus icrit/i


= 0.83/0.67 = 1.25 which is probably not
adequate given the potentially disastrous
consequences of a piping failure.
96

8. ONE DIMENSIONAL SETTLEMENT


BEHAVIOUR

8.1 One Dimensional Loading Conditions

Soils are often subjected to uniform loading over


large areas, such as shown in Figure 1, from an
embankment. Under such conditions soil which is
remote from the edges of the loaded area
undergoes vertical strain, but no horizontal strain.
That is strains, and hence surface settlement, only
occur in one-dimension.

Embankment

x
Soil layer 1
Soil layer 2 z

Rock

Figure 1 Embankment loading on a layered


soil

The accuracy of this assumption depends on the


relative dimensions of the loaded area and
thickness of the soil layer. If the area is relatively
large and the thickness of the soil layer relatively
small then the assumption of 1-D conditions will be
reasonable.

It is possible to make approximate estimates of


surface settlement using the 1-D approach even
when the loaded area is not relatively large. The
97

procedures for doing this are discussed in section


9 on the calculation of settlement.

8.2 The Oedometer

The behaviour of soil during one-dimensional


loading can be tested using a device called an
oedometer, which is shown schematically in Fig. 2.
The one-dimensional condition in which the
vertical strain, ε zz ≠ 0, and the lateral strains, ε xx =
ε yy = 0 is also referred to as confined compression.
Load Displacement
Loading cap measuring device

Cell

water
Soil sample
Porous disks

Figure 2 Schematic diagram of an


oedometer
The following points may be noted:

1. The soil is loaded under conditions of no lateral


strain (expansion), as the soil fits tightly into a
relatively rigid ring.

2. Uncontrolled drainage is provided at the top and


bottom of the specimen by porous discs (two
way drainage). In more sophisticated oedometer
apparatus control of drainage is possible.

3. A vertical load is applied to the specimen and a


record of the settlement versus time is made.
98

The load is left on until all settlement ceases


(usually 24 hours although this depends on the
soil type, impermeable clays may take longer).

4. The load is then increased (usually by a factor of


2, so the vertical stresses might be e.g. 20, 40,
80, 160 kPa). When the maximum load is
reached, the soil is unloaded in several
increments. If desired reloading can be carried
out. At each step time-settlement records are
made.

5. The relationships between voids ratio and


effective stress, and settlement and time are
found from the test. The methods by which
these are obtained will be explained in the
laboratory classes.

It is conventional to plot the void ratio versus the


logarithm of the effective stress in examining the
behaviour of soil, rather than plotting the
relationship between effective stress and strain as
is often done in materials testing. The reason for
this is that the relationship between effective
stress and voids ratio is fundamental to an
understanding of soil behaviour. The relationship
obtained is similar to that between effective stress
and strain because changes in voids ratio and
strain are simply related as shown later.

8.3 Relation of volume strain and vertical


strain

The volume strain ε v of an element of material is


defined to be the change in volume ∆V divided by
initial volume V0
99

∆V
εv = − Note: Compressive strains
Vo
are positive (1)

The volume strain is related to the vertical (axial)


strain. To show this consider Figure 3.

∆z
∆z(1− ε zz )

∆x ∆x(1− ε xx )

(a) Before Deformation (b) After Deformation

Figure 3 Deformation of a soil


Fig.3 Deformation of a soil element
element

[ ]
V = [ ∆x (1 − ε xx )] × ∆y (1 − ε yy ) ×[ ∆z (1 − ε zz )]

(2a)

 V − V0 
εv = −  
 V0 
(2b)

εv =
[ ]
∆x ∆y∆z − [ ∆x ( 1 − ε xx )] × ∆y ( 1 − ε yy ) × [ ∆z ( 1 − ε zz )]
∆x ∆y ∆z
(2c)

thus neglecting second order and higher terms

ε v = ε xx + ε yy + ε zz
(2d)

For confined compression ε xx = 0, ε yy = 0. and


thus:

ε v = ε zz (for confined
compression) (2e)
100

8.4 Relation between volume strain and


voids ratio

For most soils the skeletal material is far stiffer


than the soil composite and thus referring to
Figure 4 it can be seen that the relationship
between volume strain and voids ratio is:

 V − V0   Vs ∆ e  ∆e
εv = −  = −  = −
 V0   Vs (1 + e 0 )  1 + e0
(3a)

and thus for confined compression:

∆e
ε zz = −
1 + e0
(3b)

Vs (e 0 + ∆e)
Voids Vse0
V0 = Vs (1 + e0 )

Skeletal material Vs Vs V = Vs (1 + e0 + ∆e)

(a) Before Deformation (b) After Deformation

Figure 4 Deformation of soil


Fig. 4 Soil undergoing deformation
element
101

8.5 Behaviour of soil under one dimensional


loading

The behaviour of an initially unloaded soil under


one-dimensional conditions is illustrated in Fig. 5.

2 A

C
Voids ratio e
B
1

Log 10 (effective stress)

Figure 5 Typical effective stress, voids ratio


relationship
Fig. 5 Typical Effective stress voids ratio relationship
1. AB corresponds to initial loading of the soil.

2. BC corresponds to an unloading of the soil.

3. CD corresponds to a reloading of the soil.

4. Upon reloading the soil beyond B the soil


continues along the path that it would have
followed if loaded from A to D.

8.5.1 Preconsolidation Stress (pressure)

The preconsolidation stress, σ′ pc, is defined to be


the maximum effective stress experienced by the
102

soil. For soil at state C this would correspond to the


effective stress at point B in Fig. 5.

8.5.2 Normally consolidated soils

If the current effective stress, σ', is equal (note that


it cannot be greater than) to the preconsolidation
stress, σ′ pc, then the deposit is said to be normally
consolidated (NC)

σ ′ = σ pc
′ (normally consolidated)
(4a)

During deposition of a soil (which usually takes


place through sedimentation), the weight of the
soil (which increases with depth below the surface)
causes a decrease in void ratio. Suppose that at a
particular depth below the surface the soil is
represented by point P in Figure 6. If the soil is now
subjected to an effective stress increase under 1-D
conditions the path that will be followed in the e-
log10 σ′ plot will be along the extension of the
deposition line as shown in Fig. 6. A soil which lies
at any point on this line is called normally
consolidated, and the line is called the normal
consolidation line.

Normally consolidated soils are usually found as


recent alluvial deposits, and are mainly composed
of silt and clay sized particles. It is extremely rare
to find normally consolidated soils inland, away
from the rivers or lakes in which they were
deposited.
103

e Impossible states
Normal
P Consolidation
Line

Over-consolidated
states

log 10 (σ’)
Figure 6 The normal consolidation line

8.5.3 Overconsolidated soils

If the current effective stress σ' is less than the


preconsolidation stress, σ′ pc, then the soil is said to
be over-consolidated (OC).

σ ′ < σ pc
′ (over-consolidated)
(4b)

Note
σ ′ > σ pc
′ (not possible)
(4c)

If a soil after deposition, is normally consolidated


to point P and then unloaded (perhaps because of
erosion of the surface layers of soil) it may exist in
the state indicated by point Q in Figure 7. The path
QFR will be followed upon reloading of the soil.

It may be seen that for the same increase in


effective stress, the change in void ratio will be
much less for an overconsolidated soil (from e0 to
ef ) than it would have been for a normally
consolidated soil. Hence settlements will generally
104

be much smaller for structures built on


overconsolidated soils.

Most soils are overconsolidated to some degree;


this can be due to the effects of shrinking and
swelling of the soil on drying and rewetting,
changes in ground water levels, and unloading due
to erosion of overlying strata.

e = e0 Q
F
e = ef P

σ0′ σ′f σ′pc logarithmic scale

Figure 7 Typical effective stress, voids ratio


response

The distance from the normal consolidation line


has an important influence on the soil behaviour.
This is described numerically by the
overconsolidation ratio (OCR). The OCR is defined
as the ratio of the preconsolidation stress to the
current effective stress

σ pc

OCR =
σ′
(5)
105

Note that when the soil is normally consolidated


OCR = 1.

8.5.4 Estimation of the preconsolidation stress

A distinct change of slope is not generally


observed at the preconsolidation pressure, making
it difficult to accurately determine its value.
Empirical procedures are used to estimate the
preconsolidation stress, the most widely used
being Casagrande's construction which is
illustrated in Figure 8.
D
e

A C

E B

σp′ c log (σ’)

Figure 8 Casagrande’s construction for


estimating preconsolidation pressure

Steps in the construction are given below:

1. Determine the point of maximum curvature A.


(It’s important to draw the graph to a sensible
scale)

2. Draw a tangent to the curve at A, i.e. line AB.

3. Draw a horizontal line at A, i.e. line AC.


106

4. Draw the extension of the straight line (normally


consolidated) portion of the curve DE.

5. Where the line DE cuts the bisector (AF) of angle


CAB, is the preconsolidation stress.

For a normally consolidated soil the


preconsolidation stress will be the same as the
vertical overburden stress (due to weight of
overlying soil) existing at the depth from which the
sample was taken. Some unloading of the sample
will take place during sampling so that a
preconsolidation stress may be detected upon
reloading in the oedometer at the point where the
soil is loaded back to the stress state existing in
the ground.

An overconsolidated soil will exhibit a


preconsolidation stress which is much larger than
the overburden stress at the level from which it
was sampled.

8.5 Idealised soil behaviour

The behaviour shown in Figures 5 to 7 may be


idealised by simple linear relationships in a void
ratio, e, logarithm of effective stress, σ´, plot as
shown in Figure 9. This idealisation is based on
observations that:

1. the behaviour of most normally consolidated


soils can be approximated by straight lines for
the range of stresses that are of interest.

2. theresponse of most over-consolidated soils can


be approximated by straight lines, and further:
107

• the behaviour is assumed to be reversible,


unloading and reloading follow the same path

• the slope of the unload-reload response is


constant
e

log (σ’)
Figure 9 Idealised void ratio, effective stress
relationship

8.6 Compression and Recompression


Indexes

Figure 10 shows a portion of the e - log σ′ plot for


a normally consolidated soil.
108

I
eI

eF F

log10 (σI′) log10 (σF′ )


Figure 10 Idealised response for NC soil
Fig. 9 Idealised behaviour of a normally consolidated soil
Suppose that a soil is in an initial state I and after
loading moves to the final state F, as shown in
Figure 10.

eF − eI ∆e
Slopeof IF = =
log10 (σ ′F ) − log10 (σ ′I ) log10 (σ ′F / σ ′I )
(6a)

Because the relationship between effective stress


and voids ratio can be closely approximated by a
straight line, the slope is a constant. The slope
constant, Cc is called the compression index.

∆e
= − Cc
log10 (σ ′F / σ ′I )
(6b)

The above equation can be used to calculate the


final voids ratio from the known final effective
stress and initial conditions as follows:

e F = e I − C c log10 (σ ′F / σ ′I )
(6c)
109

A similar approach is possible if the soil is over-


consolidated and the final stress is less than the
preconsolidation stress, this is shown in Fig. 11.

Again suppose that a soil is at an initial state I and


after loading moves to a final state F, as shown in
Figure 11. As before we have:

eF − eI ∆e
Slopeof IF = =
log10 (σ ′F ) − log10 (σ ′I ) log10 (σ ′F / σ ′I )
(7a)

I
eI

eF F

log10 (σI′) log10 (σF′ )


Figure 11 Idealised response of
Fig. 10 Idealised
OCbehaviour
soil of a over consolidated soil
As the relationship between effective stress and
voids ratio is approximately linear, thus:

∆e
= − Cr
log10 (σ ′F / σ ′I )
(7b)

The constant Cr is called the recompression or


swelling index. Again this equation can be used to
determine the final voids ratio provided the final
110

effective stress and initial conditions are known, as


follows:

e F = e I − C r log10 (σ ′F / σ ′I )
(7c)

Sometimes a soil may move from an


overconsolidated state to a normally consolidated
state. Suppose the initial state of the soil is given
by point 1 in Figure 12, the point at which it
reaches the preconsolidation stress is denoted by
2 and the final state is denoted by 3. The resulting
change in voids ratio as the soil moves from the
initial state 1 to the final state 3 can be considered
to occur in two distinct stages. Stage 1 in which
the soil is oveconsolidated and stage 2 in which
the soil is normally consolidated.

(1)
e1
e2 (2)

e3 (3)

log10(σ1′) log10(σ′2 ) log10(σ′3 )

Figure 12 Response of soil moving from OC


to NC
• Stage 1 (1→2)

During stage 1 the soil is over-consolidated


and so:
111

e 2 = e1 − C r log10 (σ ′2 / σ 1′ )
(8a)

where σ′ 2= the initial value of the


preconsolidation stress σ′ pc

• Stage 2 (2→3)

During stage 2 the soil is normally


consolidated and so:

e 3 = e 2 − C c log10 (σ ′3 / σ ′2 )
(8b)

Since the soil is normally consolidated the


current state of effective stress will be the
preconsolidation stress and thus the final
value of the preconsolidation stress (σ′ pc) will
be σ′ 3

If the soil at 3, where it is normally consolidated, is


unloaded so that the effective stress drops, the
change in void ratio should be determined from
equation 7c for over-consolidated soil.

9. CALCULATION OF SETTLEMENT

9.1 Settlement of a Single Layer


112

The settlement ∆s of a single relatively thin layer, shown in


Fig. 1, can be calculated once the change in void ratio is
known.

∆S
H
z

Fig. 11 Settlement
Fig. Settlement of
ofaasoil layerlayer
single

For confined compression the horizontal strains are negligible


i.e. ε xx = 0, ε yy = 0 and thus:
∆S ∆e
εzz = = εv = −
H 1 +e
thus
∆e H
∆S = −
1 +e
(1)

the settlement of a thicker layer can be calculated by dividing


the layer into a number of sub layers as shown in Fig. 2. This
is necessary because both the initial and final effective stress
vary with depth as do the voids ratio and the OCR.

sub-layer 1

Notation
sub-layer 2
ei = voids ratio at the centre of layer i
∆e i = increase in voids ratio at the centre of layer i
Hi thickness of layer i

sub-layer n

Fig. 2intoDivision
Fig. 2 Soil profile divided a numberofofsoil layers
sub-layers into sub-layers
113

The settlement of the soil layer is calculated by calculating the


settlement of the individual sub-layers and adding them, in
doing this it is assumed that the voids ratio and the effective
stress are constant throughout the sub-layer and equal to their
values at the centre of the sub-layer.
thus
∆ei H i
For sub − layer i ∆Si = −
1 + ei
so that
∆ei H i
Total Setttlement S = ∑1n∆Si = ∑1n[− ]
1 + ei
(2)

Example - Settlement Calculation

A soil deposit, shown in Fig. 3 consists of 5 m of gravel


overlaying 8 m of clay. Initially the water table is 2 m below
the surface of the gravel. Calculate the settlement if the water
table rises to the surface of the gravel slowly over a period of
time and surface loading induces an increase of total stress of
100 kPa at the point A and 60 kPa at the point B. The
preconsolidation pressure at A is 120 kPa, and the deposit is
normally consolidated at B. The gravel has a saturated bulk
unit weight of 22 kN/m3 and a dry unit weight of 18 kN/m3 and
is relatively incompressible when compared to the clay. The
void ratio of the clay is 0.8 and the skeletal particles have a
specific gravity of 2.7. The compression index of the clay is 0.2
and the recompression index is 0.05.

In solving this problem it will be assumed that the gravel is far


less compressible than the clay and thus that the settlement of
the gravel can be neglected. The settlement of the clay layer
will be calculated by dividing it into two sub-layers
114

γ dry = 18 kN / m3
2m
Gravel 5m
γ sat = 22 kN / m3

A 4m
Clay γ sat = ?
B 4m

Fig. 3 Layered
Fig. 3 Layered soil deposit
soil deposit

In order to commence the calculations it is first necessary to


calculate the unit weight of the clay, this is shown
schematically in Fig. 4.

Voids
Vv= e*Vs =0.8 m3 Wv = γ w * Vv
= 7.84 kN
Ww + Ws 7.84 + 26.46
γ sat = = kN
Vv + Vs 0.8 + 1
Ww = Vw * γ w * G s = 19 .06 kN / m
3

Skeletal Vs=1 m3
= 26.46 kN or
material ( G s + e) γ w 3
γ sat = = 19 .06 kN / m
1+ e

Distribution of Volume Distribution of Weight

Fig. 4 Determination of Saturated Unit Weight


Ww + Ws 7.84 + 26.46
γ sat = = = 19.06 kN / m 3
Vv + Vs 0.8 + 1
or
( Gs + e ) γ w
γ sat = = 19.06 kN / m 3
1+ e

• Initial State at A
115

Total stress σzz = 2 × 18 + 3 × 22 + 2 × 19.06


= 140.12 kPa
Pore water pressure uw = 5 × 9.8 kPa = 49 kPa
(3a)
Effective stress σ′ zz = σzz - uw = 140.12 - 49 = 91.12
kPa

Notice the initial effective stress is less than σ′ pc =120 kPa thus
the clay is initially over-consolidated.

• Final State at A

Total stress σzz = 100 + 2 × 22 + 3 × 22 + 2 ×


19.06 = 248.12 kPa
Pore water pressure uw = 7 × 9.8 kPa = 68.6 kPa
(3b)
Effective stress σ′ zz = σzz - uw = 248.12 - 68.6 =
179.52 kPa

Notice that the final effective stress exceeds the initial


preconsolidation stress and thus the clay moves from being
initially over-consolidated to finally normally consolidated.

• Settlement of the first sub-layer

The soil in the first sub layer moves from being over-
consolidated to normally consolidated and so the calculation of
the change in voids ratio must be made in two stages.

Stage 1 Soil over-consolidated (σ′ < σ′ pc (initial))

∆e1 = - Cr × log10(σ′ pc (initial)/σ′ I)

Stage 2 Soil normally consolidated (σ′ = σ′ pc)


(3c)
∆e2 = - Cc × log10(σ′ F/σ′ pc (initial))
now
116

H∆e
∆S = −
1 +e
4 120 .00 179 .52
= [ 0.05 ×log 10 ( ) + 0.2 ×log 10 ( )]
1.8 91.12 120 .00
= 0.0911 m
(3d)
• Initial State at B

Total stress σzz = 2× 18 + 3 × 22 + 6 × 19.06 =


216.36 kPa
Pore water pressure uw = 9 × 9.8 kPa = 88.20 kPa
(4a)
Effective stress σ′ zz = σzz - uw = 216.36 - 88.20 =
128.16 kPa

• Final State at B

Total stress σzz = 60 + 2 × 22 + 3 × 22 + 6 ×


19.06 = 284.36 kPa
Pore water pressure uw = 11 × 9.8 kPa = 107.80 kPa
(4b)
Effective stress σ′ zz = σzz - uw = 284.36 - 107.80 =
176.56 kPa

• Settlement of the second sub-layer

The soil in the second is normally consolidated and thus:

∆e2 = - Cc × log10(σ′ F/σ′ I)


(4c)

now
H ∆e
∆S = −
1 +e
4 176 .56
= 0.2 ×log 10 ( ).
1.8 128 .16
= 0.0620 m
(4d)

• Total Settlement
117

Total settlement = 0.0911 + 0.0620 m

(5)
= 0.1531m
118

9.2 Calculation of Stress Changes

The calculation of settlement depends upon knowledge of the


initial and final effective stress within each sub layer of the
deposit. The initial effective stress state can be determined,
from knowledge of the bulk unit and the position of the water
table. The increase in total stress can be estimated using the
theory of elasticity. (Note the soil is in general not really elastic
however in the working stress range this assumption provides
reasonably accurate estimates of the stress increases due to the
applied loads)

A fundamental solution of the equations of elasticity is


Boussinesq's solution. This relates to a point load applied to the
surface of a half-space (very deep layer) and is shown
schematically in Fig. 5.

Point load of magnitude P

z
H →∞

Fig. 5 Point load acting on a half space


Fig 5 Point load on an elastic half-space
Boussinesq found that :
119

3Pz3
∆σ zz =
2 πR5
(1 + ν) Pz
∆σ xx + ∆σ yy + ∆σ zz =
πR 3
(1 + ν) P z2
∆u z = [2(1 − ν) + 2 ]
2 πER R
where
R = x2 + y 2 + z2
and
E = Young ' s mod ulus
ν = Poisson ' s ratio
(6)
∆uz = vertical displacement due to load

The symbol ∆ is used to indicate that each of the quantities in


equation (6) represents the increase in the particular quantity,
due to the applied load.

The solution for a point load is important because it can be


used to develop solutions for distributed loads by integration.
Some of these solutions are presented in the Soil Mechanics
Data Sheets.

9.3Calculation of Stress Changes

9.3.1 Stresses due to Circular foundation loads applied at the


ground surface

A circular foundation of diameter 5 m, subjected to an average


applied stress of 100 kPa is shown in Fig. 6.
120

5m

p=100 kPa

2m
z

A B
5m

Fig. 6 Circular loaded area on a deep elastic layer


Fig 6 Circular loaded area on a deep elastic layer

(a) Calculate the increase in vertical stress at point A

There is a simple analytic expression (given in the Data Sheets)


for points on the centre line under a circular load:

a2 −3 / 2
∆σ zz = p(1 − [1 + ] )
z2
(7a)

where

p = the surface stress = 100 kPa


a = the radius of the loaded area = 2.5m
z = the depth of interest = 2m

∆σ zz = 100 × (1 − [ 1 + (1.25) 2 ]−3/ 2 ) = 75.6 kPa


(7b)

(b) calculate the increase in vertical stress at point B


121

In this case there is no simple analytic expression and the


solution must be found by using the influence charts given in
the data sheets, reproduced in part in Figure 7. Note that this
chart can also be used for points on the centre line for which r =
0.
Now z/a = 2/2.5 = 0.8
r/a = 5/2.5 = 2
(8)
using the data sheets ∆σzz/p = 0.03 and so ∆σzz = 3.0 kPa

σ zz
Iσ =
p
10-3 10-2 10-1 1
0 1.00
1.25
1.5
2.0 0.0
2
2.5
5
6 3
4
7
z/a 4

6 8 Values on curves
are values of r/a
9
8
10

10

Fig. 7 Influence factors for a uniformly loaded circular area of radius a


Fig.7 Influence Factors for a Uniformly Loaded Circular Area of radius a
9.3.2 Stresses due to Rectangular foundation loads applied at
the ground surface
122

Plan B

Uniformly distributed
surface stress p

z
Point immediately
Elevation
beneath one of the
rectangle’s corners

Fig. Fig. 8 Rectangular


8 Rectangular uniform
surface loading
loading onona adeep
deepelastic
elastic layer
layer

Many loads which occur in practice are applied to foundations


that may be considered to consist of a number of rectangular
regions. It is thus of interest to be able to calculate the vertical
stress increases due to a uniformly distributed load acting on a
rectangular loaded area. This is shown schematically in Fig. 8.
The vertical stress change at a distance z below one of the
corners of the rectangular load may be determined from a chart
which is given in the data sheets and is reproduced in Fig. 9
123

8
0.25 3.0 2.0

1.0
0.20
0.8
Note m & n are
0.15 interchangeable 0.6
σ zz
Iσ =
q 0.4
0.10

0.05 0.2

0.00 m=B/z=0.0
0.01 0.1 1 10
(n=L/z)
Fig. 9 Influence factors for uniformly loaded rectangular areas
Fig. 9 Influence factors for a uniformly loaded
rectangular area

This chart can be used to determine the value of stress increase


at any point in an elastic layer, the method for doing this is
illustrated below.

9.3.2.1 Calculation of Stress below an interior point of the


loaded area

This situation is shown schematically in Fig.10. The stress


change is required at a depth z below point O.

The first step in using the influence charts is to break the


rectangular loading up into a number of components each
having a corner at O, this is relatively simple as can be seen in
Fig.(10)
124

It thus follows that at the point of interest, the stress increase


∆σzz(ABCD) is given by:

∆σ zz (ABCD ) = ∆σ zz (OXAY ) + ∆σ zz (OYBZ ) + ∆σ zz (0ZCT ) + ∆σ zz (OTDX )


(9)
D T C

X O Z Plan

A Y B

Elevation
z

O Point of interest

Fig. 10 stress increase at a point below a loaded rectangular region

Example

Suppose we wish to evaluate the increase in stress at a depth of


2m below the point O due to the rectangular loading shown in
shown in Fig. 11, when the applied stress over ABCD is 100
kPa.

D T C
2m
X O Z

3m

A Y B
3m 2m
Fig. 11 Dimensions ofFig. 11 Dimensions of rectangular loaded area
loaded area
rectangular

For rectangular loading OZCT


125

m = L/z =1
n = B/z =1
thus
Iσ = 0.175
and so
∆σzz = p Iσ = 100 × 0.175 = 17.5 kPa
(9a)

For rectangular loading OTDX

m = L/z = 1.5
n = B/z = 1
thus
Iσ = 0.194
and so
∆σzz = p Iσ = 100 × 0.195 = 19.4 kPa
(9b)

For rectangular loading OXAY

m = L/z = 1.5
n = B/z = 1.5
thus
Iσ = 0.216
and so
∆σzz = p Iσ = 100 × 0.216 = 21.6 kPa
(9c)

For rectangular loading OYBZ

m = L/z = 1.5
n = B/z = 1
thus
Iσ = 0.194
and so
∆σzz = p Iσ = 100 × 0.194 = 19.4 kPa
(9d)
126

Thus the increase in stress ∆σzz = 17.5 + 19.4 + 21.6 + 19.4


= 78.9 kPa (9e)

This must of course be added to the existing stress state prior to


loading to obtain the actual stress σzz.

9.3.2.2 Calculation of stress below a point outside the loaded


area

The stress increase at a point vertically below a point O which


is outside the loaded are can also be found using the influence
charts shown in Fig. 9.
X Z O

D C T Fig. 12 Rectangular loaded area

A B Y
Fig. 12 Rectangular loaded area ABCD and point of interest O

This is achieved by considering the stress q acting on ABCD to


consist of the following:

1. A stress +q acting over OXAY

2. A stress +q acting over OZCT

3. A stress -q acting over OZBY

4. A stress -q acting over OXDT


This is illustrated in Fig. 13.

It thus follows that at the point O, the stress increase


∆σzz(ABCD) is given by:

∆σ zz ( ABCD) = ∆σ zz (OXAY) − ∆σ zz (OYBZ) + ∆σ zz (OZCT) − ∆σ zz (OTDX)


127

and thus
(10)
σ zz ( ABCD) = q[I σ (OXAY) − I σ (OYBZ) + I σ (0 ZCT) − I σ (OTDX)]

X Z O X Z O
(q) (q) (q) (0)
D C T D C T

(q) (q) (q) (0)

A B Y A B Y
Stage 1 Stage 2

X Z O X Z O
(q) (q) (0)
D C T D C (0) T

(q) (0) (q) (0)

A B Y A B Y
Stage 3 Stage 4
Fig. 13 Decomposition of loading over a rectangular area (for stress at external point)
Fig. 13 Decomposition of Loading over a rectangular region (exterior point)
128

Example
1m

X Z O

1m
D C T

2m

A Y
B
10m
Fig. 14 Dimensions of rectangular loaded area
Fig. 14 Dimensions of rectangular loaded area

Suppose the rectangular area ABCD, shown in Fig. 14 is


subjected to a surface stress of 100 kPa AND it is required to
calculate the vertical stress increase at a point 1.5m below the
point O.

For rectangular loading OZCT

m = L/z = 0.67
n = B/z = 0.67
thus
Iσ = 0.121
and so
∆σzz = p Iσ = + 100 × 0.121 = +12.1 kPa
(11a)

For rectangular loading OTDX

m = L/z = 7.67
n = B/z = 0.67
thus
Iσ = 0.167
and so
129

∆σzz = p Iσ = -100 × 0.167 = -16.7 kPa


(11b)

For rectangular loading OXAY

m = L/z = 7.67
n = B/z = 2.00
thus
Iσ = 0.240
and so
∆σzz = p Iσ = + 100 × 0.240 = + 24.0kPa
(11c)

For rectangular loading OYBZ

m = L/z = 2
n = B/z = 0.67
thus
Iσ = 0.164
and so
∆σzz = p Iσ = -100 × 0.164 = -16.4 kPa
(11d)

Thus the increase in stress ∆σzz = 12.1 - 16.7 + 24.0 + -16.4=3.0


kPa (11e)

9.3.3 Stresses due to foundation loads of arbitrary shape


applied at the ground surface

Newmark’s chart provides a graphical method for calculating


the stress increase due to a uniformly loaded region, of
arbitrary shape resting on a deep homogeneous isotropic elastic
region.

Newmark’s chart is given in the data sheets and is reproduced


in part in Fig 15. The procedure for its use is outlined below
130

1. The scale for this procedure is determined by the depth z at


which the stress is to be evaluated, thus z is equal to the
distance OQ shown on the chart.

2. Draw the loaded area to scale so that the point of interest


(more correctly its vertical projection on the surface) is at
the origin of the chart, the orientation of the drawing does
not matter

3. Count the number of squares (N) within the loaded area, if


more than half the square is in count the square otherwise
neglect it.

4. The vertical stress increase ∆σzz = N × [scale factor(0.001)]


× [surface stress (p)]

The procedure is most easily illustrated by an example.

Example

Suppose a uniformly loaded circle of radius 2 m carries a


uniform stress of 100 kPa. It is required to calculate the vertical
stress at a depth of 4 m below the edge of the circle.

The loaded area is drawn on Newmark’s chart to the


appropriate scale (i.e. the length OQ is set to represent 4 m) as
shown in Fig. 15.

It is found that the number of squares, N = 194 and so the stress


increase is found to be

∆σzz = 194 × 0.001 × 100 = 19.4 kPa


(12)

This result can also be checked using the influence charts for
circular loading and it is then found that:

z/a = 2, r/a = 1. ∆σzz /p = 0.2 and so ∆σzz = 20 kPa


(13)
131

4m
O Q

Loaded
Area

Fig 15 Newmark’s Chart

10. ANALYSIS OF CONSOLIDATION

10.1 Introduction: the consolidation process

From the response of soils under one-dimensional


conditions it is apparent that when the effective
stress increases there will be a tendency for the
soil to compress. However, when a load is applied
to a saturated soil specimen this compression does
132

not occur immediately. This behaviour is a


consequence of the soil constituents, the skeletal
material and pore water, being almost
incompressible compared to the soil element;
deformation can only take place by water being
squeezed out of the voids. This can only occur at a
finite rate and so initially when the soil is loaded it
undergoes no volume change.

Under one dimensional conditions this implies that


there can be no vertical strain and thus no change
in vertical effective stress. For 1-D conditions we
have

− ∆ e C log(σ ′F / σ ′I )
ε zz = ε v = =
1+ e 1+ e
(1)

Hence if ε v = 0 then ∆e = 0 and σ´F = σ´I.

When the load is first applied the total stress


increases, but as shown above for 1-D conditions
there can be no instantaneous change in vertical
effective stress, this implies that the pore pressure
must increase by exactly the same amount as the
total stress as:

∆σ´ = ∆σ - ∆u
(2)

Subsequently there will be flow from regions of


higher excess pore pressure to regions of lower
excess pore pressure, the excess pore pressures
will dissipate, the effective stress will change and
the soil will deform (consolidate) with time. This is
shown schematically in Fig. 1.
133

Total Excess Pore


Stress Pressure

Time Time

Effective Settlement
Stress

Time Time

Fig. 1 Variation
Fig. of total stress
1 Variation and pore pore
of stress, pressure with time and
pressure
settlement with time

10.2 Derivation of the equation of


consolidation for 1-D conditions

If we assume that the pore fluid and soil skeleton


are incompressible, then:

Volume decrease of the soil = Volume of pore


fluid which flows out

and thus

Rate of volume decrease of soil = Rate at which


pore fluid flows out

In deriving the equations governing consolidation


we will consider only one-dimensional conditions
with purely vertical soil movements and water
flows. The solutions obtained will only be strictly
relevant to the vertical consolidation of relatively
134

thin soil layers occurring as a result of extensive


uniform loading. (This is a common situation). A
similar approach can be followed for more general
loading but the resulting equations can only be
solved numerically.

v(z, t)

Soil element ∆z

∂v
v(z + ∆z, t) = v + ∆z
∂z

Fig. 2 Flow of pore fluid into an element of


Fig. 2 Flow of pore fluid into an element of soil
soil
Referring to Fig 2 it can be seen that:

The rate at which water enters the element


= (v ( z + ∆z , t ) − v ( z , t )) A
∂v
= ∆z A
∂z

The rate of volume decrease of the element


∂ εv
= ∆z A
∂t
and thus
∂v ∂ εv
=
∂z ∂t
(3)
where
v = the pore fluid velocity,
135

ε v = the element volume strain,


A = the cross sectional area of the
element.
It will also be assumed that Darcy’s law holds and
thus that

∂h
v = −k v
∂z
(4)

In applying Darcy’s law it is only the velocity due


to the consolidation process that is of interest, and
consequently the head in (4) is the excess head
due to the consolidation process (not the total
head). The excess head is related to the excess
pore water pressure by

u
h =
γw
(5)

Note that the elevation is not involved in (5)


because it only relates the excess heads and water
pressures. From (3), (4) and (5) it follows that

∂ ∂u ∂ε v
[k v ] = −
∂z ∂z ∂t
(6)

If it is also assumed that the soil element responds


elastically to a change in effective stress then:

ε v = m v σ ′e
(7)
where

σ ′e = the change in effective stress from


the original value
136

= σe − u
(8)
with σ e = the increase in total stress over the
original value

u = the increase in pore water pressure


over the original value (excess pore
water pressure)
and
mv = the coefficient of volume decrease,

The value of mv must be determined over the


appropriate effective stress range because it
depends on the mean effective stress. This can be
seen by considering the relation between voids
ratio and effective stress:

e = A − C lo g1 0 σ ′
and hence
C dσ ′
de = −
2.3 σ ′
now
∆e C dσ ′
εv = − = = m v dσ ′
1 + e 2.3 (1 + e) σ ′
(9)

Thus mv depends on both voids ratio e, and


effective stress, σ´.

Combination of equations (6), (7) and (8) leads to


the equation of consolidation:

∂ k v ∂u ∂u ∂σ e
[ ] = mv[ − ]
∂z γ w ∂z ∂t ∂t
(10)
The equation of consolidation must be solved
subject to certain boundary conditions and initial
conditions
137

10.3 Boundary Conditions

At a boundary where the soil is free to drain the


pore water pressure will be constant and will not
change during consolidation. For such a boundary
the excess pore water pressure will be zero.

u = 0 at a permeable boundary
(11a)

At an impermeable boundary the pore water


velocity perpendicular to the boundary will be zero
and thus from Darcy’s law

∂u
∂z
= 0 at an impermeable boundary
(11b)

10.4 Initial Conditions

At the instant of loading there is no volume strain


and thus no change in vertical effective stress. At
this instant the excess pore water pressure will be
equal to the initial increase in total stress.

u = σe at the instant of loading.


(12)

10.5 The Equation of Consolidation for a


Homogeneous Soil

If the soil layer being considered is homogeneous


then equation (10) becomes:
138

∂2 u ∂u ∂σ e
cv = −
∂z 2 ∂t ∂t
(13)

where
kv
cv = is called the coefficient of
mvγ w
consolidation.

The coefficient of consolidation (cv) can be


estimated using the oedometer apparatus as can
the coefficient of volume decrease (mv). The
procedure to do this will be discussed in the
laboratory classes. It is difficult (time consuming)
to measure the permeability of clays (kv) and so
the value of permeability is usually inferred from
the values of cv and mv .

10.6 Analytic Solutions to the equations of


consolidation

10.6.1 Two-way drainage

Fig. 3 represents a layer of clay of thickness 2H


subjected to a uniform surface stress q applied at
time t = 0 and held constant thereafter. The clay
layer is free to drain at both its top and bottom
boundaries. This is called two-way drainage.
139

Uniformly distributed surcharge q

Z
2H

Fig. 3 Homogeneous
Fig 3 Homogeneous clay layer freeClay
Saturated to Layer
drain from both upper
and lower boundaries
free to drain at Upper and Lower Boundaries
The increase in stress through out the layer and
does not vary with time and so

σe = q

Equation (10) therefore becomes:

∂ 2u ∂u
cv =
∂ z2 ∂t
(14a)

The clay layer is free to drain at its upper and


lower boundary and so

u = 0 when z = 0 for t > 0


(14b)

u = 0 when z = 2H for t > 0


(14c)

Initially the excess pore pressure will just match


the increase in total stress so that there will be no
instantaneous volume change and thus:
140

u = q when t = 0 for 0 < z < 2H.


(14d)

It can be shown that the solution of equations (14


a,b,c,d) is:
∞ 1
sin(α n Z ) e − α n Tv
2
u = 2q ∑
n=0 αn

(15)

where αn = (n + ½) 
z
Z = H , a dimensionless distance
cv t
Tv = , a dimensionless time
H2

Notice that H which occurs in both dimensionless


quantities is the maximum drainage path length.

The settlement of the soil layer can be determined


by summing the vertical (= volume) strains,
giving:
2H
S = ∫ ε v dz
0
2H
= ∫ m v (q − u)dz
0
(16a)

and the variation of settlement with time can be


obtained by substituting in equation (15) which
gives the variation of u with time and depth.

2H  ∞ sin α Z 
e − α n Tv  dz
2
S = ∫ m v q 1 − 2∑ n

0  0 αn 
giving
141

 ∞ e − α n Tv 
2

S = m v q 2 H 1 − 2 ∑ 2 
 n= 0 α n 
(16b)

Noting that the final settlement of the layer, S∞ =


mv q 2H the settlement may be written:

 ∞ e − α n Tv 
2
S
= U = 1 − 2 ∑
S∞ 2 
 n=0 α n 
(16c)

where U is known as the degree of settlement

The variation of excess pore pressure within the


layer is shown in Figure 4 (also in data sheets).

Z=z/H 1 T=0.8 0.5 0.3 0.2 0.1

2
0.0 0.5 1.0
u/q
Fig. 4 Variation of excess pore pressure with
depth

The lines on Figure 4 represent the variation of


pore pressure with depth at different non-
dimensionalised times (T). These lines are known
as isochrones. It can be seen that initially the
excess pore pressure is constant (u/q = 1)
throughout the layer. With time the pore water
142

flows from the interior of the layer to the drainage


boundaries, and the excess pore pressures
dissipate until after a very long time there are no
excess pore pressures.

The variation of settlement with time is most


conveniently plotted in the form of the degree of
settlement (U) versus dimensionless time Tv, and
this is illustrated in Fig. 5 (also in data sheets)

Dimensionless Time T v
10-3 10-2 10-1 1 10
0.00
Relation of degree of
settlement and time
0.25

U 0.50

0.75

1.00

Fig. 5 Degree of settlement versus


dimensionless time

There are several useful approximations for the


degree of settlement, viz:

4 Tv
U = ( Tv ≤ 0.2 )
π

8 −π2 Tv / 4
U 1− e ( Tv > 0..2 )
π2
(17)
143

alternatively Fig. 5 may be used. It is worth


remembering that U = 0.5 when Tv = 0.197.

10.6.2 One-way drainage

Fig. 6 represents a layer of clay of thickness H


subjected to a uniform surface stress q applied at
time t = 0 and held constant thereafter. The clay
layer is free to drain at its top boundary but is
unable to drain at its base. This is called one way
drainage.
Uniformly distributed surcharge q

Z
H

Impermeable base

Fig 6 HomogeneFig. 6rateHomogeneous


ous Satu d ClayLayer resting on ansaturated clay
impermeable base layer on an impermeable
base
The increase in stress through out the layer and
does not vary with time and so

σe = q
Equation (6) therefore becomes:

∂ 2u ∂u
cv =
∂ z2 ∂t
(18a)

The clay layer is free to drain at its upper


boundary and as before
144

u = 0when z = 0 for t > 0


(18b)
at the lower boundary
∂u
= 0 when z = H for t > 0
∂z
(18c)

Initially the excess pore pressure will just match


the increase in total stress so that there will be no
instantaneous volume change and thus:

u = q when t = 0 for 0 < z < H.


(18d)

Reference to figure 4 reveals that solution (15)


also satisfies the condition

∂u
∂z
= when z = H for t > 0
0

and is thus also the solution for one way drainage (


the two way drainage problem can be viewed as
two one-way drainage problems ‘back to back’).
Further examination reveals that although the
expression for final settlement differs for the two
cases the expression for degree of settlement is
precisely the same.
145

Example - Calculation of settlement at a


given time

Figure 7 shows a soil profile, it can be assumed


that the sand and gravel are far more permeable
than the clay and so consolidation in them will
have occurred instantaneously.

Gravel

Clay 4m Final settlement=100mm


cv=0.4m 2/year

Sand

Final settlement=40mm
Clay 5m
c v=0.5m 2/year

Impermeable

Fig 7Fig. 7 Layered


layered soil deposit
Soil Deposit

It is assumed that the final settlement has for each


of the clay layers has been determined by the
methods described in the previous sections and
that their values are as indicated on figure 7. It is
required to find the settlement after 1 year

(a) Settlement of the upper Layer

In layer 1 there is two way drainage and so the


drainage path H = 2m.

cvt 0.4 × 1
Tv = = = 0.1
H2 22
146

Using Figure 5 it can be seen that U = 0.36

and thus the settlement of layer 1 = 100 × 0.36 =


36mm

(b) Settlement of the lower Layer

In layer 2 there is one way drainage and so the


drainage path H = 5m.

cv t 0.5 × 1
Tv = = = 0.02
H2 52

Using Figure 5 it can be seen that U = 0.16

and thus the settlement of layer 2 = 40 × 0.36 =


6.4mm

The total settlement after 1 year is thus = 36 +


6.4 = 42.4mm

Example - Use of scaling

An oedometer specimen reaches 50% settlement


after 2 minutes. If the specimen is 10 mm thick
calculate the time for 50% settlement of a 10 m
thick layer under conditions of one-way drainage.

In order that the test may be carried out as quickly


as possible oedometer tests are normally
conducted with two way drainage and thus the
drainage path in the oedometer = 5mm = 0.005m.

For the oedometer test

cvt cv × 2
Tv = = = 80000c v
H2 0.0052
147

For the clay layer the drainage path is 10m

cv t c ×t c ×t
Tv = 2
= v 2 = v
H 10 100

Since the degree of settlement for the two case is


the same the two values of the dimensionless
time, Tv are equal and so:

cv t
80000 c v = thus t = 8000000 min = 15.2 years
100

Example - Calculation of the coefficient of


consolidation

The data in the previous example can be used to


calculate cv. The dimensionless time for 50%
consolidation is Tv = 0.197 (from Figure 5 Tv ≈ 0.2)
thus:

m2 m2
0.197 = 80000 c v thus cv = 2.4625 ×10 −6 = 1.294
min year

11. NUMERICAL SOLUTION OF THE 1-D


CONSOLIDATION EQUATION
148

The 1-D equation of consolidation cannot be solved


analytically except for some very simple
situations. For more difficult cases it is necessary
to use approximate numerical techniques. One
numerical technique that can be used for
consolidation problems is the finite difference
approach. In this method the solution is evaluated
at a number of points at different times as
indicated on the figure below.

t=0 t=t1 t=t2


∆t
t
1
∆z
2

4
z

Fig. 1 Grid showing points at which


Fig. 1 Finite difference grid
solution calculated

11.1 Finite Difference Formulae

The 1-D consolidation equation and the boundary


conditions are approximated by finite difference
formulae. These can be derived by referring to the
figure below and taking local axes at B:

A u = uA z = -

B u= z = 0
u uB
z

C
Fig. 2 Excess
u= pore water pressure variation
z = + at∆z
time t
149

Suppose that the excess pore pressure at any time


t can be approximated by a parabola

u = a1 + a 2 z + a 3z 2
(1a)

The constants in this equation can be related to


the values of the excess pore pressures at points
A, B, C. Taking B as the origin for z gives:
uA = a 1 −a 2 ∆z +a 3 ∆z 2

uB = a1

uC = a 1 +a 2 ∆z +a 3 ∆z 2

(1b)
so that
a1 = uB
uC − uA
a2 =
2 ∆z
u A + u C − 2u B
a3 =
2 ∆z 2
(1c)

thus evaluating the slope and curvature of u at the


point B (z = 0) it is found:

 ∂u  uC − uA
 ∂z  =
B 2 ∆z

 ∂2 u  u A + u C − 2u B
 2 =
 ∂z  B ∆z 2

(1d)

11.2 Finite Difference Approximation of


Consolidation Equation

The equation of consolidation is:


150

∂2 u ∂u ∂q
cv = −
∂z 2 ∂t ∂t
(2a)

where q is the change in total stress, due to


applied loads, from the initial equilibrium situation
when the excess pore pressures were zero.

When this equation is evaluated at any point in the


soil it is equivalent to evaluating the equation at
point B, and hence the finite difference formulae
developed above can be introduced so the
equation becomes:

 ∂u   ∂q   u A + u C − 2u B 
 ∂ t  −  ∂ t  = c v  ∆z 2 
B B

(2b)

if the above equation is now integrated from times


t to t+∆t it is found that:
t + ∆t
c
∆u B = ∆q B + v2
∆z ∫[ u A + u C − 2 u B ]dt
t

(2c)

where ∆u B = u B ( t + ∆t ) − u B ( t ) and ∆q B = q B ( t + ∆t ) − q B ( t )
151

Error in approximation
F(t)

t +∆t
∫ F(t )dt ≈ F(t )∆t
t

t
t t +∆t
Fig. 3 Approximate integral evaluation
Fig 3 Approximation of integral

If the integral appearing in equation (2c) is now


approximated as indicated in Fig. 3, it is found
that:

∆ uB = ∆ q B + β[ u A ( t ) + u C ( t ) − 2 u B ( t )]
(2d)

cv ∆ t
where β =
∆ z2

Or
u B ( t +∆t) = ∆q B + u B ( t ) + β[ u A ( t ) + u C ( t ) − 2u B ( t )]
(2e)

Suppose the solution for u has been found up to


time t. The applied load will be known at time t +
∆t and so the quantity ∆q is known. This means all
the quantities on the right hand side of equation
(2e) are known and thus that u at time t + ∆t can
be calculated. Thus a knowledge of the
distribution of u at time t means that the
distribution of u at time t + ∆t can be inferred.
Now the initial distribution of u can always be
152

determined and thus the solution can be found by


‘marching’ forward in time.

11.3 Stability

There is an important restriction on the use of


equation (2e) to obtain a numerical solution of the
equation of consolidation, this is

cv ∆ t 1
β = 2

∆z 2

If this condition is violated the calculation becomes


unstable and is invalid.
153

11.4 Boundary Conditions

The solution of the equation of consolidation


depends on the boundary conditions.

11.4.1 Fully Permeable Boundary

At a free draining boundary there is no


impediment to flow and so the pore pressure
remains constant and thus the excess pore water
pressure is zero, this is illustrated in figure 4a.

Drainage Boundary u=0

Saturated soil

Fig. 4a Finite difference approximation of a


Fig. 4a Finite difference approximation of a drainage boundary
drainage boundary

11.4.2 Impermeable Boundary

Saturated soil

A .
∆z
B .
.
∆z Impermeable barrier
C

Fig. 4b Finite difference approximation of an


Fig. 4b impermeable
Finite difference approximation
boundary of an impermeable boundary
154

At an impermeable boundary, such as that


illustrated in figure 4b there can be no flow in a
direction perpendicular to the boundary. As
outlined earlier this implies:
∂u
= 0
∂z
(3a)
the finite difference analogue of this equation is
uC − uA
= 0
2 ∆z
(3b)
and hence
uC = uA
(3c)
An impermeable boundary is modelled by equating
the excess pore pressure at C to that at A. To do
this a dummy node has to be introduced at C into
the finite difference grid. This dummy node has no
affect other than to give the correct excess pore
pressure at the impermeable boundary.
Example - Numerical Solution when β =1/2

Suppose that a 4m layer of clay, shown in figure 5,


which is free to drain at its upper boundary and
rests on an impermeable base, is subjected to a
surface loading of 64 kPa.
155

q = 64 kPa

4 sub-layers
4m
cv = 2 m2/year
mv = 0.0003 m2/kN

Impermeable bedrock

Fig. 5 Clay layer subjected to a surcharge loading


Fig . 5 Clay layer subjected to a surcharge loading

If β = 0.5 the finite difference equation takes a


particularly simple form:

u A (t ) + u C (t )
u B ( t + ∆t ) = ∆q +
2
(4)

In the case under consideration the surcharge is


applied at t = 0 and remains constant thereafter
so that ∆ q = 0

The solution then proceeds as follow:

Step 1: Divide the deposit into layers - this fixes


the value of ∆z.

In this case the deposit is divided into 4 sub-


layers (all with the same thickness) and thus
∆z = 1m

Step 2: select β = 0.5 this fixes the value of ∆t

For the case under consideration:


156

c v ∆t 2 ×∆t 1
β = 2
= =
∆z 12 2
thus

∆t = 0.25 years
157

Step 3: Calculate the initial pore pressure

Because there cannot be an instantaneous


volume change it follows that

u( t = 0 ) = q(t = 0) = 64 kPa

Step 4: Introduce the dummy node to simulate


the impermeable boundary

Step 5: March the solution forward using the


finite difference equation and
introducing the boundary conditions

The solution is shown in the table below:

t(years) 0 0.2 0.5 0.7 1 1.2 1.5


5 5 5
q(kPa) 64 64 64 64 64 64 64
z=0 64 0 0 0 0 0 0
z=1m 64 64 32 32 24 24 20
z=2m 64 64 64 48 48 40 40
z=3m 64 64 64 64 56 56 48
z=4m 64 64 64 64 64 56 56
dummy 64 64 64 64 56 56 48

Step 6: Calculate settlement

The settlement is calculated as follows


H
S = ∫ε v dz
0
H
= ∫m v (q − u) dz
0
H
= m v q H − m v u dz ∫
0

(5)
158

In the above equation the integral of the excess


pore pressure cannot be evaluated exactly
because the excess pore pressures are only
calculated at the grid points. However, the integral
can be evaluated approximately using numerical
techniques. The simplest approach, and that used
here, is to use the trapezoidal method:

Thus
H
1 1
∫ u dz ≈
2
( u 0 + u 1 ) ∆z + + ( u n −1 + u n ) ∆z
2
0

 u + u n  
= ∆z  0  + u 2 + + u n −1 
 2  
where

ui = u i ( ∆z, t )
(6)

Thus after 1.5 years


H
S = m v qH − m v ∫ udz
0
0 + 56
= 0.0003 × 64 × 4 − 0.0003 × ( + 20 + 40 + 48 ) ×1
2
= 0.036 m

= 36 mm

The settlement at other times can be similarly


calculated from the excess pore pressures hence
the values can be determined as shown in the
table below.

t(years) 0 0.2 0.5 0.7 1 1.2 1.5


5 5 5
q(kPa) 64 64 64 64 64 64 64
z=0 64 0 0 0 0 0 0
z=1m 64 64 32 32 24 24 20
z=2m 64 64 64 48 48 40 40
159

z=3m 64 64 64 64 56 56 48
z=4m 64 64 64 64 64 56 56
dummy 64 64 64 64 56 56 48
Settlem 0 9.6 19. 24 28. 32. 36
ent 2 8 4
(mm)

Example - Numerical Solution when β ≠ 1/2

Suppose now the previous example is solved using


a step size of 2 months but keeping the number of
layers the same. If this is the case β =1/3. The
numerical solution proceeds as above but now
using the more complex form of the finite
difference equation, viz. equation (2e), the solution
is shown in the table below:

t(mth’s) 0.0 2.0 4.0 6.0 8.0 10. 12. 14. 16. 18.
0 0 0 0 0 00 00 00 00 00
settleme 0.0 9.6 16. 20. 23. 26. 29. 32. 34. 36.
nt(mm) 0 0 00 27 82 90 67 20 54 73
q(kPa) 64. 64. 64. 64. 64. 64. 64. 64. 64. 64.
00 00 00 00 00 00 00 00 00 00
z=0 64. 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
00 0 0 0 0 0 0 0 0 0
z=1m 64. 64. 42. 35. 30. 27. 25. 23. 21. 20.
00 00 67 56 81 65 28 44 92 61
z=2m 64. 64. 64. 56. 52. 48. 45. 42. 39. 37.
00 00 00 89 15 20 04 32 92 74
z=3m 64. 64. 64. 64. 61. 59. 56. 53. 51. 48.
00 00 00 00 63 26 63 99 39 86
z=4m 64. 64. 64. 64. 64. 62. 60. 57. 55. 52.
00 00 00 00 00 42 31 85 28 68
160

dummy 64. 64. 64. 64. 61. 59. 56. 53. 51. 48.
00 00 00 00 63 26 63 99 39 86

For z = 3 m at 12 months the calculations are

u B ( t +∆t) = ∆q B + u B ( t ) + β[ u A ( t ) + u C ( t ) − 2u B ( t )]
0
u B ( t + ∆t ) + 59.26 +
= 0.3333 × [48.20 +
62.42 - 2 × 59.26] = 56.63

The results for the two analyses are quite close.


After 18 months the settlement predicted in
example 1 is 36 mm, which compares well with the
settlement calculated in example 2, viz. 36.7mm.

Example - Variable loading

Suppose fill having unit weight 20 kN/m3 is placed


at a rate of 0.5 m/month for 12 months after which
no more load is applied, the analysis only differs
from that in the previous examples in that the
value of ∆q needs to be included in the finite
difference equation. Choosing β = 0.5 with 4 layers
gives a time step of 0.25 years as before. The
results are shown in the table below.

t(years) 0 0.2 0.5 0.7 1 1.2 1.5


5 5 5
settleme 0 4.5 13. 24. 38. 48. 56.
nt(mm) 5 75 25 938 813
q(kPa) 0 30 60 90 120 120 120
z=0 0 0 0 0 0 0 0
z=1m 0 30 45 60 71. 52. 46.
25 5 875
z=2m 0 30 60 82. 105 93. 82.
5 75 5
z=3m 0 30 60 90 116 112 105
161

.25 .5
z=4m 0 30 60 90 120 116 112
.25 .5
dummy 0 30 60 90 116 112 105
.25 .5

Note that when the load is applied gradually the


excess pore pressure at the permeable upper
boundary remains at zero. This is because there is
no instantaneous change in load.

If the calculation is repeated for the case in which


there are 5 sub-layers, and a time step of 0.1 years
is adopted, this gives β = 0.3125 and the results
are shown below:

t (years) 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
0 0 0 0 0 0 0 0 0 0 0
settleme 0.0 1.4 3.7 6.7 10. 14. 18. 23. 28. 33. 39.
nt(mm) 0 4 8 4 21 13 45 13 16 50 15
q(kPa) 0.0 12. 24. 36. 48. 60. 72. 84. 96. 108 120
0 00 00 00 00 00 00 00 00 .00 .00
z=0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
0 0 0 0 0 0 0 0 0 0 0
z=0.8m 0.0 12. 20. 27. 33. 38. 43. 47. 52. 56. 59.
0 00 25 09 04 38 27 80 04 05 85
z=1.6m 0.0 12. 24. 34. 44. 54. 62. 70. 78. 85. 92.
0 00 00 83 78 00 64 78 50 86 90
z=2.4m 0.0 12. 24. 36. 47. 58. 69. 80. 90. 99. 109
0 00 00 00 63 86 66 07 11 81 .18
z=3.2m 0.0 12. 24. 36. 48. 59. 71. 83. 94. 105 116
0 00 00 00 00 89 60 10 35 .33 .04
z=4.0m 0.0 12. 24. 36. 48. 60. 71. 83. 95. 106 117
0 00 00 00 00 00 93 72 33 .72 .85
dummy 0.0 12. 24. 36. 48. 59. 71. 83. 94. 105 116
0 00 00 00 00 89 60 10 35 .33 .04
162

t (years) 1 1.1 1.2 1.3 1.4 1.5


settleme 39. 43. 47. 50. 54. 57.
nt(mm) 145 6 495 997 236 269
q(kPa) 120 12 120 120 120 120
0
z=0 0 0 0 0 0 0
z=0.8m 59. 51. 46. 43. 40. 38.
851 5 697 186 441 191
z=1.6m 92. 87. 82. 77. 73. 70.
904 7 158 59 681 29
z=2.4m 109 10 103 99. 96. 92.
.18 6 486 068 804
z=3.2m 116 11 112 110 108 105
.04 4 .6 .44 .01 .41
z=4.0m 117 11 115 113 111 109
.85 7 .31 .61 .63 .37
dummy 116 11 112 110 108 105
.04 4 .6 .44 .01 .41

Again the settlements at 1.5 years are quite


similar. Thus although greater refinement of the
grid leads to more accurate excess pore pressures
and settlements there is, in practice, little
advantage of using β ≠ 0.5.
163

Example - Abrupt change of load

Suppose that in the case detailed in example 1 a


further surcharge of 32 kPa is added after 12
months. The solution in this case is best handled
in two stages.

Stage 1 follows exactly the path outlined in


example 1 and is detailed in the table below. Just
after 1 year the load is abruptly increased, and
since there can be no instantaneous volume strain
there can be no increase in effective stress and no
change in settlement. This means that the
increase or decrease in applied stress must be
matched by a corresponding increase or decrease
in pore water pressure. This enables the excess
pore water pressure to be calculated.

t 0 0.2 0.5 0.7 1.0


5 0 5 0
settleme 0 9.6 19. 24 28.
nt(mm) 2 8
q(kPa) 6 64. 64. 64. 64.
4 00 00 00 00
z=0 6 0.0 0.0 0.0 0.0
4 0 0 0 0
z=1m 6 64. 32. 32. 24.
4 00 00 00 00
z=2m 6 64. 64. 48. 48.
4 00 00 00 00
z=3m 6 64. 64. 64. 56.
4 00 00 00 00
z=4m 6 64. 64. 64. 64.
4 00 00 00 00
dummy 6 64. 64. 64. 56.
4 00 00 00 00
164

Stage 2 of the calculation then proceeds in the


same way as in stage 1 or in example 1. This is
shown in the table given below:

t 1 1.2 1.5 1.7 2 2.2 2.5


5 5 5
settleme 28. 37. 45. 51 56. 60. 65.
nt(mm) 8 2 6 4 75 1
q(kPa) 96 96 96 96 96 96 96
z=0 32 0 0 0 0 0 0
z=1m 56 56 36 36 29 29 24.
5
z=2m 80 72 72 58 58 49 49
z=3m 88 88 80 80 69 69 59
z=4m 96 88 88 80 80 69 69
dummy 88 88 80 80 69 69 59

12. SETTLEMENTS OF STRUCTURES

12.1 The settlement process

An important task in the design of foundations is to


determine the settlement, this is shown
schematically in Figure 1.
165

Maximum
Settlement

Soil Layer

Fig. 1Fig. 1 Settlement


Settlement of a loaded of a loaded footing
footing

As discussed earlier the skeletal soil material and


the pore water are relatively incompressible and
any change in volume can only occur due to
change in the volume of the voids. For the volume
of the voids to change, pore water must flow into
or out of a soil element. Because this cannot
happen instantaneously when a load is first
applied to a soil there cannot be any immediate
change in its volume. For one-dimensional
conditions with no lateral strain this implies that
there is no immediate vertical strain and hence
that the excess pore pressure is equal to the
change in vertical stress. However, under more
general conditions both lateral (or horizontal) and
vertical strains can occur. Immediately after load is
applied there will be no change in volume, but the
soil deformations will result in an initial settlement.
This is said to occur under undrained conditions
because no pore water has been able to drain from
the soil. With time the excess pore pressures
generated during the undrained loading will
dissipate and further lateral and vertical strains
will occur. Ultimately the settlement will reach its
long term or drained value.
166

When the load is first applied to the soil there will


be a tendency for the more highly stressed parts of
the soil to compress and thus for there to be a
reduction in the volume of the voids. The pore
water will respond to this tendency towards a
decrease in volume by undergoing an increase in
pore water pressure and so initial excess pore
water pressures will develop. Subsequently there
will be a flow of water from regions of high excess
pore water pressure to regions of low excess pore
water pressure, and the load induced excess pore
water pressures will dissipate. This is the process
of consolidation, and during this process the soil
will undergo a settlement which varies with time.
Ultimately after a long period of time all the excess
pore water pressures will have dissipated and the
settlement of the soil will cease and it will reach its
long term or drained settlement (the term drained
is used because all excess pore water pressures
have dissipated and there will be no further
drainage of water from the voids although the
voids will still remain saturated). The process of
consolidation is shown schematically in Figure 2.
It should be stated that the process described
above represents a simplification because some
soils tend to creep. For such soils there will be
additional creep settlements even though the
effective stress does not change.
167

Total
Stress

Time

Excess
Pore
Pressure

Time

Effective
Stress

Time

Fig. 2a Variation of stress and pore pressure


at a typical point under a footing

Settlement
Consolidation
settlement
Final
settlement

Initial
settlement

Time

Fig. 2b Variation of settlement with time

12.2 Analysis of Settlement under three


dimensional conditions

Previously the settlement under foundations has


been estimated assuming purely one-dimensional
conditions. However, it is clear from consideration
168

of the stress changes (predicted by the theory of


elasticity) under the centre and edges of various
loaded areas that in general the stress changes
may differ significantly from those deduced using
the purely one-dimensional assumption.
If it is hypothesised that the soil can be treated as
a linear isotropic elastic material then solutions for
the settlement can be obtained using the theory of
elasticity. This assumption involves a considerable
level of approximation which is necessary
because:

• real soil behaviour is highly non-linear


• the geometry of the foundation is often complex
• simple models enable calculations to be easily
performed

Linear isotropic elasticity is used because:

• closed form solutions which are easily evaluated


can be obtained
• complicated loadings can be synthesised from
simple components using superposition
• only 2 material constants are required from (E, ν,
G, K)
• the solutions obtained agree with intuition and
experience

12.3 Theory of Elasticity for Saturated Soils

In an isotropic elastic solid it is found that Hooke’s


law relates the changes in stress to the changes in
strain as described in equation (1):
169

∆ σ xx − ν( ∆ σ yy + ∆ σ zz )
∆ ε xx =
E
∆ σ yy − ν( ∆ σ zz + ∆ σ xx )
∆ ε yy =
E
∆ σ zz − ν( ∆σ xx + ∆ σ yy )
∆ε zz =
E
(1a)

where ∆ε xx , ∆εyy , ∆εzz denote the strains which


arise from the changes in stress ∆σ xx , ∆σyy , ∆σzz
and where E is Young’s modulus and ν is Poisson’s
ratio.

Hooke’s law in this form does not apply to soil


except for undrained conditions which will be
discussed later. For soil the correct relationship is
one between effective stress and strain as shown
below:

∆ σ ′xx − ν ′( ∆ σ ′yy + ∆ σ ′zz )


∆ εxx =
E′
∆ σ ′yy − ν ′( ∆ σ ′zz + ∆ σ ′xx )
∆ εyy =
E′
∆ σ ′zz − ν ′ ( ∆ σ ′xx + ∆ σ ′yy )
∆ εzz =
E′
(1b)

where E´ is called the effective stress, or drained,


Young’s modulus and ν´ is called the effective
stress, or drained, Poisson’s ratio, and where the
increments of effective stress are related to the
increments of total stress and the increment of
pore water pressure by:
∆ σ ′x x = ∆ σx x − ∆ u
∆ σ ′yy = ∆ σyy − ∆ u
(1c)
∆ σ ′zz = ∆ σz z − ∆ u
170

The relationship between effective stress and


strain can always be used to calculate the
deformation of soils. However, to do so it is
necessary to know both the change in total stress
and the change in pore water pressure. The
change in total stress can usually be estimated
using elastic solutions, but the change in pore
pressure is, in general, very difficult to determine.

One important case where the effective stresses


are known is in the long term. In this situation all
excess pore water pressures have dissipated and
thus the change in effective stress is equal to the
change in total stress. The settlement can then be
calculated using the effective stress, strain
relations.

Equations (1b) can be modified as follows:

∆σ ′xx − ν ′( ∆σ ′yy + ∆σ ′zz )


∆ε xx =
E′
∆σ ′xx (1 + ν ′) − ν ′( ∆σ ′yy + ∆σ ′zz + ∆σ ′xx )
=
E′
∆σ ′xx (1 + ν ′) − 3ν ′∆σ ′m
=
E′
where
( ∆σ ′yy + ∆σ ′zz + ∆σ ′xx )
∆σ ′m =
3
(2)

this alternative form of Hooke’s law is useful as will


be seen below.

12.4 Behaviour of an elastic soil under


undrained conditions

It was shown above that the long term behaviour


of soil can be analysed using Hooke’s law since all
171

excess pore pressures have dissipated and so the


effective stress equals the total stress. Another
important case which can be analysed using
Hooke’s law is immediately after loading when no
water has drained out of the soil pores and no
excess pore pressures have dissipated, i.e.
undrained behaviour. To establish this note that
under such conditions there can be no volume
change and thus:

∆ε v = ∆ε xx + ∆ε yy + ∆ε zz = 0
(3a)

The volume strain can be calculated using


equations (2) and (3a) giving:

3(1 − 2 ν ′ )∆ σ ′m
∆ εv =
E′
(3b)

If the volume strain is zero the change in mean


effective stress is zero and thus:
3(1 −2 ν′) ∆σ′

ε v = m
= 0
E′
then

∆σ′
m = ∆
σ m −∆u = 0

thus

∆u = ∆
σm

(3c)

This enables the increment in excess pore water


pressure to be expressed in terms of the total
stress. Using this relation and substitution into
equation (2) leads to the following relation
between total stress and strain:
172

(1 + ν ′ )(2 ∆ σxx − ∆ σyy − ∆ σzz )


∆ εxx =
3E′
(4)

This and similar expressions for ∆εyy and ∆εzz are


equivalent to Hooke’s law for undrained loading,
which may be written as:

∆ σ xx − ν u ( ∆ σyy + ∆ σ zz )
∆ ε xx =
Eu
∆ σ yy − ν u ( ∆ σzz + ∆ σxx )
∆ ε yy =
Eu
∆ σ zz − ν u ( ∆ σ xx + ∆ σ yy )
∆ εzz =
Eu
(5)

The quantities Eu , and νu are called the undrained


Young’s modulus and Poisson’s ratio respectively.
By comparing equations (4) and (5) it can be seen
that these quantities are related to the drained or
effective stress relations as follows:

3E ′
Eu =
2(1 + ν ′)
1
νu =
2
(6)

It is interesting to note that so far there has been


no mention of shear behaviour, for shear stresses
and strains Hooke’s law may be written as:

∆σ yz
∆γ yz =
G′
∆ σ zx
∆ γ zx =
G′
(7)
173

∆σ xy
∆γ xy =
G′

where G´ is a material property called the shear


modulus which is related to the effective stress
parameters as follows.

E′
G′ =
2(1 + ν′)
(8)

It is interesting to observe that:

E′ E Eu
G′ = = u = = Gu
2(1 + ν ′ ) 3 2(1 + ν u )
(9)

Showing that the shear modulus (and shear strain)


is unaffected by the state of drainage in the soil.

It is important to emphasise that the relation


between effective stress parameters and
undrained parameters is based on many
approximations (soil assumed elastic) and should
not be expected to be exact. Thus, although the
undrained value of Poisson’s ratio will be precisely
1/2 for a saturated soil because of
incompressibility, the undrained Young’s modulus
should be measured directly rather than
determined from the effective E´ value.

12.5 Values of the Elastic Parameters for


soils

The selection of parameters to use in elastic


analyses of settlement prediction presents
considerable difficulties in geotechnical
engineering. Soil is not a linear elastic material. In
174

selecting values for the "elastic" parameters


consideration must be given to:

The initial effective stresses in the ground.

• The values of E´,ν´ are both dependent on the


1
mean effective stress, 3 (σ ′xx + σ ′yy + σ ′zz ) , with the
moduli increasing with stress level.

The soil stress history

• OCR for clays


• Relative density (Id) for sands
• For a given stress level, the moduli will increase
with increasing OCR or Id

The strain level

• It is advisable to use an appropriate secant


modulus for the expected strain level under the
footing.

12.5.1 Values of E'

Typical values may be selected from the following


values given in the data sheets (p. 65)

Soft normally-consolidated clays ( 1400


- 4200 kPa)

Medium clays ( 4200 -


8400 kPa)

Stiff clays ( 8400 - 20000


kPa)
175

Loose normally-consolidated sands ( 7000


- 20000 kPa)

Medium normally-consolidated sands


(20000 - 40000 kPa)

Dense normally-consolidated sands


(40000 - 84000 kPa)

For over-consolidated sands, double the above


values.

12.5.2 Values of v'

Soft clay 0.35 - 0.45

Medium clay 0.30 - 0.35

Stiff Clay 0.2 - 0.3

Medium sand 0.3 - 0.35

These typical values should be used with caution.


Soils are extremely variable materials and
considerable expertise is needed to determine
accurate parameters.

Example - Strains during undrained loading

A cuboidal soil specimen is in equilibrium with a


uniform stress acting on all faces of 100 kPa, and
no pore pressure, that is u = 0. The vertical stress
is then increased by 90 kPa with the stresses on
the other faces remaining constant and with the
176

sample prevented from draining. Calculate the


vertical and lateral strains if E´ = 10 MPa and ν´ =
¼.

Initially: σ1 = σ2 = σ3 = 100 kPa; u = 0

Analysis of undrained loading can be performed in


terms of undrained parameters (Total Stress
Analysis) or drained parameters (Effective Stress
Analysis).

1. Total Stress Analysis

Calculate undrained parameters νu = 0.5,


3E′
Eu = = 12 MPa
2 (1 + ν ′ )
Now the total stress changes are ∆σxx = 0 kPa,
∆σyy = 0 kPa, ∆σzz = 90 kPa

Use Hooke’s Law in terms of Total Stress

1 ∆ σzz
∆ εzz = ( ∆ σzz − ν u ( ∆ σxx + ∆ σyy )) =
Eu Eu

1 − ν u ∆ σzz
∆ εxx = ( ∆ σxx − ν u ( ∆ σzz + ∆ σyy )) = = ∆ εyy
Eu Eu

Hence

∆εzz = 90/12000 = 0.0075

∆εxx = ∆εyy = - 0.5 × 0.0075 = -


0.00375

2. Effective stress analysis


177

Changes in effective stress are needed to evaluate


the effective Hooke’s Law relations.

Calculate ∆u = ∆σm for undrained loading (see


above)

1
= 3 (∆σ xx + ∆σ yy + ∆σ zz )
= 90/3 = 30 kPa

Hence ∆σ´xx = - 30 kPa, ∆σ´yy = - 30 kPa, ∆σ´zz


= 60 kPa

Now using Hooke’s Law

∆ σ ′xx − ν ′ ( ∆ σ ′yy + ∆ σ ′zz ) − 30 − 0.25 ( − 30 + 60)


∆ εxx = = = − 0.00375
E′ 10000
∆ σ ′yy − ν ′ ( ∆ σ ′zz + ∆ σ ′xx )
∆ εyy = = − 0.00375
E′
∆ σ ′zz − ν ′( ∆ σ ′xx + ∆ σ ′yy ) 60 − 0.25 ( − 30 × 2)
∆ εzz = = = 0.0075
E′ 10000

giving the same result as before.

Example – Strains during drained loading

If the same sample from example 1 is now allowed


to drain and consolidate, without any change to
the applied stresses, what strains will develop.

Only an effective stress analysis is relevant. Total


stress analysis cannot be used because the total
stress parameters (Eu, νu) are only relevant to
undrained loading, that is when deformation
occurs at constant volume.
178

In this example during consolidation the total


stresses remain constant. The effective stress
changes are thus ∆σ´xx = + 30 kPa, ∆σ´yy = + 30
kPa, ∆σ´zz = + 30 kPa, they are all equal to the
reduction in pore water pressure. Then from
Hooke’s Law

∆εxx = ∆εyy = ∆εzz = 0.0015

Note that the total strains due to the undrained


loading followed by consolidation are

∆εxx = ∆εyy = - 0.00375 + 0.0015 =


-0.00225

∆εzz = 0.0075 + 0.0015 = 0.009

The same total strains are obtained if the load is


applied slowly so that no pore pressures are
obtained. In this case the pore pressure change is
zero and hence the change in total stress is the
same as the change in effective stress (∆σ´xx = 0
kPa, ∆σ´yy = 0 kPa, ∆σ´zz = 90 kPa). The strains
are then given by

1 ∆ σ ′zz
∆ εzz = ( ∆ σ ′zz − υ ′ ( ∆ σ ′xx + ∆ σ ′yy )) = = 0.009
E′ E′

1 − υ ′ ∆ σ ′zz
∆ εxx = ( ∆ σ ′xx − υ ′ ( ∆ σ ′zz + ∆ σ ′yy )) = = ∆ εyy = − 0.00 2 25
E′ E′

Note that the strains are identical to those


determined as a result of undrained loading
followed by consolidation. This result is not
surprising when it is remembered that this is an
elastic analysis.
179
180

13. SETTLEMENT OF STRUCTURES

13.1 Solutions based on the theory of


elasticity

Figure 1 represents a surface footing resting on a


soil layer of depth H.

H Soil Layer

Rigid bedrock

Fig. 1 Foundation resting on a


Fig. 1 Foundation resting on a soil layer
soil layer

The settlement, s, of any point can be determined


from
H
s = ∫ ∆ε zz dz
0

(1a)
181

where for an elastic soil

(1 + ν ′ ) ∆ σ ′zz − ν ′ ( ∆ σ ′xx + ∆ σ ′yy + ∆ σ ′zz )


∆ ε zz =
E′
(1b)

and under undrained conditions:

(1 + ν u ) ∆σ zz − ν u ( ∆σ xx + ∆σ yy + ∆σ zz )
∆ε zz =
Eu
(1c)

As discussed earlier, to determine the settlement


immediately after the application of the load
equation (1c) is used, and to determine the long
term or drained settlement equation (1b) is used.
In the latter case the changes in pore water
pressure ∆u are usually zero and so the increment
in effective stress is equal to the increment in total
stress. Thus, in both cases the settlement can be
calculated if both the change in total vertical stress
∆σzz and the change in the mean total stress (∆σxx+
∆σyy+ ∆σzz ) are known.

It has been shown previously how the Boussinesq


solution for the stresses in an elastic half space
due to a point load acting on the surface can be
used to determine the stress distributions under a
variety of shapes of loaded areas (circles,
rectangles, arbitrary shapes). The same solution
can be used to determine the surface settlements,
sr as a function of the distance, r, from a point load
Q, as
182

Q(1 − ν2 )
sr =
πEr
(2)

This is illustrated in Figure 2.

Q
r

sr

H →∞
Q(1 −ν2 )
sr =
πEr

Fig.
Fig. 2 Surface
2 Surface deflection
deflection due
of a deep to alayer
elastic point load on a
deep elastic layer

Because the soil is assumed to be linear elastic it is


possible to use superposition to determine the
surface settlements for distributed loads using the
point load solution. For example, the settlement at
the centre of a circular loaded area, radius, a, with
uniform stress, q, (flexible foundation), can be
determined by considering the effect of the stress,
q, acting over an area r dθ dr (shown in Figure 3)
on the settlement at the centre. The settlement is
then given by:
183

dr

dθ a

r

Fig. 3 Stress q acting over a circular area of


radius a
a 2π
(1 − ν2 )
scentre = ∫ ∫ πEr qrd θdr
0 0
2q (1 − ν2 )a
=
E
(3)

For other positions under the circular load and for


other shapes the integration is not so
straightforward, and in many cases analytical
solutions will not be possible.

Also a limitation of this (Boussinesq) solution is


that it assumes the soil layer is infinitely deep. This
rarely occurs in practice as more generally a
relatively shallow soil layer usually overlies rock.

The procedure adopted in practice is to make use


of charted solutions that are available for a
number of commonly encountered situations.
Some of these are given in the data sheets, and
are discussed below. For other solutions the book
184

"Elastic solutions for Soil and Rock Mechanics" by


Poulos and Davis should be referred to.

13.2 Settlement under a rigid circular load

P = πa 2 p av

rigid
2a

h Soil Layer

Rigid bedrock

Fig. 4a Rigid circular footing on an elastic


layer on a rigid base

The configuration being considered is shown in


Figure 4a and the solution is presented in terms of
a settlement factor, Iρ . The settlement, s, is given
by the expression:

p av a
s = Iρ
E
(4)

where

Pav is the average stress on the footing =


Load/Area = P/(π a2)

a is the radius of the loaded area


185

E is the soil modulus

Iρ is a settlement factor read from Figure 4b


(Data Sheets page 45). Note that Iρ depends
on the value of Poisson’s ratio ν.
186

1.6
P = π a 2 p av
1.2
2a h ν = 0.0
0.2
Iρ 0.8 0.4
0.5

0.4
p av a
s = Iρ
E
0.0
0.0 0.2 0.4 0.6 0.8 1.0 0.8 0.6 0.4 0.2 0.0
h/a a/h
Fig. 4b Settlement factors for a rigid circular
Fig. 3b Settlement Factor for rigid circular footing on a layer
footing on a soil layer

Example

Determine the final settlement under a footing 3 m


in diameter which is subjected to a load of 500 kN
if it rests on a soil layer 9 m thick with properties E'
= 5 MPa, v' = 0.3.

a 15
.
= = .167
h 9
Iρ = 1.22 from figure ( 4 b)
500
p av = = 70 .7 kPa
. )2
π × (15
70 .7 ×15. ×1.22
s = = 0.026 m
5000

13.3 Settlement of square footings

The settlement under a square footing can be


estimated to sufficient accuracy by considering the
load to act over an equivalent circular area. So if
the square footing has sides of length b the
187

following equivalent pressure and radius can be


used in equation 4:

P
p av =
b2
b
a =
π

13.4 Settlement of a Circular Foundation on a


non-homogeneous soil

Soils often have a modulus that increases with


depth. The soil does not necessarily change its
nature with depth, the reason for the increase in
modulus is that the mean effective stress
increases with depth and, because the modulus
increases with the mean effective stress so the
modulus varies with depth. Often the variation
with depth is approximately linear and so can be
approximated by the relation:
E = E 0 + mz
(5)

The modulus increases linearly from E0 at the


surface as shown schematically in Figure 5..
P = πa 2 p av p

2a

E 0 + mz

Fig. 5 Circular footing on non-homogeneous


Fig. 4 Circular footing on a non-homogeneous soil
soil
188

A charted solution is available for this modulus


variation for the case of a flexible circular footing
(p constant) resting on an infinitely deep soil layer.
The settlement may be expressed in the form:
p a
s = Iρ
E0
(6)
where Iρ is the influence factor given in Figure 6
(Data Sheets p 47) and
p is the stress on the footing
a is the radius of the loaded area
E0 is the Young's modulus at the surface

2
1

ν= 0
10--1
1/3
Iρ p
10--2 1/2

2a

10--3 E 0 + mz

10--4
10---4 10---2 E0 1 102
η=
ma
ExampleFig. 6 Influence chart for flexible circular load on non-
homogeneous soil

An oil tank applies a uniform stress of 75 kPa over


a circular area with diameter 20 m. Calculate the
immediate settlement if the undrained modulus
increases linearly from 2 MPa at the surface, to 5
MPa at 10 m.
189

E = E 0 +mz

5 = 2+10 m

m = 0.3 MPa / m

E0
η =
ma
2
= = 0.67
0.3 ×10
now

Iρ ≈ 0.6 from fig . 6

thus
pa
s = Iρ
E0
75 ×10 × 0.6
= = 0.225 m
2000

13.5 Settlement under the edge of a flexible


strip load on a finite soil layer

The configuration is shown in Figure 7a. The


settlement at the edge takes the form:
p h
s = Iρ
πE
(7)

where Iρ is the influence factor given in Figure 7b


(Data sheets p 46) and

p is the stress on the strip footing

h is the depth of the soil layer

E is the Young's modulus of the soil.

The value of the settlement at other locations can


be found by superposition, as demonstrated below.
190

For a rigid strip footing the settlement can be


estimated by averaging the centre and edge
settlements of an equivalent flexible footing.
p

h Soil Layer

Rigid bedrock

Fig. 7a Flexible strip on an elastic layer on a


rigid base

2.0

1.6 ν =0.0
0.2
1.2

0.4
0.8

0.4
0.5

0.0
0.0 0.25 0.5 h/B
B/h 2.0 1.0 0.0

Fig. 7b Settlement factor for edge of flexible


strip on a soil layer
191

Example

Determine the final settlement at a point 10 m


from the centre of a 16 m wide embankment,
assuming that the embankment can be considered
as a flexible strip load which applies a surface
stress of 50 kPa. The embankment is constructed
on a soil layer 15m deep with the properties E´= 9
MPa, ν´= 0.3.

Because of the assumption of elasticity


superposition can be used. Thus the embankment
loading can be simulated as shown in Figure 8.

8m 10m 18m

Embankment 2m (+)
(-)

15m 2m

Fig. 8 Decomposition of embankment


loading to give settlement not under edge

The embankment loading consists of a strip


loading of intensity +50 kPa and width of 18 m for
which:
h 15
=
B 18
= 0.83

Iρ = 11
. from Figure 7b
192

and a strip loading of intensity -50 kPa and width


of 2 m for which:
B 2
=
h 15
= 0.13

Iρ = 0.58 from Figure 7b

Thus the settlement is:


p h +50 ×15
s1 = Iρ = ×110
.
πE π × 9000
ph −50 ×15
s2 = Iρ = × 0.58
πE π × 9000
s = s1 + s 2
+50 ×15
= × (110
. − 0.58 ) = 0.0138 m
π × 9000
193

13.6 The influence of embedment on


settlement

If a footing is embedded the settlement will be


reduced. Two cases are shown in Figure 9a for
which some solutions are available, both for a very
deep elastic layer. The settlement reduction
factors are given in Figure 9b (Data sheets p 48).
To use these solutions the settlement must be
found using the previously derived solutions for
the load resting on the surface.

Z Z

D D
(a) Uniform circular load (b) Uniform circular load
at the base of an within a deep elastic
unlined shaft layer

Fig. 9a Loads applied below the surface in a


deep elastic layer
194

1.0

Settlement of an identical surface load


0.9
Settlement of a deep load

ν
ν=0.49
=0.49 (a)
(a)
ν
ν=0.25
=0.25 (a)
(a)
0.8
ν
ν=0.00 (a)
=0.0 (a)
0.7

0.6

ν=0.50 (b)
0.5
0 5 10 15 20
Z/D

Fig. 9b Depth reduction factors for


embedded circular footings
195

13.7 Selection of Elastic parameters

The settlement of any foundation can be split into


3 components

13.7.1 Immediate or undrained settlement

This component is due to deformations in the soil


immediately after loading. As has been discussed
previously, immediately after load is applied water
has no time to drain out of the voids and so there
is no volume change. Hence any deformation must
occur at constant volume.

In practice deformation at constant volume only


occurs for relatively impermeable clayey soils that
remain undrained in the short term.

To estimate the initial settlement, si, due to the


constant volume deformation the undrained (total
stress) parameters Eu, νu = 1/2 are used in the
analyses described above.

As observed earlier when the load is applied over a


very large area the situation approaches one-
dimensional conditions, for which the initial
undrained settlement is zero.

In principle effective stress parameters could be


used to determine the settlement, but because the
excess pore pressures generated by the load vary
throughout the soil the analysis is not
straightforward, and the simple elastic formulae
cannot be used.

13.7.2 Consolidation Settlement


196

This is due to deformations arising from volume


changes which occur as a consequence of the
excess pore water pressures, which have been
generated immediately after loading, dissipating
allowing the effective stresses to come into
equilibrium with the applied loads. Finally all
excess pore water pressures will have dissipated
and the final settlement, stf, can be determined by
using E', v' in the settlement formulae developed
previously.

The settlement due to consolidation, sc, can be


determined indirectly from the final settlement stf,
and the immediate settlement, si, by:

sc = s tf − si
(8)

13.7.3 Creep deformations at constant load.

Settlements due to creep cannot be predicted


using the simple elastic formulae, and are usually
only significant for soft soil sites.

13.8 Calculation of the settlement at any


time

For relatively impermeable clayey soils, in the


short term undrained deformations occur. It is
normally assumed that construction occurs
sufficiently quickly so that no drainage occurs, and
the settlement at the end of construction is then
the immediate settlement si. For sandy soils, the
total final settlement is reached in the short term
and there is no time dependent response, thus it is
assumed that consolidation is instantaneous. Note
197

that there will be soils that have intermediate


properties, and the initial settlement will be partly
drained. The extent of the drainage (consolidation)
will depend on the boundary conditions and the
coefficient of consolidation.

For clayey soils the time settlement behaviour can


be visualised as shown in Figure 10

Construction
time
Load

Time

Const.
time Consolidation
settlement sc Total final
settlement
Settlement
Initial sTf
settlement si

Time
Fig. 10 Components of settlement

The settlement at any time t can then be


calculated from the three components described
above and it is found that:

st = si + Usc
(9a)

where U is called the degree of consolidation


198

s t −s i
U = (9b)
s Tf −s i
clearly

U = 0 when t =0

and

U = 1 when t =∞

Solutions for U versus T for a variety of boundary


conditions are given in the Data Sheets, pages 50 -
58. In general these charts use the non-
dimensionalised time factor T given by cv t / h2,
where h is the thickness of the soil layer
irrespective of the boundary conditions (Note that
this is different from the definition used for 1-D
consolidation). Solutions are given for the following
boundary conditions:

PTPB Permeable base, permeable top


boundary and permeable footing.

PTIB Impermeable base, permeable top


boundary and permeable footing.

IFIB Impermeable base, permeable top


boundary and impermeable footing.

IFPB Permeable base, permeable top


boundary and impermeable footing.

Example

Determine the immediate settlement, the final


settlement, and the settlement 1 year after the
end of construction of a rigid circular footing 5 m
in diameter which supports a load of 1.5 MN, and is
founded on a 5 m thick clay layer overlying gravel.
The clay layer has the following uniform
199

properties: E' = 5 MPa, v' = 0.2, cv = 0.5 m2/yr and


Eu = 6.25 MPa.

Step 1 Calculation of the Initial Settlement

Using Figure 4 and ν = ν u = 0.5

a 2.5
=
h 5
thus

Iρ = 0.63

The immediate settlement can now be calculated


using:
p av a
si = Iρ
Eu
with
1500
p av =
π×2.52
= 76 .39 kPa

thus
76 .39 ×2.5 ×0.63
si =
6250
= 19 .25 mm

Step 2 Calculation of the Final Settlement

Using Figure 4 and ν = ν ′ = 0.2


a
= 0.5
h
thus
Iρ = 0.95

It thus follows that


200

p av a
sTf = Iρ
E′
76 .39 × 2.5 × 0.95
=
5000
0.03629 m

Step 3 Calculation of Settlement after 1 year

(a) For the case of an Impermeable footing


(IFPB)

The consolidation settlement sc = (36.29-19.25)


mm= 17 mm

The degree of consolidation can be determined


from Figure 11, thus for:

T = cv t / h2 = 0.5 × 1 / 52 = 0.02

h/a = 2

0.0

0.2

0.4
h/a=50 20 10 5 2 1
U 0.5
0.6
0
0.8

1.0
10-5 10-4 10-3 10-2 10-1 1
cvt
T =
h2
201

Fig. 11 Consolidation response for circular


footing - case IFPB

It is found that U=0.35

This leads to a settlement after 1 year of:


s1 yr = 19 .25 +0.35 ×17

= 25 .2 mm
202

(b) For the case of a Permeable Footing


(PTPB)

0.0

0.2 0
0.5
0.4 1
2
U 5
0.6 10
20
h/a=50
0.8

1.0
10-4 10-3 10-2 10-1 1
cvt
T =
h2

Fig. 12 Consolidation response for circular


footing - case PTPB

The degree of consolidation can be determined


from Figure 12, and it is found that U=0.5 and so
the settlement after 1 year is:

s1 yr = 19 .25 + 0.50 × 17 = 27 .75 mm


203

14. SOIL STRENGTH

Soils are essentially frictional materials. They are comprised of


individual particles that can slide and roll relative to one
another. In the discipline of soil mechanics it is generally
assumed that the particles are not cemented.

One consequence of the frictional nature is that the strength


depends on the effective stresses in the soil. As the effective
stresses increase with depth, so in general will the strength.

The strength will also depend on whether the soil deformation


occurs under fully drained conditions, constant volume
(undrained) conditions, or with some intermediate state of
drainage. In each case different excess pore pressures will
occur resulting in different effective stresses, and hence
different strengths. In assessing the stability of soil
constructions analyses are usually performed to check the short
term (undrained) and long term (fully drained) conditions.

14.1 Mohr-Coulomb failure criterion

The limiting shear stress that may be applied to any plane in


the soil mass is found to be given by an equation of the form

τ = c + σn tan φ

where c = cohesion (apparent)


204

φ = friction angle

This is known as the Mohr-Coulomb failure criterion

The parameters c and φ are not generally soil constants. The


Mohr-Coulomb criterion is an empirical criterion, and the
failure locus is only locally linear. Extrapolation outside the
range of normal stresses for which it has been determined is
likely to be unreliable. The parameters depend on:

• the initial state of the soil


Overconsolidation ratio (OCR) for clays
Relative density (Id) for sands

• the type of test


Drained - slow fully drained, no excess pore water
pressures
Undrained - no drainage, excess pore water pressures
develop

• the use of total or effective stresses

In terms of effective stress the failure criterion is written

τ = c′ + σ′ n tan φ′

c′ and φ′ are referred to as the effective (drained)


strength parameters.
Soil behaviour is controlled by effective stresses, and the
effective strength parameters are the fundamental strength
parameters. But they are not necessarily soil constants.
They are fundamental in the sense that if soil is at failure
the state will always be described by an effective stress
failure criterion. The parameters can be determined from
any test provided that the pore pressures are known.

In terms of total stress the failure criterion is written

τ = cu + σn tan φu = su
205

cu, φu are referred to as the undrained (total) strength


parameters. These parameters can only be determined from
undrained tests.

The undrained strength parameters are not soil constants,


they depend strongly on the moisture content of the soil.
The total stress criterion has limited applicability as it is
only valid if soil deformation occurs without drainage.

The undrained strengths measured in the laboratory are


only relevant in practice to clayey (low permeability)
soils that initially deform without drainage, and that have
the same moisture content in-situ.

14.2 Strength Tests

The engineering strength of soil materials is often determined


from tests in either the shear box apparatus or the triaxial
apparatus.

14.2.1 The Shear Box Test

The soil is sheared along a predetermined plane by placing it in


a box and then moving the top half of the box relative to the
bottom half. The box may be square or circular in plan and of
any size, however, the most common shear boxes are square,
206

60 mm x 60 mm, and test


Normal
load
Top platen
Load cell
to
Motor measure
drive Shear
Soi
Force
l
Porous
plates

Rollers
207

A load normal to the plane of shearing may be applied to a soil


specimen through the lid of the box. Provision is made for
porous plates to be placed above and below the soil specimen.
These enable drainage to occur which is necessary if a
specimen is to be consolidated under a normal load, and if a
specimen is to be tested in a fully drained state. The soil
specimen may be submerged, by filling the containing vessel
with water, to prevent the specimens from drying out.
Undrained tests may be carried out, but in this case solid spacer
blocks rather than the porous disks must be used.

Notation

N = Normal Force
F = Tangential (Shear) Force

σn = N/A = Normal Stress


τ = F/A = Shear Stress

A = Cross-sectional area of shear plane


dx = Horizontal displacement
dy = Vertical displacement

Usually only relatively slow drained tests are performed in


shear box apparatus. For clays the rate of shearing must be
chosen to prevent excess pore pressures building up. For freely
draining sands and gravels tests can be performed quickly.
Tests on sands and gravels are usually performed dry as it is
found that water does not significantly affect the (drained)
strength.

Provided there are no excess pore pressures the pore pressure in


the soil will be approximately zero and the total and effective
stresses will be identical. That is, σn = σ´n

The failure stresses thus define an effective stress


failure envelope from which the effective (drained)
strength parameters c´, φ´ can be determined.
208

Typical test results


She
ar
load
(F)

Horizontal displacement (dx)

σ n = σ´n

At this stage we are primarily interested in the stresses at


failure. It is observed that for a set of initially similar soil
samples there is a linear failure criterion that may be expressed
as

τ = c′ + σ′ n tan φ′

From this the effective (drained) strength parameters c′ and φ′


can be determined.
209

A peak and an ultimate failure locus can be obtained from the


results each with different c´ and φ´ values. All soils are
essentially frictional materials and continued shearing results in
them approaching a purely frictional state where c′ ≈ 0.
Normally consolidated clays (OCR =1) and loose sands do not
usually show peak strengths and have c′ = 0, whereas,
overconsolidated clays and dense sands have c′ > 0. Note that
dense sands (OC clays) do not possess any true cohesion
(bonds), and the apparent cohesion results from the tendency of
soil to expand when sheared.

As a soil test the shear box is far from ideal. Disadvantages of


the test include:

Non-uniform deformations and stresses. The stresses


determined may not be those acting on the shear plane, and
no stress-strain curve can be obtained.
• There are no facilities for measuring pore pressures
in the shear box and so it is not possible to determine
effective stresses from undrained tests.
• The shear box apparatus cannot give reliable
undrained strengths because it is impossible to prevent
localised drainage away from the shear plane.

However, it has many apparent advantages:


• It is easy to test sands and gravels
• Large deformations can be achieved by reversing
the shear box. This involves pushing half of the box
backwards and forwards several times, and is useful in finding
the residual strength of a soil.
• Large samples may be tested in large shear boxes.
Small samples may give misleading results due to
imperfections (fractures and fissures) or the lack of them.
• Samples may be sheared along predetermined
planes. This is useful when the shear strengths along fissures
or other selected planes are required.
210

In practice the shear box is used to get quick and crude


estimates of the failure parameters. It is sometimes used to
obtain undrained strengths but this use should be discouraged.
14.2.2 The Triaxial Test

The triaxial test is carried out in a cell and is so named because


three principal stresses are applied to the soil sample. Two of
the principal stresses are applied to the sample by a water
pressure inside the confining cell and are equal. The third
principal stress is applied by a loading ram through the top of
the cell and therefore may be different to the other two
principal stresses. A diagram of a typical triaxial cell is shown
below.

Porous
disc
Rubber
membrane

Porous
disc
Water Water supply to soil
supply to sample
cell

A cylindrical soil specimen as shown is placed inside a latex


rubber sheath which is sealed to a top cap and bottom pedestal
by rubber O-rings. For drained tests, or undrained tests with
pore pressure measurement, porous disks are placed at the
bottom, and sometimes at the top of the specimen. For tests
where consolidation of the specimen is to be carried out, filter
211

paper drains may be provided around the outside of the specimen


in order to speed up the consolidation process.

Pore pressure generated inside the specimen during testing may


be measured by means of pressure transducers. These
transducers must operate with a very small volume change, since
fluid flowing out of the specimen would cause the pore water
pressure that was being measured to drop.
212

14.2.2.1 Stresses
F = Deviator load
σr

σr u σr = Radial stress (cell


pressure)

σa = Axial stress

F
From vertical equilibrium we have σ a = σ r + A
The term F/A is known as the deviator stress, and is usually
given the symbol q.

Hence we can write q = σa - σr = σ1 - σ3 (The axial and


radial stresses are principal stresses)

If q = 0 increasing cell pressure will result in:

• volumetric compression if the soil is free to drain. The


effective stresses will increase and so will the strength
• increasing pore water pressure if soil volume is constant (that
is, undrained). As the effective stresses cannot change it
follows that ∆u = ∆σr

Increasing q is required to cause failure

14.2.2.2 Strains

From the measurements of change in height, dh, and change in


volume dV we can determine
213

Axial strain ε a = -dh/h0


Volume strain ε v = -dV/V0

where h0 is the initial height, and V0 the initial volume. The


conventional small strain assumption is generally used.

It is assumed that the sample deforms as a right circular cylinder.


The cross-sectional area, A, can then be determined from
 dV 
1 +   1 - ε v
 V0 
A = Ao = Ao  
 1 + dh   1 - ε a
 
 h0 

1
It is important to make allowance for the changing area when
calculating the deviator stress,

q = σ1 - σ3 = F/A

14.2.2.3 Test procedure

There are many test variations. Those used most in practice are

UU (unconsolidated undrained) test.


Cell pressure applied without allowing drainage. Then
keeping cell pressure constant increase deviator load to
failure without drainage.

CIU (isotropically consolidated undrained) test.


Drainage allowed during cell pressure application. Then
without allowing further drainage increase q keeping σr
constant as for UU test.

CID (isotropically consolidated drained) test


Similar to CIU except that as deviator stress is increased
drainage is permitted. The rate of loading must be slow enough
to ensure no excess pore pressures develop.
214

As a test for investigating the behaviour of soils the triaxial test


has many advantages over the shear box test:

• Specimens are subjected to uniform stresses and


strains

• The complete stress-strain behaviour can be


investigated

• Drained and undrained tests can be performed

• Pore water pressures can be measured in undrained


tests

• Different combinations of confining and axial stress


can be applied
215

Typical results from a series of drained tests consolidated to


different cell pressures would be as follows.
q

Increasing cell
pressure

εa

The triaxial test gives the strength in terms of the principal


stresses, whereas the shear box gives the stresses on the failure
plane directly. To relate the strengths from the two tests we need
to use some results from the Mohr circle transformation of stress.
τ

σ
σ3 σ1

14.3 Mohr Circles

The Mohr circle construction enables the stresses acting in


different directions at a point on a plane to be determined,
provided that the stress acting normal to the plane is a principal
stress. The Mohr circle construction is very useful in Soil
Mechanics as many practical situations can be approximated as
plane strain problems.
216

The sign convention is different to that used in Structural


Analysis because for Soils it is conventional to take the
compressive stresses as positive.

Sign convention: Compressive normal stresses are positive


Anti-clockwise shear stresses are positive (from
inside soil element)
Angles measured clockwise positive
Let us consider the stresses acting on different planes for an
element of soil
σ1
l sinα

α l cosα
τα
σ3 l
α
α σα

(a) (b)

(a) shows the stresses on a plane at angle α to the minor


principal stress, and (b) shows the relevant lengths.

Now resolving forces gives


σ α l = σ 1 sin α l sin α + σ 3 cosα l cosα
σ1 σ
σα = (1 − cos2α ) + 3 (1 + cos2α )
2 2
( σ + σ 3 ) − ( σ 1 − σ 3 ) cos2α
σα = 1
2 2
and similarly
( σ − σ 3 ) sin2α
τ = 1
α
2

which define the Mohr circle relation


217

(τα, σα)

φ 2α σ
σ3 p
σ1

From the Mohr Circle we have

σα = p - R cos 2α
τα = R sin 2α
where
(σ1 + σ 3 ) ( σ xx + σ zz )
p = =
2 2

2
(σ1 - σ 3 ) 1
R = = ( σ xx - σ zz )2 + 4 τ 2zx
2 2

and failure occurs on a plane at an angle α from the plane on


which σ3 acts, and

π φ
α = − 
4 2

14.4 Mohr-Coulomb Failure Criterion (Principal stresses)

Failure will occur when we can find any direction such that

τα ≥ c + σα tan φ
218

φ c σ
σ3 σ1
c cot φ p

At failure from the geometry of the Mohr Circle

R = sin φ (p + c cot φ) = p sin φ + c cos φ

σ 1 + c cot φ 1 + sin φ π φ
= = tan 2  + = Nφ
σ 3 + c cot φ 1 - sin φ 4 2 

4
σ1 = Nφ σ3 + 2 c Nφ

14.4.1 Mohr-Coulomb Failure Criterion for Saturated Soil

As mentioned above it is the effective strength parameters c′ , φ′


that are the fundamental soil strength parameters. To use these
parameters the Mohr-Coulomb criterion must be expressed in
terms of effective stresses, that is

τ = c′ + σ′ n tan φ′

σ′ 1 = Nφ σ′ 3 + 2 c′ √ Nφ
1 + sin φ′
with Nφ =
1 − sin φ′
219

and the effective stresses are given by


σ′ n = σn - u
σ′ 1 = σ 1 - u
σ′ 3 = σ 3 - u

Note that the difference between the total and effective stresses is
simply the pore pressure u. Thus the total and effective stress
Mohr circles have the same diameter and are displaced along the
σ axis by the value of the pore pressure.

14.5 Interpretation of Laboratory Data

It is helpful to distinguish between drained and undrained


loading.

14.5.1 Drained loading

In drained laboratory tests the loading rate is sufficiently slow so


that all excess pore water pressures will have dissipated. From
the known pore water pressures the effective stresses can be
determined.

The behaviour of drained tests must be interpreted in terms of


the effective strength parameters c′ , φ′ , using the effective
stresses. It is possible to construct a series of total stress Mohr
Circles but the inferred total strength parameters have no
relevance to the soil behaviour.

The effective strength parameters are generally used to check the


long term (that is when all the excess pore pressures have
dissipated) stability of soil constructions. However, for sands and
gravels pore pressures dissipate rapidly and for these permeable
soils the effective strength parameters can also be used for
assessing the short term stability. In principle the effective
strength parameters can be used to check the stability at any time
for any soil type, but to do this the pore pressures in the ground
must be known and in general they are not.
220

14.5.2 Undrained loading

In undrained laboratory tests it is necessary to ensure no drainage


from the sample, or moisture redistribution within the sample
occurs. In shear box tests this requires fast rates, but because of
the more uniform conditions in the triaxial test undrained tests
can be performed more slowly simply making sure that no water
can drain from the sample.

The behaviour of undrained tests may be interpreted in terms of


the effective strength parameters c′ , φ′ , using the effective
stresses. In a triaxial test with pore pressure measurement this is
possible. The behaviour may also be interpreted in terms of the
total strength parameters cu, φu. However, if the total stress
parameters are being used they must be determined from
Unconsolidated Undrained tests if they are to be relevant to the
soil in the ground.

Let us consider the behaviour of three identical saturated soil


samples in undrained triaxial tests. No water is allowed to drain
and three different confining pressures are applied (Samples are
Unconsolidated). The Mohr circles at failure will be as follows

σ
σ ′3 σ 1′ σ3 σ1

From the total stress Mohr circles we find that φ = φ u = 0.

Because all samples are at failure the effective stress failure


condition must also be satisfied, and because all the circles have
the same radius there must be a single effective stress Mohr
221

circle. The different total stress Mohr circles indicate that the
samples must have different pore water pressures.

The explanation for the independence of the undrained strength


on the confining stress is that increasing the cell pressure
without allowing drainage has the effect of increasing the pore
pressure by the same amount (∆u = ∆σr). There is therefore no
change in effective stress. As it is the effective stresses that
control the soil behaviour the subsequent strength is unaffected.
The change in pore pressure during shearing is a function of the
initial effective stress and the moisture content. As these are
identical for the three samples an identical strength is obtained.
As will be shown later the fact that the moisture content
remains constant is the most important factor in having a
constant strength.

In some series of unconsolidated undrained tests it is found that


for different soil samples from a particular site φu is not zero, or
cu is not constant. If this occurs then either

• the samples are not saturated, or

• the samples have different moisture contents

The undrained strength cu is not a fundamental soil parameter.

The total stress strength parameters cu, φu are often used to assess
the short term (undrained) stability of soil constructions. It is
important that no drainage should occur otherwise this approach
is not valid. Therefore, for sands and gravels which drain rapidly
a total stress analysis would not be appropriate.

For soils that do not drain freely this approach is the only simple
way of assessing the short term stability, because in general the
pore water pressures are unknown.

Note however, that it is possible to measure an undrained


strength for any type of soil in the triaxial apparatus.
222

Example

In an unconsolidated undrained triaxial test the undrained


strength is measured as 17.5 kPa. Determine the cell pressure
used in the test if the effective strength parameters are c´ = 0, φ´
= 26o and the pore pressure at failure is 43 kPa.

Analytical solution

( σ 1 − σ 3 ) = ( σ 1′ − σ ′3 )
Undrained strength = 17.5 =
2 2

Effective stress failure criterion σ′ 1 = Nφ σ′ 3 + 2 c′ √ Nφ

1 + sin φ′
c´ = 0, Nφ = = 2.561
1 − sin φ′

Hence σ’ = 57.4 kPa, σ3’ = 22.4 kPa

and cell pressure (total stress) = σ3’ + u = 65.4 kPa


223

Graphical solution

o
26
τ
17.5

σ
σ ′3 σ 1′ σ3 σ1
224

15. STRESS-STRAIN BEHAVIOUR OF SOILS

15.1 The behaviour of sands

In practice sands are usually sheared under drained


conditions because their relatively high
permeability ensures that excess pore pressures
are not generated. This behaviour can be
investigated in a variety of laboratory apparatus.
We will consider the behaviour in simple shear
tests. The simple shear test is similar to the shear
box test but it has the advantage that the strain
and stress states are more uniform enabling us to
investigate the stress-strain behaviour. The name
simple shear refers to the plane strain mode of
deformation shown below:
σ
dx τ
dz

H γxz

γxz = dx/H εz = - dz/H = εv

For this deformation there are only two non-zero


strain components, these are the shear strain, γ xz
225

= dx/H, and the normal strain ε z = dz/H. The


volume strain, ε v = ε z.

For sands the two most important parameters


governing their behaviour are the Relative Density,
Id, and the effective stress level, σ′ . The Relative
density is defined by
emax - e
Id =
emax - emin

6
where emax and emin are the maximum and minimum
void ratios that can be measured in standard tests
in the laboratory, and e is the current void ratio.
This expression can be re-written in terms of dry
density as
Gs γ w
γd =
1 + e

7
and hence
1 1
-
γ dmin γd
Id =
1 1
-
γ dmin γ dmax

8
Sand is generally referred to as dense if Id > 0.6
and loose if Id < 0.3.

15.1.1 Influence of Relative Density

The influence of relative density on the behaviour


can be seen in the plots below for tests all
performed at the same normal stress.
226

τ D ense (D )
CSL

τ =Mσ ′ed iu m (M )
tanφ ′ult
e L
L o o se (L )

M
σ ′γ
εv D
D

M γ
γ

The following observations can be made:

• All samples approach the same ultimate


conditions of shear stress and void ratio,
irrespective of the initial density

• Initially dense samples attain higher peak


angles of friction (φ′ = tan-1 (τ/σ′ ) )
227

• Initially dense soils expand (dilate) when


sheared, and initially loose soils compress
228

15.1.2 Influence of Effective Stress Level

The influence of stress level can be seen in the plots


below where the two dense samples have the same
initial void ratio, e1 and similarly the loose samples
both have the same initial void ratio e2.

e
τ CSL
D2 σ1
D1
τ = σ′tanφ′ult σ2
L2

L1
σ′γ
D1
εv τ D1 CSL
D2 σ’
D2 τ = σ′tanφ′ult

L1
γ L2
L1
L2 σ′γ

The following observations can be made:

• The ultimate values of shear stress and void


ratio, depend on the stress level, but the
ultimate angle of friction (φ′ ult = tan-1 (τ/σ′ ) ult)
is independent of both density and stress level
229

• Initially dense samples attain higher peak


angles of friction (φ′ = tan-1 (τ/σ′ )), but the
peak friction angle reduces as the stress level
increases.

• Initially dense soils expand (dilate) when


sheared, and initially loose soils compress.
Increasing stress level causes less dilation
(greater compression).
230

15.1.3 Ultimate or Critical States

All soil when sheared will eventually attain a unique


stress ratio given by τ/σ′ = tan φ′ ult, and reach a
critical void ratio which is uniquely related to the
normal stress. This ultimate state is referred to as
a Critical State, defined by
dτ dσ ′ d εv
= = = 0
dγ dγ dγ
e Straight line
9 e = e0 - λ lnσ′
The locus of these critical
states defines a line
known as the Critical
State Line (CSL). This
may be represented by lnσ′
τ CSL

τ = σ′tanφ′ult τ

σ′

e σ′

e
CSL
σ′
231

At critical states soil behaves as a purely frictional


material

φ′ = φ′ ult = φ′ cs = constant = F (mineralogy,


grading, angularity)
232

15.1.4 Stress-Dilatancy Relation

During a simple shear test on dense sand the top


platen is forced up against the applied normal
stress. Work must be done against this external
force in addition to the work done in overcoming
friction between the particles. Thus the frictional
resistance of the soil may appear to be greater than
φ′ ult. Another way to demonstrate this is to consider
a "saw-tooth" analogy.

F
α N

Q = F cos α + N sin α
P = − F sin α + N cos α
Q F cos α + N sin α
=
P − F sin α + N cos α

Q
=
( F N ) + tan α
P 1 − ( F N ) tan α
233

Q F
Now = tan φ ′ and = tan φ ′ult
P N
tan φ ′ult + tan α
tan φ ′ =
1 − tan φ ′ult tan α

φ′ = φ ′ult + α

dy dε v
tan α = ≈
dx dγ
234

15.1.5 Peak Conditions

The failure conditions are normally expressed by a


Mohr-Coulomb criterion using parameters c´, φ´.
This is the approach that we will be following in
estimating the stability of soil constructions.
τ
Dense sand - Peak
strengths c’, φ’

Ultimate strength
c’ = 0, φ’ = φ’ult
φ’

c’

σ’

However, this approach obscures the fact that c´ is


only an apparent cohesion. An alternative method
of presenting the results is to determine the
maximum friction angle φ´pk which in shear box type tests
is simply given by tan-1(τ/σ´). The relation between φ´pk
and effective stress is then as shown below.
235

φ’pk Id =
1
Id =
0.5
Id =
0
φ’ult

The position of the lines in this plot is a function of


the mineralogy and angularity of the soil.

Note that even loose sand can have φ´pk > φ´ult if the
stress is low enough. This means that loose sands may expand
when sheared.
236

15.1.6 Implications for stability analysis

If you choose to use φ′ pk (or c′ , φ′ with c′ ≠ 0) in


stability calculations then you are saying that
everywhere on the critical failure surface the soil
will be dilating at failure. In most practical cases
this is unlikely to be realistic. For instance consider
the case of a retaining wall.

τ
Wall Failure C
B
Surface
A
γ

τ
B
A C
τ
A
B γ
C
γ

It is conservative to use c´ = 0 and φ´ = φ´ult for stability


analyses.
237

15.2 Behaviour of clays

The behaviour of clays is essentially identical to


that of sands. The data however is usually
presented in terms of the soils stress history (OCR)
rather than relative density.

To predict the behaviour of soil we need to combine


the CSL with our previous knowledge concerning
the consolidation behaviour. Experience has shown
that the CSL is parallel to the normal consolidation
line and lies below it in a void ratio, effective stress
plot.
e

sw ellin g
line

C S L N C L - n o rm al
con so lidatio n line

log σ’
We find that normally consolidated clays behave
similarly to loose sands and heavily over-
consolidated clays behave similarly to dense sands.
As the OCR increases there is a gradual trend
between these extremes. The response in drained
simple shear tests with σ´ constant is as follows
238

τ CSL τ CSL
OCR = 1
τ = σ′tanφ′ult τ = σ′tanφ′ult

OCR = 8

σ′ σ′γ
e εv OCR = 8

NCL
γ
CSL
OCR = 1
σ′
239

15.2.1 Undrained response

In an undrained test volume change is prevented


and therefore the void ratio must remain constant.
Because the soil always heads towards a critical
state when sheared it is possible to show the path
that will be followed in an e, σ′ plot. This is shown
below for normally consolidated (OCR=1) and
heavily over-consolidated (OCR>8) samples having
the same initial void ratio. Once the final states in
this plot are known, so too are the final states in the
τ, σ′ plot. Also if the final total stresses are known
then the excess pore pressures can be determined.

τ CSL τ CSL

τ = σ ′ tanφ ′ult OCR = τ1 = σ ′ tanφ ′ult

OCR = 8

σ′ σ ′γ
e u +ve
OCR = 1

NCL
γ
CSL
OCR = 8
-ve
σ′

• Knowledge of the Critical State Line enables an


explanation for the existence of apparent
240

cohesion (undrained strength) in frictional


materials

• It is also clear that if the moisture content


changes then so will the undrained strength,
because failure will occur at a different point on
the CSL
241

15.3 Differences between sand and clay

When considering the behaviour of sands and clays


we generally use different parameters. For sands
stress level and relative density are considered to
be the important parameters, whereas for clays the
parameters are stress level and stress history
(OCR).

However, the broad patterns of behaviour observed


for sands and clays are very similar. To understand
why different "engineering" parameters are used it
is useful to consider the positions of the
consolidation and CSL lines in the void ratio,
effective stress plot.

e
Clay

Loose
Sand
Dense

NCL NCL

0.1 1 10 100 log σ’ (MPa)

You might also like