Sunteți pe pagina 1din 334

Based on part of the GeotechniCAL reference package

by Prof. John Atkinson, City University, London

Description & classification


Basic characteristics of soils

 Soil as an engineering material


 Size range of grains
o Identification
 Shape of grains
o SAND grains
o CLAY grains
o Specific surface
 Composition of grains
 Structure or fabric

Origins, formation and mineralogy


 Origins of soils from rocks
 Weathering of rocks
 Clay minerals
 Transportation and deposition
 Loading and drainage history

Grading and composition


 Coarse soils
o Particle size tests
o Simulation
o Typical grading curves
o Grading characteristics
o Sieve analysis example
 Fine soils
o Consistency
o Simulation
o Plasticity index
o Plasticity chart
o Activity
 Specific gravity

Volume-weight properties

 Volume: the soil model


 Simulation
o Degree of saturation
o Air-voids content
 Masses of solid and water
 Densities and unit weights
 Laboratory measurements
o Water content
o Unit weight
 Field measurements

Current state of soil

1
 Deposition and erosion
 Ageing
 Density index
 Liquidity index
 Predicting stiffness and strength

BS system for description and classification


 BS description system
 Definitions of terms
 British Soil Classification System

2
Soil description and classification

 Basic characteristics of soils


 Origins, formation and mineralogy
 Grading and composition
 Volume-weight properties
 Current state of soil
 British Standard system

It is necessary to adopt a formal system of soil description and classification in order to


describe the various materials found in ground investigation. Such a system must be
comprehensive (covering all but the rarest of deposits), meaningful in an engineering context
(so that engineers will be able to understand and interpret) and yet relatively concise. It is
important to distinguish between description and classification:

Description of soil is a statement describing the physical nature and state of the soil. It can be
a description of a sample, or a soil in situ. It is arrived at using visual examination, simple
tests, observation of site conditions, geological history, etc.

Soil classification is the separation of soil into classes or groups each having similar
characteristics and potentially similar behaviour. A classification for engineering purposes
should be based mainly on mechanical properties, e.g. permeability, stiffness, strength. The
class to which a soil belongs can be used in its description.
 
  Description and classification

1-Basic characteristics of soils

 Soil as an engineering material


 Size range of grains
 Shape of grains
 Composition of grains
 Structure or fabric

Soils consist of grains (mineral grains, rock fragments, etc.)


with water and air in the voids between grains. The water and air contents are readily changed
by changes in conditions and location: soils can be perfectly dry (have no water content) or be
fully saturated (have no air content) or be partly saturated (with both air and water present).
Although the size and shape of the solid (granular) content rarely changes at a given
point, they can vary considerably from point to point.
 

First of all, consider soil as a engineering material - it is not a coherent solid material like
steel and concrete, but is a particulate material. It is important to understand the significance
of particle size, shape and composition, and of a soil's internal structure or fabric.
  

Basic characteristics of soils

1.1 Soil as an engineering material


3
The term "soil" means different things to different people: To a geologist it represents the
products of past surface processes. To a pedologist it represents currently occurring physical
and chemical processes. To an engineer it is a material that can be:

built on: foundations to buildings, bridges.


built in: tunnels, culverts, basements.
built with: roads, runways, embankments, dams.
supported: retaining walls, quays.

Soils may be described in different ways by different people for their different purposes.
Engineers' descriptions give engineering terms that will convey some sense of a soil's current
state and probable susceptibility to future changes (e.g. in loading, drainage, structure, surface
level). Engineers are primarily interested in a soil's mechanical properties: strength, stiffness,
permeability. These depend primarily on the nature of the soil grains, the current stress, the
water content and unit weight.
 
 Basic characteristics of soils

1.2 Size range of grains

 Aids to size identification

The range of particle sizes encountered in soil is very large: from boulders with a controlling
dimension of over 200mm down to clay particles less than 0.002mm (2m). Some clays
contain particles less than 1  in size which behave as colloids, i.e. do not settle in water due
solely to gravity.

BOULDERS > 200 mm


In theBritish Soil Classification System, soils are
classified into named Basic Soil Type groups
according to size, and the groups further divided
COBBLES 60 - 200 mm
into coarse, medium and fine sub-groups:
  Very coarse
soils
coarse 20 - 60 mm
G medium 6 - 20 mm
GRAVEL fine 2 - 6 mm
Coarse
coarse 0.6 - 2.0 mm
soils
S medium 0.2 - 0.6 mm
SAND fine 0.06 - 0.2 mm
coarse 0.02 - 0.06 mm
M medium 0.006 - 0.02 mm
Fine SILT fine 0.002 - 0.006 mm
soils
< 0.002 mm
C  CLAY

4
1.3  Size range of grains

Aids to size identification

Soils possess a number of physical characteristics which can be used as aids to size
identification in the field. A handful of soil rubbed through the fingers can yield the
following:

SAND (and coarser) particles are visible to the naked eye.


SILT particles become dusty when dry and are easily brushed off hands and boots.
CLAY particles are greasy and sticky when wet and hard when dry, and have to be scraped or
washed off hands and boots.
 

Shape of grains

 Shape characteristics of SAND grains


 Shape characteristics of CLAY grains
 Specific surface

The majority of soils may be regarded as either SANDS or CLAYS:

SANDS include gravelly sands and gravel-sands. Sand grains are generally broken rock
particles that have been formed by physical weathering, or they are the resistant components
of rocks broken down by chemical weathering. Sand grains generally have a rotund shape.

CLAYS include silty clays and clay-silts; there are few pure silts (e.g. areas formed by
windblown Löess). Clay grains are usually the product of chemical weathering or rocks and
soils. Clay particles have a flaky shape.

There are major differences in engineering behaviour between SANDS and CLAYS (e.g. in
permeability, compressibility, shrinking/swelling potential). The shape and size of the soil
grains has an important bearing on these differences.
 
  Shape characteristics of SAND grains

SAND and larger-sized grains are rotund. Coarse soil grains (silt-sized, sand-sized and
larger) have different shape characteristics and surface roughness depending on the amount of
wear during transportation (by water, wind or ice), or after crushing in manufactured
aggregates. They have a relatively low specific surface (surface area).

Rounded: Water- or air-worn; transported sediments


Irregular: Irregular shape with round edges; glacial sediments (sometimes sub-divided into
'sub-rounded' and 'sub-angular')
Angular: Flat faces and sharp edges; residual soils, grits
Flaky: Thickness small compared to length/breadth; clays
Elongated: Length larger than breadth/thickness; scree, broken flagstone
Flaky & Elongated: Length>Breadth>Thickness; broken schists and slates
5
 
 

Shape characteristics of CLAY grains

CLAY particles are flaky. Their thickness is very small relative to their length & breadth, in
some cases as thin as 1/100th of the length. They therefore have high to very high specific
surface values. These surfaces carry a small negative electrical charge, that will attract the
positive end of water molecules. This charge depends on the soil mineral and may be affected
by an electrolite in the pore water. This causes some additional forces between the soil grains
which are proportional to the specific surface. Thus a lot of water may be held asadsorbed
water within a clay mass.

  
Shape of grains

Specific surface

 Examples

Specific surface is the ratio of surface area per unit wight.


Surface forces are proportional to surface area (i.e. to d²).
Self-weight forces are proportional to volume (i.e. to d³).
Surface force  1 
Therefore 
self weight forces d
       area  1 
Also, specific surface = 
 * volume d

Hence, specific surface is a measure of the relative contributions of surface forces and self-
weight forces.

The specific surface of a 1mm cube of quartz ( = 2.65gm/cm³) is 0.00023 m²/N

SAND grains (size 2.0 - 0.06mm) are close to cubes or spheres in shape, and have specific
surfaces near the minimum value.

6
CLAY particles are flaky and have much greater specific surface values.

Examples of specific surface

The more elongated or flaky a particle is the greater will be its specific surface.

Click on the following examples:


cubes, rods, sheets

Examples of mineral grain specific surfaces:


 

Grain width Specific Surface


Mineral/Soil Thickness
d (m) m²/N
Quartz grain 100 d 0.0023
Quartz sand 2.0 - 0.06 d 0.0001 - 0.004
Kaolinite 2.0 - 0.3 0.2d 2
Illite 2.0 - 0.2 0.1d 8
Montmorillonite 1.0 - 0.01 0.01d 80

Structure or fabric

Natural soils are rarely the same from one point in the ground to another. The content and
nature of grains varies, but more importantly, so does the arrangement of these.

The arrangement and organisation of particles and other features within a soil mass is termed
its structure or fabric. This includes bedding orientation, stratification, layer thickness, the
occurrence of joints and fissures, the occurrence of voids, artefacts, tree roots and nodules, the
presence of cementing or bonding agents between grains.

Structural features can have a major influence on in situ properties.

 Vertical and horizontal permeabilities will be different in alternating layers of fine and
coarse soils.
 The presence of fissures affects some aspects of strength.
 The presence of layers or lenses of different stiffness can affect stability.
 The presence of cementing or bonding influences strength and stiffness.

Origins, formation and mineralogy

 Origins of soils from rocks


 Weathering of rocks
 Clay minerals
 Transportation and deposition
 Loading and drainage history

7
Soils are the results of geological events (except for the very small amount produced by man).
The nature and structure of a given soil depends on the geological processes that formed it:
breakdown of parent rock: weathering, decomposition, erosion.
transportation to site of final deposition: gravity, flowing water, ice, wind.
environment of final deposition: flood plain, river terrace, glacial moraine, lacustrine or
marine.
subsequent conditions of loading and drainage - little or no surcharge, heavy surcharge due
to ice or overlying deposits, change from saline to freshwater, leaching, contamination.
 
  Origins, formation and mineralogy

Origins of soils from rocks

All soils originate, directly or indirectly, from solid rocks in the Earth's crust:

igneous rocks
crystalline bodies of cooled magma
e.g. granite, basalt, dolerite, gabbro, syenite, porphyry

sedimentary rocks
layers of consolidated and cemented sediments, mostly formed in bodies of water (seas, lakes,
etc.)
e.g. limestone, sandstones, mudstone, shale, conglomerate

metamorphic rocks
formed by the alteration of existing rocks due to heat from igneous intrusions (e.g. marble,
quartzite, hornfels) or pressure due to crustal movement (e.g. slate, schist, gneiss).
 
  Origins, formation and mineralogy

Weathering of rocks

Physical weathering
Physical or mechanical processes taking place on the Earth's surface, including the actions of
water, frost, temperature changes, wind and ice; cause disintegration and wearing. The
products are mainly coarse soils (silts, sands and gravels). Physical weathering produces Very
Coarse soils and Gravels consisting of broken rock particles, but Sands and Silts will be
mainly consists of mineral grains.

Chemical weathering
Chemical weathering occurs in wet and warm conditions and consists of degradation by
decomposition and/or alteration. The results of chemical weathering are generally fine soils
with separate mineral grains, such as Clays and Clay-Silts. The type of clay mineral depends
on the parent rock and on local drainage. Some minerals, such as quartz, are resistant to the
chemical weathering and remain unchanged.

quartz
A resistant and enduring mineral found in many rocks (e.g. granite, sandstone). It is the
principal constituent of sands and silts, and the most abundant soil mineral. It occurs as
equidimensional hard grains.
haematite
A red iron (ferric) oxide: resistant to change, results from extreme weathering. It is
8
responsible for the widespread red or pink colouration in rocks and soils. It can form a cement
in rocks, or a duricrust in soils in arid climates.
micas
Flaky minerals present in many igneous rocks. Some are resistant, e.g. muscovite; some are
broken down, e.g. biotite.
clay minerals
These result mainly from the breakdown of feldspar minerals. They are very flaky and
therefore have very large surface areas. They are major constituents of clay soils, although
clay soil also contains silt sized particles.
 
  Origins, formation and mineralogy

Clay minerals

Clay minerals are produced mainly from the chemical weathering and decomposition of
feldspars, such as orthoclase and plagioclase,
and some micas. They are small in size and
very flaky in shape.

The key to some of the properties of clay soils,


e.g. plasticity, compressibility,
swelling/shrinkage potential, lies in the
structure of clay minerals.

There are three main groups of clay minerals:


kaolinites
(include kaolinite, dickite and nacrite) formed
by the decomposition of orthoclase feldspar
(e.g. in granite); kaolin is the principal
constituent in china clay and ball clay.
illites
(include illite and glauconite) are the commonest clay minerals; formed by the decomposition
of some micas and feldspars; predominant in marine clays and shales (e.g. London clay,
Oxford clay).
montmorillonites
(also called smectites or fullers' earth minerals) (include calcium and sodium
momtmorillonites, bentonite and vermiculite) formed by the alteration of basic igneous rocks
containing silicates rich in Ca and Mg; weak linkage by cations (e.g. Na+, Ca++) results in
high swelling/shrinking potential

Origins, formation and mineralogy

Transportation and deposition

The effects of weathering and transportation largely determine the basic nature of the soil
(i.e. the size, shape, composition and distribution of the grains). The environment into which
deposition takes place, and subsequent geological events that take place there, largely

9
determine the state of the soil, (i.e. density, moisture content) and the structure or fabric of
the soil (i.e. bedding, stratification, occurrence of joints or fissures, tree roots, voids, etc.)

Transportation
Due to combinations of gravity, flowing water or air, and moving ice. In water or air: grains
become sub-rounded or rounded, grain sizes are sorted, producing poorly-graded deposits. In
moving ice: grinding and crushing occur, size distribution becomes wider, deposits are well-
graded, ranging from rock flour to boulders.

Deposition
In flowing water, larger particles are deposited as velocity drops, e.g. gravels in river terraces,
sands in floodplains and estuaries, silts and clays in lakes and seas. In still water: horizontal
layers of successive sediments are formed, which may change with time, even seasonally or
daily.

 Deltaic & shelf deposits: often vary both horizontally and vertically.
 From glaciers, deposition varies from well-graded basal tills and boulder clays to
poorly-graded deposits in moraines and outwash fans.
 In arid conditions: scree material is usually poorly-graded and lies on slopes.
 Wind-blown Löess is generally uniformly-graded and false-bedded.

Origins, formation and mineralogy

Loading and drainage history

The current state (i.e. density and consistency) of a soil will have been profoundly influenced
by the history of loading and unloading since it was deposited. Changes in drainage
conditions may also have occurred which may have brought about changes in water content.

Loading /unloading history

Initial loading
During deposition the load applied to a layer of soil increases as more layers are deposited
over it; thus, it is compressed and water is squeezed out; as deposition continues, the soil
becomes stiffer and stronger.

Unloading
The principal natural mechanism of unloading is erosion of overlying layers. Unloading can
also occur as overlying ice-sheets and glaciers retreat, or due to large excavations made by
man. Soil expands when it is unloaded, but not as much as it was initially compressed; thus it
stays compressed - and is said to be overconsolidated. The degree of overconsolidation
depends on the history of loading and unloading.

Drainage history

Chemical changes
Some soils initially deposited loosely in saline water and then inundated with fresh water
develop weak collapsing structure. In arid climates with intermittent rainy periods, cycles of
wetting and drying can bring minerals to the surface to form a cemented soil.

10
Climate changes
Some clays (e.g. montmorillonite clays) are prone to large volume changes due to wetting and
drying; thus, seasonal changes in surface level occur, often causing foundation damage,
especially after exceptionally dry summers. Trees extract water from soil in the process of
evapotranspiration; The soil near to trees can therefore either shrink as trees grow larger, or

expand following the removal of large trees.


 
 

Description and classification

Grading and composition


 Coarse soils
 Fine soils
 Specific gravity

The recommended standard for soil classification is the British Soil Classification System,
and this is detailed in BS 5930 Site Investigation.
 

Grading and composition

Coarse soils
 Particle size tests
 Typical grading curves
 Grading characteristics
 Sieve analysis example

Coarse soils are classified principally on the basis of particle size and grading.
Very coarse BOULDERS > 200 mm
soils COBBLES 60 - 200 mm
coarse 20 - 60 mm
G
medium 6 - 20 mm
GRAVEL
Coarse fine 2 - 6 mm
soils coarse 0.6 - 2.0 mm
S
medium 0.2 - 0.6 mm
SAND
fine 0.06 - 0.2 mm

11
Coarse soils

Particle size tests


The aim is to measure the distribution of particle sizes in the sample. When a wide range of
sizes is present, the sample will be sub-divided, and separate tests carried out on each sub-
sample. Full details of tests are given in BS 1377: "Methods of test for soil for civil
engineering purposes".

Particle-size tests

Wet sieving to separate fine grains from coarse grains is carried out by washing the soil
specimen on a 60m sieve mesh.
Dry sieving analyses can only be carried out on particles > 60 m. Samples (with fines
removed) are dried and shaken through a nest of sieves of descending size.
Sedimentation is used only for fine soils. Soil particles are allowed to settle from a
suspension. The decreasing density of the suspension is measured at time intervals. Sizes are
determined from the settling velocity and times recorded. Percentages between sizes are
determined from density differences.

Particle-size analysis

The cumulative percentage quantities finer than certain sizes (e.g. passing a given size sieve
mesh) are determined by weighing. Points are then plotted of % finer (passing) against log
size. A smooth S-shaped curve drawn through these points is called a grading curve. The
position and shape of the grading curve determines the soil class. Geometrical grading
characteristics can be determined also from the grading curve.

 
Coarse soils

Typical grading curves

12
Both the position and the shape of the grading curve for a soil can aid its identity and
description.
Some typical grading curves are shown in the figure:
A - a poorly-graded medium SAND (probably estuarine or flood-plain alluvium)
B - a well-graded GRAVEL-SAND (i.e. equal amounts of gravel and sand)
C - a gap-graded COBBLES-SAND
D - a sandy SILT (perhaps a deltaic or estuarine silt)
E - a typical silty CLAY (e.g. London clay, Oxford clay)
 
  Coarse soils

Grading characteristics

A grading curve is a useful aid to soil description. Grading curves are often included in
ground investigation reports. Results of grading tests can be tabulated using geometric
properties of the grading curve. These properties are called grading characteristics

First of all, three points are located on the grading curve:


d10 = the maximum size of the smallest 10% of the sample
d30 = the maximum size of the smallest 30% of the sample
d60 = the maximum size of the smallest 60% of the sample

From these the grading characteristics are calculated:


Effective size
d10
Uniformity coefficient
Cu = d60 / d10
Coefficient of gradation
Ck = d30² / d60 d10

Both Cu and Ck will be 1 for a single-sized soil


Cu > 5 indicates a well-graded soil
Cu < 3 indicates a uniform soil
Ck between 0.5 and 2.0 indicates a well-graded soil

13
Ck < 0.1 indicates a possible gap-graded soil

Coarse soils

Sieve analysis example


The results of a dry-sieving test are given below, together with the grading analysis and
grading curve. Note carefully how the tabulated results are set out and calculated. The grading
curve has been plotted on special semi-logarithmic paper; you can also do this analysis using
a spreadsheet.

Sieve mesh Mass Percentage Percentage


size (mm) retained (g) retained finer (passing)
14.0 0 0 100.0
10.0 3.5 1.2 98.8
6.3 7.6 2.6 86.2
5.0 7.0 2.4 93.8
3.35 14.3 4.9 88.9
2.0 21.1 7.2 81.7
1.18 56.7 19.4 62.3
0.600 73.4 25.1 37.2
0.425 22.2 7.6 29.6
0.300 26.9 9.2 20.4
0.212 18.4 6.3 14.1
0.150 15.2 5.2 8.9
0.063 17.5 6.0 2.9
Pan 8.5 2.9
TOTAL 292.3 100.0

The soil comprises: 18% gravel, 45% coarse sand, 24% medium sand, 10% fine sand, 3% silt,
and is classified therefore as: a well-graded gravelly SAND

14
 
 

Grading and composition

Fine soils
 Consistency limits and plasticity
 Plasticity index
 The plasticity chart and classification
 Activity

In the case of fine soils (e.g. CLAYS and SILTS), it is the shape of the particles rather than
their size that has the greater influence on engineering properties. Clay soils have flaky
particles to which water adheres, thus imparting the property of plasticity.
 
  Fine soils

Consistency limits and plasticity

Consistency varies with the water content of the soil. The consistency of a soil can range from
(dry) solid to semi-solid to plastic to liquid (wet). The water contents at which the consistency
changes from one state to the next are called consistency limits (or Atterberg limits).

Two of these are utilised in the classification of fine soils:


Liquid limit (wL) - change of consistency from plastic to liquid
Plastic limit (wP) - change of consistency from brittle/crumbly to plastic

Measures of liquid and plastic limit values can be obtained from laboratory tests.
  Fine soils

Plasticity index (indice de plasticité)

15
The consistency of most soils in the ground will be plastic or semi-solid. Soil strength and
stiffness behaviour are related to the range of plastic consistency. The range of water content
over which a soil has a plastic consistency is termed the Plasticity Index (IP or PI).

IP = liquid limit - plastic limit


= wL - wP

Fine soils

The plasticity chart and classification

In the BSCS fine soils are divided into ten classes based on their measured plasticity index
and liquid limit values: CLAYS are distinguished from SILTS, and five divisions of plasticity
are defined:
Low plasticity wL = < 35%
Intermediate plasticity wL = 35 - 50%
High plasticity wL = 50 - 70%
Very high plasticity wL = 70 - 90%
Extremely high plasticity wL = > 90%

A plasticity chart is provided to aid classification.

Fine soils

Activity

So-called 'clay' soils are not 100% clay. The proportion of clay mineral flakes (< 2 m size) in
a fine soil affects its current state, particularly its tendency to swell and shrink with changes in
water content. The degree of plasticity related to the clay content is called the activity of the
soil.

Activity = P / (% clay particles)

Some typical values are:

Mineral Activity Soil Activity


16
Muscovite 0.25 Kaolin clay 0.4-0.5
Kaolinite 0.40 Glacial clay and loess 0.5-0.75
Illite 0.90 Most British clays 0.75-1.25
Montmorillonite > 1.25 Organic estuarine clay > 1.25

Grading and composition

Specific gravity

Specific gravity (Gs) is a property of the mineral or rock material forming soil grains.
It is defined as

Method of measurement
For fine soils a 50 ml density bottle may be used; for coarse soils a 500 ml or 1000 ml jar. The
jar is weighed empty (M1). A quantity of dry soil is placed in the jar and the jar weighed (M2).
The jar is filled with water, air removed by stirring, and weighed again (M3). The jar is
emptied, cleaned and refilled with water - and weighed again (M4).

[The range of Gs for common soils is 2.64 to 2.72]

Description and classification

Volume-weight properties

 Volumes of solid, water and air: the soil model


 Masses of solid and water: water content
 Densities and unit weights
 Laboratory measurements

17
 Field measurements

The volume-weight properties of a soil define its state. Measures of the amount of void space,
amount of water and the weight of a unit volume of soil are required in engineering analysis
and design.

Soil comprises three constituent phases:


Solid: rock fragments, mineral grains or flakes, organic matter.
Liquid: water, with some dissolved compounds (e.g. salts).
Gas: air or water vapour.

In natural soils the three phases are intermixed. To aid analysis it is convenient to consider a
soil model in which the three phases are seen as separate, but still in their correct proportions.
 
 

Volume-weight properties

Volumes of solid, water and air: the soil model

 Degree of saturation
 Air-voids content

The soil model is given dimensional values for the solid, water and air components: Total
volume,
V = Vs + Vw + Va

Since the amounts of both water and air are variable, the volume of solids present is taken as
the reference quantity. Thus, the following relational volumetric quantities may be defined:

18
Note also that:
n = e / (1 + e)
e = n / (1 - n)
v = 1 / (1 - n)

Typical void ratios might be 0.3 (e.g. for a dense, well graded granular soil) or 1.5 (e.g. for a
soft clay).

Volumes of solid, water and air: the soil model

Degree of saturation

The volume of water in a soil can only vary between zero (i.e. a dry soil) and the volume of
voids; this can be expressed as a ratio:

For a perfectly dry soil:


Sr = 0
For a saturated soil:
Sr = 1

Note: In clay soils as the amount water increases the volume and therefore the volume of
voids will also increase, and so the degree of saturation may remain at Sr = 1 while the actual
volume of water is increasing.
 
 

Volumes of solid, water and air: the soil model

Air-voids content

19
The air-voids volume, Va , is that part of the void space not occupied by water.

Va = Vv - Vw
= e - e.Sr
= e.(1 - Sr)

Air-voids content, Av
Av = (air-voids volume) / (total volume)
= Va / V
= e.(1 - Sr) / (1+e)
= n.(1 - Sr)

For a perfectly dry soil:


Av = n
For a saturated soil:
Av = 0
 

Volume-weight properties

Masses of solid and water: water content

The mass of air may be ignored. The mass of solid particles is usually expressed in terms of
their particle density or grain specific gravity.

Grain specific gravity

Hence the mass of solid particles in a soil


Ms = Vs .Gs .w
20
(w = density of water = 1.00Mg/m³)
[Range of Gs for common soils: 2.64-2.72]

Particle density
s = mass per unit volume of particles
= Gs .w

The ratio of the mass of water present to the mass of solid particles is called the water
content, or sometimes the moisture content.

From the soil model it can be seen that


w = (Sr .e .w) / (Gs .w)

Giving the useful relationship:


w .Gs = Sr .e
 

Volume-weight properties

Densities and unit weights

Density is a measure of the quantity of mass in a unit volume of material.


Unit weight is a measure of the weight of a unit volume of material.

There are two basic measures of density or unit weight applied to soils: Dry density is a
measure of the amount of solid particles per unit volume. Bulk density is a measure of the
amount of solid + water per unit volume.

The preferred units of density are:


Mg/m³, kg/m³ or g/ml.

The corresponding unit weights are:

21
Also, it can be shown that
 = d(1 + w) and
 = gd(1 + w)
 

Volume-weight properties

Laboratory measurements

 Water content
 Unit weight

It is important to quantify the state of a soil immediately it is received in the testing laboratory
and just prior to commencing other tests (e.g. shear tests, compression tests, etc.).

The water content and unit weight are particularly important, since these could change during
transportation and storage.

Some physical state properties are calculated following the practical measurement of others;
e.g. void ratio from porosity, dry unit weight from unit weight & water content.
 
  Laboratory measurements

Water content

The most usual method of determining the water content of soil is to weigh a small
representative specimen, drying it to constant weight and then weighing it again. Drying can
be carried out using an electric oven set at 104-105° Celsius or using a microwave oven.

Example: A sample of soil was placed in a tin container and weighed, after which it was
dried in an oven and then weighed again. Calculate the water content of the soil.

Weight of tin empty = 16.16 g


Weight of tin + moist soil
= 37.82 g
Weight of tin + dry soil = 34.68 g
Water content, w = (mass of water) / (mass of dry soil)
= (37.82 - 34.68) / (34.68 - 16.16)
= 0.169
Percentage water content = 16.9 %

Laboratory measurements

Unit weight

22
Clay soils: Specimens are usually prepared in the form of regular geometric shapes, (e.g.
prisms, cylinders) of which the volume is easily computed.
Sands and gravels: Specimens have to be placed in a container to determine volume (e.g. a
cylindrical can).

Example
A soil specimen had a volume of 89.13 ml, a mass before drying of 174.45 g and after drying
of 158.73 g; the water content was 9.9 %. Determine the bulk and dry densities and unit
weights.

Bulk density
 = (mass of specimen) / (volume of specimen)
= 174.45 / 89.13 g/ml
= 1.957 Mg/m³
[1 g/ml = 1 Mg/m³]

Unit weight
 = 9.81m/s² x  Mg/m³
= 19.20 kN/m³

Dry density
d = (mass after drying) / (volume)
= 158.73 / 89.13
= 1.781 Mg/m³
d =  / (1 + w)
= 1.957 / (1+0.099)
= 1.781 Mg/m³

Dry unit weight


d =  / (1 + w)
= 19.20 / (1+0.099)
= 17.47 kN/m³

Volume-weight properties

Field measurements

Measurements taken in the field are mostly to determine density/unit weight. The most
common application is the determination of the density of rolled and compacted fill, e.g. in
road bases, embankments, etc.

Note: These methods are covered in detail by BS1377. You should understand the general
principle that density is calculated from the mass and volume of a sample. How a sample of
known volume is obtained depends on the nature of the soil. You are not expected to
remember the details of each method.

23
The core cutter method
This method is suitable for soft fine grained soils.

A steel cylinder is driven into the ground, dug out and the soil shaved
off level. The mass of soil is found by weighing and deducting the
mass of the cylinder. Small samples are taken from both ends and the
water content determined.

The sand-pouring cylinder method


This method is suitable for stony soils

Using a special tray with a hole in the centre, a hole is formed in the
soil and the mass of soil removed is weighed.

The volume of the hole is calculated from the mass of clean dry
running sand required to fill the hole.

The sand-pouring cylinder is used to fill the hole in a controlled


manner. The mass of sand required to fill the hole is equal to the
difference in the weight of the cylinder before and after filling the hole,
less an allowance for the sand left in the cone above the hole.

Bulk density
 = (mass of soil) / (volume of core cutter or hole)

Description and classification

Current state of soil

 Soil history: deposition and erosion


 Soil history: ageing
 Density index (relative density)
 Liquidity index
 Predicting stiffness and strength from index properties

The state of soil is essentially the closeness of packing of the grains in the range:
Closely packed    Loosely packed
Dense  Loose
Low water content  High water content
Strong and stiff  Weak and soft

The important indicators of the current state of a soil are:


current stresses: vertical and horizontal effective stresses
current water content: effecting strength and stiffness in fine soils
liquidity index: indicates state in fine soils
density index: indicates state of compaction in coarse soils
history of loading and unloading: degree of overconsolidation

24
Engineering operations (e.g. excavation, loading, unloading, compaction, etc.) on soil bring
about changes in its state. Its initial state is the result of processes of erosion and deposition. It
is possible for the engineer to predict changes that could result from a proposed engineering
operation: changes from the soil's current state to a new future state.
 
  Current state of soil

Soil history: deposition and erosion

Original deposition
Most soils are formed in layers or lenses by deposition from moving water, ice or wind.

One-dimensional compression occurs as overlying layers are added. Vertical and horizontal
stresses increase with deposition.

Erosion
Erosion causes unloading; stresses decrease; some vertical expansion occurs.

Plastic strain has occurred; the soil remains compressed, i.e. overconsolidated.

Subsequent changes
Subsequent changes may occur in the depositional environment: further loading/unloading
due to glaciation, land movement, engineering; and ageing processes.
 
 

Current state of soil

Soil history: ageing

The term ageing includes processes that occur with time, except loading and unloading.
Ageing processes are independent of changes in loading.

Vibration and compaction


Coarse soils can be made more dense by vibration or compaction at essentially constant
effective stress

Creep
Fine soils creep and continue to compress and distort at constant effective stress after primary
consolidation is complete.

Cementing and bonding


Intergranular cementing and bonding occurs due to deposition of minerals from groundwater,
e.g. calcium carbonate; disturbance due to excavation fractures the bonding and reduces
strength.

Weathering
Physical and chemical changes take place in soils near the ground surface due to the influence
of changes in rainfall and temperature.

25
Changes in salinity
Changes in the salinity of groundwater are due to changes in relative sea and land levels, thus
soil originally deposited in sea water may later have fresh water in its pores, such soils may be
prone to sudden collapse.
 
 

Current state of soil

Density index (relative density)

The void ratio of coarse soils (sands and gravels) varies with the state of packing between the
loosest practical state in which it can exist and the densest. Some engineering properties are
affected by this, e.g.shear strength, compressibility, permeability.

It is therefore useful to measure the in situ state and this can be done by comparing the in situ
void ratio (e) with the minimum and maximum practical values (emin and emax) to give a
density index D

emin is determined with soil compacted densely in a metal mould


emax is determined with soil poured loosely into a metal mould

Density index is also known as relative density


Relative states of compaction are defined:

Density index State of compaction

0-15% Very loose

15-35 Loose

35-65 Medium

65-85 Dense

85-100% Very dense

26
Current state of soil

Liquidity index

In fine soils, especially clays, the current state is dependent on the water content with respect
to the consistency limits (or Atterberg limits). The liquidity index (L or LI) provides a
quantitative measure of the current state:

where
wP = plastic limit and
wL = liquid limit

Significant values of IL indicating the consistency of the soil are:


IL < 0  semi-plastic solid or solid
0 < IL < 1  plastic
1 < IL  liquid
 

Current state of soil

Predicting stiffness and strength from index properties

Preliminary estimates of strength and stiffness can provide a useful basis for early design and
feasibility studies, and also the planning of more detailed testing programmes. The following
suggestions have been made; they are simple, but not necessarily reliable(fiable), and should
be not be used in final design calculations.

Undrained shear strength

27
su = 170 exp(-4.6 L) kN/m²

[Schofield and Wroth (1968)]

su = (0.11 + 0.37 P) 'vo kN/m²

where 'vo = vertical effective stress in situ


[Skempton and Bjerrum (1957)]

Stiffness
The slope of the critical state line may be estimated from:
 = P .Gs / 461
[After Skempton and Northey (1953)]

The compressibility index may be estimated from:


Cc =  ln10 = P Gs / 200
(where P is in percentage units)
 

Description and classification

BS system for description and classification

 BS description system
 Definitions of terms used in description
 British Soil Classification System (BSCS)

BS 5930 Site Investigation recommends the terminology and a system for describing and
classifying soils for engineering purposes. Without the use of a satisfactory system of
description and classification, the description of materials found on a site would be
meaningless or even misleading, and it would be difficult to apply experience to future
projects.
 
  BS system for description and classification

BS description system

A recommended protocol for describing a soil deposit uses ninecharacteristics; these should
be written in the following order:

compactness
e.g. loose, dense, slightly cemented
bedding structure
e.g. homogeneous or stratified; dip, orientation
discontinuities
spacing of beds, joints, fissures
weathered state
degree of weathering
colour
28
main body colour, mottling
grading or consistency
e.g. well-graded, poorly-graded; soft, firm, hard
SOIL NAME
e.g. GRAVEL, SAND, SILT, CLAY; (upper case letters) plus silty-, gravelly-, with-fines, etc.
as appropriate
soil class
(BSCS) designation (for roads & airfields) e.g. SW = well-graded sand
geological stratigraphic name
(when known) e.g. London clay

Not all characteristics are necessarily applicable in every case.

Example:
(i) Loose homogeneous reddish-yellow poorly-graded medium SAND (SP), Flood plain
alluvium
(ii) Dense fissured unweathered greyish-blue firm CLAY. Oxford clay.
 
 

BS system for description and classification

Definitions of terms used in description

A table is given in BS 5930 Site Investigation setting out a recommended field indentification
and description system. The following are some of the terms listed for use in soil descriptions:

Particle shape
angular, sub-angular, sub-rounded, rounded, flat, elongate
Compactness
loose, medium dense, dense (use a pick or driven peg, or density index )
Bedding structure
homogeneous, stratified, inter-stratified
Bedding spacing
massive(>2m), thickly bedded (2000-600 mm), medium bedded (600-200 mm), thinly bedded
(200-60 mm), very thinly bedded (60-20 mm), laminated (20-6 mm), thinly laminated (<6
mm).
Discontinuities
i.e. spacing of joints and fissure: very widely spaced(>2m), widely spaced (2000-600 mm),
medium spaced (600-200 mm), closely spaced (200-60 mm), very closely spaced (60-20
mm), extremely closely spaced (<20 mm).
Colours
red, pink, yellow, brown, olive, green, blue, white, grey, black
Consistency
very soft (exudes between fingers), soft (easily mouldable), firm (strong finger pressure
required), stiff (can be indented with fingers, but not moulded) very stiff (indented by sharp
object), hard (difficult to indent).
Grading
well graded (wide size range), uniform (very narrow size range), poorly graded (narrow or
uneven size range).
Composite soils
In SANDS and GRAVELS: slightly clayey or silty (<5%), clayey or silty (5-15%), very
29
clayey or silty(>15%)
In CLAYS and SILTS: sandy or gravelly (35-65%)
 
  British Soil Classification System

The recommended standard for soil classification is the British Soil Classification System,
and this is detailed in BS 5930 Site Investigation. Its essential structure is as follows:
Soil group   Symbol   Recommended name
Coarse soils     Fines %  
G  GW 0-5 Well-graded GRAVEL
  GPu/GPg 0-5 Uniform/poorly-graded GRAVEL
G-F GWM/GWC 5 - 15 Well-graded silty/clayey GRAVEL

GRAVEL   GPM/GPC 5 - 15 Poorly graded silty/clayey GRAVEL


Very silty GRAVEL [plasticity sub-
GF GML, GMI... 15 - 35
group...]
Very clayey GRAVEL [..symbols as
  GCL, GCI... 15 - 35
below]
S SW 0-5 Well-graded SAND
  SPu/SPg 0-5 Uniform/poorly-graded SAND
S-F SWM/SWC 5 - 15 Well-graded silty/clayey SAND
SAND
  GPM/GPC 5 - 15 Poorly graded silty/clayey SAND
SF SML, SMI... 15 - 35 Very silty SAND [plasticity sub-group...]
  SCL, SCI... 15 - 35 Very clayey SAND [..symbols as below]
       
Liquid limit
Fine soils   >35% fines  
%
MG   Gravelly SILT
SILT M MS   Sandy SILT
ML, MI...   [Plasticity subdivisions as for CLAY]
CG   Gravelly CLAY
CS   Sandy CLAY
CL <35 CLAY of low plasticity
CLAY C CI 35 - 50 CLAY of intermediate plasticity
CH 50 - 70 CLAY of high plasticity
CV 70 - 90 CLAY of very high plasticity
CE >90 CLAY of extremely high plasticity
Organic soils O     [Add letter 'O' to group symbol]

30
Peat Pt     [Soil predominantly fibrous and organic]

Groundwater
 Introduction
 Pore water pressure
o Water table
o Elevation, pressure and total head
o Hydraulic gradient
o Effective stress
 Permeability
o Void ratio
o Stratified soil
o Seepage velocity
o Temperature
 Analytical solutions
o Steady one-dimensional flow
o Quasi-one-dimensional and radial flow
 Cylindrical flow: confined aquifer
 Cylindrical flow: groundwater lowering
 Spherical flow
o Two-dimensional flow, Laplace
o Transient flow, consolidation
 Flow nets
o Calculation of flow
o Calculation of total flow
o Boundary between layers
o Boundary conditions
o Flow through embankments
 Quick condition and piping
o Seepage force
o Critical hydraulic gradient
 Measurement
o Pore pressure: Laboratory
o Pore pressure: Field
o Permeability: Laboratory
 Permeameter
 Oedometer
o Permeability: Field
 Pumping test
 Constant head and falling head tests
 Groundwater control
o Groundwater lowering
o Groundwater exclusion
 Some case histories
o Construction below gound water level
Bristol 2000 car park
o Diaphragm cut-off wall
Sizewell B power station
o Simulation of rising groundwater
Canary Warf, London
o Ground freezing
Rheinberg, West Germany
o Deep dewatering wells
Blackwater Valley
o Complex groundwater control system
Munich airport
o Buoyancy-dependent foundation
Lecco bypass, near Milan, Italy
o Wellpoint pumping
Sevenoaks, Kent
31
o Artificial island
Trans Tokyo Bay Highway, Japan
o Flooded tunnel and ground freeze
Thames Water ring main, London
o Rising groundwater levels
The deep aquifer beneath London
o Flooded cofferdam
Ennerdale Link Road, River Hull

Based on part of the GeotechniCAL reference package


Back to home page by Prof. David Muir Wood, Bristol University

Pore water pressure


 Water table
 Elevation, pressure and total head
 Hydraulic gradient
 Effective stress

In general, the water in the voids of an element of saturated soil will be under
pressure, either due to the physical location of the soil or as a result of external forces. This
pressure is the pore water pressure or pore pressure u. It is measured relative to atmospheric
pressure.

When there is no flow, the pore pressure at depth d below the water surface is:  u = w d

Water table

 Fine-grained soils
 Coarse-grained soils
 Perched water table

The level in the ground at which the pore pressure is zero (equal to atmospheric) is defined as
the water table or phreatic surface.

When there is no flow, the water surface will be at exactly the same level in any stand pipe
placed in the ground below the water table. This is called a hydrostatic pressure condition.

The pore pressure at depth d below the water table is: :  u = w d

32
 

Water table

Fine-grained soils
In fine grained soils, surface tension effects can cause capillary water to rise above the water
table. It is reasonable to assume that the pore pressure varies linearly with depth, so the pore
pressure above the water table will be negative.

If the water table is at depth dw then the pore pressure at the ground surface is uo = - w.dw
and the pore pressure at depth z is u = w (z - dw)

Where the water table is deeper, or where evaporation is taking place from the surface,
saturation with capillary water may not occur. The height to which the soil remains saturated
with negative pore pressures above the water table is called the capillary rise.

Water table

Coarse-grained soils

Below the water table the soil can be considered to be saturated. In coarse-grained soils, water
will drain from the pores and air will therefore be present in the soil between the ground
surface and the water table.

Consequently, pore pressures above the water table can usually be ignored. Below the water
table, hydrostatic water pressure increases linearly with depth.

With the water table at depth dw


u = 0 for z < dw
u = w(z - dw) for z > dw
 

Water table

Perched water table


Where the ground contains layers of permeable soil (e.g. sands) interspersed with layers of
much lower permeability (e.g. clays) one or more perched water tables may develop and the
overall distribution of pore pressure with depth may not be exclusivelyly linear.

33
Detection of perched water tables during site investigation is important, otherwise erroneous
estimates of in-situ pore pressure distributions can arise.

Pore pressure conditions below perched water tables may be affected by local infiltration of
rainwater or localised seepage and therefore may not be in hydrostatic equilibrium.

Pore water pressure

Elevation, pressure and total head


Pore pressure at a given point (e.g. point A in the diagram) can be measured by the
height of water in a standpipe located at that point.

Pore pressures are often indicated in this way on diagrams.

The height of the water column is the pressure head (hw)

 u 
hw = w
To identify significant differences in pore pressure at different points, we need to
eliminate the effect of the points' position. A height datum is required from which
locations are measured. The elevation head (hz) of a point is its height above the
datum line. The height above the datum of the water level in the standpipe is the
total head (h).

h = hz + hw 

Pore water pressure

Hydraulic gradient
Flow of pore water in soils is driven from positions of higher total head towards positions of
lower total head. The level of the datum is arbitrary. It is differences in total head that are
important. The hydraulic gradient is the rate of change of total head along the direction of
flow.

i = h / s

In each diagram there are two points, a small distance s apart, hz1 and hz2 above datum.
34
In the first diagram, the total heads are equal. The difference in pore pressure is entirely due
to the difference in altitude of the two points and the pore water has no tendency to flow.
In the second diagram, the total heads are different. The hydraulic gradient is
i = (h2 - h1) / s
and the pore water tends to flow.

Pore water pressure

Effective stress

 All strength and stress:strain characteristics of soils can be linked to changes in effective
stress

Effective stress (') = total stress () - pore water pressure (u)
' =  - u

Groundwater

Permeability

 Void ratio
 Stratified soil
 Seepage velocity
 Temperature

Darcy's law
The rate of flow of water q (volume/time) through
cross-sectional area A is found to be proportional to hydraulic gradient i according to Darcy's
law:

v =  q   = k.i             i =  h  
A s

where v is flow velocity and k is coefficient of permeability with dimensions of velocity


(length/time).

The coefficient of permeability of a soil is a measure of the conductance (i.e. the reciprocal of
the resistance) that it provides to the flow of water through its pores.

35
The value of the coefficient of permeability k depends on the average size of the pores and is
related to the distribution of particle sizes, particle shape and soil structure. The ratio of
permeabilities of typical sands/gravels to those of typical clays is of the order of 106. A small
proportion of fine material in a coarse-grained soil can lead to a significant reduction in
permeability.

Permeability

Void ratio and permeability

Permeability of all soils is strongly influenced by the density of packing of the soil particles
which can be simply desrcibed through void ratio e or porosity n.

Sands

For filter sands it is found that k  0.01 (d10)² m/s where d10 is the effective particle size in
mm. This relationship was proposed by Hazen.

The Kozeny-Carman equation suggests that, for laminar flow in saturated soils:

where ko and kT are factors depending on the shape and tortuosity of the pores respectively, Ss
is the surface area of the solid particles per unit volume of solid material, and w and  are
unit weight and viscosity of the pore water. The equation can be written simply as

Clays

The Kozeny-Carman equation does not work well for silts and clays. For clays it is typically
found that

where Ck is the permeability change index and ek is a reference void ratio.


For many natural clays Ck is approximately equal to half the natural void ratio.

36
Permeability

Stratified soil and permeability

Consider a stratified soil having horizontal layers of thickness t1, t2, t3, etc. with coefficients of
permeability k1, k2 k3, etc.

For vertical flow, the flow rate q through area A of each layer is the same. Hence the head
drop across a series of layers is

The average coefficient of permeability is

For horizontal flow, the head drop h over the same flow path length s will be the same for
each layer. So i1 = i2 = i3 etc. The flow rate through a layered block of soil of breadth B is
therefore

The average coefficient of permeability is

Permeability

Seepage velocity(vitess de fuite)

Darcy's Law relates flow velocity (v) to hydraulic gradient (i). The volume flow rate q is
calculated as the product of flow velocity v and total cross sectional area:
q = v.A

At the particulate level the water follows a tortuous path through the pores. The average
velocity at which the water flows through the pores is the ratio of volume flow rate to the
average area of voids Av on a cross section normal to the macroscopic direction of flow. This
is the seepage velocity vs
37
Porosity of soil is related to the volume fraction of voids

Seepage velocity can be measured in laboratory models by injecting dye into the seeping pore
water and timing its progress through the soil.

Permeability

Temperature and permeability

The flow of water through confined spaces is controlled by its viscosity  and the viscosity is
controlled by temperature.

An alternative permeability K (dimensions: length² ) is sometimes used as a more absolute


coefficient depending only on the characteristics of the soil skeleton.

The values of k at 0°C and 10°C are 56% and 77% respectively of the value measured at
20°C.

Groundwater

Analytical solutions

 Steady one-dimensional flow


 Quasi-one-dimensional and radial flow
 Two-dimensional flow, Laplace
 Transient flow, consolidation

In steady-state flow, the pressures and flow rates remain constant over time. In transient
flow, the pressures and flow rates are time-dependent.

Steady one-dimensional flow is the simplest case, to which Darcy's law can be applied. This
can be extended to cases of variable aquifer thickness and radial flow. The analysis of steady
two-dimensional flow is more complex and results in flow nets.

Analytical solutions

Steady one-dimensional flow

Darcy's Law indicates the link between flow rate and hydraulic gradient. For one-dimensional
flow, constant flow rate implies constant hydraulic gradient.
Steady downward flow occurs when water is pumped from an underground aquifer. Pore
38
pressures are then lower than hydrostatic pressures.
Steady upward flow occurs as a result of artesian pressure when a less permeable layer is
underlain by a permeable layer which is connected through the ground to a water source
providing pressures higher than local hydrostatic pressures.
The fountains of London were originally driven by artesian pressure in the aquifers trapped
beneath the London clay. Pumping from aquifers over the centuries has lowered the water
pressures below artesian levels.

Analytical solutions

Quasi-one-dimensional and radial flow

 Cylindrical flow: confined aquifer


 Cylindrical flow: groundwater lowering
 Spherical flow

Where flow occurs in a confined aquifer whose thickness varies gently with position the flow
can be treated as being essentially one-dimensional.
The horizontal flow rate q is constant. For an aquifer of width B and varying thickness t, the
discharge velocity

and Darcy's Law indicates that

Hydraulic gradient varies inversely with aquifer thickness.

39
Quasi-one-dimensional and radial flow

Cylindrical flow: confined aquifer

Steady-state pumping to a well which extends the full thickness of a confined aquifer is a one-
dimensional problem which can be analysed in cylindrical coordinates: pore pressure or head
varies only with radius r.

Darcy's Law still applies, with hydraulic gradient dh/dr and area A varying with radius: A =
2r.t

where ro is the radius of the borehole and h0 the constant head in the borehole.

Quasi-one-dimensional and radial flow

Cylindrical flow: groundwater lowering

Pumping from a borehole can be used for deliberate groundwater lowering in order to
facilitate excavation. This is an example of quasi-one-dimensional radial flow with flow
thickness t=h. Then A=2r.h and

 
40
Quasi-one-dimensional and radial flow

Spherical flow

Variation of pore pressure around a point source or side (for example, a piezometer being
used for in-situ determination of permeability) is a one-dimensional problem which can be
analysed in spherical coordinates: pore pressure or head varies only with radius r.

Darcy's Law still applies, with hydraulic gradient dh/dr and area A varying with radius:
A=4r²

where r0 is the radius of the piezometer and h0 the constant head in the piezometer.

Analytical solutions

Two-dimensional flow, Laplace

 Anisotropic soil

Two-dimensional steady flow of the incompressible pore fluid is governed by Laplace's


equation which indicates simply that any imbalance in flows into and out of an element in the
x direction must be compensated by a corresponding opposite imbalance in the y direction.
Laplace's equation can be solved graphically, analytically, numerically, or analogically.
41
For a rectangular element with dimensions x, y and unit thickness, in the x direction the
velocity of flow into the element is

the negative sign being required because flow occurs down the hydraulic gradient. The
velocity of flow out of the element is

Similar expressions can be written for the y direction. Balance of flow requires that

and this is Laplace's equation. In three dimensions, Laplace's equation becomes

Two-dimensional flow, Laplace

Anisotropic soil

For a soil with permeability kx and ky in the x and y directions respectively, Laplace's equation
for two-dimensional seepage becomes

This can be solved by applying a scale factor to the x dimensions so that transformed
coordinates xt are used

42
In the transformed coordinates the equation regains its simple form

and flownet generation can proceed as usual. Calculations of flow are made using an
equivalent permeability

It may be preferable in some cases to transform the y coordinates using:

The equivalent permeability remains unchanged.

For many natural sedimentary soils seasonal variations in the depositional regime have
resulted in horizontal macroscopic permeabilities significantly greater than vertical
permeabilities. Transformation of coordinates lends itself to analysis of seepage in such
situations.

Analytical solutions

Transient flow, consolidation

Since water may be regarded as being essentially incompressible, unsteady flow may arise
when water is drawn into or expelled from pores as a result of changes in the size of pores.
This can only occur as a result of changes volume associated with changes in effective stress.

The time-dependent transient change in pore pressure that occurs as a result of some
perturbation, and associated change in effective stress is called consolidation.

One-dimensional compression tests in an oedometer define the relationship between vertical


effective stress 'v and specific volume v or void ratio e from which a one-dimensional
compliance mv can be defined

43
Then, under conditions of constant total stress, consolidation is governed by a diffusion
equation:

where cv is the coefficient of consolidation having dimensions (length²/time).

Solutions of the consolidation equation are typically presented as isochrones, i.e. variations
of pore pressure with position at successive times, but can also be converted to curves linking
settlement with time.

Groundwater

Flow nets
 Calculation of flow
 Calculation of total flow
 Boundary between layers
 Boundary conditions
 Flow through embankments

Solutions to Laplace's equation for two-dimensional seepage can be presented as flow nets.
Two orthogonal sets of curves form a flow net:
equipotentials connecting points of equal total head h

flow lines indicating the direction of seepage down a hydraulic gradient

If standpipe piezometers were inserted into the ground with their tips on a single equipotential
then the water would rise to the same level in each standpipe. (The pore pressures would be
different because of their different elevations.)
44
There can be no flow along an equipotential, because there is no hydraulic gradient, so there
can be no component of flow across a flow line. The flow lines define channels along which
the volume flow rate is constant.

Flow nets

Calculation of flow

Consider an element from a flow channel of length L between equipotentials which indicate a
fall in total head h and between flow lines b apart. The average hydraulic gradient is

and for unit width of flow net the volume flow rate is

There is an advantage in displaying or sketching flownets in the form of curvilinear 'squares'


so that a circle can be insrcibed within each four-sided figure bounded by two equipotentials
and two flow lines. Then b = L and q = kh
so the flow rate through the flow channel is the permeability multiplied by the uniform
interval h between equipotentials.

Flow nets

Calculation of total flow

For a complete problem, the flownet has been drawn with the overall head drop h divided into
Nd equal intervals:

h = h / Nd

with Nf flow channels.

Then the total flow rate per unit width is

45
It is usually convenient in sketching flownets to make Nd an integer. The number of flow
channels Nf will then generally not be an integer. In the example shown, of flow under a sheet
pile wall
Nd ;= 10, Nf = 3.5 and q = 0.35kh per unit width.

Flow nets

Boundary between layers(limites des couches)

Flow across a boundary between two layers of soil of different permeability produces a
refraction effect.

Consideration of continuity of flow and of continuity of velocity normal to the interface


shows that

It is not possible to construct a flow net with curvilinear squares on both sides of the interface
unless the head drop between equipotentials is changed in inverse proportion to the
permeability ratio.

If the ratio of permeabilities is greater than about 10, e.g. at the boundary of a drainage layer
then construction of the part of the flow net in the more permeable soil is unlikely to be
necessary.

46
 

Flow nets

Boundary conditions (conditions aux limites)

A surface on which the total head is fixed (for example, from the level of a river, pool,
reservoir) is an equipotential. A surface across which there is no flow (for example, an
impermeable soil layer or an impermeable wall) is a flow line

For the situation shown, with flow occurring under a sheet pile wall, the axis of symmetry
must also be an equipotential.

Flow nets

Flow through embankments


Seepage through an embankment dam is an example of unconfined flow bounded at the upper
surface by a phreatic surface which represents the top flow line and on which the pore pressure
is everywhere zero (atmospheric).
Total head changes and elevation changes thus match and for equal head intervals h between
equipotentials there must be equal vertical intervals between the points of intersection of
equipotentials with the phreatic surface.

47
Groundwater (eau souterraine)

Quick condition and piping

 Seepage force (débit de fuite)


 Critical hydraulic gradient

If the flow is upward then the water pressure tends to lift the soil element. If the upward water
pressure is high enough the effective stresses in the soil disappear, no frictional strength can
be mobilised and the soil behaves as a fluid. This is the quick or quicksand condition and is
associated with piping instabilities around excavations and with liquefaction events in or
following earthquakes.

Quick condition and piping

Seepage force

The viscous drag of water flowing through a soil imposes a seepage force on the soil in the
direction of flow.

Consider the actual distribution of pore water pressure around an element length L and
thickness b taken from a flownet, bounded by two equipotentials with fall in head h, and two
flow lines.

These pore water pressures are partly supporting the weight wbL of water in the element and
partly providing the seepage force. It is found that the seepage force is
J = i w b L
equivalent to a seepage pressure (force per unit volume) in the direction of flow
j = i w

Quick condition and piping

Critical hydraulic gradient

48
The quick condition occurs at a critical upward hydraulic gradient ic, when the seepage force
just balances the buoyant weight of an element of soil. (Shear stresses on the sides of the
element are neglected.)

The critical hydraulic gradient is typically around 1.0 for many soils. Fluidised beds in
chemical engineering systems rely on deliberate generation of quick conditions to ensure that
the chemical process can occur most efficiently.

 Groundwater

Measurement
 Pore pressure: Laboratory
 Pore pressure: Field
 Permeability: Laboratory
 Permeability: Field

Measurement

Laboratory measurement of pore pressure

Laboratory measurements of pore pressure are required in undrained testing where soil
properties are to be measured in terms of effective stresses and in model tests which involve
the loading or unloading of beds of clay.

Traditionally, pore pressures are measured in the drainage lines just outside a triaxial cell. It is
essential that the pore pressure measurement system should be completely free of air and as
stiff as possible so that minimal amounts of pore water flow are required in order to register
the changes in pore pressure.

In research testing, miniature pore-pressure transducers may be mounted directly on a soil


sample in order to speed up the expected response time.

 Field measurement of pore pressure

Pore pressures can be measured in the ground with different types of piezometer. The
simplest piezometer consists of a open tube or standpipe with a porous tip. The change in
measured head of water requires a large flow of water into the tube, so response is slow.

A Casagrande piezometer comprises a porous tip in a filter zone at the base of a borehole,
connected to a narrow tube. The cross sectional area of the tube is small by comparison with
the surface area of the filter, so the flow required to register a change in pressure is smaller
than in a standpipe, and the response is quicker.

Closed circuit piezometer systems are read remotely by mechanical or electrical means and
provide possibilities for de-airing the pore water circuit. The response time is dependent on
49
the length of connecting tubing.

Electrical transducers (using strain gauge or vibrating wire techniques) can be placed in the
ground. These are stiff devices which respond rapidly, but can be difficult to keep de-aired
particularly if there is a possibility of the surrounding soil becoming unsaturated.

  Laboratory measurement of permeability

 Permeameter
 Oedometer

Laboratory measurements of the permeability of soils can be made using a permeameter. For
fine-grained soils (clays), the coefficient of permeability can be estimated directly or
indirectly during one-dimensional compression tests in an oedometer.

  Laboratory measurement of permeability

Permeameter

Constant head test


Recommended for coarse-grained soils.
Steady total head drop  is measured across gauge length L, as water flows through a
sample of cross-section area A.

Falling head test


Recommended for fine-grained soils.
Total head h in standpipe of area a is allowed to fall; heads h1 and h2 are measured at times t1
and t2.
Hydraulic gradient /L varies with time.

  Oedometer

Indirect measurement
Transient consolidation phenomena are controlled by the coefficient of consolidation. With
knowledge of one-dimensional compliance mv, coefficient of permeability k can be estimated
from

50
Direct measurement
Direct measurement of permeability in oedometers is preferable. Flow pumps can be used to
maintain a constant flow rate (q) across the sample and to measure the resultant constant head
(h). The coefficient of permeability is then given by k = q.L / A.h

Field measurement of permeability

 Pumping test
 Constant head and falling head tests

Field or in-situ measurement of permeability avoids the difficulties involved in obtaining and
setting up undisturbed samples in a permeameter or oedometer and also provides information
about bulk permeability, rather than merely the permeability of a small and possibly
unrepresentative sample.

Pumping test(essai de pompage)

In a well-pumping test, the steady-state heads h1 and h2 in observation boreholes at radii r1 and
r2 are monitored at flow rate q. If the pumping causes a drawdown in an unconfined (i.e. open
surface) soil stratum then

If the soil stratum is confined and of thickness t and remains saturated then

Constant head and falling head tests with in-situ piezometers can also be used.

Constant head and falling head tests

Field tests equivalent to the laboratory constant head and falling head tests can be performed
in which controlled heads or flows are applied to piezometer tips. In general, conditions
around such piezometers are not ideally cylindrically symmetric or spherically symmetric and

51
an intake factor F (with dimensions of length) is required for each particular geometry. Values
of the intake factor may be deduced from analytical or numerical studies.

For a borehole open to its base, of diameter D, and lined to the full depth F=2.75D.
If the cased hole is through impermeable soil and the base of the casing is at the interface with
a permeable stratum F=2D.

For an intake formed by a cylindrical filter zone of diameter D and length L in an infinite
isotropic stratum

    for L/D > 4

Then for a steady state, constant head test in which a flow q is required to maintain a head h:

For a falling head test in which heads h1 and h2 are measured at times t1 and t2 in a borehole of
area A:

Groundwater control

 Groundwater lowering
 Groundwater exclusion

Failure to control groundwater adjacent to a construction project may result in

 flooding
 instability
 ground movement
 loss of bearing capacity

The available techniques fall into 4 broad categories:


 control of surface water
 removal from within the works
 extraction from the surrounding ground
 exclusion

Selection of the most appropriate method depends on cost, which depends on


 the nature of the ground
 the size of the works
 the duration of the works
 the level of acceptable risk

 
52
 

Groundwater lowering(rabattage de la nappe)

Removal of water from the ground will cause the water level to fall. How quickly and by how
much depends on the permeability of the soil and the distance between the adjacent wells.

Drainage of clays is impractical. Silt particles may be removed along with the water causing
the formation of voids in the ground and damage to pumps. The high yield in gravels may
make the method impractical.

Lowering the ground water will reduce the pore water pressure and hence increase stability.
Removing water will tend to cause settlement although the effect is likely to be small for
sandy soils in which the technique works well.

Groundwater control

Groundwater exclusion
Exclusion methods involve the installation of an impermeable barrier. This may be structural
(steel sheet piles or concrete diaphragm wall) and may form a part of the permanent work.

Other methods include

 slurry trench with bentonite or native clay


 thin grouted membrane
 other forms of grouting
 ground freezing
 compressed air (for tunnels and shafts)

Excluding water may cause a build up in pore water pressure. Heave is a particular problem
where a thin layer of impermeable soil at the base of an excavation results in a high pore
water pressure close to the base.

Groundwater

Case studies
The following brief project descriptions illustrate some issues associated with groundwater
flow. The list is not exhaustive. In each case, a reference is given to the source material.
 Diaphragm cut-off wall Sizewell B power station
 Ground freezing Rheinberg, West Germany
 Deep dewatering wells Blackwater Valley
 Complex groundwater control system Munich airport
 Buoyancy dependent foundation Lecco bypass, near Milan, Italy
 Wellpoint pumping Sevenoaks, Kent
 Artificial island Trans Tokyo Bay Highway, Japan

53
 Flooded tunnel and ground freeze Thames Water ring main, London
 Rising groundwater levels The deep aquifer beneath London
 Simulation of rising groundwater Canary Warf, London
 Flooded cofferdam Ennerdale Link Road, River Hull

Case studies

Diaphragm cut-off wall


The Sizewell B power station

Due to the groundwater conditions, construction of


the Sizewell B station could not begin until the UK's deepest ever diaphragm wall was
installed to isolate the site. The foundations would need excavations nearly 18m below the
water table, and dewatering consultants were appointed in 1984. Conventional dewatering
techniques were rejected for several reasons. The next idea was to construct a diaphragm wall,
extending into the London Clay, and linking with a cofferdam to form a 1260m-long all-
encompasing cut-off wall around the whole site.  The established techniques for diaphragm
walling had only been used down to 30m at that time, but trenching with new reverse
circulation rigs could go down to more than 100m very accurately. This solution had several
advantages, and had been selected by July 1985. The wall had to have a controlled maximum
permeability and virtually leak proof construction joints. Performance was monitored via a
network of observation wells and piezometers. After more than 4000000m³ of water had been
pumped away, the excavation was dry until the pumps were switched off in the spring of
1992. The water table had been kept at least 2m below the deepest excavation.

Reference

groundwater conditions
50m of dense sands and silts known as the Norwich Crag deposits overlay London Clay and
form a natural aquifer. The site is also uncomfortably close to the North sea.

Reasons
Preliminary calculations showed that even with 52 wells (rather than 6 used for the A station),
it would still only be possible to lower the water by 16m. The excessive draw down below
adjacent bird reserves, settlement beneath Sizewell A, heavy encrustation on the pipework
due to high iron content in the groundwater, and a cost of at least £16M, all made this option
unacceptable.

advantages
Only nine dewatering wells would be needed, and the construction period would be halved to
six months with a saving of £2M.

 
 
 
 

54
Case studies

Ground freezing
Rheinberg, West Germany

A record 15GJ/h refrigeration plant formed a tube of ice-stabilised strata reaching 528m down
to a dry zechstein formation overlying the Ruhr coalfield: the world's largest ground freeze.
The artificially stabilised section of ground, with its steel and concrete lining structure, is the
start of a 7.5m diameter shaft continuing down to 1.3km. Its prime purpose is ventilation of
the nearby mines. Proximity of the Rhine demanded that a 4m platform should be built to lift
the working area above any flood level. But the rise and fall of groundwater levels in the 14m
of terrace sands and gravels is insignificant compared with conditions deeper down. Tertiary
deposits of sands, silts and clays go down to 100m where there is about 30m chalk. Then
there is Bunter sandstone, which is very soft and charged with water right down to 511m
where the dry and stable zechstein occurs.

The 44 freeze pipe holes, with 127mm diameter casing, were drilled on a 22m diameter circle,
and the freeze was computer-monitored. A central drill hole formed a pressure relief drain for
groundwater expelled within the constricting ice front.

Reference

lining structure
When the freeze is switched off, the 30000t of steel and concrete composite tube lining the
shaft will retain its structural integrity independent of long term subsidence of the the
surrounding aquifer.

refrigeration plant
Refrigeration comes via nearly 400m³ of calcium chloride brine being delivered at -33ºC and
returned to the plant at -28ºC.

Monitoring
Three additional casings were set 120º apart and at 2m, 4m and 6m outside the freeze circle
for the computer-monitored thermometer installation assessing the progress of the freeze.
 

Case studies

Deep dewatering wells


Blackwater Valley

55
Installation of deep dewatering wells by hole punching is proving a cost-effective and
efficient way of dewatering a borrow pit on the northern contract of the Blackwater Valley
Route project. A scheme was proposed to extract sand horizons from a pit in the underlying
Bracklesham Beds, creating a home for surplus material.

The groundwater lies only a metre below ground level, and the Beds are hydrostatically
pressurised, so excavation without reducing the pressure would lead to rupturing. Following a
series of drawdown and falling head tests, a two stage scheme was proposed. The
conventional first stage produced a drawdown sufficient for excavation of the terrace gravels,
reducing ground level by 3m. Working from the reduced level at the top of the Bracklesham
Beds, the deep wells were installed. Pumping of these wells produced a drawdown sufficient
for the pit to be stepped down a further 5m. This was followed by a second series of deep
wells, finally allowing excavation to the required 15m.

Reference

Blackwater Valley Route


Surrey and Hampshire county councils' joint proposed dual carriageway bypass of
Farnborough and Aldershot.

borrow pit
The borrow pit is needed because the 5km section requires some 600000m³ of class one fill
for construction of the embankments. Importing this fill and carting away 150000m³ of
unsuitable surplus material would be less efficient.

hole punching
The deep wells were installed with a hole puncher. High pressure water and air are injected
through a central jetting pipe. This erodes the ground at the probe's base and flushes debris up
the annulus between the jetting pipe and the outer casing. At the same time, a top-acting 3t
drop weight drives the outer casing into the ground.

The pit needed to be 15m deep. Trials resulted in the sand boiling when the excavation
extended below 6m. The situation was complicated by the permeability in the overlying
gravel being ~0.001m/s, approximately 100 times greater than that in the Bracklesham Beds.

First stage
Well points were installed for pumping at the base of the gravels, around the perimeter of the
proposed pit. A garland drain was constructed around the base of the excavation to stop
overbleed from the gravels recharging the top of the Brackleshams.
56
The 300mm diameter deep wells were installed to a depth of 24m at 30m centres.
 
 

Case studies

Complex water control system


Munich airport, West Germany

Elaborate groundwater control measures are a key feature of Munich's new airport - one of the
largest greenfield construction sites in Europe. Groundwater within 1m of the surface meant a
risk of regular flooding and vulnerability to frost damage of the runways and taxiways. A
system of ditches and drains has dropped the level, but environmental constraints dictated the
water table should be restored outside the airport boundary. Restoration of diverted streams,
and a multiple level system of dirty, contaminated and clean run off drainage is required. The
solution inside the airport was to lower the groundwater table by around 2m. Two
longitudinal ditches between the runways are drained by an underground pipe link to the main
boundary interceptor. On the northern downstream boundary, the water is put back into the
ground by a series of well point injectors. Control centre computers monitor the flow
distribution and any contamination with spilt aircraft fuel etc. Glycol(used as a de-icer on the
taxiways will be intercepted by 20m wide ridged aprons of impermeable geotextile buried
under sand and gravel filters.

Reference

Geology
The subsoil is a 10m depth of Quaternary gravel overlying an impervious layer capping a 30m
thick stratum of Tertiary gravels. The gravels each have separate groundwater movements of
about 1m to 2m a day from south to north. They are valuable aquifers, and all flows need to
be maintained.

The injectors are 10m deep steel diffuser pipes 200mm in diameter
<PBiological action should break down the glycol before it reaches the edge of the membrane
and trickles down.
57
 

Case studies

Buoyancy-dependent foundations
Lecco bypass, near Milan, Italy

The Italian road authority ANAS' bypass includes twin bore 3.2km long tunnels passing
underneath Monte Barro and approach viaducts of 200m and 950m at either end. At the
longer of the two, the underlying limestone bedrock dips in a 100m deep glacial valley, now
infilled with soft and loose lake deposits. A brave design concept was proposed, in which
hydrostatic upthrust supports the viaduct's load.

While floating roads built on lightweight polystyrene blocks have been constructed over
marshland, at Lecco the viaduct is to be built on 20 huge, jet grouted cylindrical cells, bedded
25m below the ground surface. Soil within each cell is excavated from the surface using a
backhoe. The cells are lined with concrete and PVC, and a 1m thick cap completes the hollow
space.

The foundation design is dependent almost entirely upon buoyancy, provided by the cells
being water tight and remaining air filled. Yet water has leaked into some of the completed
cells and permanent pumping may be necessary to keep some of the cells dry.

Reference

jet grouted cylindrical cells


The cells are formed by vertically drilling a 10m diameter ring of 31 overlapping 1.5m
diameter jet grouted columns. Each column is reinforced by a steel pipe inserted into the soil-
grout mix before it sets. A second outer ring of jet grouting is then constructed over the
bottom 13m of the cell while a third phase of jet grouting forms a consolidated 5m deep plug
between 20m and 25m depth.

A 300mm thick reinforced concrete wall is cast in place after every 2m of excavation until the
top of the plug section has been reached. The base is then strengthened with a 2.5m thick
reinforced concrete slab and a PVC geomembrane is installed to waterproof the sides of the
cell. Then a second internal concrete wall is cast.
 
 

Case studies

Wellpoint pumping
58
Sevenoaks, Kent

In preparation for the construction of the piling mattress for a new superstore, subcontractors
commenced dewatering to lower the water table within the fine silty sands of the Lower
Greensand. The Gault clay dips down towards the northern end of the site, making dewatering
unnecessary in this area, so the wellpoints were installed along the bank of the stream at the
east boundary, and around the southern boundary. Groundwater in that area is almost at
surface level.

120 wellpoints were installed, the aim being to dewater the Greensand down to 4m, allowing
excavation for the construction of the pile mat. The wellpoints were connected to a pair of six-
inch pumps which discharge directly into the stream. Initially, the pumps were discharging at
a rate of 40 litre/sec, but after four days this had dropped to 10-15litre/sec. The Gault clay
contained a number of perched water tables, and these were sump pumped during the first
week of dewatering in order to drain them.

Reference

GEOLOGY
The geology consists of a layer of made ground overlying alluvium with some peat deriving
from the time when the area was marshland. This overlies Gault clay and an extremely fine
sand which represents the top of the Lower Greensand.

Site investigation
Site investigation consisted of eight boreholes and 16 trial pits, two of which were prevented
from going to full depth because of the high groundwater inflows.
 
 

Case studies

Artificial island
Trans Tokyo Bay Highway, Japan

59
Record breaking diaphragm walls has been trenched 135m under Tokyo bay at the artificial
island of Kawasaki - a key feature of the £5000M Trans Tokyo Bay Highway.(Note 1) The island
will serve a dual role as the main pit for tunnel driving during construction, and later as a
ventilator opening into the highway tunnel.

Loose marine sands have been stabilised by compaction from a barge before construction of
the working platforms. Note 2 The support framework functions as part of a retaining wall when
fill is placed to form a mounded island to sea level. The wall is supported and protected from
erosion by rockfill armour on the outside.

A rig worked from the platform to trench out the 98m diameter, 135m deep ring of diaphragm
wall panels, slicing down through the placed fill, then marine material, to form a cut off in
hard rock. Once the wall was watertight, the interior was excavated to form a pitNote 3 reaching
70m below sea level. Tunnel drives then started from the dry chamber.

Reference

(Note 1)
The Yokohama Bay bridge carries the new road 860m across the water from Kisarazu. The
rest of the distance to a second artificial island is spanned by a 4.4km low level viaduct. From
there the road plunges beneath the sea carried through twin 9.1km tunnels to the mainland.

(Note 2)
A 200m diameter ring of steel trestle working platforms, founded on tubular steel piles driven
into the treated material.

(Note 3)
Over 0.5M.m³ of fill and soft sediment will have to be dug out while in situ reinforced
concrete supports are erected inside.

Case studies

Flooded tunnel and ground freeze


Thames Water ring main, London

More than twice the depth of London's deepest tube station, the tunnel was being driven for
the Streatham to Brixton section of Thames Water's ring main. Groundwater pressures
approaching 400kN/m² meant that compressed air fed into the workings to stem the flow
60
broke through the legal limit of 345kN/m² for man working. The 2.5m diameter drive was
started in dry London Clay before breaking into the underlying Woolwich & Reading beds.
Saturated sand beds up to 0.5m thick needed more than the 100kN/m² air pressures predicted
to keep the tunnel dry. Even raising the air to over 200kN/m² did not prevent water ingress.

The miners were probing to see how much Woolwich & Reading bed material protected them
from the tricky Thanet Sands below, when the tunnel blew, letting 400kN/m² pressurised
groundwater blast in, bringing with it more than 10m³ of sand and exposing the Chalk below.
The volumes and rate of tunnel inflow, coupled with the fact that there is no sign of
depressurising the Thanet Sands, indicates that the water is coming from the Chalk.

Fortunately, the tunneller was stranded below open ground, so a 50m vertical shaft can be
sunk for machine recovery. The base of the 7m diameter shaft will be frozen to allow a short
horizontal drive to the machine.

Reference
 
 

Case studies

Rising groundwater levels


 Future water levels
 Effects on buildings and tunnels

The deep aquifer beneath London

Many major cities obtain water by pumping from the ground. Changed industrial practices
and water-supply systems over the years have led to much less water being taken from these
sources. As a result, water levels which had been drawn down are now rising.

During the past two centuries, the pumping from the deep aquifer lowered the groundwater
level by as much as 70m. The level is now rising, in many areas by about 1m/yr. If the rise
continues for 20 to 30 years, the water pressures in the sands and clays above the Chalk will
increase, causing ground movements in the clays. This could damage some large buildings
and tunnels and increase leakage into them. These problems can be prevented by additional
pumping, for a capital expenditure which is small compared with the potential cost of
damage.

Reference

61
Deep aquifer
The water-bearing Basal Sands and Chalk below London are referred to as the deep aquifer.
The aquifer is confined over much of the Basin by the overlying thick, relatively impermeable
layers of clay, which separate the deep aquifer from the perched groundwater in the overlying
gravels.

Future water levels


A regional groundwater computer model has been used to estimate the rates of rise likely to
occur over the next 90 years, together with the likely maximum future groundwater levels. It
shows that water levels could return almost to their original values within the next 30 to
40 years, unless action is taken.

 Additional pumping
Additional pumping of water from the aquifer in Central London not exceeding 30 mega-
litres per day would be sufficient. This is about 5% of the normal daily consumption of water
in Central London.

Capital Expenditure
The pumping would require about 30 wells at a capital cost of £10M to £30M.

Potential cost
The cumulative cost of remedial works for tunnels could exceed £10M. Repairs to large
buildings could cost tens of millions of pounds. Costs of new construction would also
increase.

Effects on buildings and tunnels

The potential effects of rising groundwater levels on buildings are:


 damage, perhaps instability caused by differential ground movements
 seepage into basements
 chemical attack on buried steel and concrete

Settlement or heave of a deep body of clay underlying a building would not necessarily
damage it. Damage is more likely where the clay is thin and the foundations or basements are
close to the aquifer. Broadly, this occurs in the lower-lying areas of Central London, east of
Westminster.

Although older buildings are unlikely to be damaged, modern structural forms, such as those
in the diagram, would be vulnerable. If water levels rise causing the clay to swell, the two
parts of the building with differing foundation levels will heave by amounts which could
differ by up to 100mm.

There are some 130km of tunnels under London which are located near the base of the
London Clay or in the sandy deposits. Lengths totalling about
40km have been identified which could suffer increased seepage,
loading and chemical attack.

Case studies

Simulation of rising groundwater

62
Canary Warf, London

Concern over rising groundwater at Canary Warf has led to the testing of a variable water
pressure trial pipe: The piezometric head is being artificially increased in the Thanet sands
pile founding strata by injecting water into five 150mm diameter wells. Bourdon gauges
connected to three piezometers monitor changes in head during testing. The experiment
models the change in pile base effective stress which is predicted if London ground water
levels continue to rise, plus the effect of unloading under the 9m deep excavation for
Founders Court car park. Trial excavation has been made redundant.

The total change in vertical effective stress at the pile base is equal to the sum of changes in
total vertical stress and water pressure. The reduction in vertical effective stress is calculated
to be about 180kPa, roughly equivalent to the 17m increase in water head induced by
injection into Thanet Sands. This compares with a current vertical effective stress of 300kPa.
The potential effect could be very serious if the factor of safety and design were inadequate.

Reference
 
 

Case studies

Flooded cofferdam
Ennerdale Link road, River Hull

The 2.5km dual carriageway road has been designed to relieve traffic congestion in
eastern Hull and provide a direct link to the docks. Its planned route beneath the River
Hull lies through an overall 361m long retained cutting, including a central 81m twin-
celled concrete box section constructed in cut-and-cover through the river bed.

63
Original site investigation suggested a 4.3m deep layer of alluvium was underlain by a
minimum 4m thick bank of impermeable boulder clay. This protected and confined a
large aquifer in the chalk beneath it. The aim was to found the tunnel box in the clay
zone without affecting the chalk. To construct the tunnel in two halves across the
river, the contractor drove the first of two sheetpiled cofferdams on the eastern bank.
With the cofferdam dewatered and being bottomed up, a 2m diameter sinkhole
suddenly appeared in the river bed close to the cofferdam's north west corner, and a
second hole formed nearby inside the structure. The excavation totally flooded
overnight.

The cofferdam was quickly plugged with tremmied concrete and following a detailed
three month second site investigation, urgent work started on jet grouting down the
three exposed faces of the cofferdam, while the possible solutions were debated.

Reference

sheetpiled cofferdams
The 140m long, 9m deep cofferdam projected 12m into the diverted river. The 22m
long piles extended down through the boulder clay to found well into the chalk
beneath.

The second investigation


The second investigation revealed the boulder clay to be thinner than expected. It may
contain perched water lenses and a prime concern is its integrity to protect the aquifer
beneath. Water in the cofferdam has come from the river, not the aquifer, and it is
possible that it has leaked through the sheetpiling rather than migrated beneath it.

Possibel solutions
Protecting the 21m wide tunnel from uplift hydrostatic pressure from the aquifer
would need extensive mass concrete loading to the 1m thick tunnel floor. Ground
anchors to hold the tunnel down are an option, but could puncture the aquifer.

The tunnel was eventually abandoned in favour of a lift bridge. In excess of £10m has
been spent on the aborted work prior to starting on the bridge.

64
Based on part of the GeotechniCAL reference package
Back to Soil Mechanics
by Prof. John Atkinson, City University, London

Basic mechanics of soils


 Analysis of stress and strain
 Strength
 Stiffness
 Material behaviour

Loads from foundations and walls apply stresses in the ground. Settlements are caused by
strains in the ground. To analyse the conditions within a material under loading, we must
consider the stress-strain behaviour. The relationship between a strain and stress is termed
stiffness. The maximum value of stress that may be sustained is termed strength.

Back to Basic mechanics of soils


Analysis of stress and strain
 Special stress and strain states
 Mohr circle construction
 Parameters for stress and strain

Stresses and strains occur in all directions and to do settlement and stability analyses it is
often necessary to relate the stresses in a particular direction to those in other directions.

normal stress
normal strain
 = z / zo
 = Fn / A
shear strain
shear stress
 = h / zo
 = Fs / A

Note that compressive stresses and strains are positive, counter-clockwise shear stress and
strain are positive, and that these are total stresses (see effective stress).

Analysis of stress and strain


Special stress and strain
states
In general, the stresses and strains in the three
dimensions will all be different.

There are three special cases which are


important in ground engineering:
General case princpal stresses

65
Axially symmetric or triaxial states
Stresses and strains in two dorections are equal.

'x = 'y and x = y


Relevant to conditions near relatively small
foundations, piles, anchors and other
concentrated loads.

Plane strain:
Strain in one direction = 0
y = 0
Relevant to conditions near long foundations,
embankments, retaining walls and other long
structures.

One-dimensional compression:
Strain in two directions = 0
x = y = 0

Relevant to conditions below wide foundations


or relatively thin compressible soil layers.

Uniaxial compression
'x = 'y = 0
This is an artifical case which is only possible
for soil is there are negative pore water
pressures.

Back to Analysis of stress and Forward to Parameters


Mohr circle strain

construction
Values of normal stress and shear stress must relate
to a particular plane within an element of soil. In
general, the stresses on another plane will be
different.

To visualise the stresses on all the possible planes, a


graph called the Mohr circle is drawn by plotting a
(normal stress, shear stress) point for a plane at every
possible angle.

66
There are special planes on which the shear
stress is zero (i.e. the circle crosses the normal
stress axis), and the state of stress (i.e. the
circle) can be described by the normal stresses
acting on these planes; these are called the
principal stresses '1 and '3 .

Analysis of stress and strain


Parameters for stress and strain
In common soil tests, cylindrical samples are used in which the axial and radial stresses and
strains are principal stresses and strains. For analysis of test data, and to develop soil
mechanics theories, it is usual to combine these into mean (or normal) components which
influence volume changes, and deviator (or shearing) components which influence shape
changes.

stress strain

p' = ('a + 2'r) / 3 ev = V/V = (a + 2r)


mean
s' =  'a + 'r) / 2 n = (a + r)

q' = ('a - 'r) es = 2 (a - r) / 3


deviator
t' =  ('a - 'r) / 2  = (a - r)

In the Mohr circle construction t' is the radius of the circle and s' defines its centre.

Note: Total and effective stresses are related to pore pressure u:


p' = p - u
s' = s - u
q' = q
t' = t

67
Back to Basic mechanics of soils
Strength
 Types of failure
 Strength criteria
 Typical values of shear strength

The shear strength of a material is most simply described as the maximum shear stress it can
sustain: When the shear stress  is increased, the shear strain  increases; there will be a
limiting condition at which the shear strain becomes very large and the material fails; the
shear stress f is then the shear strength of the material. The simple type of failure shown here
is associated with ductile or plastic materials. If the material is brittle (like a piece of chalk),
the failure may be sudden and catastrophic with loss of strength after failure.

Types of failure
Back to Strength

Materials can ‘fail’ under different loading conditions. In each case, however, failure is
associated with the limiting radius of the Mohr circle, i.e. the maximum shear stress. The
following common examples are shown in terms of total stresses:

Shearing
Shear strength = f
nf = normal stress at failure

68
Uniaxial extension
Tensile strength tf = 2f

Uniaxial compression
Compressive strength cf = 2f

Note:
Water has no strength f = 0.
Hence vertical and horizontal stresses are equal and the Mohr circle becomes a point.

Back to Strength
Strength criteria

69
A strength criterion is a formula which relates the strength of a material to some other
parameters: these are material parameters and may include other stresses.

For soils there are three important strength criteria: the correct criterion depends on the nature
of the soil and on whether the loading is drained or undrained.

In General, course grained soils will "drain" very quickly (in engineering terms)
following loading. Thefore development of excess pore pressure will not occur; volume
change associated with increments of effective stress will control the behaviour and the
Mohr-Coulomb criteria will be valid.

Fine grained saturated soils will respond to loading initially by generating excess pore
water pressures and remaining at constant volume. At this stage the Tresca criteria,
which uses total stress to represent undrained behaviour, should be used. This is the
short term or immediate loading response. Once the pore pressure has dissapated, after
a certain time, the effective stresses have incresed and the Mohr-Coulomb criterion will
describe the strength mobilised. This is the long term loading response.

 Tresca criterion
 Mohr-Coulomb (c’=0) criterion
 Mohr-Coulomb (c’>0) criterion

Back to Strength criteria Forward to Mohr-Coulomb (c’=0)


Tresca criterion
The strength is independent of the normal stress  since the response to loading simple
increases the pore water pressure and not the effective stress.

The shear strength f is a material parameter which is known as the undrained shear strength
su.

f = (a - r) = constant

Back to Strength criteria Forward to Mohr-


Mohr-Coulomb (c'=0) Coulomb (c’>0)

70
criterion
The strength increases linearly with increasing normal stress and is zero when the normal
stress is zero.
'f = 'n tan'
' is the angle of friction

In the Mohr-Coulomb criterion the material parameter is the angle of friction  and materials
which meet this criterion are known as frictional. In soils, the Mohr-Coulomb criterion
applies when the normal stress is an effective normal stress.

Back to Strength criteria


>Mohr-Coulomb (c'>0) criterion
The strength increases linearly with increasing normal stress and is positive when the normal
stress is zero.
'f = c' + 'n tan'
' is the angle of friction
c' is the 'cohesion' intercept

In soils, the Mohr-Coulomb criterion applies when the normal stress is an effective normal
stress. In soils, the cohesion in the effective stress Mohr-Coulomb criterion is not the same as
the cohesion (or undrained strength su) in the Tresca criterion.

Back to Strength
Typical values of shear strength
Undrained shear strength su (kPa)
Hard soil su > 150 kPa
Stiff soil su = 75 ~ 150 kPa
Firm soil su = 40 ~ 75 kPa
Soft soil su = 20 ~ 40kPa
Very soft soil su < 20 kPa

71
Drained shear strength c´ (kPa) ´ (deg)
Compact sands 0 35° - 45°
Loose sands 0 30° - 35°
Unweathered overconsolidated clay
critical state 0 18° ~ 25°
peak state 10 ~ 25 kPa 20° ~ 28°
residual 0 ~ 5 kPa 8° ~ 15°

Often the value of c' deduced from laboratory test results (in the shear testing apperatus) may
appear to indicate some shar strength at ' = 0. i.e. the particles 'cohereing' together or are
'cemented' in some way. Often this is due to fitting a c', ' line to the experimental data and an
'apparent' cohesion may be deduced due to suction or dilatancy.

Stress in the ground


 Total stress
 Pore pressure
 Effective stress
 Calculating vertical stress in the ground

When a load is applied to soil, it is carried by the water in the pores as well as the solid grains.
The increase in pressure within the porewater causes drainage (flow out of the soil), and the
load is transferred to the solid grains. The rate of drainage depends on the permeabilityof the
soil. The strength and compressibility of the soil depend on the stresses within the solid
granular fabric. These are called effective stresses.

Top
Total stress
 In a homogeneous soil mass
 In a soil mass below a river or lake
 In a multi-layered soil mass
 In a soil mass which is unsaturated
 In a soil mass with a surface surcharge load

72
The total vertical stress acting at a point below the ground surface is due to the weight of
everythinglying above: soil, water, and surface loading. Total stresses are calculated from the
unit weight of the soil.

Unit weight ranges are:

dry soil d 14 - 20 kN/m³ (average 17kN/m³)


saturated soil g 18 - 23 kN/m³  (average 20kN/m³)
water w  9.81 kN/m³   ( 10 kN/m³)

See Description and classification

Any change in vertical total stress (v) may also result in a change in the horizontal total
stress (h) at the same point. The relationships between vertical and horizontal stress are
complex.

total stress
Total stress in homogeneous soil
Total stress increases with depth and with unit weight: Vertical total stress at depth z,
v = .z
Simple total stress calculator 
 z v
20 3 60

The symbol for total stress may also be written z,


i.e. related to depth z.
The unit weight, , will vary with the water content of the soil.

d 

g

total stress
Total stress below a river or lake
The total stress is the sum of the weight of the soil up to the surface and the weight of water
above this: Vertical total stress at depth z,
v =  .z + w .zw
where

73
 = unit weight of the saturated soil,
i.e. the total weight of soil grains and water
w = unit weight of water
The vertical total stress will change with changes in water level and with excavation. Note
that free water (i.e. water outside the soil) applies a total stress to a soil surface.

Simple total stress calculator


 z zw v  
20 3 1 69.81

total stress
Total stress in multi-layered soil
The total stress at depth z is the sum of the weights of soil in each layer thickness above.
Vertical total stress at depth z,
v = 1d1 + 2d2 + 3(z - d1 - d2)
where
1, 2, 3, etc. = unit weights of soil layers 1, 2 , 3, etc. respectively
If a new layer is placed on the surface the total stresses at all points below will increase.
 Layer 1 2 3
Thickness 1.5 2 5

Unit weight 16 19 20

0 0
stress @ m =  kPa
Enter a value in any box (except the last) then click outside the box to see the effect

total stress
Total stress in unsaturated soil
Just above the
water table the soil will remain saturated due to
capillarity, but at some distance above the water table the soil will become unsaturated, with a
consequent reduction in unit weight (unsaturated unit weight = u)

v = w . zw + g(z - zw)

The height above the water table up to which the soil will remain saturated depends on the
grain size.
See Negative pore pressure (suction).
 
74
total stress
Total stress with a surface surcharge load
The addition of a surface surcharge load will increase the total stresses below it. If the
surcharge loading is extensively wide, the increase in vertical total stress below it may be
considered constant with depth and equal to the magnitude of the surcharge.
Vertical total stress at depth z,
v =  .z + q
For narrow surcharges, e.g. under strip and pad foundations, the induced vertical total stresses
will decrease both with depth and horizontal distance from the load. In such cases, it is
necessary to use a suitable stress distribution theory - an example is Boussinesq's theory.

Top
Pore pressure
 Groundwater and hydrostatic pressure
 Water table, phreatic surface
 Negative pore pressure (suction)
 Pore water and pore air pressure

The water in the pores of a soil is called porewater. The pressure within this porewater is
called pore pressure (u). The magnitude of pore pressure depends on:

 the depth below the water table


 the conditions of seepage flow

Pore pressure
Groundwater and hydrostatic pressure
Under hydrostatic conditions (no water flow) the
pore pressure at a given point is given by the hydrostatic
pressure:

u = w .hw
where
hw = depth below water table or overlying water surface
It is convenient to think of pore pressure represented by the column of water in an imaginary
standpipe; the pressure just outside being equal to that inside.

75
 
 

Pore pressure
Water table, phreatic surface
The natural static level of water in the ground is called the water table or the phreatic
surface (or sometimes the groundwater level). Under conditions of no seepage flow, the
water table will be horizontal, as in the surface of a lake. The magnitude of the pore pressure
at the water table is zero. Below the water table, pore pressures are positive.
u = w .hw
In conditions of steady-state or variable seepage flow, the calculation of pore pressures
becomes more complex.
See Groundwater

Pore pressure
Negative pore pressure (suction)
Below the water table, pore pressures are positive. In dry soil, the pore pressure is zero.
Above the water table, when the soil is saturated, pore pressure will be negative.
u = - w .hw

The height above the water table to which the soil is saturated is called the capillary rise, and
this depends on the grain size and type (and thus the size of pores):
· in coarse soils capillary rise is very small
· in silts it may be up to 2m
· in clays it can be over 20m

Pore pressure
Pore water and pore air pressure
Between the ground surface and the top of the saturated zone, the soil will often be partially
saturated, i.e. the pores contain a mixture of water and air. The pore pressure in a partially
saturated soil consists of two components:
· porewater pressure = uw
· pore-air pressure = ua
76
Note that water is incompressible, but air is compressible. The combined effect is a complex
relationship involving partial pressures and the degree of saturation of the soil. In Europe and
other temperate climate countries most design states are associated with saturated conditions,
and the study of partially saturated soils is considered to be a specialist subject.
 
 
Pore pressure
Pore pressure in steady state seepage conditions

In conditions of seepage in the ground there is a change in pore pressure. Consider seepage
occurring between two points P and Q.

The hydralic gradient, i, between two points is the head drop per unit length between these
points. It can be thougth of as the "potential" driving the water flow.
 

h .
u  1 
Hydralic gradient P-Q, i = -   = 
s w
s

Thus u = i . w . s

But in steady-state seepage, i = constant


Therefore the change in pore pressure due to seepage alone,  us  =  i . w . s
For seepage flow vertically downward, i is negative
For seepage flow vertically upward, i is positive.
 

Top
Effective stress
 Terzaghi's principle and equation
 Mohr circles for total and effective stress
 Importance of effective stress
 Changes in effective stress

Ground movements and instabilities can be caused by changes in total stress (such as loading
due to foundations or unloading due to excavations), but they can also be caused by changes
in pore pressures (slopes can fail after rainfall increases the pore pressures).

In fact, it is the combined effect of total stress and pore pressure that controls soil behaviour
such as shear strength, compression and distortion. The difference between the total stress and
the pore pressure is called the effective stress:

77
effective stress = total stress - pore pressure

or ´ =  - u

Note that the prime (dash mark ´ ) indicates effective stress.


 
 

Effective stress
Terzaghi's principle and equation
Karl Terzaghi was born in Vienna and subsequently became a professor of soil mechanics in
the USA. He was the first person to propose the relationship for effective stress (in 1936):

All measurable effects of a change of stress, such as compression, distortion and a


change of shearing resistance are due exclusively to changes in effective stress. The
effective stress ´ is related to total stress and pore pressure by ´ =  - u.

The adjective 'effective' is particularly apt, because it is effective stress that is effective in
causing important changes: changes in strength, changes in volume, changes in shape. It does
not represent the exact contact stress between particles but the distribution of load carried by
the soil over the area considered.
 

Effective stress
Mohr circles for total and effective stress

Mohr circles can be drawn for both total and effective stress. The points E and T represent the
total and effective stresses on the same plane. The two circles are displaced along the normal
stress axis by the amount of pore pressure (n = n' + u), and their diameters are the same.

78
The total and effective shear stresses are equal (´ = ).
 

Effective stress
The importance of effective stress
The principle of effective stress is fundamentally important in soil mechanics. It must be
treated as the basic axiom, since soil behaviour is governed by it. Total and effective stresses
must be distinguishable in all calculations: algebraically the prime should indicate effective
stress, e.g. ´

Changes in water level below ground (water table changes) result in changes in effective
stresses below the water table. Changes in water level above ground (e.g. in lakes, rivers, etc.)
do not cause changes in effective stresses in the ground below.

Effective stress
Changes in effective stress
 Changes in strength
 Changes in volume

In some analyses it is better to work in changes of quantity, rather than in absolute quantities;
the effective stress expression then becomes:
´ =  - u

If both total stress and pore pressure change by the same amount, the effective stress remains
constant. A change in effective stress will cause: a change in strength and a change in volume.

Changes in effective stress


Changes in strength
The critical shear strength of soil is proportional to the effective normal stress; thus, a change
in effective stress brings about a change in strength.

79
Therefore, if the pore pressure in a soil slope increases, effective stresses will be reduced by
' and the critical strength of the soil will be reduced by  - sometimes leading to failure.

A seaside sandcastle will remain intact while damp, because the pore pressure is negative; as
it dries, this pore pressure suction is lost and it collapses. Note: Sometimes a sandcastle will
remain intact even when nearly dry because salt deposited as seawater evaporates slightly and
cements the grains together.

Changes in effective stress


Changes in volume
Immediately after the construction of a foundation on a fine soil, the pore pressure increases,
but immediately begins to drop as drainage occurs.

The rate of change of effective stress under a loaded foundation, once it is constructed, will be
the same as the rate of change of pore pressure, and this is controlled by the permeability of
the soil.

Settlement occurs as the volume (and therefore thickness) of the soil layers change. Thus,
settlement occurs rapidly in coarse soils with high permeabilities and slowly in fine soils with
low permeabilities.

Top
Calculating vertical stress in the ground
 Simple total and effective stresses
 Effect of changing water table
 Stresses under foundations
 Short-term and long-term stresses
 Steady-state seepage conditions

The worked examples here are designed to illustrate the principles and methods dealt with in
Pore pressure, effective stress and stresses in the ground. The examples chosen are typical
and simple.

Calculating vertical stress


Simple total and effective stresses
The figure shows soil layers on a site.
Unit weights are:
dry sand: d = 16 kN/m³
saturated sand: g = 20 kN/m³

(a) At the top of saturated sand (z = 2.0 m)


Vertical total stress v = 16.0 x 2.0 = 32.0 kPa
Pore pressure u=0
Vertical effective stress ´v = v - u = 32.0 kPa
 

80
(b) At the top of the clay (z = 5.0 m)
Vertical total stress v = 32.0 + 20.0 x 3.0 = 92.0 kPa
Pore pressure u = 9.81 x 3.0 = 29.4 kPa
Vertical effective stress ´v = v - u  = 92.0 - 29.4 = 62.6 kPa

Calculation of vertical stress


Effect of changing water table
The figure shows soil layers on a site. The unit weight of the silty sand is 19.0 kN/m³ both
above and below the water table. The water level is presently at the surface of the silty sand, it
may drop or it may rise. The following calculations show the effects of this:

Water table

Calculation of vertical stress


Stresses under foundations

From an initial state, the stresses under a foundation are first changed by excavation, i.e.
vertical stresses are reduced. After construction the foundation loading increases stresses.
Other changes could result if the water table level changed.

The figure shows the elevation of a foundation to be constructed in a homogeneous soil. The
change in thickness of the clay layer is to be calculated and so the initial and final effective
stresses are required at the mid-depth of the clay.

Unit weights: sand above WT = 16 kN/m³, sand below WT = 20 kN/m³, clay = 18 kN/m³.

Calculations for
initial stresses
final stresses
Calculation of vertical stress

81
Short-term and long-term stresses
 Initially, before construction
 Immediately after construction
 Many years after construction

The figure shows how an extensive layer of fill will be placed on a certain site.

The unit weights are:


clay and sand = 20kN/m³ ,
rolled fill 18kN/m³ ,
assume water = 10 kN/m³.
Calculations are made for the total and effective stress at the mid-depth of the sand and the
mid-depth of the clay for the following conditions: initially, before construction; immediately
after construction; many years after construction.
 
 
Short-term and long-term stresses

Initially, before construction

Initial stresses at mid-depth of clay (z = 2.0m)


Vertical total stress
v = 20.0 x 2.0 = 40.0kPa
Pore pressure
u = 10 x 2.0 = 20.0kPa
Vertical effective stress
´v = v - u = 20.0kPa

Initial stresses at mid-depth of sand (z = 5.0 m)


Vertical total stress
v = 20.0 x 5.0 = 100.0 kPa
Pore pressure
u = 10 x 5.0 = 50.0 kPa

82
Vertical effective stress
´v = v - u = 50.0 kPa

Short-term and long-term stresses

Immediately after construction

The construction of the embankment applies a surface surcharge:


q = 18 x 4 = 72.0 kPa.

The sand is drained (either horizontally or into the rock below) and so there is no increase in
pore pressure. The clay is undrained and the pore pressure increases by 72.0 kPa.

Initial stresses at mid-depth of clay (z = 2.0m)


Vertical total stress
v = 20.0 x 2.0 + 72.0 = 112.0kPa
Pore pressure
u = 10 x 2.0 + 72.0 = 92.0 kPa
Vertical effective stress
´v = v - u = 20.0kPa
(i.e. no change immediately)

Initial stresses at mid-depth of sand (z = 5.0m)


Vertical total stress
v = 20.0 x 5.0 + 72.0 = 172.0kPa
Pore pressure
u = 10 x 5.0 = 50.0 kPa
Vertical effective stress
´v = v - u = 122.0kPa
(i.e. an immediate increase)
 
 

Short-term and long-term stresses


Many years after construction

83
After many years, the excess pore pressures in the clay will have dissipated. The pore
pressures will now be the same as they were initially.

Initial stresses at mid-depth of clay (z = 2.0 m)


Vertical total stress
v = 20.0 x 2.0 + 72.0 = 112.0 kPa
Pore pressure
u = 10 x 2.0 = 20.0 kPa
Vertical effective stress
´v = v - u = 92.0 kPa
(i.e. a long-term increase)

Initial stresses at mid-depth of sand (z = 5.0 m)


Vertical total stress
v = 20.0 x 5.0 + 72.0 = 172.0 kPa
Pore pressure
u = 10 x 5.0 = 50.0 kPa
Vertical effective stress
´v = v - u = 122.0 kPa
(i.e. no further change)

Calculation of vertical stress


Steady-state seepage conditions
The figure shows seepage occurring around embedded sheet piling.
In steady state, the hydraulic gradient,
i =  /  = 4 / ( 7 + 3 ) = 0.4
Then the effective stresses are:
´A = 20 x 3 - 2 x 10 + 0.4 x 10 = 44 kPa
´B = 20 x 3 - 2 x 10 - 0.4 x 10 = 36 kPa

Based on part of the GeotechniCAL reference package


Back to Compression and shear
by Prof. John Atkinson, City University, London
84
Shear strength
 Common cases of shearing
 Strength

Near any geotechnical construction (e.g. slopes, excavations, tunnels and foundations) there
will be both mean and normal stresses and shear stresses. The mean or normal stresses cause
volume change due to compression or consolidation.

The shear stresses prevent collapse and help to support the geotechnical structure. Shear stress
may cause volume change.

Failure will occur when the shear stress exceeds the limiting shear stress (strength).

Back to Shear strength


Common cases of shearing
In practice, the state of stress in the ground will be complex.

There are simple theories for two special cases.

Triaxial (axial symmetry)


Parameters used for analysis:
· deviator stress
· shear strain
· normal stress
· volumetric strain
· specific volume

85
Direct or simple shear
Parameters used for analysis:
· shear stress
· shear strain
· normal stress
· volumetric (normal) strain
· void ratio

It is not possible to draw a Mohr circle for a shear test unless stresses on vertical planes are
measured.

Strength Back to Shear strength

 Peak strength
 Critical state strength
 Residual strength

In very simple terms, the strength of soil is the maximum shear stress (f) it can sustain, or
the shear stress acting on a shear slip surface along which it is failing. There are three distinct
strengths: peak, critical (or ultimate) and residual. Shearing may be simple or direct.
Drained direct (ring) shear Drained simple shear

86
We explore the relationship between the maximum shear stress and the effective normal stress
(') by ploting a graph of f against '.

Some aspects of the behaviour show up more clearly if we normalise the data by plotting f /
' against ' / '

Back to Strength
Peak strength
 Peak strength in shear tests
 Peak strength in triaxial tests
 Peak strength and dilatancy

The peak strength is the maximum value of the shear stress or the maximum value of the ratio
of shear stress to effective mean or normal stress. For drained tests these will occur
simultaneously, for undrained tests they may occur at different points and the definition used
here is the maximum stress ratio.
 Peak strengths can only occur at shear stresses above the critical state line and at water
contents below the CSL.
 Peak states can occur anywhere in the regions above and below the CSL.
 Peak states at the same water content fall on unique smoth envelopes.
 The peak states can be represented on a graph in 3 dimensions.
 All peak states fall on a surface in this graph.

Back to Peak strength Forward to Equations

Peak strength in shear tests


The circle represent the results of a set of
shear tests on samples at the same moisture
content but different normal stresses. The
squares represent the results of a second set of
tests at a different moisture content.

We normalise the data by plotting


f / ' against ' / '

87
The basic peak states, before normalisation, fall on different curves each for a particular water
content or void ratio.

After normalisation all the peak states fall on a single unique envelope.

Equations
At a given water content or void ratio, all the peak states fall on a
single smooth envelope.
This may be represented in one of two ways:

As a power law
'p = B' ª

As a linear (Mohr-Coulomb) envelope


if the curvature is relatively small over a given range.
'p = c'p + ' tan 'p

The parameters a, B and c'p, 'p depend on the water content or void
ratio.

Even at a given water content or voids ratio, the parameters c'p and
'p depend on the range of stress for linear approximation.

88
Back to Peak strength

Peak strength in triaxial tests

The basic peak states, before normalisation, fall on different curves each for a particular water
content or specific volume.

After normalisation all the peak states fall on a single unique envelope.

equations
At a given water content or specific volume all the peak states fall on a single smooth
envelope. This may be represented in one of two ways:

As a power law

As a linear envelope
if the curvature is relatively small over a given range.
q'p = Gp + Hp p'

The parameters ,  and Gp, Hp depend on the water content or voids ratio.

Even at a given water content or voids ratio, the parameters Gp and Hp depend on the range of
stress for linear approximation.

Back to Peak strength

Peak strength and dilatancy


89
 Peak state and initial state

Stresses and displacements in a shear sample are analagous to the forces and movements of a
friction block on an inclined plane.

At critical state
 = 0 and ' = ' tan'c

The additional stress ratio (above the critical state) is due to the rate of dilation
tan  = dv/dh

Back to Peak strength and dilatancy

Peak state and initial state


The peak stress ratio depends on the initial state given by the initial overconsolidation ratio.
The maximum rate of dilation increases with overconsolidation ratio. For the same initial
overconsolidation ratio (i.e. A and A') the peak stress ratio is the same.

Back to Strength

Critical state strength


 Critical state strength in shear tests
 Critical state strength in triaxial tests
 Undrained strength

90
 Typical values of critical state strength parameters

At its critical state soil continues to distort at constant effective stress and at constant volume.
This applies for turbulent flow of the particles: if the flow becomes laminar, as in clays at
large strain, the strength falls to the residual.

When soil is at its critical state there is a unique relationship between shear stress, effective
normal stress and water content (or specific volume or void ratio). Critical states are unique
and do not depend on initial state or stress path.

Critical states correspond to shear strains typically 10% to 40%.

Critical shear stress (critical state strength) increases with increasing effective normal stress
and with decreasing water content.

The critical state line can be represented as a graph in 3 dimensions. For isotropic
compression, shear stresses are zero and the isotropic normal compression line can also be
represented.

Back to Critical state strength

Critical state strength in shear tests


The graphs show the critical state line. If you know either ' or e at the critical state you can
calculate the critical state strength '.
'c cc and eG are soil parameters.
The one-dimensional normal compression line (NCL) for zero shear stress is also shown.

91
The critical state line can be normalised with respect to the critical pressure 'c or the
equivalent void ratio el. The critical state line and the isotropic normal compression line both
reduce to single points.

Back to Critical state strength

Critical state strength in triaxial tests


The graphs show the critical state line. If you know either p' or v at the critical state you can
calculate the critical state deviation stress q'.
M  and  are soil parameters.
We should really use subscripts c and e for compression and extension as the values are
slightly different.
The isotropic normal compression line corresponds to zero deviator stress

92
The critical state line can be normalised with respect to the critical pressure 'c or the
equivalent specific
volume vl. The critical state line and the isotropic normal compression line both reduce to
single points.

Back to Critical state strength

Typical values of critical state strength parameters


The critical state parameters are basic soil parameters and they depend principally on the
nature of the soil grains. For fine grained soils the CS parameters are related to the Atterberg
limits; for coarse-grained soils they are related to the mineralogy and shape of the grains.
Typical values  G M '

high plasticity clay 0.16 2.45 0.89 23º

low plasticity clay 0.10 1.80 1.18 29º

quartz sand 0.16 3.00 1.28 32º

carbonate sand 0.34 4.35 1.65 40º

For fine-grained soils the gradient  or Cc of the critical state line is related to the Atterberg
limits by

93
Cc = (Ip x Gs) / 200
 = (Ip x Gs) / 460

For many soils the critical state lines all pass through a single point called the  (omega)
point.
v(W) = 1.25 p'(W) = 10MPa
e(W) = 0.25 '(W) = 15MPa.

Back to

Critical state strength

Undrained strength
The critical state strength is uniquely related to the water content.

If the soil is sheared without change of water content (i.e. undrained) its strength remains the
same. This is called the undrained strength su. But if the soil is not undrained and the water
content changes the strength will also change.

The undrained strength is directly related to the liquidity index IL. Some authors give slightly
different values for su but
su at PL (i.e. IL=0) is always 100 times
su at LL (i.e. IL=1)

94
Back to Strength

Residual strength
 Residual strength: equations

This is the very lowest strength which occurs after very large displacements. For sands the
residual strength is the same as the critical state strength. For clays the residual is about ½ the
critical state strength. For clays the flat clay particles become aligned parallel to the direction
of shear.

The residual strength occurs after very large (>1m) movements and is not usually relevant for
geotechnical engineering where generally ground movements must be small. However, on old
landslides there may have already been very large movements and in such cases the strength
may already be at the residual before construction starts.

Back to Residual strength

Residual strength: equations


Residual strength applies to clays after very large shear displacements when clay particles
have become aligned in well-defined shear zones or slip planes.
Drained case
 = ' tan 'r
'r = residual friction angle
In clays 'r can be less than ½'c.
For London Clay,
'c »22º and 'r»10º.
In mixed soils 'r depends on the quantity of clay present.
Undrained case
 = sur
sur = undrained residual strength
(depends on water content)

Produced by Dr. Leslie Davison, University of the West of England, Bristol, May 2000
in association with Prof. Sarah Springman, Swiss Federal Technical Institute, Zurich

95
Foundations
 Load-settlement behaviour
 Types of foundation
 Bearing capacity
 Settlement
 Foundation design
 Ground improvement

The foundation of a structure is in direct contact with the ground and transmits the load of the
structure to the ground. Foundations may be characterised as shallow (pad, strip or raft) or
deep (piles, piers or caissons). When designing foundations, two principal criteria must be
satisfied:

Bearing capacity
There must be an adequate factor of safety against collapse (plastic yielding in the soil and
catastrophic settlement or rotation of the structure).

Settlement
Settlements at working loads must not cause damage, nor adversely affect the serviceability of
the structure.

There are other considerations that may be relevant to specific soils, foundation types and
surface conditions.

Foundations

Load-settlement behaviour
 General shear failure
 Local shear failure
 Punching shear failure
 Factors affecting mode of failure

The application of a load on a foundation causes some settlement. The three main stages of
the load-settlement curve are:
Relatively elastic vertical compression

96
(O-A) The load-settlement curve is almost straight.

Local shear failure

(B) Local yielding causes some upward and outward movement of the soil and results in
slight surface heave.
General shear failure
(C) Large settlements are produced as plastic yielding is fully developed within the soil.

Bearing capacity failure can occur in three different modes: general shear failure, local shear
failure, or punching shear failure. Local or punching shear are characterised by relatively
large settlements and the ultimate bearing capacity is not clearly defined. In these cases
settlement is the major factor in the foundation design.

Load-settlement behaviour

General shear failure


When a load (Q) is gradually applied on a foundation, settlement occurs which is almost
elastic to begin with. At the ultimate load, general shear failure occurs when a plastic yield
surface develops under the footing, extending outward and upward to the ground surface, and
catastrophic settlement and/or rotation of the foundation occurs. The load per unit area at this
point is called the ultimate bearing capacity (qf) of the foundation.

Load-settlement behaviour
97
Local shear failure

In moderately compressible soils, and soils of medium relative density, significant vertical
settlement may take place due to local shear failure, i.e. yielding close to the lower edges of
the footing. The yield surfaces often do not reach the surface. Several yield developments
may occur accompanied by settlement in a series of jerks. The bearing pressure at which the
first yield takes place is referred to as the first-failure pressure (qf(1)) - the term first-failure
load (Qf(1)) is also used.

Load-settlement behaviour

Punching shear failure


In weak compressible soils, and soils of low relative density, considerable vertical settlement
may take place with the yield surfaces restricted to vertical planes immediately adjacent to the
sides of the foundation; the ground surface may be dragged down. After the first yield has
occured the load-settlement curve will steepen slightly, but remain fairly flat. This is referred
to as a punching shear failure.

98
Load-settlement behaviour

Factors affecting modes of failure


According to experimental results from foundations resting on sands (Vesic, 1973), the mode
of failure likely to occur in any situation depends on the size of the foundation and the relative
density of the soil.

Other factors might be:


permeability: relating to drained/undrained behaviour
compressibility: similar to RD
shape: e.g. strips can only rotate one way
interaction between adjacent foundations and other structures
relative stiffnesses of soil and footing/structure
incidence and relative magnitude of horizontal loadings or moments
presence of stiffer or weaker underlying layers.

Foundations

Settlement
 Total and differential settlement criteria
 Imposed stress distribution below foundations
 Settlements on clays and silty clays
 Settlements on sands and gravels

The are three components making up the final settlement quantity:

Immediate settlement (i)


· elastic deformation with no change in water content
· occurs rapidly during the application of load
· quite small quantity in dense sands/gravels and stiff/hard clays
Consolidation settlement (c)
· decrease in voids volume as porewater is squeezed out of the soil
· occurs slowly according to the permeability
· only significant in clays and silts
Secondary settlement or creep (a)
· due to gradual changes in the particulate structure of the soil
· occurs very slowly, long after consolidation is completed
· most significant in soft organic soils and peats

Thus, final settlement,  = i + rc + ra

99
Reliable predictions of settlement require a thorough assessment of ground conditions,
including measurements of soil properties. Extensive ground investigations and statistically-
reliable testing programmes can be expensive and time-consuming, and thus not economically
viable for the routine design of shallow foundations.

Settlement

Total and differential settlement criteria


 Definitions of quantities
 Cracking and ease of repair

The principal undesirable effects of settlement in buildings are cracking due to angular
distortion; tilting due to differential settlement and excessive downward displacement. The
settlement of foundations must be limited to satisfy three main sets of criteria:
Cracking or tilting that affects visual appearance
· cracking damage can be related to the ease of repair
· a vertical tilt (w) of > 1/250 is unpleasantly noticeable
· horizontal dveviation of >1/100 and deflection/span ratos are also noticeable
Cracking or other damage that adversely affects serviceability
· cracking of wall and floor components can affect serviceability and be costly to repair
· differential displacements may affect the functionality of lifts, conveyors, cranes, internal
traffic movement, aligment of drains, etc.
Damage that affects structural integrity or stability
· Angular distortion (D/L) in the foundation produces consequential distortion and increases in
stresses in the structure above
· Structural damage is unlikely to occur if /L is < 1/150
· Values of /L > 1/1000 can cause cracking and overstressing.

Total and differential settlement criteria

Definitions of quantities

The following definitions and quantities have been suggested by Burland and Wroth (1975):
Settlement (  or s): the downward displacement of a point in a foundation

100
Differential settlement () the displacment of one point with respect to another
e.g BA = displacement of B with respect to A = B - A
Angular strain ( ): the change in slope from horizontal at a point
Tilt angle (): the vertical angle of displacement of a unit of structure
Relative deflection (): maximum displacement between two points relative to a straight line
drawn between them
Angular distortion (/L): the ratio of the relative deflection between two points and the
horizontal distance between them
Sagging occurs when the relative deflection is downward, i.e.  is +ve
Hogging occurs when the relative deflection is upward, i.e.  is -ve

Total and differential settlement criteria

Cracking and ease of repair


Jennings and Kerrich (1962) proposed a classification of damage due to the cracking based on
the ease of repair.

degree of
type of damage crack width (mm)
damage
0 negligible Barely visible or none < 0.1
1 very slight Fine cracks repairable by normal decoration 1-5
2 slight Cracks easily filled, may be visible externally. >5
Doors & windows stick slightly
3 moderate Cracks need raking out & filled by a mason 5-15
Doors & windows sticking Services fractured; water (more than 3)
ingress
4 severe Breaking out & replacing of masonry; floors sloping; 15-25
walls leaning/bulging; doors & windows not (depends on
functioning; number)
floor beams displaced; services disrupted, serious water
ingress
5 very severe Rebuilding or major repair required; beams lose >25
bearing; (depends on
walls require shoring; number)
danger of instability or collapse.

101
Settlements on clay

Imposed stress distribution below foundations


 Boussinesq's solution
 Circular foundation
 Strip foundation
 Rectangular foundation
 Approximate methods

The distribution of stresses in the ground under a foundation due to applied loading is not
uniform. Changes in vertical stress decrease with both depth and horizontal distance from the
load; but can be predicted with reasonable accuracy using elasticity theory (and sometimes
simpler approximate methods).

Imposed stress distribution below foundations

Boussinesq's solution
According to Boussinesq (1885), the vertical stress v in the ground due to a point load Q is

where z is the depth below the load and r is the horizontal distance from the load.

Simple solutions are available for stresses below strip footings, below the centre of circular
footings and below a corner of a rectangular footing. The latter can be used to calculate the
stress at any point by dividing the rectangle into two or more rectangles and summing the
stresses due to each part. For example, to find the stress under the centre of a (B x L)
rectangular base, find the stress under the corner of a (1/2B x 1/2L) rectangle, then multiply
by 4.

Imposed stress distribution below foundations

Stress below a circular foundation


102
The vertical stress at a depth z below the centre of circular base of radius R is
v = q .Ic
where q is the bearing pressure and

The stress value below the centre is a maximum for a given depth. The expressions for
stresses off-centre are much more complex. Over deep soil layers, the average value will be
between 0.85 and 0.6 of the centre-line value according to the stiffness of the footing.

Imposed stress distribution below foundations

Stress below a strip foundation


The vertical stress at a depth z below a uniformly loaded strip footing of width B=2b is
v = q .Is
where q is the bearing pressure and
Is = [ + sin .cos(  + 2 ) ] / 

103
Imposed stress distribution below foundations

Stress below a rectangular foundation


The vertical stress at a depth z below the corner of a rectangular subject to uniform pressure is
v = q. IR
where q is the bearing pressure and

Imposed stress distribution below foundations

Approximate methods
For settlement calculations (but not necessarily for other geotechnical problems), sufficient
accuracy is usually obtained by assuming a simplified pressure distributuion. The two
methods given below are in common use.

Tomlinson's method
Tomlinson suggests using an
approximate pressure distribution
for small foundations on stiff clay:

Settlement,  = ( 1.5 B) x ( 0.55 qn) x mv

where
B is the breadth of the foundation
qn is the net bearing pressure

104
2 : 1 distribution method
The vertical stress on horizontal planes is assumed to remain uniform, but decrease linearly
with depth below the foundation, thus:

For a strip footing:

For a rectangular footing:


where: qn = applied net bearing pressure
B = breadth,
L = length
z = depth from underside of footing

Settlement

Settlements on clays and silty clays


 Calculation of immediate settlement
 Calculation of consolidation settlement

Settlements are time-dependent and may take several years to complete. The soil properties
are obtained principally from laboratory tests. Relative proportions of final settlement that
occur as immediate and consolidation settlement vary according to the stress history and state
of the soil.

Soft normally consolidated clays and silts


i » 0.1roed
where oed = consolidation settlement from oedometer mv
The final settlement (excluding creep),
 =  + oed = 1.1 roed

Stiff overconsolidated clays


Both immediate and consolidation settlement are incorporated in the oedometer-measured
modulus.
Thus, final settlement,  = oed
Proportions:
i » 0.5oed - 0.6oed
c » 0.5oed - 0.4oed

105
Settlements on clays and silty clays

Calculating immediate settlement


 Typical values of Poisson's ratio
 Typical values of undrained stiffness
 Typical values of influence factor

The net immediate settlement is assumed to be due entirely to elastic volume change upon
loading.

The following expression is based on elastic theory.

Immediate settlement
where qn = net bearing pressure
B = breadth of foundation
 = Poisson's ratio
Eu = undrained stiffness modulus

Ip = influence factor

Calculating immediate settlement

Typical values of Poisson's ratio

Soil Poisson's ratio ()

Undrained saturated clays/silts u = 0.5

Stiff sandy or silty clays ´ = 0.2 - 0.4

Medium to loose sands ´ = 0.4

Dense sands ´ = 0.2 - 0.45

106
Calculating immediate settlement

Typical values of undrained stiffness


It is general practice to obtain values for Eu from undrained laboratory tests. However, other
empirical relationships can also be used:

From oedometer test results


The drained stiffness modulus,
E´ = 1/mv
and then
Eu = 1.5E´/(1 + ´)
where ´ = drained value of Poisson's ratio.

From undrained shear strength


In soils, stiffness moduli are strain-dependent. The strains associated with normal foundations
at working loads are generally less than 0.1%. Some empirical relationships have been noted,
for example, for London clay,
Eu /su = 150 to 500

Calculating immediate settlement

Typical values of influence factor


Influence factors Ip for vertical displacement due to elastic compression of a layer of semi-
infinite thickness

Shape Flexible Rigid


  Centre Corner Average  
Circle 1.00 0.64 0.85 0.79
Rectangle    
L/B = 1.0 1.122 0.561 0.946 0.82
L/B = 1.5 1.358 0.679 1.148 1.06
L/B = 2.0 1.532 0.766 1.300 1.20
L/B = 3.0 1.783 0.892 1.527 1.42
L/B = 4.0 1.964 0.982 1.694 1.58
L/B = 5.0 2.105 1.052 1.826 1.70
L/B = 10.0 2.540 1.270 2.246 2.10
L/B = 100.0 4.010 2.005 3.693 3.47

107
Settlements on clays and silty clays

Calculating consolidation settlement


 Typical values of volume compressibility

The oedometer settlement of a soil layer can be calculated from:


oed = mv.´.HL
where
mv = coefficient of volume compressibilty measured in an oedometer test
´ = the average change in vertical effective stress in the layer
HL = thickness of the layer

In layered deposits, the sum of layer-settlements should be take down to a depth where
´ < 0.1qn

Oedometer test results tend to overestimate the actual settlement; the following correction
factors should be applied:
c = g.oed
where g = a factor suggested by Skempton and Bjerrum (1957)

Type of clay mg

Very sensitive; soft alluvial, estuarine and 1.0-1.2


marine clays

Normally consolidated clay 0.7-1.0

Overconsolidated clays, 0.5-0.7


e.g. London Clay, Weald, Kimmeridge,
Oxford and Lias clays.

Heavily overconsolidated clays, 0.2-0.5


e.g. Keuper Marl, glacial till

108
Calculating consolidation settlement

Typical values of volume compressibility


The values given below should be taken as guideline figures only, for preliminary design and
first estimates; values based on reliable tests should be used in final design calculations.

Type of soil Degree of mv


commpressibility (m²/MN)

Heavilly overconsolidated clays Very low


< 0.05
compressibilty

Overconsolidated clays Low compressibilty 0.05 - 0.1

Weathered overconsolidated Medium


0.1 - 0.3
clays compressibilty

Normally consolidated clays High compressibilty 0.3 - 1.5

Organic alluvial clays and silts; Very high


> 1.5
peats compressibilty

109
Settlement

Settlements on sands and gravels


Settlements on sands and gravels take place almost immediately. Obtaining undisturbed
samples for laboratory testing is difficult. The soil properties are estimated from in situ test
results.

Standard penetration test


Relationships have been suggested between N-values and the settlement; one exmaple is that
of Burland and Burbidge (1985), a simplified version of which is:
Immediate settlement,

where qn´ = net effective bearing pressure


B = breadth of foundation
N = average SPT N-value in the layer below of thickness B

Cone penetration test


For overconsolidated sands, estimates of E´ may be made from the measured cone resistance
(qc):
When qc < 50 MN/m², E' - 5qc
When qc >= 50 MN/m², E' - 250 MN/m²

Pressuremeter tests
The shear modulus G is obtained from pressuremeter results.
Then, E´= 2G(1 + ´)

Foundations

Foundation design

110
Design is an iterative process. Designers use their experience to estimate the dimensions, then
check whether the design is safe. If it is not safe, or the check indicated that it may be possible
to make economies, then they modify the dimensions and repeat the calculations. For
example:

Use presumed bearing values to obtain a first estimate of the size.

Calculate the ultimate bearing capacity (qf) at which collapse will occur.

Obtain the allowable bearing pressure from

Divide the design load by this allowable pressure to obtain a required area.

Select appropriate dimensions.

Calculate the likely settlement for this size of foundation.

Check that the predicted settlement due to this allowable bearing pressure is likely to be
acceptable.

Foundations

Types of foundation
 Shallow foundations
 Deep foundations

Shallow foundations (sometimes called 'spread footings') include pads ('isolated footings'),
strip footings and rafts.
Deep foundations include piles, pile walls, diaphragm walls and caissons.

Types of foundation

111
Shallow foundations
 Pad foundations
 Strip foundations
 Raft foundations

Shallow foundations are those founded near to the finished ground surface; generally where
the founding depth (Df) is less than the width of the footing and less than 3m. These are not
strict rules, but merely guidelines: basically, if surface loading or other surface conditions will
affect the bearing capacity of a foundation it is 'shallow'. Shallow foundations (sometimes
called 'spread footings') include pads ('isolated footings'), strip footings and rafts.
Shallows foundations are used when surface soils are sufficiently strong and stiff to support
the imposed loads; they are generally unsuitable in weak or highly compressible soils, such as
poorly-compacted fill, peat, recent lacustrine and alluvial deposits, etc.

Shallow foundations

Pad foundations
Pad foundations are used to support an individual point load such as that due to a structural
column. They may be circular, square or reactangular. They usually consist of a block or slab
of uniform thickness, but they may be stepped or haunched if they are required to spread the
load from a heavy column. Pad foundations are usually shallow, but deep pad foundations can
also be used.

Shallow foundations

112
Strip foundations
Strip foundations are used to support a line of loads, either due to a load-bearing wall, or if a
line of columns need supporting where column positions are so close that individual pad
foundations would be inappropriate.

Shallow foundations

Raft foundations
Raft foundations are used to spread the load from a structure over a large area, normally the
entire area of the structure. They are used when column loads or other structural loads are
close together and individual pad foundations would interact.

A raft foundation normally consists of a concrete slab which extends over the entire loaded
area. It may be stiffened by ribs or beams incorporated into the foundation.

Raft foundations have the advantage of reducing differential settlements as the concrete slab
resists differential movements between loading positions. They are often needed on soft or
loose soils with low bearing capacity as they can spread the loads over a larger area.

Types of foundation

Deep foundations
 Piles

Deep foundations are those founding too deeply below the finished ground surface for their
base bearing capacity to be affected by surface conditions, this is usually at depths >3 m
below finished ground level. They include piles, piers and caissons or compensated
foundations using deep basements and also deep pad or strip foundations. Deep foundations
can be used to transfer the loading to a deeper, more competent strata at depth if unsuitable
soils are present near the surface.

Piles are relatively long, slender members that transmit foundation loads through soil strata of
low bearing capacity to deeper soil or rock strata having a high bearing capacity. They are
used when for economic, constructional or soil condition considerations it is desirable to
transmit loads to strata beyond the practical reach of shallow foundations. In addition to
supporting structures, piles are also used to anchor structures against uplift forces and to assist
structures in resisting lateral and overturning forces.

Piers are foundations for carrying a heavy structural load which is constructed insitu in a deep
excavation.

113
Caissons are a form of deep foundation which are constructed above ground level, then sunk
to the required level by excavating or dredging material from within the caisson.

Compensated foundations are deep foundations in which the relief of stress due to
excavation is approximately balanced by the applied stress due to the foundation. The net
stress applied is therefore very small. A compensated foundation normally comprises a deep
basement.

Deep foundations

Piles
 Types of pile
 Types of construction
 Factors influencing choice
 Pile groups

Piled foundations can be classified according to


the type of pile
(different structures to be supported, and different ground conditions, require different types
of resistance) and
the type of construction
(different materials, structures and processes can be used).

Piles

Types of pile
 End bearing piles
 Friction piles
 Settlement reducing piles
 Tension piles
 Laterally loaded piles
 Piles in fill

Piles are often used because adequate bearing capacity can not be found at shallow enough
depths to support the structural loads. It is important to understand that piles get support from
both end bearing and skin friction. The proportion of carrying capacity generated by either
end bearing or skin friction depends on the soil conditions. Piles can be used to support
various different types of structural loads.

Types of pile

End bearing piles


114
End bearing piles are those which terminate in hard, relatively impenetrable material such as
rock or very dense sand and gravel. They derive most of their carrying capacity from the
resistance of the stratum at the toe of the pile.

Types of pile

Friction piles

Friction piles obtain a greater part of their carrying capacity by skin friction or adhesion. This
tends to occur when piles do not reach an impenetrable stratum but are driven for some
distance into a penetrable soil. Their carrying capacity is derived partly from end bearing and
partly from skin friction between the embedded surface of the soil and the surrounding soil.

Types of pile

Settlement reducing piles

115
Settlement reducing piles are usually incorporated beneath the central part of a raft foundation
in order to reduce differential settlement to an acceptable level. Such piles act to reinforce the
soil beneath the raft and help to prevent dishing of the raft in the centre.

Types of pile

Tension piles
Structures such as tall chimneys, transmission towers and jetties can be subject to large
overturning moments and so piles are often used to resist the resulting uplift forces at the
foundations. In such cases the resulting forces are transmitted to the soil along the embedded
length of the pile. The resisting force can be increased in the case of bored piles by under-
reaming. In the design of tension piles the effect of radial contraction of the pile must be taken
into account as this can cause about a 10% - 20% reduction in shaft resistance.

Types of pile

Laterally loaded piles


Almost all piled foundations are subjected to at least some degree of horizontal loading. The
magnitude of the loads in relation to the applied vertical axial loading will generally be small
and no additional design calculations will normally be necessary. However, in the case of
wharves and jetties carrying the impact forces of berthing ships, piled foundations to bridge
piers, trestles to overhead cranes, tall chimneys and retaining walls, the horizontal component
is relatively large and may prove critical in design. Traditionally piles have been installed at
an angle to the vertical in such cases, providing sufficient horizontal resistance by virtue of
the component of axial capacity of the pile which acts horizontally. However the capacity of a
vertical pile to resist loads applied normally to the axis, although significantly smaller than
the axial capacity of that pile, may be sufficient to avoid the need for such 'raking' or 'battered'
piles which are more expensive to install. When designing piles to take lateral forces it is
therefore important to take this into account.
116
Types of pile

Piles in fill

Piles that pass through layers of moderately- to poorly-compacted fill will be affected by
negative skin friction, which produces a downward drag along the pile shaft and therefore an
additional load on the pile. This occurs as the fill consolidates under its own weight.

Piles

Types of pile construction


 Displacement piles
 Non-displacement piles

Displacement piles cause the soil to be displaced radially as well as vertically as the pile shaft
is driven or jacked into the ground. With non-displacement piles (or replacement piles), soil is
removed and the resulting hole filled with concrete or a precast concrete pile is dropped into
the hole and grouted in.

Types of pile construction

Displacement piles
 Totally preformed displacement piles
 Driven and cast-in-place displacement piles
 Helical (screw) cast-in-place displacement piles
 Methods of installation

Sands and granular soils tend to be compacted by the displacement process, whereas clays
will tend to heave. Displacement piles themselves can be classified into different types,
depending on how they are constructed and how they are inserted.

Displacement piles

117
Totally preformed displacement piles
These can either be of precast concrete;
· full length reinforced (prestressed)
· jointed (reinforced)
· hollow (tubular) section
or they can be of steel of various section.

Displacement piles

Driven and cast-in-place displacement piles


This type of pile can be of two forms. The first involves driving a temporary steel tube with a
closed end into the ground to form a void in the soil which is then filled with concrete as the
tube is withdrawn. The second type is the same except the steel tube is left in place to form a
permanent casing.

Displacement piles

Helical (screw) cast-in-place displacement piles


This type of construction is performed using a special type of auger. The soil is however
compacted, not removed as the auger is screwed into the ground. The auger is carried on a
hollow stem which can be filled with concrete, so when the required depth has been reached
concrete can be pumped down the stem and the auger slowly unscrewed leaving the pile cast
in place.

Displacement piles

Methods of installation
 Dropping weight
 Diesel hammer
 Vibratory methods of pile driving
 Jacking methods of insertion

Displacements piles are either driven or jacked into the gound. A number of different methods
can be used.

Methods of installation

118
Dropping weight
The dropping weight or drop hammer is the most commonly used method of insertion of
displacement piles. A weight approximately half that of the pile is raised a suitable distance in
a guide and released to strike the pile head. When driving a hollow pile tube the weight
usually acts on a plug at the bottom of the pile thus reducing any excess stresses along the
length of the tube during insertion.

Variants of the simple drop hammer are the single acting and double acting hammers.
These are mechanically driven by steam, by compressed air or hydraulically. In the single
acting hammer the weight is raised by compressed air (or other means) which is then released
and the weight allowed to drop. This can happen up to 60 times a minute. The double acting
hammer is the same except compressed air is also used on the down stroke of the hammer.
This type of hammer is not always suitable for driving concrete piles however. Although the
concrete can take the compressive stresses exerted by the hammer the shock wave set up by
each blow of the hammer can set up high tensile stresses in the concrete when returning. This
can cause the concrete to fail. This is why concrete piles are often prestressed.

Methods of installation

Diesel hammer
Rapid controlled explosions can be produced by the diesel hammer. The explosions raise a
ram which is used to drive the pile into the ground. Although the ram is smaller than the
weight used in the drop hammer the increased frequency of the blows can make up for this
inefficiency. This type of hammer is most suitable for driving piles through non-cohesive
granular soils where the majority of the resistance is from end bearing.

Methods of installation

Vibratory methods of pile driving


Vibratory methods can prove to be very effective in driving piles through non cohesive
granular soils. The vibration of the pile excites the soil grains adjacent to the pile making the
119
soil almost free flowing thus significantly reducing friction along the pile shaft. The vibration
can be produced by electrically (or hydraulically) powered contra-rotating eccentric masses
attached to the pile head usually acting at a frequency of about 20-40 Hz. If this frequency is
increased to around 100 Hz it can set up a longitudinal resonance in the pile and penetration
rates can approach up to 20 m/min in moderately dense granular soils. However the large
energy resulting from the vibrations can damage equipment, noise and vibration propagation
can also result in the settlement of nearby buildings.

Methods of installation

Jacking methods of insertion


Jacked piles are most commonly used in underpinning existing structures. By excavating
underneath a structure short lengths of pile can be inserted and jacked into the ground using
the underside of the existing structure as a reaction.

Types of pile construction

Non-displacement piles
 Small diameter bored cast-in-place piles
 Large diameter bored cast-in-place piles
 Partially preformed piles
 Grout or concrete intruded piles

With non-displacement piles soil is removed and the resulting hole filled with concrete or
sometimes a precast concrete pile is dropped into the hole and grouted in. Clays are especially
suitable for this type of pile formation as in clays the bore hole walls only require support
close to the ground surface. When boring through more unstable ground, such as gravels,
some form of casing or support, such as a bentonite slurry, may be required. Alternatively,
grout or concrete can be intruded from an auger rotated into a granular soil. There are then
essentially four types of non displacement piles.

This method of construction produces an irregular interface between the pile shaft and
surrounding soil which affords good skin frictional resistance under subsequent loading.

Non-displacement piles

Small diameter bored cast-in-place piles

120
These tend to be 600mm or less in diameter and are usually constructed by using a tripod rig.
The equipment consists of a tripod, a winch and a cable operating a variety of tools. The basic
tools are shown in this diagram.

In granular soils, the basic tool consists of a heavy cylindrical shell with a cutting edge and a
flap valve at the bottom. Water is necessary to assist in this type of excavation. By working
the shell up and down at the bottom of the bore hole liquefaction of the soil takes place (as
low pressure is produced under the shell as the liquified soil is rapidly moved up) and it flows
into the shell and can be winched to the surface and tipped out. There is a danger when boring
through granular soil of over loosening the material at the sides of the bore. To prevent this a
temporary casing should be advanced by driving it into the ground.

In cohesive soils, the borehole is advanced by repeatedly dropping a cruciform-section tool


with a cylindrical cutting edge into the soil and then winching it to the surface with its burden
of soil. Once at the surface the clay which adheres to the cruciform blades is paired away.

Non-displacement piles

Large diameter bored cast-in-place piles

121
Large boreholes from 750mm up to 3m diameter (with 7m under-reams) are possible by using
rotary drilling machinery. The augering plant is usually crane or lorry mounted.

A spiral or bucket auger as shown in this diagram is attached to a shaft known as a Kelly bar
(a square section telescopic member driven by a horizontal spinner). Depths of up to 70m are
possible using this technique. The use of a bentonite slurry in conjunction with bucket auger
drilling can eliminate some of the difficulties involved in drilling in soft silts and clays, and
loose granular soils, without continuous support by casing tubes. One advantage of this
technique is the potential for under reaming. By using an expanding drilling tool the diameter
at the base of the pile can be enlarged, significantly increasing the end bearing capacity of the
pile. However, under-reaming is a slow process requiring a stop in the augering for a change
of tool and a slow process in the actual under-reaming operation. In clay, it is often preferable
to use a deeper straight sided shaft.

Non-displacement piles

Partially pre-formed piles


This type of pile is particularly suitable in conditions where the ground is waterlogged, or
where there is movement of water in an upper layer of the soil which could result in cement
being leached from a cast-in-place concrete pile. A hole is bored in the normal way and
annular sections are then lowered into the bore hole to produce a hollow column.
Reinforcement can then be placed and grout forced down to the base of the pile, displacing
water and filling both the gap outside and the core inside the column.

 
122
Non-displacement piles

Grout- or concrete-intruded piles


The use of continuous flight augers is becoming a much more popular method in pile
construction. These piles offer considerable environmental advantages during construction.
Their noise and vibration levels are low and there is no need for temporary borehole wall
casing or bentonite slurry making it suitable for both clays and granular soils. The only
problem is that they are limited in depth to the maximum length of the auger (about 25m).
The piles are constructed by screwing the continuous flight auger into the ground to the
required depth leaving the soil in the auger. Grout (or concrete) can then be forced down the
hollow shaft of the auger and then continues building up from the bottom as the auger with its
load of spoil is withdrawn. Reinforcement can then be lowered in before the grout sets.

An alternative system used in granular soils is to leave the soil in place and mix it up with the
pressured grout as the auger is withdrawn leaving a column of grout reinforced earth.

Piles

Factors influencing choice of pile


 Location and type of structure
 Ground conditions
 Durability
 Cost

There are many factors that can affect the choice of a piled foundation. All factors need to be
considered and their relative importance taken into account before reaching a final decision.

Factors influencing choice of pile

Location and type of structure


For structures over water, such as wharves and jetties, driven piles or driven cast-in-place
piles (in which the shell remains in place) are the most suitable. On land the choice is not so
straight forward. Driven cast-in-place types are usually the cheapest for moderate loadings.
However, it is often necessary for piles to be installed without causing any significant ground
heave or vibrations because of their proximity to existing structures. In such cases, the bored
cast-in-place pile is the most suitable. For heavy structures exerting large foundation loads,
large-diameter bored piles are usually the most economical. Jacked piles are suitable for
underpinning existing structures.

123
 

Factors influencing choice of pile

Ground conditions
Driven piles cannot be used economically in ground containing boulders, or in clays when
ground heave would be detrimental. Similarly, bored piles would not be suitable in loose
water-bearing sand, and under-reamed bases cannot be used in cohesionless soils since they
are susceptible to collapse before the concrete can be placed.

Factors influencing choice of pile

Durability
This tends to affect the choice of material. For example, concrete piles are usually used in
marine conditions since steel piles are susceptible to corrosion in such conditions and timber
piles can be attacked by boring molluscs. However, on land, concrete piles are not always the
best choice, especially where the soil contains sulphates or other harmful substances.

Factors influencing choice of pile

Cost
In coming to the final decision over the choice of pile, cost has considerable importance. The
overall cost of installing piles includes the actual cost of the material, the times required for
piling in the construction plan, test loading, the cost of the engineer to oversee installation and
loading and the cost of organisation and overheads incurred between the time of initial site
clearance and the time when construction of the superstructure can proceed.

Piles

Pile groups
Piles are more usually installed in groups, rather than as single piles. A pile group must be
considered as a composite block of piles and soil, and not a multiple set of single piles. The
capacity of each pile may be affected by the driving of subsequent piles in close proximity.
Compaction of the soil between adjacent piles is likely to lead to higher contact stresses and
thus higher shaft capacities for those piles. The ultimate capacity of a pile group is not always
dependent on the individual capacity of each pile. When analysing the capacity of a pile group
3 modes of failure must be considered.
· Single pile failure
· Failure of rows of piles
· Block failure
124
The methods of insertion, ground conditions, the geometry of the pile group and how the
group is capped all effect how any pile group will behave. If the group should fail as a block,
full shaft friction will only be mobilised around the perimeter of the block and so any increase
in shaft capacity of individual piles is irrelevant. The area of the whole base of the block must
be used in calculating the end bearing capacity and not just the base areas of the individual
piles in the group. Such block failure is likely to occur if piles are closely spaced or if a
ground-contacting pile cap is used. Failure of rows of piles is likely to occur where pile
spacing in one direction is much greater than in the perpendicular direction.

Bearing capacity
 Failure mechanisms and derivation of equations
 Bearing capacity of shallow foundations
 Presumed bearing values
 Bearing capacity of piles

The ultimate load which a foundation can support may be calculated using bearing capacity
theory. For preliminary design, presumed bearing values can be used to indicate the
pressures which would normally result in an adequate factor of safety. Alternatively, there is a
range of empirical methods based on in situ test results.

The ultimate bearing capacity (qf) is the value of bearing stress which causes a sudden
catastrophic settlement of the foundation (due to shear failure).

The allowable bearing capacity (qa) is the maximum bearing stress that can be applied to the
foundation such that it is safe against instability due to shear failure and the maximum
tolerable settlement is not exceeded. The allowable bearing capacity is normally calculated
from the ultimate bearing capacity using a factor of safety (Fs).

When excavating for a foundation, the stress at founding level is relieved by the removal of
the weight of soil. The net bearing pressure (qn) is the increase in stress on the soil.
qn = q - qo
qo = D
where D is the founding depth and  is the unit weight of the soil removed.
 
 
 

Failure mechanisms and derivation of Bearing capacity


equations
 Upper and lower bound solutions
 Semi-circular slip mechanism
 Circular arc slip mechanism

125
 A relatively undeformed wedge of soil below the foundation forms an active Rankine
zone with angles (45º + '/2).
 The wedge pushes soil outwards, causing passive Rankine zones to form with angles
(45º - '/2).
 The transition zones take the form of log spiral fans.

For purely cohesive soils ( = 0) the transition zones become circular for which Prandtl had
shown in 1920 that the solution is
qf = (2 + ) su = 5.14 su
This equation is based on a weightless soil. Therefore if the soil is non-cohesive (c=0) the
bearing capacity depends on the surcharge qo. For a footing founded at depth D below the
surface, the surcharge qo = . Normally for a shallow foundation (D<B), the shear strength of
the soil between the surface and the founding depth D is neglected.
radius of the fan r = r0 .exp[.tan'].
 is the fan angle in radians (between 0 and /2)
' is the angle of friction of the soil
ro = B/[2 cos(45+'/2)]

Failure mechanisms and derivation of


equations

Upper and lower bound


solutions
The bearing capacity of a soil can be investigated using the limit theorems of ideal rigid-
perfectly-plastic materials.

The ultimate load capacity of a footing can be estimated by assuming a failure mechanism and
then applying the laws of statics to that mechanism. As the mechanisms considered in an
upper bound solution are progressively refined, the calculated collapse load decreases.

As more stress regions are considered in a lower bound solution, the calculated collapse load
increases.

Therefore, by progressive refinement of the upper and lower bound solutions, the exact
solution can be approached. For example, Terzaghi's mechanism gives the exact solution for a
strip footing.
 
 
 
 

Failure mechanisms and derivation of


equations

126
Semi-circular slip
mechanism
Suppose the mechanism is assumed to have a semi-circular slip surface. In this case, failure
will cause a rotation about point O. Any surcharge qo will resist rotation, so the net pressure (q
- qo) is used. Using the equations of statics:
Moment causing rotation
= load x lever arm
= [(q - qo) x B] x [½B]
Moment resisting rotation
= shear strength x length of arc x lever arm
= [s] x [.B] x [B]
At failure these are equal:
(q - qo ) x B x ½B = s x .B x B
Net pressure (q - qo ) at failure
= 2  x shear strength of the soil
This is an upper-bound solution.

Failure mechanisms and derivation of equations


Circular arc slip mechanism
Consider a slip surface which is an arc in cross section, centred above one edge of the base.
Failure will cause a rotation about point O. Any surcharge qo will resist rotation so the net
pressure (q - qo) is used. Using the equations of statics:
 
 
Moment causing rotation
= load x lever arm
= [ (q - qo) x B ] x [B/2]
Moment resisting rotation
= shear strength x length of arc x lever arm
= [s] x [2R] x [R]
At failure these are equal:
(q - qo) x B x B/2 = s x 2  R x R
Since R = B / sin  :
(q - qo ) = s x 4 /(sin )²
The worst case is when
tan=2 at  = 1.1656 rad = 66.8 deg
The net pressure (q - qo) at failure
= 5.52 x shear strength of soil

Bearing capacity
Bearing capacity of shallow foundations
 Bearing capacity equation (undrained)
 Bearing capacity equation (drained)
 Factor of safety

127
The ultimate bearing capacity of a foundation is calculated from an equation that
incorporates appropriate soil parameters (e.g. shear strength, unit weight) and details about the
size, shape and founding depth of the footing. Terzaghi (1943) stated the ultimate bearing
capacity of a strip footing as a three-term expression incorporating the bearing capacity
factors: Nc, Nq and N, which are related to the angle of friction (´).

qf =c.Nc +qo.Nq + ½g.B .Ng

For drained loading, calculations are in terms of effective stresses; ´ is > 0 and N c, Nq and
N are all > 0.
For undrained loading, calculations are in terms of total stresses; the undrained shear
strength (su); Nq = 1.0 and N = 0

c = apparent cohesion intercept


qo = D  (i.e. density x depth)
D = founding depth
B = breadth of foundation
 = unit weight of the soil removed.

Bearing capacity equation Bearing capacity of shallow


foundations
(undrained)
Skempton's equation is widely used for undrained clay soils:
qf = su .Ncu + qo
where Ncu = Skempton's bearing capacity factor, which can be obtained from a chart
or by using the following expression:
Ncu = Nc.sc.dc
where sc is a shape factor and dc is a depth factor.
Nq = 1,  N = 0, Nc = 5.14
sc = 1 + 0.2 (B/L)   for B<=L
dc = 1+ Ö(0.053 D/B )  for D/B < 4

Bearing capacity equation Bearing capacity of shallow


foundations
(drained)
 Bearing capacity factors
 Shape factors
 Depth factors

Terzaghi (1943) stated the bearing capacity of a foundation as a three-term expression


incorporating the bearing capacity factors
Nc, Nq and N.
He proposed the following equation for the ultimate bearing capacity of a long strip footing:
qf =c.Nc +qo.Nq + ½g.B .N
This equation is applicable only for shallow footings carrying vertical non-eccentric loading.
For rectangular and circular foundations, shape factors are introduced.
128
qf = c .Nc .sc + qo .Nq .sq + ½  .B .N .sg
Other factors can be used to accommodate depth, inclination of loading, eccentricity of
loading, inclination of base and ground. Depth is only significant if it exceeds the breadth.
 
 
 
 

Bearing capacity equation (drained)


Bearing capacity factors
The bearing capacity factors relate to the drained angle of friction ('). The c.Nc term is the
contribution from soil shear strength, the qo.Nq term is the contribution from the surcharge
pressure above the founding level, the ½.B..Ng term is the contribution from the self weight
of the soil. Terzaghi's analysis was based on an active wedge with angles ' rather than
(45+'/2), and his bearing capacity factors are in error, particularly for low values of '.
Commonly used values for Nq and Nc are derived from the Prandtl-Reissner expression giving

Exact values for Ng are not directly obtainable; values have been proposed by Brinch Hansen
(1968), which are widely used in Europe, and also by Meyerhof (1963), which have been
adopted in North America.
Brinch Hansen:
N = 1.8 (Nq - 1) tan'
Meyerhof:
N = (Nq - 1) tan(1.4 ')

Bearing capacity equation (drained)


Shape factors
Terzaghi presented modified versions of his bearing capacity equation for shapes of
foundation other than a long strip, and these have since been expressed as shape factors.
Brinch Hansen and Vesic (1963) have suggested shape factors which depend on '. However,
modified versions of the Terzaghi factors are usually considered sufficiently accurate for most
purposes.
  sc sq s
square 1.3 1.2 0.8
circle 1.3 1.2 0.6
rectangle (B<L) 1+ 0.2(B/L) 1+ 0.2(B/L) 1 - 0.4(B/L)
B = breadth, L = length
 
 
 
 

Bearing capacity equation (drained)


Depth factors

129
It is usual to assume an increase in bearing capacity when the depth (D) of a foundation is
greater than the breadth (B). The general bearing capacity equation can be modified by the
inclusion of depth factors.
qf = c.Nc.dc + qo.Nq.dq + ½ B..d
for D>B:
dc = 1 + 0.4 arctan(D/B)
dq = 1 + 2 tan('(1-sin')² arctan(B/D)
d = 1.0
for D=<B:
dc = 1 + 0.4(D/B)
dq = 1 + 2 tan('(1-sin')² (B/D)
d = 1.0

Bearing capacity of shallow foundations


Factor of safety
A factor of safety Fs is used to calculate the allowable bearing capacity qa from the ultimate
bearing pressure qf. The value of Fs is usually taken to be 2.5 - 3.0.

The factor of safety should be applied only to the increase in stress, i.e. the net bearing
pressure qn. Calculating qa from qf only satisfies the criterion of safety against shear failure.
However, a value for Fs of 2.5 - 3.0 is sufficiently high to empirically limit settlement. It is for
this reason that the factors of safety used in foundation design are higher than in other areas of
geotechnical design. (For slopes, the factor of safety would typically be 1.3 - 1.4).

Experience has shown that the settlement of a typical foundation on soft clay is likely to be
acceptable if a factor of 2.5 is used. Settlements on stiff clay may be quite large even though
ultimate bearing capacity is relatively high, and so it may be appropriate to use a factor nearer
3.0.
 
 
 

Bearing capacity
Presumed bearing values
For preliminary design purposes, BS 8004 gives presumed bearing values which are the
pressures which would normally result in an adequate factor of safety against shear failure for
particular soil types, but without consideration of settlement.
Category Types of rocks and soils Presumed bearing value
Non-cohesive soils Dense gravel or dense sand and gravel >600 kN/m²
Medium dense gravel, 
  <200 to 600 kN/m² 
or medium dense sand and gravel
  Loose gravel, or loose sand and gravel  <200 kN/m²
  Compact sand >300 kN/m²
  Medium dense sand 100 to 300 kN/m²
<100 kN/m² depends on 
  Loose sand
degree of looseness
130
Cohesive soils Very stiff bolder clays & hard clays 300 to 600 kN/m²
  Stiff clays 150 to 300 kN/m² 
  Firm clay  75 to 150 kN/m² 
  Soft clays and silts  < 75 kN/m² 
  Very soft clay  Not applicable
Peat   Not applicable
Made ground   Not applicable

Presumed bearing values for Keuper Marl

Presumed bearing
Weathering Zone Description
value
Fully weathered IVb Matrix only as cohesive soil
IVa Matrix with occasional pellets less than 3mm 125 to 250 kN/m²
Partially III Matrix with lithorelitics up to 25mm 250 to 500 kN/m²
weathered Angular blocks of unweathered marl with
II 500 to 750 kN/m²
virtually no matrix
Unweathered 1 Mudstone (often not fissured) 750 to 1000 kN/m²

Bearing capacity
Bearing capacity of piles
 Driven piles in non-cohesive soil
 Bored piles in non-cohesive soil
 Driven piles in cohesive soil
 Bored piles in cohesive soil
 Carrying capacity of piles in a layered soil
 Effects of ground water

The ultimate bearing capacity of a pile used in design may be one three values:
the maximum load Qmax, at which further penetration occurs without the load increasing;
a calculated value Qf given by the sum of the end-bearing and shaft resistances;
or the load at which a settlement of 0.1 diameter occurs (when Qmax is not clear).

For large-diameter piles, settlement can be large, therefore a safety factor of 2-2.5 is usually
used on the working load.

A pile loaded axially will carry the load:


partly by shear stresses (s) generated along the shaft of the pile and
partly by normal stresses (qb) generated at the base.
The ultimate capacity Qf of a pile is equal to the base capacity Qb plus the shaft capacity Qs.
Qf  =  Qb + Qs   =  Ab . qb + (As . s)
where Ab is the area of the base and As is the surface area of the shaft within a soil layer.

131
Full shaft capacity is mobilised at much smaller displacements than those related to full base
resistance. This is important when determining the settlement response of a pile. The same
overall bearing capacity may be achieved with a variety of combinations of pile diameter and
length. However, a long slender pile may be shown to be more efficient than a short stubby
pile. Longer piles generate a larger proportion of their full capacity by skin friction and so
their full capacity can be mobilised at much lower settlements.

The proportions of capacity contributed by skin friction and end bearing do not just depend on
the geometry of the pile. The type of construction and the sequence of soil layers are
important factors.
 
 
 
 

Bearing capacity of piles


Driven piles in non-cohesive soil
 Ultimate pile capacity
 Standard penetration test
 Cone penetration test

Driving a pile has different effects on the soil surrounding it depending on the relative density
of the soil. In loose soils, the soil is compacted, forming a depression in the ground around the
pile. In dense soils, any further compaction is small, and the soil is displaced upward causing
ground heave. In loose soils, driving is preferable to boring since compaction increases the
end-bearing capacity.

132
In non-cohesive soils, skin friction is low because a low friction 'shell' forms around the pile.
Tapered piles overcome this problem since the soil is recompacted on each blow and this gap
cannot develop.

Pile capacity can be calculated using soil properties obtained from standard penetration
tests or cone penetration tests. The ultimate load must then be divided by a factor of safety
to obtain a working load. This factor of safety depends on the maximum tolerable settlement,
which in turn depends on both the pile diameter and soil compressibility. For example, a
safety factor of 2.5 will usually ensure a pile of diameter less than 600mm in a non-cohesive
soil will not settle by more than 15mm.

Although the method of installing a pile has a significant effect on failure load, there are no
reliable calculation methods available for quantifying any effect. Judgement is therefore left
to the experience of the engineer.
 
 
 
 

Driven piles in non-cohesive soil


Ultimate pile capacity
The ultimate carrying capacity of a pile is:
Qf = Qb + Qs
The base resistance, Qb can be found from Terzaghi's equation for bearing capacity,
qf = 1.3 c Nc + qo Nq + 0.4  B N
The 0.4  term may be ignored, since the diameter is considerably less than the
depth of the pile.
The 1.3 c Nc term is zero, since the soil is non-cohesive.
The net unit base resistance is therefore
qnf = qf - qo = qo (Nq -1)
and the net total base resistance is
Qb = qo (Nq -1) Ab
The ultimate unit skin friction (shaft) resistance can be found from
qs = Ks .'v .tan
where 'v = average vertical effective stress in a given layer
 = angle of wall friction, based on pile material and ´
Ks = earth pressure coefficient
Therefore, the total skin friction resistance is given by the sum of the layer resistances:
Qs = (Ks .'v .tan .As)
The self-weight of the pile may be ignored, since the weight of the concrete is almost equal to
the weight of the soil displaced.
Therefore, the ultimate pile capacity is:
Qf = Ab qo Nq + (Ks .'v .tan .As)

Values of Ks and  can be related to the angle of internal friction (´) using the following table
according to Broms.

133
Ks
Material  
low density high density
steel 20° 0.5
1.0
concrete 3/4 ´ 1.0 2.0
timber 2/3 ´ 1.5 4.0

It must be noted that, like much of pile design, this is an empirical relationship. Also, from
empirical methods it is clear that Qs and Qb both reach peak values somewhere at a depth
between 10 and 20 diameters.

It is usually assumed that skin friction never exceeds 110 kN/m² and base resistance will not
exceed 11000 kN/m².
 
 

 
 

Driven piles in non-cohesive soil


Standard penetration test
The standard penetration test is a simple in-situ test in which the N-value is the mumber of
blows taken to drive a 50mm diameter bar 300mm into the base of a bore hole.

Schmertmann (1975) has correlated N-values obtained from SPT tests against effective
overburden stress as shown in the figure.
The effective overburden stress =  the weight of material above the base of the borehole - the
wight of water
e.g. depth of soil = 5m, depth of water = 4m, unit weight of soil = 20kN/m³, 'v = 5m x
20kN/m³ - 4m x 9.81kN/m³  60 kN/m²

Once a value for ´ has been estimated, bearing capacity factors can be determined and used
in the usual way.

Meyerhof (1976) produced correlations between base and frictional resistances and N-values.
It is recommended that N-values first be normalised with respect to effective overburden
stress:

Normalised N = Nmeasured x 0.77 log(1920/´v)


Pile type  Soil type Ultimate base resistance Ultimate shaft resistance
134
qb (kPa) qs (kPa)
Gravelly sand  40(L/d) N
Driven 2 Navg
Sand but < 400 N
Sandy silt  20(L/d) N
   
Silt but < 300 N
 
13(L/d) N
Bored Gravel and sands
but < 300 N
Navg
Sandy silt 13(L/d) N
   
Silt but < 300 N
L = embedded length
d = shaft diameter
Navg = average value along shaft
 
 

Driven piles in non-cohesive soil


Cone penetration test
End-bearing resistance
The end-bearing capacity of the pile is assumed to be equal to the unit cone resistance (qc).
However, due to normally occurring variations in measured cone resistance, Van der Veen's
averaging method is used:
qb = average cone resistance calculated over a depth equal to three pile diameters above to one
pile diameter below the base level of the pile.
Shaft resistance
The skin friction can also be calculated from the cone penetration test from values of local
side friction or from the cone resistance value using an empirical relationship:
At a given depth, qs = Sp. qc
where Sp = a coefficient dependent on the type of pile
 
 
Type of pile Sp
Solid timber )
Pre-cast concrete )
0.005 - 0.012
Solid steel driven )
Open-ended steel 0.003 - 0.008

Bearing capacity of piles


Bored piles in non-cohesive soil
The design process for bored piles in granular soils is essentially the same as that for driven
piles. It must be assumed that boring loosens the soil and therefore, however dense the soil,
the value of the angle of friction used for calculating Nq values for end bearing and  values
for skin friction must be those assumed for loose soil. However, if rotary drilling is carried out
under a bentonite slurry ' can be taken as that for the undisturbed soil.
 
  

135
Bearing capacity of piles
Driven piles in cohesive soil
Driving piles into clays alters the physical characteristics of the soil. In soft clays, driving
piles results in an increase in pore water pressure, causing a reduction in effective stress;.a
degree of ground heave also occurs. As the pore water pressure dissipates with time and the
ground subsides, the effective stress in the soil will increase. The increase in 'v leads to an
increase in the bearing capacity of the pile with time. In most cases, 75% of the ultimate
bearing capacity is achieved within 30 days of driving.

For piles driven into stiff clays, a little consolidation takes place, the soil cracks and is heaved
up. Lateral vibration of the shaft from each blow of the hammer forms an enlarged hole,
which can then fill with groundwater or extruded porewater. This, and 'strain softening',
which occurs due to the large strains in the clay as the pile is advanced, lead to a considerable
reduction in skin friction compared with the undisturbed shear strength (su) of the clay. To
account for this in design calculations an adhesion factor, , is introduced. Values of  can be
found from empirical data previously recorded. A maximum value (for stiff clays) of 0.45 is
recommended.

The ultimate bearing capacity Qf of a driven pile in cohesive soil can be calculated from:
Qf = Qb + Qs

where the skin friction term is a summation of layer resistances


Qs = (  .su(avg) .As)

and the end bearing term is


Qb = su .Nc .Ab

Nc = 9.0 for clays and silty clays.


 
 
 
 

Bearing capacity of piles


Bored piles in cohesive soil
Following research into bored cast-in-place piles in London clay, calculation of the ultimate
bearing capacity for bored piles can be done the same way as for driven piles. The adhesion
factor should be taken as 0.45. It is thought that only half the undisturbed shear strength is
mobilised by the pile due to the combined effect of swelling, and hence softening, of the clay
in the walls of the borehole. Softening results from seepage of water from fissures in the clay
and from the un-set concrete, and also from 'work softening' during the boring operation.

The mobilisation of full end-bearing capacity by large-diameter piles requires much larger
displacements than are required to mobilise full skin-friction, and therefore safety factors of
2.5 to 3.0 may be required to avoid excessive settlement at working load.
 
 
 
 

136
Bearing capacity of piles
Carrying capacity of piles in layered soil
When a pile extends through a number of different layers of soil with different properties,
these have to be taken into account when calculating the ultimate carrying capacity of the pile.
The skin friction capacity is calculated by simply summing the amounts of resistance each
layer exerts on the pile. The end bearing capacity is calculated just in the layer where the pile
toe terminates. If the pile toe terminates in a layer of dense sand or stiff clay overlying a layer
of soft clay or loose sand there is a danger of it punching through to the weaker layer. To
account for this, Meyerhof's equation is used.

The base resistance at the pile toe is


qp = q2 + (q1 -q2)H / 10B but £ q1

where B is the diameter of the pile, H is the thickness between the base of the pile and the top
of the weaker layer, q2 is the ultimate base resistance in the weak layer, q1 is the ultimate base
resistance in the strong layer.

 
 
 

Bearing capacity of piles


Effects of groundwater
The presence and movement of groundwater affects the carrying capacity of piles, the
processes of construction and sometimes the durability of piles in service.

Effect on bearing capacity


In cohesive soils, the permeability is so low that any movement of water is very slow. They
do not suffer any reduction in bearing capacity in the presence of groundwater.
In granular soils, the position of the water table is important. Effective stresses in saturated

137
sands can be as much as 50% lower than in dry sand; this affects both the end-bearing and
skin-friction capacity of the pile.

Effects on construction
When a concrete cast-in-place pile is being installed and the bottom of the borehole is below
the water table, and there is water in the borehole, a 'tremie' is used.

With its lower end lowered to the bottom of the borehole, the tremmie is filled with concrete
and then slowly raised, allowing concrete to flow from the bottom. As the tremie is raised
during the concreting it must be kept below the surface of the concrete in the pile. Before the
tremie is withdrawn completely sufficient concrete should be placed to displace all the free
water and watery cement. If a tremie is not used and more than a few centimetres of water lie
in the bottom of the borehole, separation of the concrete can take place within the pile,
leading to a significant reduction in capacity.

A problem can also arise when boring takes place through clays. Site investigations may show
that a pile should terminate in a layer of clay. However, due to natural variations in bed levels,
there is a risk of boring extending into underlying strata. Unlike the clay, the underlying beds
may be permeable and will probably be under a considerable head of water. The 'tapping' of
such aquifers can be the cause of difficulties during construction.

Effects on piles in service


The presence of groundwater may lead to corrosion or deterioration of the pile's fabric.
In the case of steel piles, a mixture of water and air in the soil provides conditions in which
oxidation corrosion of steel can occur; the presence of normally occurring salts in
groundwater may accelerate the process.
In the case of concrete piles, the presence of salts such as sulphates or chlorides can result in
corrosion of reinforcement, with possible consequential bursting of the concrete. Therefore,
adequate cover must be provided to the reinforcement, or the reinforcement itself must be
protected in some way. Sulphate attack on the cement compounds in concrete may lead to the
expansion and subsequent cracking. Corrosion problems are minimised if the concrete has a
high cement/aggregate ratio and is well compacted during placement.

Foundations
 Load-settlement behaviour
 Types of foundation
138
 Bearing capacity
 Settlement
 Foundation design
 Ground improvement

The foundation of a structure is in direct contact with the ground and transmits the load of the
structure to the ground. Foundations may be characterised as shallow (pad, strip or raft) or
deep (piles, piers or caissons). When designing foundations, two principal criteria must be
satisfied:

Bearing capacity
There must be an adequate factor of safety against collapse (plastic yielding in the soil and
catastrophic settlement or rotation of the structure).

Settlement
Settlements at working loads must not cause damage, nor adversely affect the serviceability of
the structure.

There are other considerations that may be relevant to specific soils, foundation types and
surface conditions.

Foundations

Load-settlement behaviour
 General shear failure
 Local shear failure
 Punching shear failure
 Factors affecting mode of failure

The application of a load on a foundation causes some settlement. The three main stages of
the load-settlement curve are:
Relatively elastic vertical compression
(O-A) The load-settlement curve is almost straight.
Local shear failure
(B) Local yielding causes some upward and outward movement of the soil and results in
slight surface heave.
General shear failure
(C) Large settlements are produced as plastic yielding is fully developed within the soil.

139
Bearing capacity failure can occur in three different modes: general shear failure, local shear
failure, or punching shear failure. Local or punching shear are characterised by relatively
large settlements and the ultimate bearing capacity is not clearly defined. In these cases
settlement is the major factor in the foundation design.

Load-settlement behaviour

General shear failure


When a load (Q) is gradually applied on a foundation, settlement occurs which is almost
elastic to begin with. At the ultimate load, general shear failure occurs when a plastic yield
surface develops under the footing, extending outward and upward to the ground surface, and
catastrophic settlement and/or rotation of the foundation occurs. The load per unit area at this
point is called the ultimate bearing capacity (qf) of the foundation.

140
Load-settlement behaviour

Local shear failure


In moderately compressible soils, and soils of medium relative density, significant vertical
settlement may take place due to local shear failure, i.e. yielding close to the lower edges of
the footing. The yield surfaces often do not reach the surface. Several yield developments
may occur accompanied by settlement in a series of jerks. The bearing pressure at which the
first yield takes place is referred to as the first-failure pressure (qf(1)) - the term first-failure
load (Qf(1)) is also used.

Load-settlement behaviour

Punching shear failure


In weak compressible soils, and soils of low relative density, considerable vertical settlement
may take place with the yield surfaces restricted to vertical planes immediately adjacent to the
sides of the foundation; the ground surface may be dragged down. After the first yield has
occured the load-settlement curve will steepen slightly, but remain fairly flat. This is referred
to as a punching shear failure.

Load-settlement behaviour

Factors affecting modes of failure


According to experimental results from foundations resting on sands (Vesic, 1973), the mode
of failure likely to occur in any situation depends on the size of the foundation and the relative
density of the soil.

Other factors might be:


permeability: relating to drained/undrained behaviour
compressibility: similar to RD
shape: e.g. strips can only rotate one way
interaction between adjacent foundations and other structures
relative stiffnesses of soil and footing/structure
incidence and relative magnitude of horizontal loadings or moments
presence of stiffer or weaker underlying layers.

141
 

Foundations

Settlement
 Total and differential settlement criteria
 Imposed stress distribution below foundations
 Settlements on clays and silty clays
 Settlements on sands and gravels

The are three components making up the final settlement quantity:

Immediate settlement (i)


· elastic deformation with no change in water content
· occurs rapidly during the application of load
· quite small quantity in dense sands/gravels and stiff/hard clays
Consolidation settlement (c)
· decrease in voids volume as porewater is squeezed out of the soil
· occurs slowly according to the permeability
· only significant in clays and silts
Secondary settlement or creep (a)
· due to gradual changes in the particulate structure of the soil
· occurs very slowly, long after consolidation is completed
· most significant in soft organic soils and peats

Thus, final settlement,  = i + rc + ra

Reliable predictions of settlement require a thorough assessment of ground conditions,


including measurements of soil properties. Extensive ground investigations and statistically-
reliable testing programmes can be expensive and time-consuming, and thus not economically
viable for the routine design of shallow foundations.

Settlement

Total and differential settlement criteria


 Definitions of quantities
 Cracking and ease of repair

The principal undesirable effects of settlement in buildings are cracking due to angular
distortion; tilting due to differential settlement and excessive downward displacement. The
settlement of foundations must be limited to satisfy three main sets of criteria:
Cracking or tilting that affects visual appearance
· cracking damage can be related to the ease of repair
· a vertical tilt (w) of > 1/250 is unpleasantly noticeable
· horizontal dveviation of >1/100 and deflection/span ratos are also noticeable
Cracking or other damage that adversely affects serviceability

142
· cracking of wall and floor components can affect serviceability and be costly to repair
· differential displacements may affect the functionality of lifts, conveyors, cranes, internal
traffic movement, aligment of drains, etc.
Damage that affects structural integrity or stability
· Angular distortion (D/L) in the foundation produces consequential distortion and increases in
stresses in the structure above
· Structural damage is unlikely to occur if /L is < 1/150
· Values of /L > 1/1000 can cause cracking and overstressing.

Total and differential settlement criteria

Definitions of quantities

The following definitions and quantities have been suggested by Burland and Wroth (1975):
Settlement (  or s): the downward displacement of a point in a foundation
Differential settlement () the displacment of one point with respect to another
e.g BA = displacement of B with respect to A = B - A
Angular strain ( ): the change in slope from horizontal at a point
Tilt angle (): the vertical angle of displacement of a unit of structure
Relative deflection (): maximum displacement between two points relative to a straight line
drawn between them
Angular distortion (/L): the ratio of the relative deflection between two points and the
horizontal distance between them
Sagging occurs when the relative deflection is downward, i.e.  is +ve
Hogging occurs when the relative deflection is upward, i.e.  is -ve

Total and differential settlement criteria

Cracking and ease of repair


Jennings and Kerrich (1962) proposed a classification of damage due to the cracking based on
the ease of repair.
143
degree of
type of damage crack width (mm)
damage
0 negligible Barely visible or none < 0.1
1 very slight Fine cracks repairable by normal decoration 1-5
2 slight Cracks easily filled, may be visible externally. >5
Doors & windows stick slightly
3 moderate Cracks need raking out & filled by a mason 5-15
Doors & windows sticking Services fractured; water (more than 3)
ingress
4 severe Breaking out & replacing of masonry; floors sloping; 15-25
walls leaning/bulging; doors & windows not (depends on
functioning; number)
floor beams displaced; services disrupted, serious water
ingress
5 very severe Rebuilding or major repair required; beams lose >25
bearing; (depends on
walls require shoring; number)
danger of instability or collapse.

Settlements on clay

Imposed stress distribution below foundations


 Boussinesq's solution
 Circular foundation
 Strip foundation
 Rectangular foundation
 Approximate methods

The distribution of stresses in the ground under a foundation due to applied loading is not
uniform. Changes in vertical stress decrease with both depth and horizontal distance from the
load; but can be predicted with reasonable accuracy using elasticity theory (and sometimes
simpler approximate methods).

Imposed stress distribution below foundations

Boussinesq's solution
According to Boussinesq (1885), the vertical stress v in the ground due to a point load Q is

144
where z is the depth below the load and r is the horizontal distance from the load.

Simple solutions are available for stresses below strip footings, below the centre of circular
footings and below a corner of a rectangular footing. The latter can be used to calculate the
stress at any point by dividing the rectangle into two or more rectangles and summing the
stresses due to each part. For example, to find the stress under the centre of a (B x L)
rectangular base, find the stress under the corner of a (1/2B x 1/2L) rectangle, then multiply
by 4.

Imposed stress distribution below foundations

Stress below a circular foundation


The vertical stress at a depth z below the centre of circular base of radius R is
v = q .Ic
where q is the bearing pressure and

The stress value below the centre is a maximum for a given depth. The expressions for
stresses off-centre are much more complex. Over deep soil layers, the average value will be
between 0.85 and 0.6 of the centre-line value according to the stiffness of the footing.

Imposed stress distribution below foundations

Stress below a strip foundation


145
The vertical stress at a depth z below a uniformly loaded strip footing of width B=2b is
v = q .Is
where q is the bearing pressure and
Is = [ + sin .cos(  + 2 ) ] / 

Imposed stress distribution below foundations

Stress below a rectangular foundation


The vertical stress at a depth z below the corner of a rectangular subject to uniform pressure is
v = q. IR
where q is the bearing pressure and

Imposed stress distribution below foundations

Approximate methods
For settlement calculations (but not necessarily for other geotechnical problems), sufficient
accuracy is usually obtained by assuming a simplified pressure distributuion. The two
methods given below are in common use.

146
Tomlinson's method
Tomlinson suggests using an
approximate pressure distribution
for small foundations on stiff clay:

Settlement,  = ( 1.5 B) x ( 0.55 qn) x mv

where
B is the breadth of the foundation
qn is the net bearing pressure

2 : 1 distribution method
The vertical stress on horizontal planes is assumed to remain uniform, but decrease linearly
with depth below the foundation, thus:

For a strip footing:

For a rectangular footing:


where: qn = applied net bearing pressure
B = breadth,
L = length
z = depth from underside of footing

Settlement

Settlements on clays and silty clays


 Calculation of immediate settlement
 Calculation of consolidation settlement

Settlements are time-dependent and may take several years to complete. The soil properties
are obtained principally from laboratory tests. Relative proportions of final settlement that
occur as immediate and consolidation settlement vary according to the stress history and state
of the soil.

Soft normally consolidated clays and silts


i » 0.1roed
where oed = consolidation settlement from oedometer mv
The final settlement (excluding creep),
 =  + oed = 1.1 roed

147
Stiff overconsolidated clays
Both immediate and consolidation settlement are incorporated in the oedometer-measured
modulus.
Thus, final settlement,  = oed
Proportions:
i » 0.5oed - 0.6oed
c » 0.5oed - 0.4oed

Settlements on clays and silty clays

Calculating immediate settlement


 Typical values of Poisson's ratio
 Typical values of undrained stiffness
 Typical values of influence factor

The net immediate settlement is assumed to be due entirely to elastic volume change upon
loading.

The following expression is based on elastic theory.

Immediate settlement
where qn = net bearing pressure
B = breadth of foundation
 = Poisson's ratio
Eu = undrained stiffness modulus

Ip = influence factor

Calculating immediate settlement

Typical values of Poisson's ratio

Soil Poisson's ratio ()

Undrained saturated clays/silts u = 0.5

Stiff sandy or silty clays ´ = 0.2 - 0.4

Medium to loose sands ´ = 0.4

148
Dense sands ´ = 0.2 - 0.45

Calculating immediate settlement

Typical values of undrained stiffness


It is general practice to obtain values for Eu from undrained laboratory tests. However, other
empirical relationships can also be used:

From oedometer test results


The drained stiffness modulus,
E´ = 1/mv
and then
Eu = 1.5E´/(1 + ´)
where ´ = drained value of Poisson's ratio.

From undrained shear strength


In soils, stiffness moduli are strain-dependent. The strains associated with normal foundations
at working loads are generally less than 0.1%. Some empirical relationships have been noted,
for example, for London clay,
Eu /su = 150 to 500

Calculating immediate settlement

Typical values of influence factor


Influence factors Ip for vertical displacement due to elastic compression of a layer of semi-
infinite thickness

Shape Flexible Rigid


  Centre Corner Average  
Circle 1.00 0.64 0.85 0.79
Rectangle    
L/B = 1.0 1.122 0.561 0.946 0.82
L/B = 1.5 1.358 0.679 1.148 1.06

149
L/B = 2.0 1.532 0.766 1.300 1.20
L/B = 3.0 1.783 0.892 1.527 1.42
L/B = 4.0 1.964 0.982 1.694 1.58
L/B = 5.0 2.105 1.052 1.826 1.70
L/B = 10.0 2.540 1.270 2.246 2.10
L/B = 100.0 4.010 2.005 3.693 3.47

Settlements on clays and silty clays

Calculating consolidation settlement


 Typical values of volume compressibility

The oedometer settlement of a soil layer can be calculated from:


oed = mv.´.HL
where
mv = coefficient of volume compressibilty measured in an oedometer test
´ = the average change in vertical effective stress in the layer
HL = thickness of the layer

In layered deposits, the sum of layer-settlements should be take down to a depth where
´ < 0.1qn

Oedometer test results tend to overestimate the actual settlement; the following correction
factors should be applied:
c = g.oed
where g = a factor suggested by Skempton and Bjerrum (1957)

Type of clay mg

Very sensitive; soft alluvial, estuarine and 1.0-1.2


marine clays

Normally consolidated clay 0.7-1.0

Overconsolidated clays, 0.5-0.7


e.g. London Clay, Weald, Kimmeridge,
Oxford and Lias clays.

150
Heavily overconsolidated clays, 0.2-0.5
e.g. Keuper Marl, glacial till

Calculating consolidation settlement

Typical values of volume compressibility


The values given below should be taken as guideline figures only, for preliminary design and
first estimates; values based on reliable tests should be used in final design calculations.

Type of soil Degree of mv


commpressibility (m²/MN)

Heavilly overconsolidated clays Very low


< 0.05
compressibilty

Overconsolidated clays Low compressibilty 0.05 - 0.1

Weathered overconsolidated Medium


0.1 - 0.3
clays compressibilty

Normally consolidated clays High compressibilty 0.3 - 1.5

Organic alluvial clays and silts; Very high


> 1.5
peats compressibilty

Settlement

151
Settlements on sands and gravels
Settlements on sands and gravels take place almost immediately. Obtaining undisturbed
samples for laboratory testing is difficult. The soil properties are estimated from in situ test
results.

Standard penetration test


Relationships have been suggested between N-values and the settlement; one exmaple is that
of Burland and Burbidge (1985), a simplified version of which is:
Immediate settlement,

where qn´ = net effective bearing pressure


B = breadth of foundation
N = average SPT N-value in the layer below of thickness B

Cone penetration test


For overconsolidated sands, estimates of E´ may be made from the measured cone resistance
(qc):
When qc < 50 MN/m², E' - 5qc
When qc >= 50 MN/m², E' - 250 MN/m²

Pressuremeter tests
The shear modulus G is obtained from pressuremeter results.
Then, E´= 2G(1 + ´)

Foundations

Foundation design
Design is an iterative process. Designers use their experience to estimate the dimensions, then
check whether the design is safe. If it is not safe, or the check indicated that it may be possible
to make economies, then they modify the dimensions and repeat the calculations. For
example:

Use presumed bearing values to obtain a first estimate of the size.

Calculate the ultimate bearing capacity (qf) at which collapse will occur.

Obtain the allowable bearing pressure from

Divide the design load by this allowable pressure to obtain a required area.

Select appropriate dimensions.


152
Calculate the likely settlement for this size of foundation.

Check that the predicted settlement due to this allowable bearing pressure is likely to be
acceptable.

Ground improvement
 Engineering properties
 Drainage
 Pre-consolidation
 Compaction
 Grouting
 Geo-textiles

Where poor ground conditions make traditional forms of construction expensive, it may be
economically viable to attempt to improve the engineering properties of the ground before
building on it. This can be done by reducing the pore water pressure, by reducing the volume
of voids in the soil, or by adding stronger materials.
 
 

Engineering properties
 Compression
 Consolidation
 Shear strength
 Permeability

The properties of soil which most affect the cost of construction are strengthand
compressibility. Both can be improved by reducing the volume of the voids in the soil mass.
Water must be displaced from saturated soils in order to reduce the volume of the voids. This
may take months if the permeability of the soil is low.
 
 

Back to Engineering properties

Compression
Soil which is highly compressible is prone to volume change when a load is applied. This
leads to settlement. Fine-grained soils which have been compressed and then allowed to swell,
experience a smaller volume change when re-compressed. Loosely-compacted coarse-grained
soils may exhibit little change in volume under static loads, but become unstable and exhibit
large volume changes when either vibrated or flooded and then
drained.
 
 

153
Back to Engineering properties

Consolidation
The sudden application of a load to a saturated soil produces an immediate increase in
porewater pressure. Over time, the excess porewater pressure will dissipate, the effective
stress in the soil will increase and settlement will increase. Since shear strength is related to
effective stress, it may be necessary to control the rate of construction to avoid a shear failure.
This was the case, for example, when approach embankments were constructed on soft
alluvium, for the bridge which carried the M180 motorway over the River Trent near
Scunthorpe. The rate at which the excess water pressure dissipates, and settlement occurs,
depends on the permeability of the soil, the amount of water to be expelled and the distance
the water must travel.

Back to Engineering properties

 
Shear strength
Collapse will occur if the shear stress along a potential
failure surface exceeds the shear strength of the soil. Shear
strength depends on the effective normal stress, which
depends on the porewater pressure. Undrained loading
causes an increase in porewater pressure equal to the
change in the total normal stress so that there is no increase
in strength to match the change in the shear stress.

The shear strength can be increased either by decreasing the


water pressure or reducing the void ratio of the soil to
produce a peak strength which exceeds the critical shear
stress.

 
 
 
 

154
Back to Engineering properties

Permeability
Fine-grained soils have a lower permeability than coarse-grained soils, thus excess porewater
pressures take longer to dissipate. Consolidation reduces the void ratio of the soil and further
decreases the permeability. Real soils are not hydraulically isotropic: the natural orientation of
particles in soils which have been consolidated vertically tends to produce a horizontal
permeability which is greater than the vertical permeability.

Thin horizontal layers of coarse-grained soil in a mass of fine-grained soil may dramatically
increase the horizontal permeability while having little effect on the vertical permeability. It is
possible to increase the drainage rate without changing the permeability of the bulk of the soil
by introducing layer drains (sandwicks) or fracturing the soil. The most effective way to
reduce seepage into an excavation, through or under a dam, or away from contaminated
ground is to create a low permeability zone perpendicular to the direction of flow.
 
 

Back to Ground improvement

Drainage
Pumping water out of the ground will cause a local
lowering of the ground water level and a decrease in
water pressure. Both will return to their natural state
when pumping stops. The rate of drawdown and the
radius of influence depend on the permeability of the
soil: Low permeability implies slow drawdown and
large radius. Decreasing the water pressure increases
the effective stress, which increases the shear strength
and causes settlement.

The introduction of a grid of vertical drains, connected


by layer of highly permeable soil, reduces the distance
water has to travel through the natural soil and
facilitates horizontal flow. This limits the excess water
pressure generated during and after construction and
increases the rate of settlement.

Back to Ground improvement

Pre-consolidation

155
Settlement due to an applied pressure occurs over a period of time. A proportion of the final
settlement can be achieved prior to construction by pre-loading the soil. The larger the pre-
load, the less time it will take to achieve the final settlement. Pre-consolidating the ground in
this way tends to be an expensive solution compared with the use of piles to support localised
loads such as columns. Pre-consolidation may be a cost-effective way of reducing the
settlement due to lightly distributed loads from roads or warehouse or supermarket floors
provided that material is readily available to provide the pre-loading. Pre-consolidation is
normally designed to take 6 - 9 months.

Back to Ground improvement

Compaction
 Compaction of fill
 Dynamic compaction
 Vibro-compaction/replacement

Compaction is a dynamic process, reducing the volume of soil by expelling air. The moisture
content is not altered significantly under normal circumstances. (Water may migrate a short
distance from the point of application but is forced to return when compaction is applied to
the adjacent soil). Compaction is most effective when applied to a thin layer because the
energy dissipates with distance. Vibration is the most effective method of compacting loose
coarse-grained soils.
 
 

Back to Compaction

Compaction of fill
Fills are normally compacted in layers between 300mm and 600mm thick. For granular soils,
a motor on the back of the roller is used to rotate an eccentric mass causing the roller to
vibrate. For fine-grained soils, the roller may be fitted with blunt spikes known as sheep's feet.
Sheep's foot rollers produce a kneeding action which changes the shape of clods of soil and
displaces air from the spaces between the clods.
 
 

Back to Compaction

156
Dynamic compaction
Dynamic compaction involves lifting and dropping a heavy weight several times in one place.
The process is repeated on a grid pattern across the site. Trials in the UK indicate that the
masses in the range 5 to 10 tonnes and drops in the range 5 to 10m are effective for
compacting loose sand but not clay. Masses up to 190 tonnes and drops of 25m are used by
TLM (Technique Louis Ménard) in France. Such heavy compaction causes fractures through
which water can flow. This, according to the proponents of the system, enables fine-grained
soils to be compacted. Heavy compaction tends to annoy the neighbours, which limits its use
in built-up areas.
compactive energy per blow = m.g.h
where m = mass, g = gravitational constant, h = drop.

estimated depth of compaction = n.(m.h)


where n is an empirical constant between 0.3 and 1 depending on the grain size distribution
and degree of saturation (0.5-1 for sands, 0.3-0.5 for silts and clayey soils).

Back to Compaction

Vibro-compaction/replacement
Both vibro-compaction and vibro-replacement use a vibrating poker to make a hole in the
ground. Soil is displaced sideways, not removed from the ground.

vibro-compaction
in coarse-grained soils the poker may be removed slowly while still vibrating. This causes the
sides of the hole to collapse and results in a depression in the ground surface.

vibro-replacement
in fine-grained soils it is usual to fill the hole with coarse aggregate (up to 50mm). The poker
may be used to compact the stone column in layers. A typical column might be 5m deep and
500mm diameter. A line of columns at say 3m centres can be used to support a reinforced
concrete ground beam effectively producing a piled foundation.
 
 

Back to Ground improvement

Grouting

157
Injecting cementitious material into a soil mass tends to reduce permeability, cause swelling
and may increase strength.

Grout injection into fractured rock which forms the foundation of a dam is possibly the oldest
and best known application. Grout injection has been used successfully to strengthen and
reduce permeability of soil around a basement excavation below the water table. It has also
been used to control the settlement of structures adjacent to tunnel excavations in London:
predicted settlements of 60mm, which would have caused extensive damage to old buildings,
were limited to 10mm.

Silty soils with high water contents are unsuitable for embankment construction in their
natural state because they are difficult to compact. They can be improved by mixing hydrated
lime with the soil.
 
 

Back to Ground improvement

Geo-textiles
Geo-textiles can be used for:
segregation of layers
Rock-fill laid on soft ground to form a road or embankment base can be prevented from
punching into the soil below using a geotextile underlay.
tensile strength
Horizontal membranes can be used to provide tensile re-inforcement and reduce
settlement. There are two primary difficulties:
(i) aligning the mebrane in the direction of the principal tensile stress, which is
probably not horizontal, and
(ii) the fact that geotextiles have a low modulus of elasticity and are plastic and
therefore tend to creep.
a drainage layer
Either as a water-conductor or as a filter to reduce the migration of fine particles into a
granular soil drain.
an impermeable barrier
To prevent or control the flow of contaminated groundwater from or in land-fill sites.

Further work is requred to complete the interal links and popup references (98 broken links out of 384)
Based on part of the GeotechniCAL reference package
back to home page by Tony Price, Warwick University

Slope stability
 Causes of instability
 Mechanics of slopes
 Analysis of translational slip
 Analysis of rotational slip
 Site investigation
158
 Remedial measures

Soil or rock masses with sloping surfaces, either natural or constructed, are subject to forces
associated with gravity and seepage which cause instability. Resistance to failure is derived
mainly from a combination of slope geometry and the shear strength of the soil or rock itself.

The different types of instability can be characterised by spatial considerations, particle size
and speed of movement. One of the simplest methods of classification is that proposed by
Varnes in 1978:

I. Falls
II. Topples
III. Slides rotational and translational
IV. Lateral spreads
V. Flows in Bedrock and in Soils
VI. Complex

Falls
In which the mass in motion travels most of the distance through the air. Falls include: free
fall, movement by leaps and bounds, and rolling of fragments of bedrock or soil.

Topples
Toppling occurs as movement due to forces that cause an over-turning
moment about a pivot point below the centre of gravity of the unit. If
unchecked it will result in a fall or slide.

The potential for toppling can be identified using the graphical construction on a stereonet.
The stereonet allows the spatial distribution of discontinuities to be presented alongside the
slope surface. On a stereoplot toppling is indicated by a concentration of poles "in front" of
the slope's great circle and within ± 30º of the direction of true dip.

159
Lateral Spreads
Lateral spreads are disturbed lateral extension movements in a fractured mass. Two subgroups
are identified:

A. Where the spread is without a well-defined controlling basal shear surface or zone of
plastic flow.
B. In which extension of rock or soil results from liquefaction or plastic flow of subjacent
material.

Flows
Two subtypes are identified:

A. In Bedrock
Where flows include spatially continuous deformation and superficial as well as deep creep.
They also involve extremely slow deep creep and extremely slow and generally non-
accelerating differential movements among relatively intact units. Movements may:

 be along shear surfaces that are apparently not connected,


 result in folding, bending or bulging, or
 roughly simulate those of viscous fluids in distribution of velocities.

B. In Soils
In which the movement within the displaced mass is such that the form taken by moving
material, or the apparent distribution of velocities and displacements, resemble those of
viscous fluids. The slip surfaces within the moving material are usually not visible or are
short-lived. The boundary between the moving mass and material may be a sharp surface or
differential movement or a zone of distributed shear. Movement ranges from extremely rapid
to extremely slow.

Complex
Complex movement is by a combination of one or more of the five other principal types of
movement described by Varnes' Classification. Many landslides are complex, although one
type of movement generally dominates over the others at certain areas within a slide or at a
particular time.

back to Slope stability

Causes of instability
 Influence of factors on the factor of safety
 Classification of causes of instability
 Further reading

Starting from the general definition for the factor of safety,


Fs = f /  = (shear stress at failure) / (shear stress)

160
Terzaghi divided landslide causes into external causes which result in an increase in shearing
stress and internal causes which result in a decrease of the shearing resistance.

Varnes pointed out that there are a number of external or internal causes which may be
operating either to reduce the shearing resistance or increase the shearing stress. There are
also causes affecting simultaneously both terms of the factor of safety ratio.

The influence of different contributory factors on the factor of safety of a slope varies in time
due to a variety of factors.

back to Causes of instability

Classification of causes of instability


It must be appreciated that the causes of instability are often complex and any attempt at
classification will be approximate and incomplete. The Working party on World Landslide
Inventory have proposed a list of causal factors grouped under four main headings:

 Ground conditions
 Geomorphological processes
 Physical processes
 Man-made processes

Ground conditions

 Plastic weak material


 Sensitive material
 Collapsible material
 Weathered material
 Sheared material
 Jointed or fissured material
 Adversely oriented mass discontinuities (including bedding, schistosity, cleavage,
faults, unconformities, flexural shears, sedimentary contacts)
 Contrast in permeability and its effects on ground water
 Contrast in stiffness (stiff, dense material over plastic materials)

Geomorphological processes

 Tectonic uplift
161
 Volcanic uplift
 Glacial rebound
 Fluvial erosion of the slope toe
 Wave erosion of the slope toe
 Glacial erosion of the slope toe
 Erosion of the lateral margins
 Subterranean erosion (solution, piping)
 Deposition loading the slope crest
 Vegetation removal (by erosion, forest fire, drought)

Physical factors

 Intense, short period, rainfall


 Rapid melt of deep snow
 Prolonged high precipitation
 Rapid drawdown following floods, high tides or breaching of natural dams
 Earthquake
 Volcanic eruption
 Breaching of crater lakes
 Thawing of permafrost
 Freeze and thaw weathering
 Shrink and swell weathering of expansive soils

Man-made processes

 Excavation of the slope or at its toe


 Loading of the slope or at its crest
 Drawdown (of reservoirs)
 Irrigation
 Defective maintenance of drainage system
 Water leakage from services (water supplies, sewers, stormwater drains)
 Vegetation removal ( deforestation)
 Mining and quarrying (open pits or underground galleries)
 Creation of dumps of very loose waste
 Artificial vibration (including traffic, pile driving, heavy machinery)

back to Causes of instability

Influence of different factors on instability

162
Reference
It can be seen that short-term variations in factor of safety may occur due to seasonal
variations in groundwater levels while longer term trends may reflect the influences of
weathering or longer term changes in groundwater conditions. This approach is useful in
emphasising that landslides (and slope instability in general) may not be attributable to a
single causal factor.

From the physical point of view it may be useful to visualise slopes as existing in one of the
following three stages:
Stable: the margin of stability is sufficiently high to withstand all destabilising forces.
Marginally stable: likely to fail at some time in response to destabilising forces reaching a
certain level of activity.
Actively unstable: slopes where destabilising forces produce continuous or intermittent
movements.

These three stability stages provide a useful framework for understanding the causal factors of
instability and classifying them into two groups on the basis of their function:

Preparatory causal factors - which make the slope susceptible to movement without actually
initiating it and thereby tending to place the slope in a marginally stable state.

Triggering causal factors - which initiate movement. These causal factors shift the slope from
a marginally stable state to an actively unstable state.

Examples of External causes Resulting in Increased Shearing Stress

 Geometrical changes
 Unloading the slope toe
 Loading the slope crest
 Shocks and vibrations
 Drawdown, and
 Changes in water regime

Examples of Internal Causes Resulting in Decreased Shearing Resistance


163
 Progressive failure,
 Weathering,
 Seepage erosion

back to Causes of instability

Further Reading
Brunsden D., "Mass movement, in Processes in Geomorphology", ed. C. Embleton and J.
Thornes, Edward Arnold, London, 1979, pp. 130-186.

Brunsden, D., "Landslides and the International Decade for Natural Disaster Reduction: do
we have anything to offer. Landslides Hazard Mitigation", Royal Academy of Engineering,
June 1993, London, 1995, pp 8-18.

Crozier, M.J., "Landslides - Causes, consequences and environment", Croom Helm, London,
1986, pp 252.

Cruden, D.M., "A simple definition of a landslide", Bulletin IAEG, No. 43, 1991, pp 27-29.

McRoberts E.C., Morgenstern, N.R., "The stability of thawing slopes", Canadian


Geotechnical Journal, Vol. 11, 447-469.

O'Shea B.E., Ruapehu and the Tangiwai disaster, New Zealand Journal of Science and
Technology, B, 36, 1974, pp 174-189.

Popescu, M., "Landslides in overconsolidated clays as encountered in Eastern Europe",


Proceedings 4th International Symposium on Landslides, Toronto, Vol. 1, 1984, pp 83-106.

Popescu, M., "A suggested method for reporting landslide causes." Bull IAEG, No. 50, Oct.
1994, pp 71-74.

Terzaghi, K., "Mechanisms of landslides", Geological Society of America, Berkely Volume,


1950, pp 83-123.

Varnes, D.J. "Slope movements and types and processes." In: Landslides Analysis and
Control, Transportation Research Board Special Report 176, 1978, pp 11-33.

Working Party on World Landslide Inventory, "A suggested method for reporting a
landslide", Bulletin EEG, No. 41, 1990, pp 5-12.

Working Party on World Landslide Inventory, "A suggested method for a landslide
summary", Bulletin IAEG, No. 43, 1991, pp 101-110.

164
Mechanics
 Loads
 Pore pressure
 Earthquakes
 Peak, critical state and residual strength
 Stress changes in slopes
 Choice of strength parameters
 Choice of factor of safety

In every slope there are forces which tend to promote downslope movement and opposing
forces which tend to resist movement.

A general definition of the factor of safety (Fs) of a slope results from comparing the
downslope shear stress () with the shear strength (f) of the soil along an assumed or known
rupture surface:
Fs = f / 

back to Mechanics

Loads
 Neutral Point Concept

Changes produced by loading or unloading are widely recognised as being mechanisms by


which instability can occur. The effects of variations in loading can be considered using the
concept of the 'neutral point' developed by Hutchinson in 1977.

back to Loads

Neutral Point Concept


 Location of the drained and undrained neutral points
 Interpreting the neutral point concept
165
 Further Reading

Essentially the neutral point method considers the ratio of the new to old factors of safety
(Fs1 / Fs0) for a potential (or existing) failure surface, as a load is placed at different points on
the ground above it.

If placed toward the toe of the failure surface, the load will produce an increase in the factor
of safety (i.e. the loading is beneficial and Fs1 / Fs0 increases), but if placed toward the head of
the failure surface, it produces a decrease in the factor of safety (i.e. loading is detrimental and
Fs1 / Fs0 decreases).

There is clearly an intermediate position where the loading causes no change in the factor of
safety for the failure surface and Fs1 / Fs0 = 1.0, this is termed the 'neutral point'.

back to Neutral Point Concept

Location of the drained and undrained neutral points


It can be shown that two extreme neutral points exist, depending upon whether the loading is
applied under drained or undrained conditions. For the undrained condition ( =1.0), the
neutral point (Nu) occurs at the point of zero inclination of the failure surface, whilst for
drained conditions ( =0 ) the neutral point (Nd) occurs where the inclination of the failure
surface  is given by

Drainage conditions between these two extremes (1.0 > > 0 ) have a corresponding
intermediate neutral point (Ni).

Similar effects can be observed for the removal of loads, by for example making cuts in the
slope.

Fso = factor of safety before addition or removal of load.


Fs1 = factor of safety after addition or removal of load.
 = local dip of slip surface (units: degrees).
' = angle of friction in terms of effective stress (units: degrees).
= pore pressure coefficient.
166
Nu = neutral point for undrained condition.
Nd = neutral point for drained condition.
Ni = neutral point for intermediate drained conditions.

back to Neutral Point Concept

Interpreting the neutral point concept


The effects due to the location of load vary according to its position in relation to the drained
and undrained neutral points. Three zones describing the effect of adding loads can therefore,
be defined:

· Zone A : Always detrimental,


· Zone B : Detrimental in the short term; beneficial in the long term, and
· Zone C : Always beneficial.

Three zones can also be defined for the effects of reduced loading (i.e. cutting):

· Zone A : Always beneficial,


· Zone B : Beneficial in the short term; detrimental in the long term, and
· Zone C : Always detrimental.

The simple concept of the neutral point must, however, be applied with care. For example, the
engineer must recognise that the addition or removal of load may cause a change in the
position of the critical failure surface (and hence the neutral points). Particular care must be
exercised when dealing with potential deep/shallow slips, since the addition of load may, in
these cases, cause improvements in the stability condition of a shallow slip surface, whilst
reducing the stability condition of a deeper slip surface.

back to Neutral Point Concept

167
Further reading
Hutchinson, J.N., "Assessment of the effectiveness of corrective measures in relation to
geological conditions and types of slope movement." General Report to Theme 3. Symposium
on Landslides and other Mass Movements, Prague, September 1977. Bulletin, International
Association of Engineering Geology, No. 16, 1977, pp. 131-155. Reprinted (1978) in
Norwegian Geotechnical Institute Publication, No. 124, pp. 1-25.

Hutchinson, J.N. , "Engineering in a landscape." Inaugural Lecture, 9 October 1979, Imperial


College of Science and Technology, University of London, London, England. 1983.

Hutchinson, J.N.,"An influence line approach to the stabilisation of slopes by cuts and fills."
Canadian Geot. J., 21, 2, 1984, pp 363-370.

back to Mechanics

Pore pressures
 End of construction
 Steady seepage
 Rapid drawdown
 Further reading

In order to estimate the factor of safety Fs for a slope in terms of effective stress (i.e. in the
long-term condition), the pore water pressure must be known. This is frequently the greatest
source of inaccuracy in slope stability work, since the determination of the most critical
conditions of pore water pressure is complex and costly. The following three sets of
conditions are usually considered for constructed slopes:
· End of construction
· Steady seepage
· Rapid drawdown

In natural slopes, the distribution of pore water pressure may be highly complicated due to
changes in soil type, anisotrophy etc. The pore pressures are detemined from site
measurements using observation wells or piezometers. Monitoring must be continued over
long periods of time in order to define the worst or most critical conditions. Methods for the
in-situ measurement of pore water pressure are described by Clayton, Matthews and Simons.

Clayton, C.R.I., Matthews, M.C., and Simons, N.E.,1995. Site Investigation. Blackwell
Science. 584 pp.

back Pore pressures

168
End of construction
Analyses of the short-term condition of stability are normally performed in terms of total
stress, with the assumption that any pore water pressure set up by the construction activity
will not dissipate at all. In many earth dams or large embankments, however, the construction
period is relatively long, and some dissipation of the excess pore water pressure is likely.
Under these conditions, a total stress analysis would yield a value of F s on the low side,
possibly resulting in un-economic design.

At any point, the pore water pressure is given by


u = uo + 
The pore pressure coefficient is defined as
=  / 1
Hence,
u = uo + 1

The pore pressure ratio is defined as


ru = u / h
Then,

Generally,
1 = 
Therefore,

The pore pressure coefficient may be determined in the triaxial apparatus. In embankments
etc., if the soil is placed at a water content below the optimum water content, the value of uo
may be close to zero, and in this case ru = .
In order to increase the rate of dissipation of pore water pressure (and hence reduce the value
of ru), drainage layers may be incorporated in the embankment.

uo = initial pore water pressure


 = change in pore water pressure due to a change in the major principal stress 1
 = unit weight of soil
h = depth of element
= pore pressure coefficient
ru = pore pressure ratio
hw = head of water
w = unit weight of water

back to Pore pressures

Steady seepage
169
Under conditions of steady seepage, the pore water pressure can be obtained from the flow
net. The pore water pressure at a point on the actual or assumed slip surface is obtained from
the value of the equipotential passing through the point.

Referring to the figure, the pore water pressure at point P is obtained by constructing the
equipotential line through P and is then equal to the head given by hw.

u = w . hw

Then,

In homogeneous conditions, ru may attain values approaching 0.45. Internal drainage layers
will produce lower values.

back to Pore pressures

Rapid drawdown
In earth dams, rapid reductions in the water level produce significant and potentially
dangerous changes in pore water pressure. This occurs because the water in the soil tends to
flow back into the reservoir through the upstream face. In this scenario, even a period of some
weeks may bring about a 'rapid' change in the pore water pressure distribution.

The pore water pressures under conditions of rapid drawdown are determined using the
following procedure.

In the figure, consider the point P on a trial slip surface. Under conditions of steady seepage,
the pore water pressure at P is obtained from the flow net, giving:
uo =  (h + hw - h' )

It is assumed that the total major stress at P is given by the overburden pressure,
1 =  h

If under drawdown, the phreatic surface falls to a level below hw, then the change in total
major principal stress is given by
1 = - w hw
The corresponding change in pore water pressure is
 = - w hw

Therefore, immediately following drawdown, the pore water pressure at P is


u = uo + 
= w [ h + hw - h' ] - w hw
= w [ h + hw (1 - ) - h' ]

170
And, if
ru = u / 

Then,

back to Pore pressures

Further reading
Bromhead, E.N. 1992. The Stability of Slopes. Blackie. 411 pp.

Clayton, C.R.I., Matthews, M.C., and Simons, N.E., 1995. Site Investigation. Blackwell
Science. 584 pp.

Whitlow, R. 1995. Basic Soil Mechanics. Longman. Scientific and Technical. 559 pp.

back to Mechanics

Earthquakes
 Earthquakes and slope stability
 Earthquake mechanism
 Mercalli scale
 Further reading

Earthquakes result from the sudden release of elastic strain energy stored in the Earth's crust.
Stress accumulates locally from various causes until it exceeds the strength of the rocks, when
slip occurs by brittle failure on dislocations or fractures known as faults. Earthquakes give
rise to two types of surface displacement: permanent offsets on the fault itself, and transient
displacement resulting from the propagation of seismic waves away from the source. A small
movement on a fault may produce a considerable shock because of the energy involved.
Earthquakes range from slight tremors which do little damage, to severe shocks which can
cause widespread damage including the initiation of landslides, collapse of buildings and
fracturing of supply mains and lines of transport.

The surface displacement associated with earthquakes may range from a few centimetres up
to several meters. For example in the disastrous earthquake which affected San Francisco in
1906, the motion of the ground was mainly horizontal along the San Andreas fault with one
side moving approximately 4.6m relative to the other.

 
171
back to Earthquakes

Earthquakes and slope stability


 Stability analysis including seismic effects
 Shear strength parameters in seismic analysis

Earthquakes can affect slope stability in three ways:

Earthquakes produce horizontal and vertical accelerations in soil masses. The horizontal
accelerations may reach as much as 0.5g (where g is gravitational acceleration), altering the
distribution of forces in hillslopes in a manner equivalent to a temporary steepening of the
slope.

Rapid repeated stress fluctuations (due to cyclic loading and unloading) can induce changes in
the pore fluid pressures which may lead to liquefaction.

Earthquake shaking may change the applicable shear strength properties.

The detailed discussion of the effects of earthquakes on slope stability is outside the scope of
this section. It is clear however that earthquakes do influence slope stability and can lead to
slope failures on a catastrophic scale (see, for example, Seed).

Most of the analytical work concerned with the stability of slopes subjected to earthquakes
has been performed in connection with earth and rock-fill dams where there is great potential
for damage and loss of life if failure occurs. Much of this work is also applicable to natural
slopes; an excellent review of the factors involved in determining the stability of slopes under
earthquake shock is given by Seed.

Seed, H.B. 1968. Landslides during earthquakes due to soil liquefaction. A.S.C.E., Journal of
the Soil Mechanics and Foundations Division, 94, 5, 1055-1122.
Seed, H.B. 1979. Considerations in the earthquake - resistant design of earth dams.
Geotechnique, 29, 3, 215-263.

back to Earthquakes and slope stability

Stability analysis including seismic effects

172
The simplest way of including seismic effects is to perform a limit equilibrium analysis where
the forces induced by the earthquake accelerations are treated a horizontal force. Vertical
forces may also be caused by the earthquake but these are ignored in this simple form of
analysis. The principle of this method is illustrated opposite, where a horizontal force Fh due
to the earthquake is assumed to act through the centre of gravity of the soil involved in the
predicted or actual failure. It is assumed that:

Fh = kw = k mg
where m is the mass of the soil.

Thus the seismic coefficient k is a measure of the acceleration of the earthquake in terms of g.

For a purely cohesive soil, the factor of safety Fs is given by

where su is the undrained strength of the soil.

This approach may also be used in conjunction with the method of slices for soils possessing
cohesion and friction. For example, Sarma presented a procedure where the earthquake
induced forces are considered by applying a horizontal force kWi to the slice where Wi is the
weight of slice i. The method then involves the computation of the critical horizontal
acceleration (i.e. value of k) required to bring the soil above the slip surface into a state of
limiting equilibrium: this critical acceleration can then be used as an index of stability.

An alternative method of analysis, using seismic coefficients has also been presented by
Sarma.

Although the pseudo-static methods of analysis have many limitations, they are widely used.
One of the problems occurs in estimating the value of the seismic coefficient k to be used in
the analysis. This coefficient depends on the accelerations caused by the earthquake and may
be difficult to define.

Recently, dynamic methods of analysis have been developed which allow for the inclusion of
the full time history of the earthquake acceleration. Such methods generally allow the
displacements of a potential slide mass along an assumed failure surface to be estimated. A
dynamic finite element approach has been described by Seed.

Sarma, S.K., 1973. Stability analysis of embankments and slopes. Geotechnique, 23, 3, 423-
433.
173
Sarma, S.K. 1979. Stability analysis of embankments and slopes. A.S.C.E., Journal of the
Geotechynical Engineering Division, 105, 12, 1511-1524.
Seed, H.B. 1966. A method for the earthquake resistant design of earth dams. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, 92, 1, 13-41.

back to Earthquakes and slope stability

Shear strength parameters in seismic analysis


Generally, in the pseudo-static methods of analysis, the shear strength parameters used are
those measured in conventional shear strength tests. This assumption appears to be justified
by the relatively few examples where problems have arisen.

However, in cases where seismic loading may cause movement along an existing
discontinuity (such as a fault or old slip surface), significant decreases in the value of f ' have
been reported by Skempton.
Skempton, A.W. 1985. Residual strength of clays in landslides, folded strata and the
laboratory. Geotechnique, 35, 1, 3-18.

back to Earthquakes

Earthquake mechanism
Elastic strain energy may accumulate in rocks anywhere in the Earth's crust. Currently the
plate tectonic model is used to explain that the most seismically active regions are at the
'active' margins of plates. Thus, many earthquake centres are located along two belts of the
earth's surface: one belt extends around the coastal regions of the Pacific, from the East Indies
through the Phillipines, Japan, the Aleution Islands and down the western coasts of North and
South America; the other runs from Central Europe through the eastern Mediterranean to the
Himalayas and the East Indies, where it joins the first belt (see figure opposite). Both of these
belts are in places parallel to relatively young fold-mountain chains (e.g. the Andes), where
much faulting is associated with the crumpled rocks. Many volcanoes are also situated along
the active earthquake belts. Many shocks also occur in zones of submarine fault activity such
as the mid-Atlantic ridge, and in fault-zones on continents such as the Rift Valley system of

174
Africa.

back to Earthquakes

Mercalli scale of earthquake intensity


The intensity of an earthquake is estimated from the effects produced in the affected area. The
most common scale of intensity used is the Mercalli Scale, which has twelve grades as
follows:

i) Detected only by instruments

ii) Felt by persons at rest

iii) Felt indoors; vibration like passing of a truck

iv) Hanging objects swing; windows, door rattle

v) Felt outdoors; some objects displaced; pendulum clocks


stop

vi) Strong: felt by all, many frightened; weak masonry


cracked

vii) Damage to some buildings; weak chimney fall; waves on

175
ponds

viii) Destructive: much damage to buildings, ground cracks,


flow of springs affected

ix) General damage (including reservoirs), buried pipes


broken

x) Disastrous: framed buildings destroyed, rails bent,


landslides

xi) Few structures left standing, fissures opened in ground

xii) Catastrophic: damage nearly total, ground twisted and


warped.

A more detailed description of the Mercalli scale, and of other ways for quantifying
earthquakes is given by Skipp and Emreaseys.

Skipp, B.O. and Emreaseys, N.N., 1987. Engineering seismology, in Ground Engineer's
Reference Book, ed. by F.C. Bell, Butterworths, London.

back to Earthquakes

Further reading
Close, U. and McCormick, D. 1922. Where the mountain walked. Nat. Geograph. Mag., 41,
445-464.

Hutchinson, J.N. and DEL PRETE, M. 1985. Landslides at Calitri, southern Apennines
reactivated by the earthquake of 23rd November 1980. Geologia Applicata e Idrogeologia, 20,
9-38.

Murphy, W. 1995. The geomorphological controls on seismically triggered landslides during


the 1908 Straits of Messina earthquake, Southern Italy. Quart. Journal of Eng. Geol. 28, 61-
74.

Pain, C.F. 1972. Characteristics and geomorphic effects of earthquake-initiated landslides in


the Adelbert Range, Papua New Guinea. Eng. Geol,. 6, 261-274

176
Sarma, S.K., 1973. Stability analysis of embankments and slopes. Geotechnique, 23, 3, 423-
433.

Sarma, S.K., 1975. Seismic stability of earth dams and embankments. Geotechnique, 25, 4,
743-761.

Sarma, S.K. 1979. Stability analysis of embankments and slopes. A.S.C.E., Journal of the
Geotechynical Engineering Division, 105, 12, 1511-1524.

Seed, H.B. 1966. A method for the earthquake resistant design of earth dams. A.S.C.E.,
Journal of the Soil Mechanics and Foundations Division, 92, 1, 13-41.

Seed, H.B. 1967. Slope stability during earthquakes. A.S.C.E., Journal of the Soil Mechanics
and Foundations Division, 93, 4, 299-323.

Seed, H.B. 1968. Landslides during earthquakes due to soil liquefaction. A.S.C.E., Journal of
the Soil Mechanics and Foundations Division, 94, 5, 1055-1122.

Seed, H.B. 1979. Considerations in the earthquake - resistant design of earth dams.
Geotechnique, 29, 3, 215-263.

Skempton, A.W. 1985. Residual strength of clays in landslides, folded strata and the
laboratory. Geotechnique, 35, 1, 3-18.

Skipp, B.O. and Emreaseys, N.N., 1987. Engineering seismology, in Ground Engineer's
Reference Book, ed. by F.C. Bell, Butterworths, London.

Walker, B. and Fell, R., 1987. Soil slope instability and stabilisation. Belkema, Rotterdam.
440 pp.

back to Mechanics

Peak, critical state and residual strength


 Typical results for a drained test on clay
 Typical values of strength parameters
 The significance of residual strengths
 Measurement of residual shear strength
 Further reading

When a soil is subjected to shear, an increasing resistance is built up. For any given applied
effective pressure, there is a limit to the resistance that the soil can offer, which is known as
the peak shear strength p. Frequently the test is stopped immediately after the peak strength
has been clearly defined. The value p has been referred to, in the past, as simply the shear
strength of the clay, under the given effective pressure and under drained conditions.

177
If the shearing is continued beyond the point where the maximum value of the shear strength
has been mobilised, it is found that the resistance of the clay decreases, until ultimately a
steady value is reached, and this constant minimum value is known as the residual strength r
of the soil. The soil maintains this steady value even when subjected to very large
displacements.

In the absence of pre-existing failures the choice is between the peak or critical state strength.
In uncemented soils the peak strength is associated with dilation and occurs at relatively small
strains or displacements of the order of 1% or 1mm. The critical state strength is the shearing
resistance for constant volume straining and occurs at strains or displacements of the order of
10% or 10mm. In many slopes ground movements and strains are relatively large and exceed
the small movements required to mobilise the peak state. In addition, there is evidence that the
peak value reflects the nature of the laboratory test procedure and gives an unconservative
result in slope stability analysis.

back to Peak and residual strength

Typical results for a drained test on clay

The figure shows a typical plot for a shear test which has been taken to displacements large
enough to mobilise the residual strength. The decrease in shear strength from the peak to the
residual condition is associated with orientation of the clay particles along shear planes.

Further tests could be made on the same clay but under differing effective pressures. The
results previously described would again be obtained, and from a number of tests it would be
noticed that the peak and residual shear strengths would define envelopes in accordance with
the Coulomb-Terzaghi relationship, as shown. Thus, the peak strengths can be expressed as:
p =c' + ' tan'
and the residual strengths can be expressed as:
r = c'r +  tan'r
 
For critical state analysis the strength parameter su or 'c should be used.

178
p = peak shear strength
r = residual shear strength
c' = apparent cohesion
c'r = residual apparent cohesion
' = angle of shearing resistance
'r = residual angle of shearing resistance
' = applied effective pressure

back to Peak and residual strength

Typical values of strength parameters


 
Peak shear strength Residual shear strength Critical state

Soil c' f' c'r f'r f'c


 
kN/m² degrees kN/m² degrees degrees

London clay, 35 20 0 13 20
brown

London clay, blue 25 23 0 14 22

Upper Lias clay 20 24 0 15 22

Cucarasha shale 30 23 0   8 22

Upper Siwalik, 40 22 0 18 20
Jari

Boulder clay, 10 32 0 30 32
Selset

back to Peak and residual strength

The significance of residual strengths


The residual shear strength condition is of considerable practical importance since, if the soil
in situ already contains slip planes or shear surfaces, then the strength operable on these

179
surfaces will be less than the peak strength, and if sufficient displacement has taken place, the
strength may be as low as the residual strength.

There are a number of circumstances, as a result of which shearing of the soil may already
have taken place, and the principal processes, summarised by Morgenstern et al., are:

landsliding,
tectonic folding,
valley rebound,
glacial shove,
periglacial phenomena and
non-uniform swelling.

The identification of the existence of shear surfaces is a problem of great importance during
any site investigation, particularly where mass movements are involved.

back to Peak and residual strength

Measurement of residual shear strength


 Ring shear

It is generally accepted that the residual shear strength of a soil is independent of stress
history effects, not influenced by specimen size, and rate-dependent to only a small extent
unless very rapid rates of shearing are used. The major difficulty in determining the residual
shear strength lies in the fact that large displacements may be necessary to achieve the
required degree of orientation of the particles.

The methods of measuring residual shear strength in the laboratory are given below. The most
satisfactory methods, in many ways, are to obtain undisturbed samples which contain a
natural slip surface and then test them either in the shear box or triaxial apparatus so that
failure occurs by sliding along the existing slip plane. Alternatively, an artificial slip plane can
be produced by cutting the specimen with a thin wire-saw. Much of the early work on
determining the residual shear strength of soils in the laboratory was performed using multi-
reversal type tests in the shear box on previously un-sheared material.

The results of tests to measure residual shear strength in the shear box and triaxial apparatus
have been reported by Skempton and Petley. There are practical difficulties with each of these
tests, and they also have the major disadvantage that none of them permits the complete
shear-stress-displacement relationship to be obtained.

Methods for measuring residual shear strength

Shear box
(a) Tests on natural shear surfaces
(b) Reversal-type tests
(c) Cut-plane tests
180
Triaxial
(a) Tests on natural shear surfaces
(b) Cut-plane tests

Ring shear

back to Measurement of residual shear strength

Ring shear

The large displacements required to define the complete shear-stress-displacement


relationship can be obtained by using the ring-shear (or torsional shear) apparatus. The
apparatus, shown diagrammatically, consists of two pairs of metal rings which hold an
annular sample. The sample is subjected to a normal stress and then one pair of rings
(normally the lower pair) is subjected to rotation. It is therefore a form of direct shear test, and
failure occurs along a predetermined plane, as with the shear box. This type of apparatus was
probably first used by Hvorslev and Tiedemann. More recent designs of the ring-shear
apparatus have been described by Bishop et al. and Bromhead.
# Hvorslev, M.J. (1973). Über die Festigkeitseigenschaften gestörter bindiger Böden.
Ingenior Skriftor A., Copenhagen, 45.
# Tiedemann, B. (1937). Über die Schubfestigkeit bindiger Böden. Bautechnik, 15,
# Bishop, A.W., Green, G.R., Garga, V.K., Andresen, A., and Brown, J.D. (1971). A new
ring-shear apparatus and its application to the measurement of residual strength.
Geotechnique, 21, 273-328.
# Bromhead, E.N. (1979). A simple ring shear apparatus. Ground Eng., 12, 40-44.

back to Peak and residual strength

181
Further reading
Bishop, A.W., Green, G.R., Garga, V.K., Andresen, A., and Brown, J.D. (1971). A new ring-
shear apparatus and its application to the measurement of residual strength. Geotechnique, 21,
273-328.

Bromhead, E.N. (1979). A simple ring shear apparatus. Ground Eng., 12, 40-44.

Hvorslev, M.J. (1973). Über die Festigkeitseigenschaften gestörter bindiger Böden. Ingenior
Skriftor A., Copenhagen, 45.

Morgenstern, N.R., Blight, G.R., Janbu, N., and Resendiz, D. (1977). Slopes and excavations,
9th Int. Conf. Soil Mech. and Found. Eng., 12, 547-604.

Petley, D.J. (1984). Ground investigation, sampling and testing for studies sof slope
instability. In Slope Instability, edited by D. Brunsden and D.B. Prior, John Wiley and Sons
Ltd., Chichester.

Skempton, A.W. (1964). Long-term stability of clay slopes. Geotechnique, 14, 75-102.

Skempton, A.W., and Hutchinson, J.N. (1969). Stability of natural slopes and embankment
foundations, State-of-the-Art Report. 7th Int. Conf. Soil Mech. Found. Eng., Mexico, 291-
335.

Skempton, A.W., and Petley, D.J. (1967). The strength along structural discontinuities in stiff
clay. Proc. Geot. Conf. on Shear Strength of Natural Soils and Rocks. Oslo, 2, 3-20.

Terzaghi, K. (1943). Theoretical Soil Mechanics, Wiley, New York.

Tiedemann, B. (1937). Über die Schubfestigkeit bindiger Böden. Bautechnik, 15,

back to Mechanics

Stress changes in slopes


 Changes in stress during undrained slope excavation
 Comparison of cuttings with embankments
 General case

Any sudden changes in loading on a slope will lead to changes in the pore pressures. Natural
slopes are usually eroded very slowly and the soil is essentially drained throughout the
process. This means that the pore pressures are governed by steady seepage from the ground
towards the excavation and their magnitude can be obtained from a flownet. Man-made slopes
182
are constructed relatively quickly, and in soils with low permeability, such as clay, there will
be inadequate time during the construction period for the pore pressures to adjust to the new
loading conditions, and the soil will then be essentially undrained.

back to Stress changes in slopes

Changes in stress during undrained slope excavation


The changes in total and effective stress during undrained slope excavation are shown below.
In this simple example, the excavation is assumed to be kept full of water so that the initial
and final pore pressures are identical.

In (a) the total stresses on an element on a slip surface are  and , and the pore pressure is
indicated by the height of water in the standpipe. Obviously, before any excavation, the water
in the standpipe is level with the phreatic surface, and the initial pore pressure can be
represented by uo. This initial stress state is represented on (b) by the points A (in terms of
total stress) and A' (in terms of effective stress).

As excavation proceeds, the value of  decreases and  increases due to the increase in slope
height and/or angle. The total stress path is AB. The effective stress path is A'B',
corresponding to undrained loading at constant water content, as indicated in (c).

The precise stress path A'B' in (c) will depend upon the characteristics of the soil and whether
the soil is normally or over consolidated.

Since the process of excavation leads to reductions in loading, the pore pressure immediately
after construction (u) is less than the initial pore pressure (uo), and so the initial excess pore
pressure is negative (i.e. the level of the water in the standpipe is below the phreatic surface as
shown in (d). As time passes, the total stress remains unchanged at B (because there is no
change in the geometry of the slope). The negative excess pore pressures dissipate, leading to
an increase in pore pressure. The normal effective stresses on the failure surface will decrease
and swelling takes place as indicated in (b) and (c). The final state at C' corresponds to a

183
steady state pore pressure after swelling has been
completed. In this simple example, the excavation is kept
full of water, and thus the initial and final pore pressures are
equal (i.e. uo = uc).

Failure of the slope will take place if the states of all


elements along the slip surface reach the failure line. If B'
reaches the failure line, then the slope fails during the
course of undrained excavation (i.e. during construction); whilst if C' reaches the failure line,
it does so some time after construction has been completed. The distance of the points B' and
C' from the failure line is a measure of the factor of safety of the slope.

These changes in the values of total stress and pore pressure with time are shown here.

For the case considered above, the initial (uo) and final (uc) values of pore pressure are
identical, but in general, this may not be the case.

For cutting slopes, the effective stress reduces with time, leading to leass shear resistance and
a lower factor of safety. Consequently the critical time in the life of these slopes is likely to be
in the long-term when the pore pressures have come into equilibrium with the steady seepage
flownet.

back to Stress changes in slopes

Comparison of cuttings with embankments

It is of interest to compare the behaviour of cutting slopes with that of embankments. In the
case of embankments, the construction process leads to an increase in pore pressure in the
foundation soil, which then dissipates with time. Thus one of the critical stability conditions is
likely to be at the end of construction, and will involve failure through the foundation. The
184
relationship between total stress, pore pressure and time for this condition is indicated in the
figure. In the embankment material itself, since the fill is unsaturated, the initial pore
pressures are negative, and the initial states at B and B', are similar to the cut slopes.

back to Stress changes in slopes

General case
The more general case is illustrated below.

Here the long-term pore pressures are controlled by the steady state flownet: the general
principles, however, remain unchanged.

back to Mechanics

Choice of strength parameters


The choice between the undrained strength su and the drained strength is relatively simple and
straightforward. For temporary slopes and cuts in fine-grained soils with low permeability the
undrained strength su should be used and a total stress analysis performed. This analysis is
only valid whilst the soil is undrained. The stability will deteriorate with time as the pore
pressures increase and the soil swells and softens.

For any permanent slope the critical conditions are at the end of swelling when pore pressures
have reached equilibrium with a steady state seepage flownet or with hydrostatic conditions.
In this case an effective stress strength is appropriate and the pore pressures are calculated
separately.

185
back to Mechanics

Choice of factor of safety


The factor of safety should take account of uncertainties in the determinations of the loads
(including the unit weight), the soil strengths and particularly the pore pressure or drainage
conditions and the consequences of failure. The greatest uncertainty is in the determination of
steady state pore pressures in drained analyses or in the assumption of constant volume (and
hence constant strength) in undrained analyses. There is no single value of Fs that can be
recommended for slope stability calculations. Typical values are often in the range 1.25 to
1.35, but can fall outside this range. The table below gives a range of suggested methods of
analyses and typical factors of safety.

 
Typical Fs range

Temporary cuttings and embankments: 1.1 - 1.3


using undrained strength su and total
stresses

Permanent cuttings: using critical 1.2 - 1.4


strength 'c and effective stresses

Embankment foundation: undrained su 1.2 - 1.5


or drained '

Embankment fill: drained ' for 1.2 - 1.4


compacted soil and effective stresses

Reactivated landslip: residual strength 'r (natural value)

Slides
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

186
A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:

187
Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

Analysis of translational slip

Drained soil with parallel flow


In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
maximum slope angle (max) has a magnitude significantly lower than the angle of friction
(').

s = effective unit weight of soil


s = unit weight of saturated soil

back to Analysis of translational slip

General equation
 Derivation

188
A translational slip analysis may be used for slip surfaces
with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety


against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
s = unit weight of saturated soil

back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:

Ignoring any side forces acting on the elements, the stress conditions will be identical at every
point along the slip surface. Therefore the maximum shear strength is given by the Mohr-
Coulomb equation:

and the stresses are determined by resolving the load due to an element.

189
back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip


plane and converting to a stress, gives

back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore water pressure at the slip surface is

Analysis of translational slip

Stability of vertical cuts


 Derivation

It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
190
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

Stability of vertical cuts

Derivation: stability of vertical cuts


Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)
The theoretical value of h is (2su/)
and then H = (2su/)

191
 

back to Analysis of translational slip

Translational slip in rock slopes


 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the spatial arrangement of the
discontinuities within the mass and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although others may exist as 'release
surfaces'.

Wedge failure
In which slip occurs on two discontinuities and is governed by their line of intersection.

The engineer needs some means of graphically representing these discontinuities, if he or she
is to be able to spot potential failure mechanisms. One graphical method uses stereonets to
analyse the spatial arrangement of the planar discontinuities and slope surface.

The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434

192
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

back to Translational slip in rock slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

Possible plane failure is suggested by a stereonet plot, if a pole concentration lies close to the
pole of the slope surface and in the shaded area corresponding to the above rules.

back to Translational slip in rock slopes

Wedge
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

On the stereonet plot these conditions are indicated by the intersection of two discontinuity
great circles within the shaded crescent formed by the friction angle and the slope's great
circle. Note that this intersection can also be located by finding the pole P12 of the great circle
which passes through the pole concentrations P1 and P2.

back to Analysis of translational slip

Further reading: translational

193
Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural discontinuities in
stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

Weeks, A.G., 1969. "The stability of natural slopes in south-east England as affected by
periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

back to Slope stability

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational
194
The shear strength of the soil is a function of the normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
 Movement along geological boundary

back to Analysis of rotational slip

Ordinary Method of slices


 Derivation of Ordinary Method

This method is also refered to as "Fellenius' Method" and the "Swedish Circle Method".
Consider the geometry of the trial slip circle shown in the diagram. The slipping mass is
divided into slices in the normal way and these are numbered 1, 2, 3, etc, for ease of
identification. For a potential slip circle, the factor of safety against slip is given by:

This solution tends to give a conservative value for the factor of safety, of between 5 and
20%. This can be expensive and therefore a more rigorous approach is favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

195
Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

back to Ordinary method of slices

Derivation of Ordinary Method of slices


 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and therefore
the problem is statically indeterminate. Even so, this problem can be overcome by simply
ignoring the effects of the side forces, i.e.

Then substitute for


xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

back to Derivation of ordinary method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

196
n = c' + (n - un) tan'

or, for a unit thickness:

n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to the slip surface, we can
show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin n].tan'

back to Analysis of rotational slip

Bishop rigorous method


 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and gives a more
accurate solution to the idealised geometry of circular slip. The importance of such forces can
be demonstrated by considering a slice directly below the center of rotation: this slice has an
independent safety factor of infinity since  = 0 at that point. Bishop's solution requires that
the factor of safety is constant along the complete slip circle. For a potential slip circle, the
factor of safety against slip is given by:

or

197
However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

back to Bishop rigorous method

Derivation of Bishop rigorous method


 Determining the shear force for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

back to Derivation of bishop rigorous method


198
Determining the shear force for each slice
From Coulomb's equation, the shear strength mobilised along the failure surface for slice
n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n /Fs) tan' ) sin n

giving,

and thus,

back to Analysis of rotational slip

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil

Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

199
The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

Solving the Bishop simplified equation

Example of table calculations


   
1 2 3 ...
n

200
 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

area = bn .zn = approximated area of slice

201
Wn = weight of the slice for a unit thickness, calculated by multiplying the
slice area by the unit weight of the soil
n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

exp(2) = [1+ (tan' tan n) / Fs] / sec n

back to Bishop simplified method

Short term analysis for an undrained soil


 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we consider the case of
a cohesive soil and analyse its stability in the short term, we can substitute the total stress soil
parameters into the Bishop simplified equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

202
R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Short term analysis for an undrained soil

Locating the centre of the slip circle


Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

1:5 11 25 37

For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

back to Analysis of rotational slip


203
Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope
and hence no shear strength can occur over that length. The angle  must be
reduced accordingly (see diagram). Furthermore, water in the crack will supply an additional
hydrostatic force, acting to reduce the factor of safety. This can be incorporated into the
analysis by treating it as an additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

Analysis of rotational slip

Methods of analysis
 Limit equilibrium procedure

In practice, limit equilibrium methods of analysis are generally adopted, in which it is


considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.
204
In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.

For undrained loading, the ultimate strength of the soil is given by


 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

Methods of analysis

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.

back to Analysis of rotational slip

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

205
Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  = 0 analysis of the
Stability of clays". Proc. 2nd Int. Conf. Soil Mechs (Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Slides
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


206
 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:

Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

207
Analysis of translational slip

Drained soil with parallel flow


In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
maximum slope angle (max) has a magnitude significantly lower than the angle of friction
(').

s = effective unit weight of soil


s = unit weight of saturated soil

back to Analysis of translational slip

General equation
 Derivation

A translational slip analysis may be used for slip surfaces with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
208
s = unit weight of saturated soil

back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:

Ignoring any side forces acting on the elements, the stress conditions will be identical at every
point along the slip surface. Therefore the maximum shear strength is given by the Mohr-
Coulomb equation:

and the stresses are determined by resolving the load due to an element.

back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip plane and converting to a
stress, gives

209
 

back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore water pressure at the slip surface is

Analysis of translational slip

Stability of vertical cuts


 Derivation

It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

Stability of vertical cuts

Derivation: stability of vertical cuts

210
Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)
The theoretical value of h is (2su/)
and then H = (2su/)

211
back to Analysis of translational slip

Translational slip in rock slopes


 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the


spatial arrangement of the discontinuities within the mass
and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although
others may exist as 'release surfaces'.

Wedge failure
In which slip occurs on two discontinuities and is governed
by their line of intersection.

The engineer needs some means of graphically representing


these discontinuities, if he or she is to be able to spot
potential failure mechanisms. One graphical method uses
stereonets to analyse the spatial arrangement of the planar
discontinuities and slope surface.

The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

back to Translational slip in rock slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

212
Possible plane failure is suggested by a stereonet plot, if a pole concentration lies
close to the pole of the slope surface and in the shaded area corresponding to the
above rules.

back to Translational slip in rock slopes

Wedge
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

On the stereonet plot these conditions are indicated by the intersection of two discontinuity
great circles within the shaded crescent formed by the friction angle and the slope's great
circle. Note that this intersection can also be located by finding the pole P12 of the great circle
which passes through the pole concentrations P1 and P2.

back to Analysis of translational slip

Further reading: translational


Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

213
Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural
discontinuities in stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

Weeks, A.G., 1969. "The stability of natural slopes in south-east England as


affected by periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

back to Slope stability

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational

The shear strength of the soil is a function of the normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
214
 Movement along geological boundary

back to Analysis of rotational slip

Ordinary Method of slices


 Derivation of Ordinary Method

This method is also refered to as "Fellenius' Method" and the "Swedish Circle
Method". Consider the geometry of the trial slip circle shown in the diagram.
The slipping mass is divided into slices in the normal way and these are
numbered 1, 2, 3, etc, for ease of identification. For a potential slip circle, the
factor of safety against slip is given by:

This solution tends to give a conservative value for the factor of safety, of between 5 and
20%. This can be expensive and therefore a more rigorous approach is favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

back to Ordinary method of slices

Derivation of Ordinary Method of slices


 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:
215
Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and
therefore the problem is statically indeterminate. Even so, this problem can be overcome
by simply ignoring the effects of the side forces, i.e.

Then substitute for


xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

back to Derivation of ordinary method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

or, for a unit thickness:

n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to the slip surface, we can
show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin n].tan'

216
 

back to Analysis of rotational slip

Bishop rigorous method


 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and
gives a more accurate solution to the idealised geometry of circular slip. The
importance of such forces can be demonstrated by considering a slice directly
below the center of rotation: this slice has an independent safety factor of infinity since  = 0
at that point. Bishop's solution requires that the factor of safety is constant along the complete
slip circle. For a potential slip circle, the factor of safety against slip is given by:

or

However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

217
Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

back to Bishop rigorous method

Derivation of Bishop rigorous method


 Determining the shear force for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

back to Derivation of bishop rigorous method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n /Fs) tan' ) sin n

giving,
218
and thus,

back to Analysis of rotational slip

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil

Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

219
Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

Solving the Bishop simplified equation

Example of table calculations


   
1 2 3 ...
n

 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...

220
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

area = bn .zn = approximated area of slice

Wn = weight of the slice for a unit thickness, calculated by multiplying the slice area by the
unit weight of the soil

n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

exp(2) = [1+ (tan' tan n) / Fs] / sec n

back to Bishop simplified method

Short term analysis for an undrained soil


221
 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we


consider the case of a cohesive soil and analyse its stability in the short term,
we can substitute the total stress soil parameters into the Bishop simplified
equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Short term analysis for an undrained soil

Locating the centre of the slip circle


Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

222
slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

1:5 11 25 37

For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

back to Analysis of rotational slip

Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope and hence no
shear strength can occur over that length. The angle  must be reduced accordingly (see
diagram). Furthermore, water in the crack will supply an additional hydrostatic force, acting
to reduce the factor of safety. This can be incorporated into the analysis by treating it as an
additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

223
hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

Analysis of rotational slip

Methods of analysis
 Limit equilibrium procedure

In practice, limit equilibrium methods of analysis are generally adopted, in which it is


considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.

In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.

For undrained loading, the ultimate strength of the soil is given by


 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

224
Methods of analysis

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.

back to Analysis of rotational slip

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

225
Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  =
0 analysis of the Stability of clays". Proc. 2nd Int. Conf. Soil Mechs
(Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes


and embankment foundations". 7th Int. Conf. Soil Mech. and Found.
Engrg. (Mexico), State-of-the-Art Vol., 291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Slides
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


226
 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:

Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

Analysis of translational slip

Drained soil with parallel flow


In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
227
maximum slope angle (max) has a magnitude significantly
lower than the angle of friction (').

s = effective unit weight of soil


s = unit weight of saturated soil

back to Analysis of translational slip

General equation
 Derivation

A translational slip analysis may be used for slip surfaces with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
s = unit weight of saturated soil

back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:
228
Ignoring any side forces acting on the elements, the stress conditions
will be identical at every point along the slip surface. Therefore the
maximum shear strength is given by the Mohr-Coulomb equation:

and the stresses are determined by resolving the load due to an element.

back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip plane and converting to a
stress, gives

back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore water pressure at the slip surface is

229
Analysis of translational slip

Stability of vertical cuts


 Derivation

It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

Stability of vertical cuts

Derivation: stability of vertical cuts


Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
230
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W
and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)
The theoretical value of h is (2su/)
and then H = (2su/)

back to Analysis of translational slip

Translational slip in rock slopes


 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the spatial arrangement of the
discontinuities within the mass and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although others may exist as 'release
surfaces'.

231
Wedge failure
In which slip occurs on two discontinuities and is governed by their line of
intersection.

The engineer needs some means of graphically representing these discontinuities, if


he or she is to be able to spot potential failure mechanisms. One graphical method
uses stereonets to analyse the spatial arrangement of the planar discontinuities and
slope surface.

The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

back to Translational slip in rock slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

Possible plane failure is suggested by a stereonet plot, if a pole concentration lies close
to the pole of the slope surface and in the shaded area corresponding to the above
rules.

back to Translational slip in rock slopes

Wedge
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

232
On the stereonet plot these conditions are indicated by the intersection of two
discontinuity great circles within the shaded crescent formed by the friction angle and
the slope's great circle. Note that this intersection can also be located by finding the
pole P12 of the great circle which passes through the pole concentrations P1 and P2.

back to Analysis of translational slip

Further reading: translational


Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural discontinuities in
stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

Weeks, A.G., 1969. "The stability of natural slopes in south-east England as affected by
periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

 
233
back to Slope stability

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational

The shear strength of the soil is a function of the normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
 Movement along geological boundary

back to Analysis of rotational slip

Ordinary Method of slices


 Derivation of Ordinary Method

This method is also refered to as "Fellenius' Method" and the "Swedish Circle Method".
Consider the geometry of the trial slip circle shown in the diagram. The slipping mass is
divided into slices in the normal way and these are numbered 1, 2, 3, etc, for ease of
identification. For a potential slip circle, the factor of safety against slip is given by:

234
This solution tends to give a conservative value for the factor of safety, of
between 5 and 20%. This can be expensive and therefore a more rigorous approach is
favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

back to Ordinary method of slices

Derivation of Ordinary Method of slices


 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and therefore
the problem is statically indeterminate. Even so, this problem can be overcome by simply
ignoring the effects of the side forces, i.e.

235
Then substitute for
xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

back to Derivation of ordinary method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

or, for a unit thickness:

n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to the slip surface, we can
show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin


n].tan'

back to Analysis of rotational slip

Bishop rigorous method


 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and gives a more
accurate solution to the idealised geometry of circular slip. The importance of such forces can
be demonstrated by considering a slice directly below the center of rotation: this slice has an
independent safety factor of infinity since  = 0 at that point. Bishop's solution requires that
236
the factor of safety is constant along the complete slip circle. For a potential slip circle, the
factor of safety against slip is given by:

or

However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

back to Bishop rigorous method

Derivation of Bishop rigorous method


 Determining the shear force for each slice

237
The factor of safety against rotational slip is defined as the ratio of the
restraining moments to the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

back to Derivation of bishop rigorous method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n /Fs) tan' ) sin n

giving,

and thus,

238
back to Analysis of rotational slip

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil

Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

 
239
Solving the Bishop simplified equation

Example of table calculations


   
1 2 3 ...
n

 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

240
area = bn .zn = approximated area of slice

Wn = weight of the slice for a unit thickness, calculated by multiplying the slice area by the
unit weight of the soil

n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

exp(2) = [1+ (tan' tan n) / Fs] / sec n

back to Bishop simplified method

Short term analysis for an undrained soil


 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we consider the case of
a cohesive soil and analyse its stability in the short term, we can substitute the total stress soil
parameters into the Bishop simplified equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

241
Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Short term analysis for an undrained soil

Locating the centre of the slip circle


Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

242
1:5 11 25 37

For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

back to Analysis of rotational slip

Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope and hence no
shear strength can occur over that length. The angle  must be reduced accordingly (see
diagram). Furthermore, water in the crack will supply an additional hydrostatic force, acting
to reduce the factor of safety. This can be incorporated into the analysis by treating it as an
additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

Analysis of rotational slip

Methods of analysis
 Limit equilibrium procedure

243
In practice, limit equilibrium methods of analysis are generally adopted, in which it is
considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.

In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.

For undrained loading, the ultimate strength of the soil is given by


 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

Methods of analysis

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.
244
 

back to Analysis of rotational slip

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  = 0 analysis of the
Stability of clays". Proc. 2nd Int. Conf. Soil Mechs (Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Slides
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

245
A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:

246
Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

Analysis of translational slip

Drained soil with parallel flow


In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
maximum slope angle (max) has a magnitude significantly lower than the angle of friction
(').

s = effective unit weight of soil


s = unit weight of saturated soil

247
back to Analysis of translational slip

General equation
 Derivation

A translational slip analysis may be used for slip surfaces with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
s = unit weight of saturated soil

back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:

Ignoring any side forces acting on the elements, the stress conditions will be identical at every
point along the slip surface. Therefore the maximum shear strength is given by the Mohr-
Coulomb equation:

248
and the stresses are determined by resolving the load due to an element.

back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip plane and converting to a
stress, gives

249
back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore water pressure at the slip surface is

 Analysis of translational slip

Stability of vertical cuts


 Derivation

It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

Stability of vertical cuts

Derivation: stability of vertical cuts


Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

250
Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)
The theoretical value of h is (2su/)
and then H = (2su/)

back to Analysis of translational slip

Translational slip in rock slopes


251
 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the spatial arrangement of the
discontinuities within the mass and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although others may exist as 'release
surfaces'.

Wedge failure
In which slip occurs on two discontinuities and is governed by their line of intersection.

The engineer needs some means of graphically representing these discontinuities, if he or she
is to be able to spot potential failure mechanisms. One graphical method uses stereonets to
analyse the spatial arrangement of the planar discontinuities and slope surface.

The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

 
252
back to Translational slip in rock slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

Possible plane failure is suggested by a stereonet plot, if a pole concentration lies close to the
pole of the slope surface and in the shaded area corresponding to the above rules.

back to Translational slip in rock slopes

Wedge
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

On the stereonet plot these conditions are indicated by the intersection of two discontinuity
great circles within the shaded crescent formed by the friction angle and the slope's great
circle. Note that this intersection can also be located by finding the pole P 12 of the great circle
which passes through the pole concentrations P1 and P2.

253
back to Analysis of translational slip

Further reading: translational


Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural discontinuities in
stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

Weeks, A.G., 1969. "The stability of natural slopes in south-east England as affected by
periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

254
back to Slope stability

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational

The shear strength of the soil is a function of the normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
 Movement along geological boundary

back to Analysis of rotational slip

Ordinary Method of slices


 Derivation of Ordinary Method

255
This method is also refered to as "Fellenius' Method" and the "Swedish Circle Method".
Consider the geometry of the trial slip circle shown in the diagram. The slipping mass is
divided into slices in the normal way and these are numbered 1, 2, 3, etc, for ease of
identification. For a potential slip circle, the factor of safety against slip is given by:

This solution tends to give a conservative value for the factor of safety, of between 5 and
20%. This can be expensive and therefore a more rigorous approach is favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

back to Ordinary method of slices

Derivation of Ordinary Method of slices


 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

256
Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and therefore
the problem is statically indeterminate. Even so, this problem can be overcome by simply
ignoring the effects of the side forces, i.e.

Then substitute for


xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

back to Derivation of ordinary method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

or, for a unit thickness:

n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to the slip surface, we can
show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin n].tan'

257
 

back to Analysis of rotational slip

Bishop rigorous method


 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and gives a more
accurate solution to the idealised geometry of circular slip. The importance of such forces can
be demonstrated by considering a slice directly below the center of rotation: this slice has an
independent safety factor of infinity since  = 0 at that point. Bishop's solution requires that
the factor of safety is constant along the complete slip circle. For a potential slip circle, the
factor of safety against slip is given by:

or

However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

258
c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

back to Bishop rigorous method

Derivation of Bishop rigorous method

 Determining the shear force for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

259
back to Derivation of bishop rigorous method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n /Fs) tan' ) sin n

giving,

and thus,

back to Analysis of rotational slip

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil
260
Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

Solving the Bishop simplified equation

Example of table calculations

261
   
1 2 3 ...
n

 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

area = bn .zn = approximated area of slice

262
Wn = weight of the slice for a unit thickness, calculated by multiplying the slice area by the
unit weight of the soil

n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

exp(2) = [1+ (tan' tan n) / Fs] / sec n

back to Bishop simplified method

Short term analysis for an undrained soil


 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we consider the case of
a cohesive soil and analyse its stability in the short term, we can substitute the total stress soil
parameters into the Bishop simplified equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

263
Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Short term analysis for an undrained soil

Locating the centre of the slip circle


Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

264
1:5 11 25 37

For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

back to Analysis of rotational slip

Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope and hence no
shear strength can occur over that length. The angle  must be reduced accordingly (see
diagram). Furthermore, water in the crack will supply an additional hydrostatic force, acting
to reduce the factor of safety. This can be incorporated into the analysis by treating it as an
additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

265
Analysis of rotational slip

Methods of analysis
 Limit equilibrium procedure

In practice, limit equilibrium methods of analysis are generally adopted, in which it is


considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.

In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.

For undrained loading, the ultimate strength of the soil is given by


 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

Methods of analysis

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

266
2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.

back to Analysis of rotational slip

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  = 0 analysis of the
Stability of clays". Proc. 2nd Int. Conf. Soil Mechs (Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Slides
267
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:


268
Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

Analysis of translational slip

Drained soil with parallel flow


In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
maximum slope angle (max) has a magnitude significantly lower than the angle of friction
(').

269
s = effective unit weight of soil
s = unit weight of saturated soil

back to Analysis of translational slip

General equation
 Derivation

A translational slip analysis may be used for slip surfaces with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
s = unit weight of saturated soil

270
back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:

Ignoring any side forces acting on the elements, the stress conditions will be identical at every
point along the slip surface. Therefore the maximum shear strength is given by the Mohr-
Coulomb equation:

and the stresses are determined by resolving the load due to an element.

back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip plane and converting to a
stress, gives

 
271
back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore water pressure at the slip surface is

Analysis of translational slip

Stability of vertical cuts


 Derivation

It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

 
272
Stability of vertical cuts

Derivation: stability of vertical cuts


Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)
The theoretical value of h is (2su/)
and then H = (2su/)

273
 

back to Analysis of translational slip

Translational slip in rock slopes


 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the spatial arrangement of the
discontinuities within the mass and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although others may exist as 'release
surfaces'.

Wedge failure
In which slip occurs on two discontinuities and is governed by their line of intersection.

The engineer needs some means of graphically representing these discontinuities, if he or she
is to be able to spot potential failure mechanisms. One graphical method uses stereonets to
analyse the spatial arrangement of the planar discontinuities and slope surface.

The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

274
back to Translational slip in rock
slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

Possible plane failure is suggested by a stereonet plot, if a pole concentration lies close to the
pole of the slope surface and in the shaded area corresponding to the above rules.

back to Translational slip in rock slopes

Wedge

275
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

On the stereonet plot these conditions are indicated by the intersection of two discontinuity
great circles within the shaded crescent formed by the friction angle and the slope's great
circle. Note that this intersection can also be located by finding the pole P12 of the great circle
which passes through the pole concentrations P1 and P2.

back to Analysis of translational slip

Further reading: translational


Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural discontinuities in
stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

276
Weeks, A.G., 1969. "The stability of natural slopes in south-east England as affected by
periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

back to Slope stability

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational

The shear strength of the soil is a function of the normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
 Movement along geological boundary

back to Analysis of rotational slip

Ordinary Method of slices


 Derivation of Ordinary Method

277
This method is also refered to as "Fellenius' Method" and the "Swedish Circle Method".
Consider the geometry of the trial slip circle shown in the diagram. The slipping mass is
divided into slices in the normal way and these are numbered 1, 2, 3, etc, for ease of
identification. For a potential slip circle, the factor of safety against slip is given by:

This solution tends to give a conservative value for the factor of safety, of between 5 and
20%. This can be expensive and therefore a more rigorous approach is favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

back to Ordinary method of slices

Derivation of Ordinary Method of


slices
 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and therefore
the problem is statically indeterminate. Even so, this problem can be overcome by simply
ignoring the effects of the side forces, i.e.
278
Then substitute for
xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

back to Derivation of ordinary method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

or, for a unit thickness:

n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to


the slip surface, we can show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin n].tan'

279
back to Analysis of rotational slip

Bishop rigorous method


 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and gives a more
accurate solution to the idealised geometry of circular slip. The importance of such forces can
be demonstrated by considering a slice directly below the center of rotation: this slice has an
independent safety factor of infinity since  = 0 at that point. Bishop's solution requires that
the factor of safety is constant along the complete slip circle. For a potential slip circle, the
factor of safety against slip is given by:

or

However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

280
 

back to Bishop rigorous method

Derivation of Bishop rigorous method


 Determining the shear force for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

back to Derivation of bishop rigorous method

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n /Fs) tan' ) sin n

giving,

281
and thus,

back to Analysis of rotational slip

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil

Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

282
Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

Solving the Bishop simplified equation

Example of table calculations


   
1 2 3 ...
n

 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...

283
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

area = bn .zn = approximated area of slice

Wn = weight of the slice for a unit thickness, calculated by multiplying the slice area by the
unit weight of the soil

n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

exp(2) = [1+ (tan' tan n) / Fs] / sec n

284
back to Bishop simplified method

Short term analysis for an undrained soil


 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we consider the case of
a cohesive soil and analyse its stability in the short term, we can substitute the total stress soil
parameters into the Bishop simplified equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Short term analysis for an undrained soil

285
Locating the centre of the slip circle
Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

1:5 11 25 37

For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

back to Analysis of rotational slip

Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope and hence no
shear strength can occur over that length. The angle  must be reduced accordingly (see
diagram). Furthermore, water in the crack will supply an additional hydrostatic force, acting
to reduce the factor of safety. This can be incorporated into the analysis by treating it as an
286
additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

Analysis of rotational slip

Methods of analysis
 Limit equilibrium procedure

In practice, limit equilibrium methods of analysis are generally adopted, in which it is


considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.

In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.
287
For undrained loading, the ultimate strength of the soil is given by
 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

Methods of analysis

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.

back to Analysis of rotational slip

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

288
Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  = 0 analysis of the
Stability of clays". Proc. 2nd Int. Conf. Soil Mechs (Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Slides
Movement involves shear displacement along one or more surfaces, or within a relatively
narrow zone, which are visible or may reasonably be inferred. Two subgroups are identified
as:

A. Rotational
Where movement results from forces that cause a turning moment about a
point above the centre of gravity of the unit. The surface of rupture
concaves upwards.
B. Translational
Where movement occurs predominantly along more or less planar or gently
undulatory surfaces. Movement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded
deposits, or by the contact between firm bedrock and overlying detritus.

back to Slope stability

Analysis of translational slip


 Drained soil with zero flow
 Drained soil with parallel flow
 General equation
289
 Stability of vertical cuts
 Translational slip in rock slopes
 Further reading: translational

Translational or infinite slope movement predominantly occurs along more or less planar or
gently undulatory surfaces. Displacement is frequently, structurally controlled by
discontinuities and variations in shear strength between layers of bedded deposits, or by the
contact between firm bedrock and overlying detritus.

See also the case studies:

 Shallow slab slide


 Quick clay slide

back to Analysis of translational slip

Drained soil with zero flow


In this case, the soil cohesion is zero, and the slope is dry or fully submerged, so no ground
water seepage occurs to generate pore water pressures.
(In the general equation c' = 0 and m = 0.)

The Factor of Safety against slip reduces to:

Note that the factor of safety is independent of the mass of the soil, the length of the slip and
the depth of the slip surface. The limiting condition occurs when the slope angle () has the
same magnitude as the angle of friction (').

back to Drained soil with zero flow

Derivation: Drained soil with zero flow


 

Analysis of translational slip

Drained soil with parallel flow


290
In this case, the soil cohesion is zero, and there is flow parallel to and coincident with the
ground surface.
(In the general equation, c' = 0 and m = 1.)

The factor of safety against slip reduces to:

Note that the factor of safety is independent of the length of the slip and the depth of the slip
surface. Because, in this case, the factor of safety is dependent upon the mass of the soil, the
maximum slope angle (max) has a magnitude significantly lower than the angle of friction
(').

s = effective unit weight of soil


s = unit weight of saturated soil

back to Analysis of translational slip

General equation
 Derivation

A translational slip analysis may be used for slip surfaces with small depth / length ratios.
This allows the end effects to be neglected.

For a potential slip surface in any soil, the factor of safety against slip is given by:

1 = a unit width of the slipping mass


b = length of an element on plan (unit: m)
L = slope length of an element (unit: m)
m = ratio of zw / z
Having a range of values 0 to 1.
Zero corresponds to a dry slope or one that is submerged with no seepage.
291
u = pore water pressure due to seepage (unit: kN/m²)
W = vertical load due to the element (unit: kN)
z = vertical depth from the slope surface to the slip plane (unit: m)
zw = vertical depth from the phreatic surface to the slip plane (unit: m)
s = unit weight of saturated soil

back to General equation

Derivation of general equation


 Stresses along the slip surface
 Pore pressure

The factor of safety against slip is defined in terms of the ratio of this maximum shear
strength to the disturbing shear stress:

Ignoring any side forces acting on the elements, the stress conditions will be identical at every
point along the slip surface. Therefore the maximum shear strength is given by the Mohr-
Coulomb equation:

and the stresses are determined by resolving the load due to an element.

292
back to Derivation of ordinary equation

Stresses along the slip surface


Given that the soil is saturated below the phreatic surface,

Resolving W into components parallel and perpendicular to the slip plane and converting to a
stress, gives

back to Derivation of General Equation

Pore pressure
From a consideration of the flownet, the pore
water pressure at the slip surface is

Analysis of translational slip

Stability of vertical cuts


 Derivation

293
It is impossible to make a vertical cut in a drained soil - this is easily demonstrated by the use
of dry sand. In soils which are undrained, however, a vertical cut can be made since the
negative pore pressures set up by the unloading due to the excavation will generate positive
effective stresses.

If there is no tension crack present, the theoretical height of the cut is given by:
H = (4 su / g)
If a tension crack is anticipated, its theoretical value, h, is (2su / ), giving a maximum height
of cut as
H = (2su / )

We should note that even if H is kept smaller than these theoretical values, local over
stressing may occur near the base of the cut. As H increases towards the theoretical maximum
value, the plastic zones extend, and significant deformations will take place.
 = unit weight of soil

su = undrained strength of soil

Stability of vertical cuts

Derivation: stability of vertical cuts


Consider the vertical cut of height H shown in the figure. Assume that the soil is undrained
and the strength can be represented by:  = su.

Consider the collapse mechanism shown, where failure occurs along the plane surface AB,
inclined at  to the horizontal. BC represents a vertical tension crack of depth h. The mass of
soil represented by ABCD is in equilibrium under the action of 3 forces, namely:
W = weight of ABCD,
S = shear strength along BC,
R = normal reaction on BC.

From the triangle of forces


W = R cos + S sin
R sin = S cos
Now,
W = 0.5(H - h) (H+h) cot
and,
S = su.AB
where,
AB = (H-h) cosec
Eliminating R from these equations and substituting for W and S gives:
H = (4su/) - h
If there is no tension crack, i.e. h = 0, then,
H = (4su/)

294
The theoretical value of h is (2su/)
and then H = (2su/)

back to Analysis of translational slip

Translational slip in rock slopes


 Plane failure
 Wedge failure

Translational slides in rock masses are dependent upon the spatial arrangement of the
discontinuities within the mass and their relationship to the geometry of the slope. Two
arrangements are considered:

Plane failure
In which slip is controlled by a single discontinuity, although others may exist as 'release
surfaces'.

Wedge failure
In which slip occurs on two discontinuities and is governed by their line of intersection.

The engineer needs some means of graphically representing these discontinuities, if he or she
is to be able to spot potential failure mechanisms. One graphical method uses stereonets to
analyse the spatial arrangement of the planar discontinuities and slope surface.
295
The construction of stereonets is beyond the scope of this reference, good descriptions are
available in
Goodman, R.E., "Introduction to Rock Mechanics",
2nd ed., Wiley, 1989, pp 417-434
Hoek, E. and Bray, J., "Rock Slope Engineering",
3rd ed., The Institution of Mining & Metallurgy, 1981
Priest, S.D., "Hemispherical Projection Methods in Rock Mechanics",
George Allen & Unwin, 1985

back to Translational slip in rock slopes

Plane Failure
Plane failure occurs due to sliding along a single discontinuity. The conditions for sliding are
that:

· the strikes of both the sliding plane and the slope face lie parallel (±20°) to each other.
· the failure plane "daylights" on the slope face.
· the dip of the sliding plane is greater than '.
· the sliding mass is bound by release surfaces of negligible resistance.

Possible plane failure is suggested by a stereonet plot, if a pole concentration lies close to the
pole of the slope surface and in the shaded area corresponding to the above rules.

296
Wedge
Wedge failure occurs due to sliding along a combination of discontinuities. The conditions for
sliding require that  is overcome, and that the intersection of the discontinuities "daylights"
on the slope surface.

On the stereonet plot these conditions are indicated by the intersection of two discontinuity
great circles within the shaded crescent formed by the friction angle and the slope's great
circle. Note that this intersection can also be located by finding the pole P12 of the great circle
which passes through the pole concentrations P1 and P2.

back to Analysis of translational slip

Further reading: translational


Chandler, R.J., 1970. "A shallow slab slide in the Lias clay near Uppingham, Rutland".
Geotechnique, 20¸ 253-260.

Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood,
Staffordshire".Quart. J.Eng. Geol., 5, 19-41.

Esu, F. 1966. Short-term stability of slopes in unweathered jointed clays. Geotechnique, 16,
321-328.

Hutchinson, J.N. 1961. A landslide on a thin layer of quick clay at Furre, Central Norway.
Geotechnique, 11, 69-94.

Hutchinson, J.N. 1967. The free degradation of London clay cliffs. Proc. Geotechnical Conf.
(Oslo) 1, 113-118.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent" Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements", Geotechnique, 21, 353-358.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture,
Geotechnique, 14, 77-101.

Skempton, A.W., 1966. "Bedding-plane slip, residual strength and the Vaiont landslide".
Geotechnique, 16, 82-84.

297
Skempton, A.W. and Hutchinson, J.N., 1969. "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and Petley, D.J., 1967. "The shear strength along structural discontinuities in
stiff clays". Proc. Geot. Conf. (Oslo), 2, 29-46.

Weeks, A.G., 1969. "The stability of natural slopes in south-east England as affected by
periglacial activity", Quart. J. Eng. Geol. 2, 49-62.

Analysis of rotational slip


 Methods of analysis
 Ordinary Method of slices
 Bishop rigorous method
 Bishop simplified method
 Tension cracks
 Further reading: rotational

The shear strength of the soil is a function of the


normal stress s :

 = c' + ' tan '

For this reason the method of analysis must take account of the changes in overburden
pressure along the length of the slip circle. The procedure requires that a series of trial circles
are chosen and analysed in the quest for the circle with the minimum factor of safety. Each
circle is divided into vertical strips and the factor of safety is determined by considering the
forces acting on each strip.

When specifying strips, care must be exercised to avoid:


(a) A dip angle (n) of zero magnitude, since this gives an infinite factor of safety for that
slice.
(b) A steep dip angle (n) such that negative normal forces are calculated. This case
corresponds to the formation of a tension crack and checks should be performed to ensure this
condition does not exist.

See also the case studies:


 Circular slide
 Coastal landslide
 Movement along geological boundary

298
Ordinary Method of slices
 Derivation of Ordinary Method

This method is also refered to as "Fellenius' Method" and the "Swedish Circle Method".
Consider the geometry of the trial slip circle shown in the diagram. The slipping mass is
divided into slices in the normal way and these are numbered 1, 2, 3, etc, for ease of
identification. For a potential slip circle, the factor of safety against slip is given by:

This solution tends to give a conservative value for the factor of safety, of between 5 and
20%. This can be expensive and therefore a more rigorous approach is favoured.

n = the positive or negative dip angle of the tangent line at the centre of the slice base (unit:
degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN)

299
Derivation of Ordinary Method of slices
 Determining the shear foce for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn gives,

However, the magnitude and position of the side forces (X and E) are unknown and therefore
the problem is statically indeterminate. Even so, this problem can be overcome by simply
ignoring the effects of the side forces, i.e.

Then substitute for


xn = Rsinn and
un = (Wn run) / (Ln cosn)
to give the ordinary equation.

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

or, for a unit thickness:

300
n = c' + [(Nn / Ln) - un] tan'

Therefore the shear force acting along the base of the slice is:

Sn = n .Ln
= Ln {c' + [(Nn / Ln) - un] tan'}

By considering the forces acting on the slices and resolving normal to the slip surface, we can
show that

Nn = (Wn + Xn+1 - Xn) cos n - (En+1 - En) sin n

giving

Sn = c'Ln + (Wn .cos n - un).tan' + [(Xn+1 - Xn).cos n - (En+1 - En).sin n].tan'

Bishop rigorous method



 Derivation of Bishop rigorous equation

Bishop derived an expression which takes account of the interslice forces and gives a more
accurate solution to the idealised geometry of circular slip. The importance of such forces can
be demonstrated by considering a slice directly below the center of rotation: this slice has an
independent safety factor of infinity since  = 0 at that point. Bishop's solution requires that
the factor of safety is constant along the complete slip circle. For a potential slip circle, the
factor of safety against slip is given by:

301
or

However, the determination of the interslice forces is labourious and therefore Bishop's
simplified method is prefered.

n = the dip angle of the tangent line at the centre of the slice base (unit: degrees).

' = angle of friction in terms of effective stress (unit: degrees).

c' = cohesion in terms of effective stress (unit: kN/m²).

Ln = base length of a slice (unit: m).

run = pore pressure ratio for slice.

Wn = vertical load due to the slice (unit: kN).

Xn = vertical interslice force (unit: kN).

un = pore pressure at base of slice (unit: kN/m²).

302
Derivation of Bishop rigorous method
 Determining the shear force for each slice

The factor of safety against rotational slip is defined as the ratio of the restraining moments to
the disturbing moments:

Substituting for Sn and xn gives the rigorous solution.

Determining the shear force for each slice


From Coulomb's equation, the shear strength mobilised along the failure surface for slice n is:

n = c' + (n - un) tan'

therefore the shear force acting along the base of a slice of unit thickness is:

Sn = n .Ln
= c'.Ln + N'n .tan'

Resolving the forces on the slice vertically gives:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + Sn sin n

But this is true for Sn = c' Ln + N'n tan' at the limiting state only. Therefore, the factor of
safety is introduced into this expression for all other states (Sn / Fs) giving:

Wn + Xn+1 - Xn = (N'n + un Ln) cos n + ((c'Ln /Fs) + (N'n

/Fs) tan' ) sin 

303
giving,

and thus,

Bishop simplified method


 Solving the Bishop simplified equation
 Short term analysis for an undrained soil

Although the rigorous method is more "accurate" it is a time consuming process and can be
easily simplified by ignoring the X terms. The errors associated with such a simplification
have been shown to be small.

The values for run vary from slice to slice but, unless zones of high pore pressure exist, an
average value weighted according to area can be used throughout. Like the Ordinary Method
the factor of safety calculated is conservative. The value is underestimated by about 2%, but
can be as large as 7%. The nature of the equation, makes solution using a computer
programme or spreadsheet desirable.

304
back to Bishop simplified method

Solving the Bishop simplified equation


 Example of table calculations

Note that n may have both positive and negative values.

In order to solve this equation, which has Fs on both sides, the right hand value is first
estimated using the Ordinary Method and then the left hand value is calculated. This new
value is then used and the procedure repeated until the two values converge.

Create a table of calculations for each slice. Note that this calculates the factor of safety for
this circle, not the slope. There may be another circle or failure mechanism with a lower
factor of safety.

Solving the Bishop simplified equation

Example of table calculations

   
1 2 3 ...
n

 
(m²) ... ... ... ...
area

 
(kN) ... ... ... ...
Wn

 
(deg) ... ... ... ...
an

305
   
sinn ... ... ... ...

Wn sinn (kN) ... ... ... ... S(1)

 
bnBISHOPSIMPLEQT_B (m) ... ... ... ...

 
(kN) ... ... ... ...
exp(1)

   
... ... ... ...
exp(2)

exp(1)/exp(2) (kN) ... ... ... ... S(2)

             

Fs = (2) / (1)

n = number of the slice under consideration

bn = the plan width of the slice

area = bn .zn = approximated area of slice

Wn = weight of the slice for a unit thickness, calculated by multiplying the slice area by the
unit weight of the soil

n = the dip of the slip circle at the centre of the base of the slice, most easily obtained by
drawing a line from the centre of rotation to the centre of the slice at the slip circle and
measuring the angle it makes with the vertical (remember its sign!)

exp(1) = c'bn + Wn(1-run)tan'

306
exp(2) = [1+ (tan' tan n) / Fs] / sec n

Short term analysis for an undrained soil


 Locating the centre of the slip circle

This method is commonly referred to as the "Total Stress Analysis". If we consider the case of
a cohesive soil and analyse its stability in the short term, we can substitute the total stress soil
parameters into the Bishop simplified equation. Such that,

c' Þ su
' Þ u = 0

and the equation reduces to

substituting for Ln = bn secn and sin n = xn / R, gives

307
Substitute for n = R, where  is in radians. Then because the resistance to slip is constant
along the slip circle, the slices are redundant and the slip can be treated as one mass (ABCD),
giving

Where W acts through the centre of gravity of the slipping mass.

su = cohesion in terms of total stress (units: kN/m²).

R = radius of the slip circle (unit: m).

 = angle subtended by the slip circle (unit: radians).

W = vertical load of the slipping mass (unit: kN).

x = moment arm of the slipping mass (unit: m).

Locating the centre of the slip circle


Choosing the centre for the circle can be daunting at first! In 1936, Fellenius proposed the
following method for locating the centre of a circle passing through the toe of the slope:

slope bº a1 º a2 º

1:58 60 29 40

1:1 45 28 37

1:1.5 34 26 35

1:2 27 25 35

1:3 18 25 35

1:5 11 25 37

308
For deeper circles, the centre of rotation is generally vertically above the mid-point of the
slope.

Tension cracks
In undrained conditions, a tension crack may develop at the top of the slope and hence no
shear strength can occur over that length. The angle  must be reduced accordingly (see
diagram). Furthermore, water in the crack will supply an additional hydrostatic force, acting
to reduce the factor of safety. This can be incorporated into the analysis by treating it as an
additional disturbing moment,
= Fw zw.

The depth of the tension crack is given by:

u = angle of friction in terms of total stress [undrained] (unit: degrees).

su = cohesion in terms of total stress [undrained] (unit: kN/m²).

hc = depth of tension crack (unit: m).

 = unit weight of soil (unit: kN/m³).

Fw = resultant force for water pressure in tension crack acting at ²/3 depth of water (unit: kN).

zw = moment arm of resultant force Fw (unit: m).

Methods of analysis

309
In practice, limit equilibrium methods of analysis are generally adopted, in which it is
considered that failure is on the point of occurring along an assumed or a known failure
surface. The shear strength required to maintain a condition of limiting equilibrium is
compared with the available shear strength of the soil, giving the average factor of safety
along the failure surface.

The problem is normally considered in two dimensions, with the conditions of plane strain
being assumed, but a rule of thumb states that the factor of safety in 3-D is 10% greater. This
is usually ignored, giving an additional safety margin.

The limit equilibrium method can be used in cases of undrained or drained loading, provided
that the appropriate shear strength parameters are used. It is important to note that these
strengths define the ultimate collapse states. In order to design safe structures or to limit
ground movements, they may be reduced.

In using the limit equilibrium method, the geometry of the assumed slip surfaces must form a
mechanism that will allow collapse to occur, but since they may be of any shape, they do not
necessarily meet all the conditions of compatibility. In addition, the overall conditions of
equilibrium of forces on blocks within the mechanism must be satisfied although the local
states of stress within the blocks are not investigated.

For undrained loading, the ultimate strength of the soil is given by


 = su where su is the undrained shear strength.

For drained loading, where pore pressures can be determined from hydrostatic groundwater
conditions or from a steady seepage flownet, the strength is given by:
 = ' tan ' = ( - u) tan '
where ' is the appropriate angle of shearing resistance.

Limit equilibrium method procedure


The steps in calculating a limit equilibrium solution are as follows:

1. Draw an arbitrary collapse mechanism of slip surfaces. This mechanism may consist of a
combination of straight lines or curves.

2. Determine the static equilibrium of the mechanism by resolving forces or moments and
hence calculate the strength mobilised in the soil.

3. Compare the strength mobilised with the available shear strength of the soil, and hence
define an average value for the factor of safety for the mechanism considered.

4. Repeat the procedure for other mechanisms and thus find the critical mechanism which
defines the minimum value for the factor of safety.
310
 

Further reading: rotational


Early, K.R. and Skempton, A.W., 1972. "The landslide at Walton's Wood, Staffordshire".
Quart. J. Eng. Geol. 5, 19-41.

Hutchinson, J.N. 1969. "A reconsideration of the coastal landslides at Folkestone Warren,
Kent", Geotechnique, 19, 6-38.

Hutchinson, J.N. and Bhandari, R.K., 1971. "Undrained loading; a fundamental mechanism of
mudflows and other mass movements, Geotechnique 21, 353-358.

Ireland, H.O., 1954. "Stability analysis of the Congress Street open cut in Chicago".
Geotechnique, 4, 163-168.

Kjaernsli, B. and Simons, N., 1962. "Stability investigations of the north bank of the
Drammen River". Geotechnique, 12, 147-167.

Sevaldson, R.A., 1956. "The slide at Lodalen". Geotechnique, 6, 167-182.

Skempton, A.W., 1964. "Long-term stability of clay slopes", Fourth Rankine Lecture.
Geotechnique, 14, 77-101.

Skempton, A.W. and Brown, J.D., 1961. "A landslide in boulder clay at Selset, Yorkshire",
Geotechnique, 11, 280-293.

Skempton, A.W. and Golder, H.Q., 1948. "Practical examples of the  = 0 analysis of the
Stability of clays". Proc. 2nd Int. Conf. Soil Mechs (Rotterdam),2, 63-70.

Skempton, A.W. and Hutchinson, J.N., 1969, "Stability of natural slopes and embankment
foundations". 7th Int. Conf. Soil Mech. and Found. Engrg. (Mexico), State-of-the-Art Vol.,
291-340.

Skempton, A.W. and La Rochelle, P., 1965. "The Bradwell slip, a short term failure in
London Clay". Geotechnique, 15. 221-242.

Site investigation
In relation to slope stability, the main aims of site investigation are:

to obtain an understanding of the development and nature of natural slopes, and of the
processes which have contributed to the formation of different natural features;
to assess the stability of various forms of slopes under given conditions;
to assess the risk of instability in natural or artificial slopes, and to quantify the influence of
engineering works or other modifications to the stability of an existing slope;
311
to facilitate the redesign of failed slopes, and the planning and design of prevention and
remedial measures;
to analyse slope failures which have occurred and to define the causes of failure;
to assess the risk of special external factors on the stability of slopes, e.g. earthquakes.

Organisation of site investigation


Any site or ground investigation is performed under the constraints of time, money and the
complexity and variability of the geological environment. It is therefore often necessary to
reach a compromise between the precise details at a site and the time and money available.
The investigation must be tackled in a logical and scientific manner. In addition, site
investigation is a skilled operation, and must be entrusted only to suitably trained operators.

Site investigations can be considered under 3 main headings:


desk study,
field study,
laboratory work.
Here the discussion will be concentrated upon topics which are unique or have special
importance in the context of slope stability.

Desk study
The aim here is to obtain all available information with regard to the site and its geological
environments. It will involve a search through records, maps, (topographical and geological),
and any other information which is relevant to the geology, history and present condition of
the site. A useful list of sources of information and the procedure to be followed in carrying
out the desk study has been given by Dumbleton and West.

It is helpful at this stage to attempt a preliminary analysis of the geology by preparing sections
etc. - this exercise may help to define where further information is required. A visit to the site
must also be made to confirm observations and predictions already made.

Dumbleton, NLJ., and West, G. (1971). Preliminary Sources of Information for Site
Investigations in Britain. RRL Report No. LR403, Transport and Road Research Laboratory,
Crowthorne, Berks.

Field study

312
 Geomorphological mapping
 Trial pits

The basic aims of the field study are to record accurately the topography of the site, to
determine the precise nature of the geological deposits underlying the site and to determine
their engineering properties, either by the collection of good quality samples which can be
tested subsequently in the laboratory, or by performing tests in-situ.

Conventional surveying techniques may provide sufficient information to permit cross-


sections etc. to be prepared. In some cases, useful input can also be obtained from air-photos.

In the context of mass movements on slopes, however, a recent development has been the
introduction and application of geomorphological mapping. The employment of such
techniques has been shown to usefully precede and supplement standard geotechnical and
geological investigations.

In civil engineering it is quite usual in the case of light structures to limit the subsurface
investigation to only a few relatively shallow trial pits. It should, however, be recognised that
trial pits have an important function in investigations for other, more important structures and,
particularly, on sites where landsliding and other forms of mass movement have already taken
place or may be expected to occur in the future.

An essential part of any investigation concerns the groundwater conditions pertaining on the
site, and the accurate measurement of water pressures in the ground. It must be stressed that
great care is needed in order to obtain reliable data on this topic, which is of comparable
importance in assessing the stability of a slope as the determination of the shear strength
properties.

Geomorphological mapping
The mapping is performed using the techniques described by Waters and Savigear. It is based
on identifying breaks and changes of slope, and the resulting delimitation of slope units. The
direction and value of maximum slope, when measured across each unit, is recorded. In the
case of large units or in areas characterised by complex forms, the number of slope
measurements per unit is increased. An example of the detail and impression of topographical
form which can be obtained using this technique can be seen here. If necessary, a clearer
visual impression of the topography can be obtained by using the slope information to prepare
a slope category map, as described by Brunsden and Jones.

Further descriptions of the development of geomorphological mapping and its application in


engineering projects have been given by Brunsden et al.

# Waters, R.S. (1958). Morphological mapping. Geography, 10-17.

# Savigear, R.A.G. (1965). A technique of morphological mapping. Mapping Assoc. Amer.


Geogr., 55, 514-38
313
# Brunsden, D., and Jones, D.K.C. (1972). The morphology of degraded landslide slopes in
South West Dorset. Quart. J. Eng. Geol., 5, 205-222.

# Brunsden, D., Doornkamp, J.C., Fookes, P.G., Jones, D.K.C., and Kelly, J.M.H. (1975).
Large scale geomorphological mapping and highway engineering design. Quart. J. Eng.,
Geol. 8,227-254.

Trial pits
Mobile rubber-tyred excavators can be used to excavate trial pits to depths of about 4-5m, and
they are economical since hiring can be made on a time basis. Deeper trial pits may require
the use of tracked excavators, which are more expensive since they have to be transported to
and from the site by low-loader. Great care must be taken when using trial pits to avoid the
risks associated with collapse of the sides of the pits. As a general rule, the spoil from the pit
must be placed well clear of the top of the pit, and adequate bracing used to provide stability
for the sides. deep trial pits with full timbering have been used for specialist purposes (see, for
example, Hutchinson et al.), but this is an expensive operation. It has been found to be
economical to use a rotary power auger in order to sink deep inspection shafts in soils or soft
314
rocks, but these shafts generally need to be not smaller than 1 m diameter if the strata are to
be examined or tests conducted in situ.

The major benefit derived from the use of trial pits, compared with boreholes, is that they
permit a physical examination of the soils en masse to be carried out in their natural habitat. It
is then possible to establish the degree of variation which may be found in a particular soil
and it is also possible to search for discontinuities which are frequently damaged or disturbed
during a sampling operation in a conventional borehole. An example of the information which
can be found in a trial pit is shown here. Judicious positioning of a series of trial pits, or the
excavation of trenches, are of great use in the investigation of mass-movement processes,
particularly since it is possible to locate the surfaces along which movements have occurred.
Block samples can be taken to include these shell surfaces and appropriate tests, performed in
the laboratory, will detertmne the shear strength along them. An accurate stability analysis can
then be performed.

Hutchinson, J.N., Somerville, S., and Petley, D.J. (1973). A landslide in periglacially
disturbed Etruria Marl at Bury Hill, Staffordshire. Quart. J. Eng. Geol., 6, 377-404.

315
Laboratory work
The object of performing laboratory tests is to obtain information, additional to that obtained
from in situ tests, on the composition and properties of the materials encountered on any site.
Laboratory tests can be grouped under three main headings:

tests for classification and identification;


tests for engineering properties;
tests for special purposes in engineering construction.

The first group include tests to determine the particle-size distribution of the material, index
property tests (Liquid and Plastic Limits), specific gravity tests, and tests to determine the
bulk density and water content of the soils. Since these are very common tests they will not be
discussed further here (see British Standard 1377).

The second group of tests includes those to determine the engineering properties of the soils,
i.e. permeability, compressibility and shear strength.

In the third group are special tests devised for earthworks and roads and airfields. They have
little relevance for mass-movement studies.

BS 1377 (1990). Methods of Testing Soils for Civil Engineering Purposes, British Standards
Institution, London.

Measurement of shearing resistance


The accurate measurement of the shearing resistance or shear strength of a material is
essential in attempting to predict future instability or to assess the present or past stability
condition. As stated previously, shear strength tests must be performed on samples of the
highest quality if reliable information is to be obtained. Even when this condition is satisfied,
however, there may still be cases where the shear strength measured in the laboratory differs
from that mobilised in situ.

Laboratory tests can broadly be divided into two types, depending primarily on the pore
pressures set up within the sample during the test and whether dissipation of these pore
pressures is prevented or permitted. Tests can therefore be categorised as either 'undrained' or
'drained'. In undrained tests, the pore pressures set up during the test are not permitted to
dissipate, and the test may be performed relatively quickly. The existence of these pore
pressures - which may or may not be monitored - influences the behaviour of the soil to a
marked extent. It is generally considered that the results obtained from undrained tests are
applicable to short-term stability conditions. In drained tests adequate time is allowed for the
dissipation of pore pressures, so tests are much longer than most undrained tests. The results
of these tests can be used to assess the long-term stability in slopes and cuttings.

Shear strength properties of soils are defined by two parameters, apparent cohesion c and the
angle of shearing resistance f . In undrained tests the parameters are expressed in terms of
total stresses, whereas in drained tests the parameters are denoted by c' and '. A summary of
316
problems which can be analysed in terms of total or effective stresses has been given by
Bishop and Henkel.

Bishop, A.W., and Henkel, D.J. (1962). The Measurement of Soil Properties in the Triaxial
Test, Edward Arnold, London.

Shear box tests


The shear box was probably the first type of apparatus used for the measurement of the
shearing resistance of soils. The apparatus, which is shown in the figure, consists essentially
of a square brass box split horizontally at the level of the centre of the soil specimen which is
held between metal grills and porous stones. The horizontal force acting on the upper part of
the box is gradually increased until the specimen fails in shear. The shear force at failure s f is
divided by the cross-sectional area A to give the shearing stress t f at failure. The vertical
stress n is provided by a vertical load on the sample, normally by dead-weights and a lever
system. The horizontal load is applied by pushing the lower part of the box by means of an
electric motor and gearbox. Volume changes are monitored by a dial gauge mounted to show
the vertical movement of the top loading platen.

The size of the shear box normally used for tests on fine-grained soils is 60 mm square, and
the sample is approximately 20 mm thick. For soils containing gravel, a shear box 300 mm
square is frequently used; in dealing with some soils even larger specimens may be required
since, as a rough rule, the maximum particle tested should not exceed one-eighth of the length
of the shear box.

Tests in the shear box are relatively simple to perform, but the test is open to a number of
criticisms. The most important of these are:

 it may be difficult to install an undisturbed sample in the apparatus;


 the stress distribution across the sample is complex;
 failure occurs along a plane dictated by the design of the apparatus;
 the area under shear reduces during the test;
 there is no direct control over drainage conditions in the sample.

Typical results from tests on well-graded sand are illustrated here.

317
 

Triaxial tests
The triaxial compression test is the most widely used technique to determine the shear
strength of soils. The apparatus is shown diagramatically in the figure. The sample, which is
cylindrical, is tested inside a perspex cylinder filled with water under pressure. The sample
under test is enclosed in a thin rubber membrane to seal it from the surrounding water. The
pressure in the cell is raised to the desired value, and the sample is then brought to failure by
applying an additional vertical stress.

One of the major advantages of the triaxial apparatus is the control provided over drainage
from the sample. When no drainage is required (i.e. in undrained tests), solid end caps are
used. When drainage is required, the end caps are provided with porous plates and drainage
channels. It is also possible to monitor pore-water pressures during a test. Full details of the
basic apparatus and refinements, and procedures for a wide range of tests in the triaxial
apparatus, are given by Bishop and Henkel.

For cohesive soils, the size of sample normally used in the triaxial apparatus is 38 mm
diameter and 76 mm long. When gravel is present, for example in boulder clay, larger
samples may be used, the most common being 100 mm diameter and 200 mm long. For
coarse gravelly soils, rockfill and artificially prepared granular material such as railway
ballast, even larger samples are required if realistic values of the shearing strength are to be
obtained. This is also true for fissured cohesive soils, where the sample tested must be of
sufficient size to contain a truly representative collection of all the structural features which
may affect the shear strength.
318
To obtain the shear strength parameters of the soil, a number of specimens (normally at least
three) are tested at different values of cell pressure. For each test, the vertical stress s 3 at
failure are determined and are used to plot a Mohr circle. The envelope to these circles then
defines the shear strength parameters.

It is important that the values of the shear strength parameters c' and ' are obtained from the
Mohr's circles obtained by tests on similar material. In markedly heterogeneous materials, it
may be difficult to obtain sufficient samples for testing, and the technique of 'multi-stage'
testing may be employed. This form of test is normally perforated on 100 mm diameter
samples. The sample is initially tested at a particular cell pressure and the vertical stress is
increased until failure is approached. At this point the cell pressure is increased, and shearing
resumes until failure is again approached under the new cell pressure. The process is repeated
a number of times. There has been some criticism of this type of test, but it does appear to
give reasonably acceptable results if the test is performed with care.

319
 

320
Measurement of residual shear strength
 Residual shear strength
 Methods of measurement of residual strength

When a soil is subjected to shear, an increasing resistance is built up. For any given applied
effective pressure, there is a limit to the resistance that the soil can offer, which is known as
the peak shear strength sp. Frequently the test is stopped immediately after the peak strength
has been clearly defined. The value sp has been referred to, in the past, as simply the shear
strength of the clay, under the given effective pressure and under drained conditions.

If the shearing is continued beyond the point where the maximum value of the shear strength
has been mobilised it is found that the resistance of the clay decreases, until ultimately a
steady value is reached, and this constant minimum value is known as the residual strength sr
of the soil. The soil maintains this steady value even when subjected to very large
displacements.

Typical results for a drained test on clay, taken to displacements large enough to mobilise the
residual strength, are shown below.

Further tests could be made on the same clay but under differing effective pressures. The
results previously described would again be obtained, and from a number of tests it would be
noticed that the peak and residual shear strengths would define envelopes in accordance with
the Coulomb-Terzaghi relationship, a. Thus the peak strengths can be expressed as:
sp = c' + ' tan '

and the residual strengths can be expressed as:

sr = cr' + ' tan r'

The decrease in shear strength from the peak to the residual condition is associated with
orientation of the clay particles along shear planes.

c' = apparent cohesion


c'r = residual apparent cohesion
321
' = angle of shearing resistance
'r = residual angle of shearing resistance
sp = peak shear strength
sr = residual shear strength
' = applied effective stress

Residual shear strength


The residual shear strength condition is of considerable practical importance since, if the soil
in situ already contains slip planes or shear surfaces, then the strength operable on these
surfaces will be less than the peak strength, and if sufficient displacement has taken place, the
strength may be as low as the residual strength.

There are a number of circumstances, as a result of which shearing of the soil may have taken
place, and the principal processes, summarised by Morgenstern et al , are:

 landsliding,
 tectonic folding,
 valley rebound,
 glacial shove,
 periglacial phenomena, and
 non-uniform swelling.

The identification of the existence of shear surfaces is a problem of great importance during
any site investigation, particularly where mass movements are involved.

It is generally accepted ( Skempton and Hutchinson) that the residual shear strength of a soil
is independent of stress history effects, not influenced by specimen size, and rate-dependent to
only a small extent. The major difficulty in determining the residual shear strength lies in the
fact that large displacements may be necessary to achieve the required degree of orientation of
the particles.

# Morgernstern, N.R., Blight, G.R., Janbu, N., and Resendiz, D. (1977). Slopes and
excavations, 9th Int. Conf. Soil Mech. and Found. Eng., 12, 547-604.

# Skempton, A.W., and Hutchinson, J.N. (1969). Stability of natural slopes and embankment
foundations. State-of-the-Art Report. 7th Int. Conf. Soil Mech. Found. Eng., Mexico, 291335.

Methods of measurement of residual strength


 Ring shear

322
The methods of measuring residual shear strength in the laboratory are given in the table. The
most satisfactory methods, in many ways, are to obtain undisturbed samples which contain a
natural slip surface and then test them either in the shear box or triaxial apparatus so that
failure occurs by sliding along the existing slip plane. Alternatively, an artificial slip plane can
be produced by cutting the specimen with a thin wire-saw. Much of the early work on
determining the residual shear strength of soils in the laboratory was performed using multi-
reversal type tests in the shear box on previously unsheared material ( Skempton). The results
of tests to measure residual shear strength in the shear box and triaxial apparatus have been
reported by Skempton and Petley. There are practical difficulties with each of these tests, and
they also have the major disadvantage that none of them permits the complete shear-stress-
displacement relationship to be obtained.

Shear box
(a) Tests on natural shear surfaces
(b) Reversal-type tests
(c) Cut-plane tests
Triaxial
(a) Tests on natural shear surfaces
(b) Cut-plane tests
Ring shear

# Skempton, A.W. (1964). Long-term stability of clay slopes. Geotechnique, 14, 75-102.

# Skempton, A.W., and Petley, D.J. (1976). The strength along structural discontinuities in
stiff clay. Proc. Geot. Conf. on Shear Strength of natural Soils and Rocks. Oslo, 2, 3-20.

Ring shear
The large displacements required to define the complete shear-stress-displacement
relationship can be obtained by using the ring-shear (or torsional shear) apparatus. The
apparatus consists of two pairs of metal rings which hold an annular sample. The sample is
subjected to a normal stress and then one pair of rings (normally the lower pair) is subjected
to rotation. It is therefore a form of direct shear test, and failure occurs along a predetermined
plane, as with the shear box. this type of apparatus was probably first used by Hvorslev and
Tiedemann. More recent designs of the ring-shear apparatus have been described by Bishop et
al. and Bromhead.

323
#

Difficulties
The aim of laboratory testing is to define a shear strength which is applicable to the field
situation. Unfortunately, there are many reasons why laboratory tests may give values for
shear strength different from those which apply in the field, and the main reasons are
considered below.

The most obvious source of error is bad or indifferent sampling. it has already been
emphasised that samples must not only be truly representative, but they must also be of the
highest quality and this entails the use of well-designed apparatus which is in good condition
and operated by skilled personnel.

As a general rule sampling disturbance will tend to reduce the strength of the soil, and, as a
further generalisation, it is likely that the shear strength parameters in terms of total stresses,
(i.e. simple undrained tests) will be more greatly affected than the parameters in terms of
effective stresses. The effect of sampling disturbance on the stress-strain relationship for
brittle and ductile soils is given in the figure.

Most triaxial tests are performed on samples with a vertical axis, and the majority of shear
box tests are performed so that failure occurs along a horizontal plane. In the field, however, a
failure plane may be appreciably curved. In clays, anisotropy is likely to occur as a
consequence of their mode of formation, and the presence of discontinuities such as joints and
fissures which may exhibit some degree of preferred orientation. Some results indicating the
anisotropy of undrained strength in London Clay are presented in the table. In terms of
effective stress, it has been found that results of tests where shearing occurs in a horizontal
direction are lower than of tests where failure occurs at other orientations, but there is a
paucity of information on this topic.

To obtain realistic results from laboratory tests, it is essential that the tests are performed on
samples which are sufficiently large to be representative of the in situ state. For intact clays, it
is likely that typical laboratory specimens (i.e. 38 mm diameter for triaxial tests) are adequate
for practical purposes, but in boulder clays or fissured clays, this may not be true. For London
Clay, for example, it appears that the in situ undrained strength is around 65-70 per cent of the
strength measured on a conventional 38 mm diameter sample, and accurate laboratory
estimates of strength would only be obtained from tests on larger samples (possibly up to 300
mm diameter). Again there is a lack of information on the effect of sample size on the
324
effective stress parameters c' and f ' of stiff fissured clay. Marsland and Butler report the
following results for Barton clay:

38 mm diameter samples: c'= 11kPa, ¢ =24deg


76 mm and 125 mm diameter samples: c'= 7kPa, ¢ =23.5deg

A discussion of other factors which may lead to discrepancies between field and laboratory
shear strengths has been given by Skempton and Hutchinson.

# Ratio of undrained strength of London Clay parallel to bedding cB and in compression


specimens with their axis normal to bedding cN (after Skempton and Hutchinson)
c
site clay size of specimens B Reference
cN
Maldon brown London Clay, shallow 38 x 76 mm 0.88 Bishop and Little (1967)
  brown London Clay, shallow 100 x 200 mm 0.86 Bishop and Little (1967)
Walton blue London Clay, shallow 38 x 76 mm 0.78 Bishop (1948)
Wraysbury blue London Clay, shallow 38 x 76 mm 0.75 Agarwal (1967)
  blue London Clay, shallow 300 x 600 mm 0.76 Agarwal (1967)
Ashford blue London Clay, deep 38 x 76 mm 0.83 Ward et al. (1965)

Marsland A., and Butler, F.G. (1967). Strength measurements in stiff fissured Barton Clay
from Fawley, Hampshire, Proc. Geot. Conf. on Shear Strength of Natural Soils a Rocks,
Oslo., 1, 139-146.

Further reading
Agarwal, K.B. (1967). The influence of size and orientation of samples on the strength of
London Clay. Ph.D thesis, University of London, unpublished.

Bishop, A.W. (1948). Some factors involved in the design of a large earth dam in the Barnes
Valley. Proc. 2nd Int. Conf. Soil Mech. and Found. Eng., Rotterdam, 2, 13-18.

Bishop, A.W., Green, G.R., Garga, V.K., Andresen, A., and Brown, J.D. (1971). A new ring-
shear apparatus and its application to the measurement of residual strength. Geotechnique, 21,
273-328.

Bishop, A.W., and Henkel, D.J. (1962). The Measurement of Soil Properties in the Triaxial
Test, Edward Arnold, London.
325
Bishop, A.W., and Little, A.L. (1967). The influence of size and orientation of the sample on
the apparent strength of the London Clay at Maldon, Essex, Proc. Geot. Conf., Oslo, 1, 8996.

BS 1377 (1990). Methods of Testing Soils for Civil Engineering Purposes, British Standards
Institution, London.

Bromhead, E.N. (1 979), A simple ring shear apparatus. Ground Eng., 12, 40-44.

Broms, B.B. (1980). Soil sampling in Europe; state of the art. Journ. Geot. Eng. Div.,

Brunsden, D., Doornkamp, J.C., Fookes, P.G., Jones, D.K.C., and Kelly, J.M.H. (1975).
Large scale geomorphological mapping and highway engineering design. Quart. J. Eng.,
Geol. 8, 227-254.

Brunsden, D., and Jones, D.K.C. (1972). The morphology of degraded landslide slopes in
South West Dorset. Quart. J. Eng. Geol., 5, 205-222.

Dumbleton, NLJ., and West, G. (1971). Preliminary Sources of Informati on for Site
Investigations in Britain. RRL Report No. LR403, Transport and Road Research Laboratory,
Crowthorne, Berks.

Hutchinson, J.N., Somerville, S., and Petley, D.J. (1973). A landslide in periglacially
disturbed Etruria Marl at Bury Hill, Staffordshire. Quart. J. Eng. Geol., 6, 377-404.

Hvorslev, M.J. (1937). Uber die Festigkeitseigenschaften gestorter bindiger Boden. Ingenior
Skriftor A., Copenhagen, 45.

Marsland A., and Butler, F.G. (1967). Strength measurements in stiff fissured Barton Clay
from Fawley, Hampshire, Proc. Geot. Conf. on Shear Strength of Natural Soils a Rocks,
Oslo., 1, 139-146.

Morgernstern, N.R., Blight, G.R., Janbu, N., and Resendiz, D. (1977). Slopes and
excavations, 9th Int. Conf. Soil Mech. and Found. Eng., 12, 547-604.

Savigear, R.A.G. (1965). A technique of morphological mapping. Mapping Assoc. Amer.


Geogr., 55, 514-38

Skempton, A.W. (1964). Long-term stability of clay slopes. Geotechnique, 14, 75-102.

Skempton, A.W., and Hutchinson, J.N. (1969). Stability of natural slopes and embankment
foundations. State-of-the-Art Report. 7th Int. Conf. Soil Mech. Found. Eng., Mexico, 291335.

Skempton, A.W., and Petley, D.J. (1967). The strength along structural discontinuities in stiff
clay. Proc. Geot. Conf. on Shear Strength of natural Soils and Rocks. Oslo, 2, 3-20.

Tidemann, B. (1937). Uber die Schubfestigkeit bindiger Boden. Bautechnik 15,

Ward, W.H., Marsland, A., and Samuels, S.G. (1965). Properties of the London Clay at the
Ashford Common shaft: in situ and undrained strength tests. Geotechnique 15, 321-344.

Waters, R.S. (1958). Morphological mapping. Geography, 10-17.

326
 

Remedial measures

The factor of safety of a slope in soil possessing cohesion and friction can be written as

If, for a particular slope, the computed or actual factor of safety Fs is inadequate, clearly Fs
can be increased by:
increasing the numerator (i.e. [c'l+(Wcos-ul)tan] ), or
decreasing the denominator (i.e.  ), or
a combination of the above.

The main methods for achieving this increase in Fs, are: replacement; modification of slope
geometry; drainage; use of restraining structures.
Detailed reviews of the wide range of remedial methods used in improving the stability of
slopes are given by Hutchinson and Zaruba and Mencl.

Hutchinson, J.N. 1977. Assessment of the effectiveness of corrective measures in relation to


geological conditions and types of slope movement. Bulletin of the Int. Assoc. Eng. Geol., 16,
131-155.

Zaruba, Q. and Mencl, V., 1982. Landslides and their control. Elsevier, Amsterdam;
Academia, Prague.

Drainage
 Trench drains

Drainage is one of the most widely used methods for improving stability. Clearly surface
water must be removed and build-up of water pressures in tension cracks prevented.
Subsurface drainage must be designed to reduce the water pressures acting on actual or
potential slip surfaces; in this way, the value of the pore pressure (u) is reduced, thereby
producing an increase in the factor of safety.

327
Several methods exist for subsurface drainage, including:
trench drains
horizontal drains
vertical drains (or wells)
galleries

Drainage may also be achieved by the use of electro-osmosis and by planting suitable
vegetation.

Of these various methods, trench drains are frequently the cheapest and most widely used
method. They are applicable to slips of moderate depth, but for deeper failures other methods
may be more appropriate.

back to Drainage

Trench drains
 Diagram of trench drains
 Design of trench drains

Trench drains are normally constructed by machine; the drains are typically 0.5 to 1.0m wide
and up to 7 or 8m deep. They are back-filled with suitable free-draining material with a
porous pipe at the base to collect and remove the water. Provision to prevent clogging must be
incorporated in the design. Ideally, the drain should penetrate through the slip surface (such
drains are referred to as "counterfort" drains) and then in addition to the improvement in
stability as a result of reduced pore water pressured on the slip surface, some additional
restraint is achieved by the replacement of the weak slipped material by the stronger material
in the drain.

328
#

Design of trench drains


 Example drainage design

An approximate method for designing trench and counterfort drains has been developed by
Hutchinson using finite element analyses and assuming two-dimensional steady-state flow.
Hutchinson used the results of the analysis to define the efficiency () of the drains and to
relate the efficiency to the ratio s/ho where s is the spacing of the drains and h o is the depth of
the drains beneath the groundwater level (see diagram). He suggested that Lines G and H can
be taken as reasonable upper and lower bounds for the design of drains. For the purposes of
preliminary design, curve G can be regarded as a conservative lower bound. Hutchinson also
presented data from the long-term performance of drains at 6 sites, with encouraging results.

329
 

back to Design of trench drains

Example drainage design


 Solution without drains
 Solution with drains

Consider a slope inclined at 9º to the horizontal in which a translational failure has occurred.
The slip surface exists at a depth of 4m below ground level, and the phreatic surface has been
located at 0.5m below ground level. If the shear strength parameters on the shear surface are
given by c'r = 0 and f 'r = 16º, what is the factor of safety of the slope in its present condition
and what will be the factor of safety if trench drains 4m deep are installed at 10m spacings?

Take  = 20kN/m³ w = 10 kN/m³

The introduction of these drains in the slope increases the factor of safety from 1.02 to 1.34.

Example drainage design

Solution without drains


For a plane translational slip,

If c' = 0, then

330
In this example,  = 20kN/m³ , w = 10 kN/m³, z = 4m, h = 3.5m, 'r = 16º and  = 9º

Therefore, without drains,

Example drainage design

Solution with drains


The depth of the drains below the water table, ho = 3.5m and their spacing, s = 10m.

Therefore s / ho = 10 / 3.5 = 2.9

On the figure opposite, Hutchinson suggested that Line G can be taken as a reasonable upper
bound for the design of drains. So, using Line G in this case,

for s / ho = 2.9, we obtain = 0.48

is the average efficiency of the drains, and is given by:

Thus for = 0.48 and ho = 3.5m,

= 1.82m

The new value of Fs can be computed by substituting h = in the appropriate equation:

331
 

back to Remedial measures

Restraining structures
Retaining structures such as piles, walls and anchors may be used to improve stability. It must
be appreciated that the forces and moments to which these structures are subjected may be
very large and hence careful design is essential. The detailed design of these structures is
outside the scope of this section. Useful discussions are given by Zaruba and Mencl,
Bromhead and Leventhal and Mostyn.

# Bromhead, E.N. 1992. The stability of slopes. Blackie, London.

# Leventhal, A.R. and Mostyn, G.R. 1987, Slope stabilisation techniques and their
application. In Slope Instability and Stabilisation, ed. by B. Walker and R. Fell, Balkema,
Rotterdam.

back to Remedial measures

Modification of slope geometry


Changing the geometry of a slope to improve stability can involve the following:
excavation to unload the slope,
filling to load the slope,
reducing the overall height of the slope.

Where excavation and/or filling are used as remedial measures, it is essential that they are
correctly positioned, and use should be made of the Neutral Point Concept.

332
#

back to Remedial measures

Replacement
Where the slip surface is not unduly deep, removal of all (or part) of the slipped material and
replacement provides a relatively simple and straightforward remedial measure. The removed
soil may be replaced by free-draining material (in which case some additional benefit may be
achieved by drainage) or by the recompacted slip debris. If shear surfaces exist at shallow
depth, they can be destroyed by digging out, remoulding and recompacting. A recent
development has seen the incorporation of geotextile reinforcement within the replaced
material.

back to Remedial measures

Further reading
333
Bromhead, E.N. 1992. The stability of slopes. Blackie, London.

Chandler, R.J. 1991. Slope stability engineering. Thomas Telford, London.

Hutchinson, J.N. 1977. Assessment of the effectiveness of corrective measures in relation to


geological conditions and types of slope movement. Bulletin of the Int. Assoc. Eng. Geol., 16,
131-155.

Leventhal, A.R. and Mostyn, G.R. 1987, Slope stabilisation techniques and their application.
In Slope Instability and Stabilisation, ed. by B. Walker and R. Fell, Balkema, Rotterdam.

Zaruba, Q. and Mencl, V., 1982. Landslides and their control. Elsevier, Amsterdam;
Academia, Prague.

334

S-ar putea să vă placă și