Sie sind auf Seite 1von 138

APPLIED GEOTECHNICS

CVEN3201

Student Notes
2013
PREAMBLE
These notes cover most of the material given in the second part of the undergraduate subject
CVEN3201 Applied Geotechnics and Engineering Geology taught at the School of Civil
and Environmental Engineering, the University of New South Wales.

In this part of the subject students become familiar with some topics in geotechnical
engineering. It is a design course which aims to develop an understanding of site investigation
methods, concept of safety margins and the methods used widely in geotechnical design,
design of shallow and deep foundations and methods of estimating their settlements and
deformations, evaluation of lateral earth pressures and design of retaining structures, and
evaluation of stability of slopes.

These notes are not intended to take the place of attending lectures, text books or further
reading. The content of these notes and relevance to the course are subjected to change during
the semester, and cannot be relied upon in lieu of attendance.

There may be some errors in the typing of the current draft notes; it would be greatly
appreciated if any of them is brought to the attention of the lecturer.

H. A. Taiebat
April 2013
TABLE OF CONTENTS

Geotechnical Design Criteria 3


Strength Criterion 3
Serviceability Criterion 6
Bearing Capacity of Shallow Foundation 9
Types of Shallow Foundations 9
Geotechnical Design of Shallow Foundations 10
Bearing Capacity of Shallow Foundations 11
General Bearing Capacity Equation based on Hansens Theory 14
Allowable Bearing Pressure 18
Choice of Factor of Safety 18
Footings on Layered Soil 23
Selection of Material Parameters 23
Presumptive Bearing Capacity 25
Settlement of Shallow Foundations 27
Stresses under Loaded Footings 27
Components of Settlements 32
Calculation of Settlement 32
Settlement Calculation Based on Theory of Elasticity 33
Choice of Elastic Stiffness Parameters 37
Consolidation Settlements 40
Secondary Consolidation Settlements (Creep) 47
Footings on Cohesionless Soils 51
Introduction 51
Plate Loading Tests 51
Penetration Testing 52
Prediction of Footing Capacity and Settlement Based on In-situ Test Results 54
Appendices 59
Presumptive Bearing Capacity 59
References: 61
Pile Foundations 63
Classification of Piles 64
Design Requirements 66
Axial Capacity of Single Piles 67
Pile Driving Formulae 70
Vertical Settlement of Single Piles 71
Capacity of Pile Groups 74
Vertical Settlement of Pile Groups 75
CVEN3201 Applied Geotechnics ii Table of Contents

Lateral Capacity of Single Piles 78


Lateral Deflection of Single Piles 84
Safety Margin 87
References 88
Lateral Earth Pressure 89
Lateral Earth Pressure at Rest 89
Rankines Theory of Lateral Earth Pressure 89
Coulombs Theory of Earth Pressure 94
Drained or Undrained Analysis (Total or Effective Stress Analysis) 96
Effects of Water 96
Retaining Walls 99
Design Requirements and Safety Margin 99
Design of Gravity Retaining Walls 103
Design of Embedded Retaining Walls 106
Anchors 112
Design for serviceability limit states 112
References 113
Past Exam Questions 115
2008 115
2009 117
2010 119
2011 121
2012 123
Tutorial Questions 125
GEOTECHNICAL DESIGN CRITERIA
In geotechnical engineering rational design methods were established over 50 years ago by
Terzaghi and Peck, and published in their book Soil Mechanics in Engineering Practice.
Since then these design methods have been refined and improved, following better
understanding of foundation behaviour and the factors which govern this behaviour, providing
practitioners with the broad design techniques and design criteria.
Any geotechnical design deals with various loads and different soil strength parameters. Both
loading and strength have various degrees on variability and uncertainty. The design criteria
aim to ensure that the load applied to the
Load
soil is below that which would (a) cause
failure of the soil, a strength criterion, or Failure load
(b) cause excessive settlement or
movement of the soil, a serviceability
criterion. Both strength criterion and Strength
serviceability criterion limit the load that
can be applied to a foundations or a
Serviceability
retaining structure.
For many structures such as towers,
bridges, buildings and power stations, Settlement
Settlement limit
serviceability considerations may be
more important that strength criterion
and hence the foundation design is governed by design to limit settlement rather than soil
pressure. Structures, such as earth fill, earth dams, levees, braced sheeting and retaining walls,
can tolerate large settlements and hence ultimate bearing resistance governs.

Strength Criterion
To satisfy strength criterion, geotechnical engineers generally adopt one of the following
methods:
Global factor of safety method;
Load and resistance factor design method;
Partial factor of safety design method;
Probabilistic method.
Following is a brief description of the methods.

Global factor of safety method:


This is the first method introduced for a safe design and used by geotechnical engineering
practitioners for most of the 20th century when designing against failure. It is still used for the
design of most geotechnical problems such as shallow foundations, pile foundations, and
retaining walls.
The design criterion can be described as follows:
Ru
Pi
F
where Ru is the ultimate load capacity or strength of the foundation, F is a safety factor, and Pi
are different components of applied loading. The definition of F varies with the type of
problem, for example consider the following examples:
F = Ultimate pressure / Applied pressure (Footings)
F = Restoring moment / Overturning moment (Retaining walls)
F = Shear stress at failure / Applied shear stress (Slope stability).
CVEN3201 Applied Geotechnics 4 Geotechnical Design Criteria

The value of F is chosen based usually on experience and precedence. Attempts were also
made to relate factors of safety to statistical parameters of the ground and the foundation
types.
The choice of a suitable value of factor of safety, F, depends on several factors. The following
table shows typical values of F for buildings and structures. It can be seen that its value may
vary from 2 to 4 depending upon the circumstances.
F value
Category Typical Structure Load Character
Thorough site Limited site
investigation investigation
Railway bridges
Max. design load likely
A Warehouses 3 4
to occur often
Silos
Highway bridges Max. design load only
B Industrial buildings expected to occur on 2.5 3.5
Public buildings rare occasions
Max. design load does
C Residential 2 3
not occur
Values of F for buildings and structures

The following table also gives typical ranges of F for other types of geotechnical engineering
works. These values are typically reduced by 25% for temporary structures.

Works Range of F
Earthworks: dams, slopes, fills 1.2 - 1.6
Braced excavations 1.2 - 1.5
Retaining walls, excavation, offshore foundations 1.5 - 2.0
Foundations on land 2.0 3.0
Piled footings 2.0 - 5.0
Seepage forces - uplift, heaving 1.5 - 2.0
Typical values of F for geotechnical engineering works

Load and Resistance Factor Design (LRFD) Method:


This method is based on a limit state design in geotechnical engineering. It is the results of an
attempt to close the gap between the geotechnical engineering and the structural engineering,
where load factors have been used as part of the design. One approach within the limit state
design category is the LRFD approach, which can be represented by the following design
criterion:
R u a i Pi
where is strength reduction factor applied to the ultimate capacity, Ru, calculated using the
best estimate geotechnical parameters, and ai are load factors. One of the advantages of this
method is that the values of load factors are usually specified in codes or standards that also
used by structural engineers. Typical load combinations are shown in the following table for
both ultimate and serviceability load conditions (Standards Australia). Other combinations are
also specified, including liquid and earth pressure loadings.
Case Load factors in different combinations
Dead + Live 1.25G + 1.5Q and 0.8G + 1.5Q
Dead + Live + Wind 1.25G + Wu + 0.4Q and 0.8G + Wu
Dead + Live + Wind 1.25G + 1.6E + 0.4Q and 0.8G + 1.6E
G : Dead Load, Q: Live load, Wu: Wind load, E: Earthquake load
Typical Load combinations and load factors
CVEN3201 Applied Geotechnics 5 Geotechnical Design Criteria

The strength reduction factor, , is also specified in standards. For example the following
table shows the range of values for strength reduction factor, , used in design of piles by
Australian Standard (AS2159-1995).
Method of assessment of ultimate geotechnical strength Range of
Static load testing to failure 0.7 0.9
Dynamic load testing to failure 0.5 - 0.85
Static analysis using CPT data 0.45 0.65
Static analysis using laboratory data for cohesive soils 0.45 0.55
Dynamic analysis using different formula 0.45 0.65
Rage of strength reduction factors in pilling code

The Australian Standard also provides guide for application of different values of the strength
reduction factors, as shown in the following table.
Circumstances where lower can be used Circumstances where higher can be used
Limited site investigation Comprehensive site investigation
Simple method of calculation More sophisticated deign method
Average geotechnical properties used Conservative geotechnical properties used
Use of published correlation for design Use of site-specific correlation for design
parameters parameters
Limited construction control Careful construction control
General guide for selecting strength reduction factors

The LRFD approach is sometimes referred to as the American Approach to limit state
design, because of its increasing popularity in North America. This method has also been
adopted in Australian standard for pilling, AS2159-1995.

Partial Factor of Safety Design Method:


In this approach, the design criterion for stability is:
R a i Pi
where R is design resistance which is calculated using the design strength parameters
obtained by reducing the characteristic strength values of the soil with partial factors of safety.
The partial factor of safety approach is sometimes referred to as the European Approach to
limit state design because of the considerable extent of its application in parts of that
continent. This method is also implemented in Australian Standard for earth retaining
structures, AS4678-2002. As an example, the partial strength reduction factors used in
AS4678 are given in the following table.

Soil or fill properties: c and


Strength factor
Fill Class I Fill Class II Uncontrolled In-situ
(98%) (95%) Fill Soil
u 0.95 0.90 0.75 0.85
Strength
uc 0.90 0.75 0.50 0.70
u 1.0 0.95 0.90 1.00
Serviceability
uc 1.0 0.85 0.65 0.85
Soil or fill property: cu
Strength uc 0.6 0.5 0.3 0.5
Serviceability uc 0.9 0.8 0.5 0.75
Strength reduction factors in AS4678
CVEN3201 Applied Geotechnics 6 Geotechnical Design Criteria

Probabilistic Method:
This method is based on the application of probability theory to geotechnical engineering.
This method has not been widely accepted by most designers but can be used for unusual
projects, where the other three methods are not valid or deemed to be conservative, such as
cases in geotechnical earthquake engineering.
Note that while the global factor of safety has been the dominant method in the 20 th century,
there is an increasing trend toward the application of the LRFD and the partial factor of safety
design methods among the geotechnical engineering practitioners.

Serviceability Criterion
In design of foundations one of the questions that needs to be addressed is what settlement
will the structure tolerate? This clearly depends on the type of structure. For example a steel
framed structure clad with galvanised iron is flexible and it would cope with quite high
settlements without damage. On the other hand a brick structure is relatively rigid and would
suffer damage easily. Settlements can be reduced by proper design of the foundations but
generally cannot be eliminated. The settlement of a foundation should be limited so that the
functions of the structure are not compromised. Also the visual appearance such as crack
width, tilt and inclination, etc. should be acceptable to users.
When evaluating settlement it is important to realise that differential settlement is the most
important component as it is that damages structures. If a foundation settles uniformly under
the entire structure, it normally does not have any adverse effect on the stability and strength
of the structure, though it may create problem will services. Most foundations, especially
flexible footings, have non-uniform settlements due to non-uniform loading or variable
subsoil conditions. Therefore limits need to be placed on differential settlement (or angular
distortion). However limits are normally placed on total settlement as well, as this gives a
measure of protection against excessive differential movement.
The tolerable settlements of structures depend on parameters such as:
- The type and function of buildings.
- Construction materials (more ductile materialfor example, steelcan tolerate larger
movements that concrete and brick).
- Joints in floors and brickwork.
- Detailing over doors, windows and other crack-prone areas.
- What the owner will tolerate.
- Settlement relative to footpaths, services (e.g., sewer pipes).
The tolerable settlements for foundations resting on clay are larger than for foundations
resting on sand:
- Sand is less homogeneous than clay due to presence of (large) gravel.
- The longer time interval during which settlement occurs in clay can be important long time
spans allow the structure to adjust and better resist differential movement.
A number of authors and organizations have surveyed a large amount of existing buildings
and suggested tolerable settlements based on the study of settlement and structural response.
The estimated settlements from analysis under service loads should be smaller than the
tolerable values. Values of settlement limits for a reinforced concrete framed structure
typically range from 50-100 mm for maximum settlement and 0.0025 L-0.004 L for
differential settlement, where L is the span length of the frame or the distance between two
adjacent foundations.
In Australia there is no code that limits settlement. Designers must choose their own limits,
bearing in mind that Australian engineering is conservative and it would be unusual to allow
settlements greater than 50 mm for buildings and a common allowance is only 25 mm. The
CVEN3201 Applied Geotechnics 7 Geotechnical Design Criteria

Australian Standard for Residential Slabs and Footings (AS2870) sets very tight settlement
limits for domestic buildings, as shown in the following table:

Type of structure Maximum (mm) Differential


Clad frame 40 1/300
Articulated brick veneer 30 1/400
Brick veneer 20 1/600
Articulated full brick 15 1/800
Full brick 10 1/2000
Settlement limits for domestic buildings

Bjerrum (1963) recommended various limits to the differential settlements of buildings, as


given in the following table:

Type of concern Diff. settlement


Limit where structural damage of general buildings is to be feared.
Safe limit for flexible brick walls of H/L<0.25. L/150
Considerable cracking in panel walls and non-articulated brick walls.
Limit where tilting of high structures visible. L/250
Limit where first cracking in panel walls occurs. L/300
Safe limit for no cracking in buildings. L/500
Limit of danger for frames with diagonals. L/600
Limit for machinery sensitive to settlement L/750
Recommended limits for differential settlements

The following table provides rages of allowable total settlements based on various movements
of different structures (Sowers, 1962).

Type of movement Limiting factor Max settlement (mm)


Drainage 150-300
Total settlement
Access 300-600
Masonry walled structure 25-50
Prevention of
Framed structure 50-100
differential settlement
Chimneys, silos, mats 75-300
Tower, chimneys 0.004L
Rolling of trucks etc. 0.01L
Tilting Staking of goods 0.01L
Crane rails 0.003L
Drainage of floors 0.01-0.02L
High brick walls 0.0005-0.001L
Plaster cracking 0.001L
RC framed building 0.0025-0.004L
Differential movement
RC framed with curtain walls 0.003L
Steel frame continuous 0.002L
Simple steel frame 0.005L
Range of allowable settlements for various movements

The following table also shows some of recommendations for various structures and the
damage or concern associated with them and criterion for the settlement and acceptable range
of settlements. Note that when applying the criteria for settlement, consideration shall be
given to the settlements which may have already taken place before the construction. For
CVEN3201 Applied Geotechnics 8 Geotechnical Design Criteria

example, if the concern is related to architectural finish, then assessment is required only of
the total and differential settlements that are likely to occur after the finishes are in place.

Type of structure Type of damage/concern Criterion Limiting value


Structural damage Differential 1/150 - 1/250
1500 (1/1000-1/1400
Framed buildings and Walls & partitions cracking Differential
for end bays)
reinforced load bearing
walls Visual appearance Tilt 1/300
50-75 mm (sands)
Connection to services settlement
75-135 mm (clays)
Tilt (after lift
Operation of lift & elevators 1/1200 1/2000
Tall buildings, installation)
structures with 1/2500 (L/H = 1)
Cracking by sagging Deflection ratio
un-reinforced load 1/1250 (L/H = 5)
bearing walls 1/5000 (L/H = 1)
Cracking by hogging Deflection ratio
1/2500 (L/H = 5)
Ride quality Settlement 100 mm
Bridges general Structural distress Settlement 63 mm
Horizontal
Function 38 mm
movement
Bridgesmultiple span Structural damage Differential 1/250
Bridgessingle span Structural damage Differential 1/200
Recommended limits for various movements of buildings
BEARING CAPACITY OF SHALLOW FOUNDATION
Foundations are part of the structures that transfer loads imposed by super-structure to the
ground. Generally the resultant imposed loads are compressive, but in some cases, such as
foundations under the bracing bays of steel structures, tensile forces may be applied to the
foundation which requires special treatment. The focus of this section is on foundations under
compressive loads.
Different types of foundations have been developed to accommodate different ground
conditions. Shallow foundations transfer loads to the upper layer of soil which has sufficient
capacity to carry the imposed load. If the upper soil layer is not strong enough, deep
foundations, such as piles, will be used to transfer the imposed loads to the stronger layers
deep into the ground. The behaviour and therefore the method of design for shallow
foundations are different from those of deep foundations. This chapter presents methods of
estimating the bearing capacity of shallow foundations.
The definition of shallow foundations varies in different literature. Some assume a footing is
shallow if its depth of embedment is less than 4 times its width, some take a constant limit for
depth of embedment of shallow foundations, e.g. 3m. Terzaghi defines a shallow foundation
as one which has a depth of embedment less than or equal to its minimum planar dimension.
In the context of these notes a shallow foundation is taken as one with a limited thickness so
that it transfers the imposed load to the ground solely through its base; i.e., any side resistance
is insignificant and is ignored.
Shallow foundations are the more traditional type of foundations, and typically the most
economical form of foundations.
P (kN)

Types of Shallow Foundations


Spread footings Df
Spread footings generally support individual columns.
The shape of the load bearing area can be square, qo (kPa)
circular, or rectangular. If the footing is rectangular, the
shorter dimension of the footing is addressed as breadth
or B, and the longer dimension as L. For square footings
B will be equal to L.
P L
If the load applied on a spread footing is concentric, the
pressure beneath the footing base will be uniform, q, and
can be calculated as:
B
P
q
Area P (kN/pmr)

Strip footings
The length of these footings is much longer than their Df
width. They provide support for line loads, i.e., for load
bearing walls or closely spaced columns. The major qo (kPa)
applications of strip footings are in domestic or
commercial buildings of limited height.
A strip footing is regarded as infinitely long. However,
the geotechnical design of this type of footings is based P
on a unit length while the ratio of width to length is
always taken as zero (B/L=0). If intensity of the line load
is P (kN per metre run of the footing), the pressure
B
CVEN3201 Applied Geotechnics 10 Bearing Capacity of Shallow Foundation

beneath the footing, q, is calculated as:


P
q
B
where B is the width (the shorter dimension) of the footing.

P1 P2 P3 P4
Mat foundations
A mat/raft foundation is a flat slab occupying the plan
area of the building. When over half the plan area would
q
be covered by spread footings, a raft foundation may
become an economical alternative.
Raft or mat foundations are used for soils having low
shearing strength or where soil conditions are variable Columns
and erratic. The pressure under a mat footing is calculated
as sum of the forces of all columns divided by the total
area of the footing. In geotechnical design often mat
foundations are treated as inverse slabs carrying the
pressure q and supported by the columns.

Factors affecting selection of footing types


The choice of a foundation type interacts with the choice of the structure type. However, in
many cases the geotechnical conditions will ultimately control the type of foundations.
The most important factors to be considered in selecting a foundation type include:
Total load (dead, live, earthquake):
Small loads require small base area of foundation. Spread footings may be sufficient;
Large loads more likely require raft foundation or their load should be carried to strong
stratum or bedrock with piles.
Geotechnical conditions of the site, strength and compressibility of soil and rock:
Soft soils may require use of raft or pile foundations.
Depth to rock:
When bedrock is close, the depth of foundation is often extended to the rock.
Type of structure (which dictates total and differential allowable settlements).
Costs of labour and materials.
Other factors include presence (or absence) of groundwater, available construction
technology, shrink-swell characteristics of soil, etc.
In practice, it is often the case that more than one type of foundations are suitable based on
experience. Some simple calculations have to be carried out to estimate the dimensions of the
different choices. The most economical solution is generally retained.

Geotechnical Design of Shallow Foundations


As mentioned in the previous section, the design criteria for a footing aims to ensure that the
load applied to the soil is below that which would cause:
a) Failure of the soil, a strength/stability criterion, or
b) Excessive settlement or movement of the soil, a serviceability criterion.
Both strength criterion and serviceability criterion limit the load that can be applied to a
foundations or a retaining structure. The geotechnical design of foundations requires an
analysis of both the foundations stability and the expected settlement under the applied loads.
CVEN3201 Applied Geotechnics 11 Bearing Capacity of Shallow Foundation

In a strength/stability analysis, the ultimate bearing capacity of the foundation, qu, is


estimated. The ultimate bearing capacity is the maximum pressure the foundation can apply to
the soil before it undergoes shear failure with large continuing displacement with little or no
increase in load. A factor of safety, typically between 2.5 to 4, will be applied to qu in order to
evaluate the allowable bearing pressure applicable to the foundation base based on strength
criterion. This ensures that the foundation will not fail in practice.
In a serviceability analysis, immediate and long term settlements of the foundation are
estimated under the applied service load. The predicted settlements must be less than the
tolerable values for the structure.
The allowable bearing capacity, qa, may be determined as the minimum of the allowable
pressures obtained from strength criterion and serviceability criterion

Bearing Capacity of Shallow Foundations


If a footing is subjected to excessive load, some of the soil supporting it fails under the load
and the footing may experience a bearing capacity failure and/or excessive settlement. The
bearing capacity of a footing is the limiting load that the footing can support.
For shallow footings, excessive loading may produce three different types of shear failure:
General shear failure: This is characterised by the
formation of continuous failure planes in the ground
which shows heave on the surface adjacent to the
footing. This may accompanied by rapid tilting of the
footing. This mode of failure is typical for a footing on
dense sand or undrained loading of a footing on
overconsolidated (stiff) clay.
Punching failure: This type of failure causes large
settlement of the footing but shows only minimal
deformation at ground surface. The soil below the
footing compresses or consolidates while shear failure
is confined to areas close to the vertical planes around
the perimeter of the footing. The soil beyond the edges
of the foundation is not affected and no bulging occurs.
This type of failure is typical for loose sand and silt and
normally consolidated (soft) clay. The deeper the base
of the footing, the more likely that punching shear
failure develops under excessive loading, rather than
general shear failure.
Local shear failure: This mode of failure is in between
the general shear failure and the
punching failure, and is typical for
medium dense to loose sand of high
compressibility potential.
The theoretical bearing capacity of
foundations is based on ultimate limit
state analysis. The assumption in such
an analysis is that the material is rigid-
plastic, i.e., infinitely stiff before
yielding and infinitely soft after yielding
or failure. Based on these assumptions
the maximum failure load can be
determined based on equilibrium. Bearing capacity failure of a silo foundation (Tschebotarioff, 1951.)
CVEN3201 Applied Geotechnics 12 Bearing Capacity of Shallow Foundation

There are two methods for theoretical solutions to the problem of bearing capacity of a
foundation failed in a general shear failure:
The lower bound approach in which a stress field in the ground is assumed which is in a
state of failure. A lower bound bearing capacity is calculated based on compatibility and
equilibrium state of stress which nowhere exceeds the failure state. Failure cannot occur
for any load lower than a lower bound value.
The upper bound approach in which a failure mechanism is assumed which is based on
displacement compatibility. The failure load is calculated by equating the work done by
external loads to the work done by internal stresses at an increment of plastic
displacement at collapse. Failure must occur for any load greater than an upper bound
value.
The true bearing capacity will be between those obtained by an upper bound method and
lower bound method. As the mechanisms considered in upper bound and lower bound
solutions are refined, the failure loads calculated from both methods approach the true bearing
capacity.
To illustrate the calculation of the bearing capacity of a strip footing based on the lower
bound method, consider the simple stress field shown in the following figure. The soil has a
unit weight of , a friction angle , and a cohesion c. A surcharge of qo is applied at the
surface of the soil adjacent to
the footing. The ultimate B
Frictionless discontinuity
bearing pressure on the qu
qo qo
foundation will be qu.
When the soil is at a state of
a
failure, the Mohr-Coulomb Active Passive
criterion must be satisfied H
everywhere. The Mohr-
Coulomb criteria can be
written as: b
1 N 3 2c N

where N= tan2(45+/2) = (1 + c cot )/(3 + c cot ), 1 and 3 are the major and minor
principal stresses.
Underneath the footing the vertical stress will be greater than the horizontal stress, that is
v = 1 and h = 3, while in the passive zone the horizontal stress is greater than vertical
stress, i.e., v = 3 and h = 1. Table below shows the state of stresses on the active and
passive zones of the foundation soil at any depth z.
Active zone Passive zone
Vertical stress v = qu + z v = qo + z
Mohr-Coulomb q z c cot c cot
N u N h
failure criterion h c cot q o z c cot
1
Horizontal stress h (q u zc cot) c cot h N (q o zc cot)c cot
N

At the frictionless discontinuity beneath the edge of the footing, the horizontal forces from the
two failure zones must be equal. Integration of the stresses over the length of the
discontinuity, H, results in:
1 H 2 H 2
q u H c H cot N q o H cH cot
N 2 2
Therefore, the bearing capacity of the foundation, qu, is calculated from the above equation as:
CVEN3201 Applied Geotechnics 13 Bearing Capacity of Shallow Foundation

H 2

q u c cot N2 1 q o N2
2

N 1

This solution gives a lower bound to the true bearing capacity because of the simplified stress
distribution assumed in the soil. Note that in the above equation there are three terms, all
functions of , added together to form qu. Apart from the friction angle of the soil, , the first
term is affected by the cohesion of the soil, the second term by the surcharge applied to the
soil surface adjacent to the footing, and the third term by the least dimension of the footing, B,
and the unit weight of the soil beneath the footing, . Similar terms will appear in a more
sophisticated assumption for stress distribution at failure.
The oldest method of estimating the bearing capacity of foundations was developed by
Terzaghi in 1943 (although Prandtl in 1921 studied the punching effect of a long metal tool
pushed into the surface of smooth metal mass that possessed cohesion). Initially Terzaghis
method was applicable to centrally loaded horizontal strip footings, but later was modified to
include footings of other shapes. Meyerhof in 1953 expanded application of the bearing
capacity calculation to a more generalised form to include the effects of load eccentricity and
Hansen in 1970 and Vesic in 1973 produced generalised forms of bearing capacity equations
that include different shapes, load eccentricity, load inclination, sloping ground, etc.

Terzaghi bearing capacity equations


The failure mechanism assumed by
Terzaghi is shown in the figure
opposite. Only half of the symmetric
mechanism is shown. There are three
zones in the mechanism:
A wedge zone immediately below
footing represents an active soil
being pushed into the ground.
A radial shear zone bounded by a
spiral.
A passive zone being pushed
laterally and upward.
The solution obtained originally by Terzaghi was for a strip footing on the surface of a
homogeneous soil. For footing embedded in the ground the effect of overburden pressure is
taken into account via surcharge qo. The solution was modified to include footings of square
and circular shapes.
The equations derived by Terzaghi for different footing shapes are:
Continuous footing: qu = c Nc + qo Nq + 0.5 B N
Square footing: qu = 1.3 c Nc + qo Nq + 0.4 B N
Circular footing: qu = 1.3 c Nc + qo Nq + 0.3 B N
Where:
qu = ultimate bearing capacity;
c = cohesion;
qo = surcharge, i.e., overburden pressure at the level of footing base;
= unit weight of soil;
B = the least dimension of rectangular footings or diameter of circular footings;
Nc, Nq & N = bearing capacity factors which are functions of and given in the following
table.
CVEN3201 Applied Geotechnics 14 Bearing Capacity of Shallow Foundation

General Bearing Capacity Equation based on Hansens Theory


The ultimate limit state analysis for foundations of other shapes becomes much more
complicated than that for strip footing. Other researchers have improved the original
Terzaghis equations, by introducing additional empirical factors. This leaded to the so-called
general bearing capacity equations. Hansen (1970) and Vesic (1973) produced general
bearing capacity equations which are very similar in forms, but with different empirical
factors.
The general form of Hansens equation is:
qu = c Nc sc dc ic gc bc + qo Nq sq dq iq gq bq + 0.5 B N s d i g b
For u0 conditions (i.e., undrained loading of saturated clays) the equation is simplified to:
qu = 5.14 cu ( 1 + scu + dcu - icu - bcu - gcu ) + qo
where:
s = shape factor; takes account of the shape of the footing;
d = depth factor; the depth to the footing base below the surface;
i = inclination factor; where loads act other than normal to the footing;
g = ground factor; the slope of the ground on which the footing is built;
b = base factor; for a footing with an inclined base.
Nc = (Nq -1)/tan.
Nq = tan2(45+/2) e( tan)
N = 1.5(Nq-1)tan

Nc, Nq & N are called the bearing capacity factors, their values for selected friction angles are
given in the following table. These factors are different from those suggested by Terzaghi.
Additional values given in table below, i.e., 2tan (1-sin)2, are useful in calculating the shape
and depth factors. The shape, depth, inclination, ground, and base factors (i.e., s, d, i, g, and b)
can be calculated from the information given in the table on the next page.

(o) Nc Nq N 2 tan (1-sin )2


0 5.14 1.0 0 0
5 6.5 1.6 0.1 0.15
10 8.3 2.5 0.4 0.24
15 11.0 3.9 1.2 0.29
20 14.8 6.4 2.9 0.32
25 20.7 10.7 6.8 0.31
30 30.1 18.4 15.1 0.29
35 46.1 33.3 33.9 0.25
40 75.3 64.2 79.5 0.21
45 133.9 134.9 200.8 0.17
Hansen's bearing capacity factors

The use of inclination factors, ic, iq, and ig, is not without problems. If the computed
inclination factors do not appear reasonable, alternative methods, for example those
developed by Vesic (1973), shall be used.
There are several other relationships to define different bearing capacity factors. The bearing
capacity calculated by any set of these relationships would be equally valid. Among these
relationships are those proposed by Vesic (1973) and Mayerhof (1963). Note that all the
relationships give no more than estimates and none of the several sets of relationships has a
significant advantage over any other in terms of a best prediction. It is a good practice to use
at least two methods and compare the calculated values of the bearing capacities.
CVEN3201 Applied Geotechnics 15 Bearing Capacity of Shallow Foundation

Factors for use with Hansen's bearing capacity equation

Shape
Depth factors Inclination factors Ground factors Base factors
factors
D
d cu 0.4 D B
B' B i cu 0.51 1
H o o
s cu 0.2 g cu b cu
L' D A f ca 147 147
d cu 0.4 tan 1 D B
B
D
N B d c 1 0.4 D B 1 iq
sc 1 q B ic iq o o
gc 1 bc 1
N c L d c 1 0.4 tan1
D
D B
Nq 1 147 147
B
D
d q 1 2 tan (1 sin ) 2
B
b q exp 2 tan
5
B D B
g q 1 0.5 tan
0.5H
sq 1 tan i q 1
5

L 1 D V A c cot is in radians.
d q 1 2 tan (1 sin ) tan
2 f a
B
D B
For horizontal ground
5
0.7H
i 1
V A c
f a cot
B d = 1.0 for all For sloping ground: b exp 2 tan
s 1 0.4 g 1 0.5 tan
5
5
L o is in radians.
0.7 H
450
i 1
V A f c a cot

Note: Do not use shape factors in combination with load inclination factors. Use depth factors
and load inclination factors only in combination, or shape factors with depth factors.

Following are the definitions of the parameters used in the table above:
c = soil cohesion
= soil friction angle
D = depth of footing in ground.
L = effective length of footing, L-2eL.
B = effective width of footing, B-2eB.
Af = effective contact area of footing, BL.
where:
eB = eccentricity of load with respect to the centre of footing in B direction.
eL = eccentricity of load with respect to the centre of footing in L directions.
ca = adhesion to base, must be less than soil cohesion, c.
H = horizontal component of loading parallel to footing base. H V tan + Af ca.
V = vertical component of loading perpendicular to footing base.
tan = coefficient of friction between footing and base soil. Use = for concrete poured on
the ground.
& = as shown in the figure above with positive directions shown; + 90o.
Note that iq & i must be greater than zero.
CVEN3201 Applied Geotechnics 16 Bearing Capacity of Shallow Foundation
eB P
Effects of eccentric loading
Many vertical loads are not applied at the centre on
the foundation or are combined with moments. The
effects of eccentric loading can be included using a B
method commonly known as the effective width P eB P
rule, after Mayerhof (1953). For a strip footing of M eB=M/P
width B which is subjected to a vertical load, V, =
applied with an eccentricity, eB, to the foundation,
the effective width, B, is calculated as: B B
B = B 2eB
For rectangular footing of length L and subjected to an eccentricity of eL in L direction, the
effective length, L, can be similarly defined as:
L = L 2 eL
The effective area of the base of the footing is then calculated as:
A = BL
The effective width, or the effective area in case of square or rectangular footings subjected to
double eccentricities, will be used to calculate the bearing capacity of the footings. All the
relevant factors, such as inclination factor, shape factor, etc. shall be calculated using the
effective dimensions of the footings.
Note that when eccentricity is in the direction of the longer dimension of a footing, L, the
effective value of that dimension may become smaller than the effective value in the shorter
dimension. In this case the shorter effective value is always used at the effective width of the
footing, i.e., L= max {(L-2eL) & (B-2eB)} and B= min {(L-2eL) & (B-2eB)}.
The presence of eccentricity reduces the available bearing capacity of a foundation
significantly. Therefore, it is wise to locate columns and piers so that the eccentricity is
minimised whenever possible.

Lateral earth resistance


Large horizontal loads on a footing can be transferred to the ground through footing base,
footing sides, and by tie beams connected to adjacent foundations. The lateral earth resistance
which could be mobilized on footing sides can be considered as a favourable action for the
footing, but often ignored in the design. Lateral earth resistance may be reduced by future
excavation close to the footing. However, if the lateral earth resistance is going to be
considered as part of resistive action against horizontal load, the following points should be
taken into account:
- In cohesionless soil the horizontal displacement needed to mobilize an earth resistance
depends on the relative density of the soil and must be consistent with tolerable displacements
of the footing.
- In normally consolidated soft cohesive soils, which are subjected to long term creep,
significant earth resistance may not develop in front of a footing.
- In normally consolidated soils with large clay content, permanent horizontal loads should be
balanced by base resistance only. However, temporary horizontal loads may be balanced by
lateral earth resistance using the undrained shear strength of the soil.
In evaluating the bearing capacity of foundations subjected to large horizontal load, it is
reasonable to check the capacity against sliding first. The design value of the component of
total force acting parallel to the foundation base must be smaller, with a suitable safety
margin, than the sum of the shearing resistances between the foundation base and the ground,
and the resisting force caused by earth pressure on the side of the footing that can be
mobilized within an appropriate range of horizontal displacement for the footing.
CVEN3201 Applied Geotechnics 17 Bearing Capacity of Shallow Foundation

Note that in frictional soil, the base shearing resistance is a function of the applied vertical
load; the larger the vertical load, the greater the horizontal base resistance. Therefore, only the
variable component of the vertical load that occurs concurrently with the maximum horizontal
load may be considered when evaluating the horizontal resistance of the footing base.

Drained and undrained analysis (total and effective stress analysis)


It is important to distinguish between drained and undrained analysis in geotechnical
engineering, and in particular in calculation of the bearing capacity of foundations. Drained
analysis is carried out using effective stresses and the effective strength parameters, c and ,
obtained by laboratory tests under drained conditions. Whereas undrained analysis is carried
out using total stresses and the undrained strength parameters, cu and u. In an effective stress
analysis the state of pore pressure in the soil must be known.
In general, an effective stress analysis is applicable to cases where pore pressure is known
within the soil. These include cases where the soil is sand and gravel and is free to drain, or
the rate of application of the load is low compared to the permeability of the soil, or the
analysis is carried out a long time after application of the load on soils with low permeability
when all excess pore pressures are dissipated.
A total stress analysis is applicable to cases where the state of pore water pressure in the soil
is not easy to determine. These include short term behaviour of footings on clay, or when the
rate of application of the load is high compared to the permeability of the soil.

Effects of ground water


Drained analysis:
In this case the effective overburden pressure, qo, is used in the second term of the bearing
capacity equations. The third term in the bearing capacity equations is affected by the unit
weight of the soil below the base of the footing, . If water table is above the footing base, the
soil below the base is fully saturated and the submerged unit weight, = t - w is used in the
equations. On the other hand if the water table is sufficiently below the base, to a depth
greater than the footing width, B, it does not have any effect on the bearing capacity and the
total unit weight, t, or the dry unit weight, d, of the soil will be used in the third term.
However, for cases where the depth of water table below the footing base, d, is less than the
breadth of the footing, i.e., d<B, the unit weight of the soil must be adjusted by a linear
interpolation between the submerged and dry unit
weight of the soil. In these cases the design unit
weight used in the third term of the bearing capacity B

equation is taken as:


d
d design w dB
design w for 0<dB B
B
Undrained analysis:
In this case the foundation soil is assumed to be saturated and the total unit weight, t, is used
in the third term of the bearing capacity equations. The total overburden pressure, q o, is used
in the second term. Note that if water table is above the ground level its effect should be
included in the calculation of the overburden pressure, qo, at the level of the footing base.

Effects of soil density


As highlighted before, the failure of loose sand can be categorized as punching failure or local
shear failure. The failure mechanisms used to derive the bearing capacity equations, where the
general shear failure occurs, are not applicable to punching or local shear failures and
therefore these equations are not valid. In case of punching or local shear failures the shear
CVEN3201 Applied Geotechnics 18 Bearing Capacity of Shallow Foundation

strength of a smaller area of the soil is mobilized, which results in a smaller bearing capacity
for the foundations.
In order to extend application of the general form of the bearing capacity equations given
earlier for loose sand, it is recommended that the strength parameters of the sand is reduced to
2/3 of their measured values, i.e., the design values of friction angle and cohesion are
calculated as:
2c 2
cdesign and design tan 1 ( tan )
3 3
where c and are the measured shear strength parameters for loose sand. The design value of
cohesion, cdesign, is used in the bearing capacity equation and the bearing capacity factors, Nc,
Nq, and N, are determined using the modified friction angle, design.

Allowable Bearing Pressure


To obtain the value of the allowable bearing capacity, a factor of safety is applied to the
ultimate bearing capacity.
The bearing capacity equations give the gross ultimate bearing capacity, qu, which is the
capacity of the soil beneath the base of the footing. However, to obtain the allowable net
bearing capacity, the factor of safety is applied to the net ultimate bearing capacity, q u(net),
which is the capacity of the footing above the ground level. Generally, the unit weight of the
footing is approximated to be the same as the unit weight of the replaced soil. Therefore, the
pressure on the footing base due to the weight of the footing itself and the weight of the soil
above the footing (if there is any) can be approximated as D, where D is the depth of the
footing base and is the unit weight of the soil. The allowable net bearing pressure, qa(net), and
the allowable load applicable on the footing, Pa, are calculated as:
q q D
q a ( net) u ( net) u and Pa = qa(net) A
F F

Choice of Factor of Safety


As stated in the section Geotechnical Design, there are two limit states that must be
satisfied in any foundation design; one based on a strength criterion which corresponds to the
failure of a foundation, and the other based on a serviceability criterion which corresponds to
the development of unacceptable settlements or distortions.
Design of shallow foundations for strength criterion has traditionally been based on a global
factor of safety, F. The selection of a suitable value of F requires considerations such as types
of loading, level of confidence on soil properties obtained form site investigation and/or
laboratory tests, the importance of the building and the consequence of failure. The following
table can be used as a guide in selection of a global factor of safety, F, for buildings and
structures.

Factor of safety, F
Category Typical Structure Load Character
Thorough SI Limited SI
Railway bridges, Maximum design load likely to
A 3 4
Warehouses, Silos occur often
Highway bridges, Maximum design load only
B Industrial buildings, expected to occur on rare 2.5 3.5
Public buildings occasions
Maximum design load does not
C Residential 2 3
occur
Values of the global factor of safety, F, for buildings and structures
CVEN3201 Applied Geotechnics 19 Bearing Capacity of Shallow Foundation

To achieve a more uniform margin of safety for different types of structures, loadings, and
soil conditions, and to develop an approach compatible with the ultimate strength design
method used in structural engineering, design based on the load and resistance factor design
(LRFD) or based on the partial factors of safety method is increasingly gaining more
popularity among geotechnical engineering professionals. In both methods different loads are
increased by different load factors suggested by codes and standards. In the load and
resistance factor design the ultimate bearing capacity is multiplied by a factor less than unity,
in a range of 0.5 to 0.8, to account for uncertainties in evaluation of the shear strength
parameters adopted in its estimation as well as approximations inherent in any estimation, and
its significance in the type of problem at hand. These factors are taken into account in the
partial factors of safety method by reducing different components of shearing strength, i.e.,
cohesion and friction angle, by different reduction factors before using them in calculation of
the bearing capacity. For example, the Canadian Foundation Engineering Manual (1985)
suggested application of a reduction factor of 0.5 for cohesion and a reduction factor of 0.8
for tan. Therefore, the strength parameters used in calculation of the bearing capacity of
shallow foundations will be cdesign = 0.5c and design= tan-1(0.8 tan). The design capacity
must be less than the sum of the applied factored loads. Note that no further factor of safety
will be used in these two methods.
CVEN3201 Applied Geotechnics 20 Bearing Capacity of Shallow Foundation

Example B1 0.5m
P=400kN
What is the factor of safety of the square footing shown
in the figure opposite? The dimension of the footing is 0.5m
1.5m1.5m. An eccentric column applies a vertical load m m
1m
of 400 kN to the footing. The footing is embedded 1.5 1.5
0.5 m in a layer of sand which has a friction angle of
= 35o and a unit weight of t = 20 kN/m3. The water
table is 1m below the ground level.
CVEN3201 Applied Geotechnics 21 Bearing Capacity of Shallow Foundation

Example B2
Check the stability of the continuous bridge pier shown in the
figure opposite. The bearing pressure at the base of the pier is
350 kPa. The water table in the preliminary design was P=700kN
assumed to be deeper than 2m below the footing base.
Gravel 2m
During a flood water rose up to 2.5m above the base and then
1.5m of the gravel was washed away. The properties of the soil
layers can be assumed as: 2m
Sand
Gravel: d = 16 kN/m3; sat = 18 kN/m3.
Dense sand: t = 18 kN/m3; c = 0, = 30o.
CVEN3201 Applied Geotechnics 22 Bearing Capacity of Shallow Foundation

Example B3
Design a combined footing for columns A P2=500 kN
P1=300 kN
and B shown in the figure opposite.
Assume a factor of safety of 3 in your
design. Column A has a dimension of 4m
0.3m0.3m, is located at the boundary of
the land and carries a vertical load of 300kN. Column B is at a distance of 4m (c/c) from
column A and carries a load of 500kN. The foundation soil is cohesionless with a friction
angle of =35o and a unit weight of =18kN/m3. You may ignore the effects footing
embedment.
CVEN3201 Applied Geotechnics 23 Bearing Capacity of Shallow Foundation

Footings on Layered Soil


The bearing capacity equations presented in the previous sections are derived for the simple
case of uniform soil where the strength, stiffness, and unit weight of the soil does not vary
with depth. In reality most footings are founded on soils with varying strength or with layers
of different materials. There are limited solutions available for calculation of the bearing
capacity of foundations for cases where the soil strength varies with depth, or where two
layers of material exist beneath the footing. Layers of different materials may be identified
when the difference in shearing strength of the layers become significant, say with more than
5o difference in the friction angle.
In cases where the founding soil consists of layers of material with different properties, the
bearing capacity is often calculated based of the following simple rules:
Using the strength properties of the weakest layer in the founding soil, especially if the
additional cost is not significant. This is due to the fact that in most cases of layered soil
foundations, serviceability criterion controls the design rather than bearing capacity.
Using the strength properties of the softer layer if it is on top of a harder one.
Using an average strength parameter values obtained in proportion to the length segments
of the potential failure line. The average unit weight of the soil layers within a failure zone
must be obtained in proportion to the area of different soil layers constrained in the failure
zone.
More rigorous design would be based on assessment of the bearing capacities of each
layer. If a footing rests on a firm strong layer
located above a soft weak layer and the weak
layer is close to the base of the footing (figure
opposite), the footing may break through the
firm layer into the soft deposit (punch through Layer A
Firm & Strong
problem). The size of the footing must be 60o
selected as such that the pressure on the weak
layer, qB, does not exceed the allowable bearing Layer B qB
Soft & Weak
pressure for that soil deposit.
The pressure on top of different layers can be Stress distribution
calculated by a method that will be presented later Layer C
in the text. Less accurately, the total footing load
can be assumed to be uniformly distributed over the base of a truncated pyramid whose sides
slope from the footing at an angle of 60o with respect to the horizontal (or with a batter of 2 in
the horizontal direction to 1 in the vertical direction).
If the upper boundary of the soft layer, i.e., top of layer B, is located at a considerable depth
below the base of the footing, failure by breaking into the soft layer may not occur as the
pressure is distributed over a wide area. However, the settlement due to the soft layer may be
excessive and need to be evaluated. In many cases the settlement criterion controls the
dimension of the footing.
The presence of a rigid layer, such as rook, or soft layer, such as soft clay, close to the base of
a footing affects the bearing capacity of the foundation. The depth of influence, within which
the rigid/soft layer has some effects on the capacity, depends on the width of the footing and
the friction angle of the soil. If a rigid or soft layer is within the depth of influence, the
bearing capacity factors need to be modified before they are used in the general bearing
capacity equations. Such modifications are given in many text books.

Selection of Material Parameters


The ultimate bearing capacity equations are derived starting from the limit state analysis with
the Mohr-Coulomb failure criterion. The external loads are equalised by the distributed
CVEN3201 Applied Geotechnics 24 Bearing Capacity of Shallow Foundation

stresses on the failure surfaces, which are total stresses. However, the resulting equations can
be used both as total stress formula and as effective stress formula like the Mohr-Coulomb
criterion depending on the drainage conditions of the soils. The soil parameters have to be
chosen correspondingly.
Because the hydraulic conductivity of granular soils is high, the excess pore water pressure
generated by loading is dissipated quickly. Hence, fully drained states exist at all stages
during and after construction and therefore the stresses on the failure surfaces are always
equal to effective stresses. The relevant soil strength parameter is the effective friction angle
. The cohesion c is generally zero for granular soils.
Because the hydraulic conductivity of cohesive soils (clays and silts) is very small, the excess
pore water pressure needs a long time to dissipate. Hence, undrained states exist during and
immediately after construction. The Mohr-Coulomb failure criterion is formulated in total
stresses. The relevant parameters are undrained shear strength parameters, cu, and u (with u
is generally assumed to be zero). The long-term stability of foundations on cohesive soils can
be checked, if required, assuming that the soil is in drained condition. The relevant parameters
are effective friction angle and cohesion c (with c is generally taken as zero). This
procedure is not often carried out as the strength of soils generally increases during
consolidations.
Some representative values of strength parameters of different soils are given in the following
table which can be used as guidelines for material parameters in the absence of the soil
properties when material data from geotechnical site investigation are limited.

Cohesionless Soils
Unit weight w Friction angle
Soil type
(kN/m3) ()
Loose gravel with sand content 16 19 28 30
Medium dense gravel with low sand content 18 20 30 36
Dense to very dense gravel with low sand content 19 21 36 45
Loose well graded sandy gravel 18 20 28 30
Medium dense clayey sandy gravel 19 21 30 35
Dense to very dense clayey sandy gravel 21 22 35 40
Loose, coarse to fine sand 17 20 28 30
Medium dense, coarse to fine sand 20 21 30 35
Dense to very dense, coarse to fine sand 21 22 35 40
Loose, fine and silty sand 15 17 28 30
Medium dense, fine & silty sand 17 19 30 35
Dense to very dense, fine and silty sand 19 21 35 40
Cohesive Soils
Unit weight w Undrained
Soil type
(kN/m3) cohesion, cu (kPa)
Soft plastic clay 16 19 10 25
Firm plastic clay 17.5 20 25 50
Stiff plastic clay 18 21 50 100
Soft slightly plastic clay 17 20 10 25
Firm slightly plastic clay 18 21 25 50
Stiff slightly plastic clay 21 22 50 100
Stiff to hard, glacial clay 20 23 100 +
Organic clay 14 17
Peat 10.5 14
Representative values for soil strength parameters
CVEN3201 Applied Geotechnics 25 Bearing Capacity of Shallow Foundation

Presumptive Bearing Capacity


Presumptive bearing capacities are defined as the minimum bearing capacities that can be
presumed to be available based on the foundation material types. They are usually given in
building codes, e.g., New York State Building Construction Code, British Code of Practice
for Foundations, etc. Presumptive bearing capacities are the allowable values and are usually,
but not always, conservative. They are based on generalised descriptions of the foundation
material or classification and therefore should be used only in primary design where the size
of a foundation needs to be estimated before more accurate calculation of the bearing capacity
could be carried out. The presumptive bearing capacity often provides fairly acceptable
guidelines for design of foundations for minor construction projects. For large construction
projects, they should never be taken as a substitute for a proper engineering analysis of
bearing capacity.
Good knowledge of the presumptive bearing capacities in a local area helps an engineer in
making prompt decisions. The following table gives an example of the presumptive bearing
capacity for some soil in NSW. Examples of the presumptive bearing capacity from some
other countries are given in the appendices at the end of this section of the notes.

Allowable bearing
Foundation type
pressure (kPa)
Made ground (Fill) 0-300
Soft clay or loam 100
Confined wet sand 150
Medium clay or sandy clay 200
Hard dry clay and dense sand 300
Soft shale 400
Weathered rock or medium shale 600
Shale rock at the boundary 650
Soft sandstone, free of defects to a depth of 450 mm and with a
total seam thickness not exceeding 20 mm for the next 450 mm of 1300
depth, where the footing is 900 mm or more from the boundary
Soft sandstone, free of defects to a depth of 450 mm and with a
total seam thickness not exceeding 20 mm for the next 450 mm of 850
depth, where the footing is at the boundary
Medium sandstone, free of defects to a depth of 600 mm and with
a total seam thickness not exceeding 20 mm for the next 600 mm 2100
of depth, where the footing is 1200 mm or more from the boundary
Medium sandstone, free of defects to a depth of 600 mm and with
a total seam thickness not exceeding 20 mm for the next 600 mm 1400
of depth, where the footing is at the boundary
Hard sandstone, free of defects to a depth of 900 mm and with a
total seam thickness not exceeding 20 mm for the next 900 mm of 3200
depth, where the footing is 1800 mm or more from the boundary
Hard sandstone, free of defects to a depth of 900 mm and with a
total seam thickness not exceeding 20 mm for the next 900 mm of 2100
depth, where the footing is at the boundary
Hard igneous rock free from gas holes 4300
Massive crystalline bed rock 8500
Presumptive bearing capacity of foundations from New South Wales Local
Government Act and Regulations, Ordinance 70

Note that presumptive bearing pressures have been given for different bearing strata, for
which the material characteristics and properties are given in descriptive terms, often very
vague and without specification of physical properties of the soil in question. The underlying
CVEN3201 Applied Geotechnics 26 Bearing Capacity of Shallow Foundation

strata are assumed to have no effect on safe bearing capacity. Also important factors such as
size, shape, and depth of foundation and the position of the water table are normally assumed
to have no effect on bearing capacity. Therefore, they should never be taken as a substitute for
a proper engineering analysis of bearing capacity.
SETTLEMENT OF SHALLOW FOUNDATIONS
The serviceability criterion for design of shallow foundations requires calculation of
settlement induced by loads applied to the footings. In order to calculate the settlement, the
increase in the vertical stress at different points under a foundation due to different loadings
needs to be evaluated. In this section some methods for calculation of the increase in stresses
under various footings will be given first followed by description of methods most commonly
used in calculation of settlements.

Stresses under Loaded Footings


Various textbooks provide charts and tables for determining the increase in stresses in the soil
due to various loadings. The increase in stresses is usually expressed as a percentage of the
applied stress on the footing. When a footing is very wide, compared to the depth and
thickness of the soil layer, the increase in stresses would be more uniform with depth. As
most footings cover a finite area, the increase in the vertical stress will not be uniform with
depth as it dissipates laterally away from the footing.
The calculation of the increase in soil pressure due to loading is mainly based on the theory of
elasticity, which assumes that:
The soil mass is a homogeneous, elastic, and isotropic half-space.
The footing is flexible so that the distribution of pressure on soil beneath the footing is
known.
Elastic solutions have been available for point loads, line loads, and extended to include
square, rectangular, circular and strip footings. These solutions are even used for layered soils,
and often for rigid footings, and their accuracy is regarded as acceptable. A summary of these
solutions will be given in the following section.

Stress due to a concentrated point load Q


For a concentrated load Q applied at the surface of an elastic half
space, the change in the vertical stress, z, at a point located at
R z
depth z below and at radial distance R from the load is given by
Boussinesq in 1885 as: z
r
3Qz3
z
2R 5
Although the solution of the point load is rarely used in engineering applications, it is the
fundamental solution to construct other solutions for more complicated cases.
There are a number of other solutions to this problem, such as Westergaards solution which
unlike that of Boussinesq is a function of Poissons ratio of the soil. These solutions can be
found in various text books.

Stress due to a uniform line load q per unit length


Integration of the Boussinesqs equation in one direction results
in the change in the vertical stress due to a line load at a point
located at depth z below and at radial distance R from the line
load. The line load is assumed to be uniform with an intensity of
q per unit length. z R
z
2qz 3
z x
R 4
CVEN3201 Applied Geotechnics 28 Settlement of Shallow Foundations

Stress due to a uniform strip loading Uniform pressure q


The changes in the vertical and horizontal stresses at a point
due to a flexible strip loading with a uniform pressure q are
given as:

z sin cos( 2)
q

x sin cos( 2)
q
z
Note that if the projection of the point on the surface is outside x
the loaded area is positive, otherwise negative.

Stress under a uniformly loaded circular area


The change in the vertical stress at a point located at depth z below the centre of a uniformly
loaded flexible circular area having a radius r and carrying a uniform pressure q is given by:
2
3

z q1
1

1 (r / z) 2

For other points under a circular footing, the change in the vertical stress can be calculated as:
z = (q Ic)/100
where q is the pressure on the footing and Ic can be obtained from the following chart.

Influence factor, Ic, for the change in the vertical stress under circular footings
CVEN3201 Applied Geotechnics 29 Settlement of Shallow Foundations

Stress under a corner of a uniformly loaded rectangular area


The change in the vertical stress under a corner of a rectangular area carrying a uniform
pressure q can be obtained from the following equation.
z = q Ir
Ir is called influence factor and can be obtained from the following equation or from its
graphical representation on Fadums chart. In the equation and the chart m and n represent
the dimensions of the rectangular area normalised with z, that is the depth of the point of
interest below the corner of the rectangle (see the sketch in the Fadums chart). Note that n
and m are interchangeable. The loaded rectangle is assumed to be flexible.

1 2mn m 2 n 2 1 m 2 n 2 2
1 2mn m n 1
2 2
Ir tan
4 m 2 n 2 m 2 n 2 1 m 2 n 2 1 m 2 n 2 m 2 n 2 1

Fadum's chart
0.26
n=
mz
0.24 2.0
Uniform
nz 1.4
Pressure q
0.22
1.0
0.20 z
0.8
0.18

0.16 0.6

0.14 0.5
Ir

0.12 0.4

0.10
0.3

0.08
0.2
0.06

0.04 0.1

0.02

0.00
0.1 1 10
m
Influence factor, Ir, for stress change under a corner of rectangular footings

The principle of superposition can be used to find the vertical stress below some points other
than a corner of a loaded rectangle. In these cases the change in the vertical stress is
determined by taking the algebraic sum of the effects of a suitable combination of rectangular
areas. For example, the stress below an interior point of a loaded area can be determined by
CVEN3201 Applied Geotechnics 30 Settlement of Shallow Foundations

dividing the loaded area into 4 smaller rectangles which share a corner above the point of
interest.

Stress under a uniformly loaded irregular area


The change in the vertical stress at a point under a uniformly loaded irregular shape can be
obtained from the Newmarks chart given below. The procedure for using this chart is that a
scaled outline of the loaded area is drawn on the chart using the scale provided. The scale for
the dimensions of the loaded area is fixed by setting the length of the line OQ (given on the
chart) equal to the depth at which the stress is to be calculated, i.e., z is set equal to the
distance OQ. The point beneath which the stress is to be calculated is placed over the centre
of the chart. The number of squares within the scaled outline of the loaded area is counted as
N. The change in the vertical stress at the desired point is given by:
z = 0.001 q N
where q is the uniform pressure applied on the flexible area.

Newmarks chart
CVEN3201 Applied Geotechnics 31 Settlement of Shallow Foundations
Strip footing
Example S1
Calculate the change in the vertical stress at a point 3m
3m
below the edge of a 3m wide strip footing carrying a load
of 150kN pmr.
o
3m

Example S2
Consider a rectangular footing resting on the surface of the soil as 3m
shown in the figure opposite. Calculate the change in the vertical
stress at a point 2m below point o due to 100 kPa uniform pressure o
on the footing. 2m

3m 2m
CVEN3201 Applied Geotechnics 32 Settlement of Shallow Foundations

Components of Settlements
Settlement of soil is time dependent. The time required to achieve the total settlement in
granular cohesionless soil is relatively short, as the settlement occurs immediately after
application of the load or, for fine sand and silt, during construction period. Settlements of
saturated clays occur over a relatively long period of time as the excess pore pressure needs to
dissipate. For fine grain soils the applied forces are immediately transferred to the pore water
and the pore water pressures increase. Therefore the effective stress in the soil is initially
unchanged and thus there is little immediate settlement. As the excess pore pressures dissipate
with time the effective stresses increase which result in settlement.
The total settlement of a saturated soil consists of three different components:
Immediate settlement, Si, which occurs within a short period of time after application of
the load. If the foundation consists of clay, this settlement occurs at constant volume and
it is mostly due to shear deformation. In sandy soils, almost all settlements occur
immediately after application of the load and therefore immediate settlement is the most
dominant component of settlement in sandy soils. Most part of this settlement is
recoverable and therefore is often called elastic settlement.
Primary consolidation settlement, Sc, which occurs as excess pore pressures dissipate and
therefore is time dependent. This settlement is usually the most dominant component of
settlement for fine grain soils and clays and can be many times larger than their initial
settlement. Only part of the consolidation settlement is recoverable. The results of
oedometer test can be used to determine the amount and rate of consolidation settlement.
The consolidation settlement at time t is referred to as Sct and the final consolidation
settlement which occurs long time after application of the load is denoted as Scf.
Secondary consolidation settlement or creep, Ss, is also time dependent and occurs very
slowly due to rearrangement of soil particles under constant effective stress. This
settlement is irreversible and, depending on stress level, it can be significant in soft clay
and soils with a large organic content. The secondary consolidation settlement is usually
estimated form empirical formulas and experimental data.
For cohesionless soils and unsaturated clays the immediate settlement predominates with
perhaps some creep. The consolidation settlement predominates for saturated cohesive soils
unless the soil is very organic, in which case the creep term may predominate.
The total final settlement is defined as:
Stf = Si + Scf + Ss
The creep settlement, Ss, is insignificant for many soils and may be ignored in calculation of
total settlement.
Sav = 0.5 (Scentre + Sedge)

Calculation of Settlement
There are many methods developed to estimate foundation settlements, they can provide
approximate values of settlements due mainly to the complex nature of soil deformation and
soil-structure interaction. Even complex analyses, such as finite element analyses, may not
predict settlements accurately since they still rely on the soil deformability parameters
(modulus of elasticity or compressibility index), approximation of which is the major
difficulty in geotechnical engineering.
CVEN3201 Applied Geotechnics 33 Settlement of Shallow Foundations

The settlement of a rigid footing is different from that


of flexible footing. When a flexible footing is
subjected to a uniformly distributed load, the contact
pressure will be uniform but the settlement will vary
on the footing surface. The settlement at the centre of
a flexible footing, Scentre, is larger than the settlement
at the edge, Sedge. The average settlement of a flexible
footing, Sav, is calculated as:
Sav = 0.5 (Scentre + Sedge)
A fully rigid footing settles uniformly but the contact pressure beneath the footing will be
non-uniform. There are few theoretical solutions available to estimate the settlement of rigid
footings. However, often the settlement of a rigid footing is estimated to be equal to the
average settlement of an equivalent flexible footing, or 75% of the settlement at the centre on
the equivalent flexible footing.
Most of the solutions developed to evaluate the settlement of foundations are based on theory
of elasticity; a brief description of the theory is given below.

Settlement Calculation Based on Theory of Elasticity


The immediate settlement Si and the total
final settlement Stf can be calculated from y y xy
theory of elasticity, using appropriate values xx xx
of stiffness parameters. The settlements can
xy
be calculated by integration of the vertical
strains developed below the foundation. x x
(a) (d)
An isotropic body is one in which the
behaviour on an element within the body yy
y z xz
does not depend on the orientation of the
element. Suppose an element of an isotropic
elastic material, shown in the figure below, xz
yy
subjected to increases in both normal stress x x
and shear stress. The response to this (b) (e)
loading can be found by summing the zz
z yz
responses of the six components of the z
loading. Consider component (a) in the
figure below, it is clear from symmetry that yz
the components of shear strain yz, zx, xy zz
are all zero and also that yy = zz. Hookes (c)
x
(f)
y
law for uniaxial behaviour states that:
xx xx xx
xx yy zz
E , E , E
where E and are material constants called Young's modulus and Poisson's ratio,
CVEN3201 Applied Geotechnics 34 Settlement of Shallow Foundations

respectively.
A consideration of the component (b) leads to the conclusion that the only non-zero strain
components are:
yy yy yy
xx yy zz
E , E , E
Similarly it is found that the response to the component (c) leads to the non-zero strains:
zz zz zz
xx yy zz
E , E , E
The response to the combined normal stresses is thus:
xx ( yy zz )
xx
E
yy ( xx zz )
yy
E
zz ( xx yy )
zz
E
The shear strain increments, xy, yz, zx occur due to shear stress increments xy (d),
yz (e), zx (f) can be calculated by the following relations:
xy yz zx
xy , yz , zx
G G G
where G is a material property called the shear modulus. The shear modulus can be related to
Young's modulus and Poisson's ratio; for isotropic elastic material it becomes:
E
G
2(1 )
The vertical displacement can be calculated by integrating the vertical strain over the soil
layers. This procedure is useful for calculation of settlement of shallow foundations on non-
homogeneous or layered soil strata.
h h
zz ( xx yy )
S zz dh = dh
E
where E and are the appropriate values of the Youngs modulus and Poisson ratio of the soil
skeleton; effective values, E and , for calculation of the long term settlement, Stf, as well as
immediate settlement of granular soils deforming under drained conditions, and the undrained
values, Eu and u=0.5, for immediate settlement of cohesive soils under undrained conditions.
For layered soils the above integration changes to:
layers
zz ( xx yy )

layers
S zz h = h
E
where xx, yy, zz are values of change in stresses evaluated at the middle of each layer,
E and are stiffness parameters for the layer, and h is the corresponding thickness of the
layer.
If the soil foundation is homogeneous, with reasonably constant values of stiffness
parameters, the elastic settlements may be calculated from the above method and presented in
forms of charts or simple relations. Most of these solutions are either homogeneous soils, or
for soils whose stiffness increases linearly with depth. Examples of these solutions are given
below.
CVEN3201 Applied Geotechnics 35 Settlement of Shallow Foundations

Settlement of Strip and Rectangular Footings


For a rectangular footing having a width B, a length L, an embedment depth D, which is
founded on a layer of homogeneous soil extended H below the foundation to an
incompressible layer, the average settlement can be calculated by:
qB(1 2 )
Sav 0 1
E
In the above equation q is the applied pressure on the footing, E and are the Youngs
modulus and Poissons ratio of the soil, 0 and 1 are influence factors which take into
account different aspect ratios and depth of the footing and thickness of the soil layer. The
influence factors can be obtained from the following charts.

3.0
Rectangular footing with Length L
q 100

2.5 D 50
L/B=
B
H 20
2.0
L/B=10
1 1.5 qB(1 2 ) 5
S 01
E
1.0 2

1
0.5

0
0.1 0.2 0.5 1 2 5 10 20 50 100 200 1000
H/B
1.0
Rectangular footing with width B
0.9
and length L
0.8 L/B=1 2
0 5 10
0.7 20
50
0.6 100 200
0.5
0.1 0.2 0.5 1 2 5 10 20 50 100 200 1000
D/B
Settlement of Circular Footings
For a circular footing founded at depth D below the surface of a homogeneous soil with a total
thickness of h, the average settlement can be calculated by:
qR
Sav 0 1
E
where R is the radius of the footing, 0 and 1 are influence factors which can be obtained
from the following charts.
CVEN3201 Applied Geotechnics 36 Settlement of Shallow Foundations

1.6
Circular footing with radius R

1.4 q

D
1.2
2R h
1.0
=0
1 0.8 0.2
qR
S 0 1 0.4
0.6 E
0.5

0.4

0.2

0
0.0 0.2 0.4 0.6 0.8 1.0 0.8 0.6 0.4 0.2 0.0
h/R R/h
1.00
0.95 Circular footing with radius R

0.90
0 =0.5
0.85
=0.25
0.80
=0
0.75
0 10 20 30 40
D/2R
The solutions based on the theory of elasticity exist only for very simple shapes of footing and
geometries of underlying layers. Since the estimation of the settlement is based on the theory
of elasticity, the principle of superposition is valid. This implies that settlement of the soil due
to a number of foundations is equal to the sum of the settlements calculated due to individual
foundations. Therefore, a footing of complicated geometry may be broken up into few
rectangular sub-footings and the settlements of each sub-footing are calculated. Then the
settlement of the footing can be estimated as the sum of the settlements of the sub-footings.

Elastic settlement of layered soils


The immediate settlement of a foundation on layered soils
Layer 1, E1, 1 H1
can be calculated approximately following the step-by-step
superposition procedure:
a) The settlement in layer 1, S1, is calculated assuming that
the second layer is rigid. The material properties of layer 1 Layer 2, E2, 2
are used in this calculation. H2
b) The settlement of layer 2 is calculated as:
S2 = Sh,p2 S1,p2
where:
Sh,p2 is the settlement of both layers, assuming that the whole soil has the properties of layer 2.
In this case the depth of soil layer is (H1+H2) and the properties of the soil are E2 and 2.
S1,p2 is the settlement of layer 1 assuming that it has the properties of layer 2, E2 and 2.
CVEN3201 Applied Geotechnics 37 Settlement of Shallow Foundations

c) The settlement of the footing on the two layers will be the sum of the settlements of layer 1
and layer 2, i.e.:
S = S1 + S2
In the above superposition procedure it is assumed that the stress distribution in the soil layers
is not affected by the different deformation properties of different layers. This superposition
technique can be extended to sites with more than two layers.

Choice of Elastic Stiffness Parameters


The key to the successful prediction of settlement by elastic theory, which relies on elastic
stiffness parameters E and , is the estimation of appropriate values of the soil stiffness
parameters. The stress-strain relationship of soils is not elastic. The modulus of elasticity of
soil, Es or E, is an idealised value for the purpose of estimating the settlement of a footing
based on elastic theory, which assumes that the stress level in the soil under the service load
applied to the footing is in the elastic range. Under heavier load a secant modulus may be
used to describe the behaviour of the foundation soil, assuming that the stress-strain response
is linear over a limited stress range. Provided that the soil stiffness parameters, and in
particular E, can be properly determined over the correct ranges of initial effective stress and
stress increments, the use of elastic theory can lead to satisfactory settlement predictions.
The choice of the elastic parameters E and also depends on the soil type and the drainage
conditions.
For sandy soils drainage occurs almost instantaneously and therefore the drained
parameters, E and should be used in calculation of the settlement. The settlement
obtained in this way will be the immediate settlement, Si.
The Youngs modulus for sand is best determined from correlations with simple in-situ
tests such as the Standard Penetration Test (SPT) or Plate Load Test. The Youngs
modulus can also be related to the coefficient of volume change, mv:

E
1 1 2
1 mv
The coefficient of volume change, mv, can be obtained from the results of oedometer tests
as:
1 e
mv
1 e o
For clay the immediate settlement occurs under undrained conditions where no drainage
occurs and thus the volume of the soil will be constant. Therefore the undrained Youngs
modulus, Eu, should be used together with a Poissons ratio of =0.5 to model the constant
volumetric behaviour of the soil under undrained conditions. The settlement calculated in
this way will be the immediate settlement, Si.
Eu can be obtained from laboratory tests. It can also be related to E by:
3E
Eu
2(1 )
For clay the total final settlement occurs along time after application of the load where all
excess pore pressures are dissipated. Therefore the drained parameters, E and , should
be used to calculate the total final settlement. The settlement calculated in this way
includes both immediate settlement and consolidation settlement of the footing, Stf. The
final consolidation settlement, Scf, the time dependent consolidation settlement, Sct, and
the time dependent total settlement can be calculated as:
Scf = Stf Si Sct = U (Stf Si) Stt = Si + U (Stf Si)
CVEN3201 Applied Geotechnics 38 Settlement of Shallow Foundations

where U is the degree of consolidation for the soil layer at time t.


The following table (extracted from Bowles, 1996) gives a number of correlations between
the elastic Youngs modulus of soils, Es (kPa), and in-situ test data that may be used when no
test data is available.
Soil SPT CPT
Es = 500(N+15) Es = (2 to 4) qc
Sand (normally consolidated) Es = 7000N
Es = (15000 to 22000) ln N Es = (1+D2r) qc
Sand (saturated) Es = 250(N+15)
Es = 40000+1050N Es = (6 to 30) qc
Sand (overconsolidated)
Es(OCR) = Es(NC) (OCR)
Es = 1200 (N+6)
Gravelly sand Es = 600(N+6) if N15
Es = 600(N+6)+2000 if N>15
Clayey sand Es =320(N+15) Es = (3 to 6) qc
Silts, sandy silt, or clayey silt Es =300(N+6) Es = (1 to 2) qc
Soft clay or clay silt Es = (3 to 8) qc

The following table also shows the approximate range of the elastic parameters for various
soils that may be used when no test data is available (after Das, 1999).
Type of soil ES (MPa) Poissons ratio, S
Loose sand 10.35 24.15 0.20 0.40
Medium dense sand 17.25 27.60 0.25 0.40
Dense sand 34.50 55.20 0.30 0.45
Silty sand 10.35 17.25 0.20 0.40
Sand and gravel 69.00 172.50 0.15 0.35
Soft clay 4.1 20.7
Medium clay 20.7 41.4 0.20 0.50
Stiff clay 41.4 96.6
CVEN3201 Applied Geotechnics 39 Settlement of Shallow Foundations

Example S3
Calculate the settlement of a 2m4m rigid footing constructed at depth 2m below the ground
level on a thick layer of homogeneous sand. The footing carries a pressure of 200kPa. The
stiffness properties of the sand layer can be assumed as: E=40MPa and =0.3.
What would be the settlement if a stiff bed rock exists at depth 10m below the footing base?
CVEN3201 Applied Geotechnics 40 Settlement of Shallow Foundations

Consolidation Settlements
One dimensional consolidation settlement of large footings
For cases where the soil is uniformly loaded over a wide area or when the dimension of the
loaded area is substantially larger than the thickness of the soil layer, the stress can be
assumed to be uniform across the thickness of the soil. Settlement of a uniformly loaded layer
of clay can be calculated using one-dimensional consolidation theory. Generally the soil
beneath the footing is divided into a number of sub-layers of finite thickness through which
the stress and soil properties do not vary significantly. The settlement of the soil is the sum of
the settlements of the sub-layers that can be calculated separately.
The deformation parameters of soil can be obtained from oedometer tests by which the
compression index, Cc, the rebound index, Cr, and the preconsolidation pressure, pc, are
determined.
For normally consolidated clay, the final consolidation settlement of a layer of soil with an
initial thickness of Ho and an initial void ratio of eo can be calculated as:
Ho
Scf Cc log f
1 eo i
where i is the initial effective stress at the middle of the layer, f = i + is the final
effective stress, is the change in the effective stress due to the loading calculated at the
middle of the layer, and Cc is the compression index of the soil. A soil is normally
consolidated if the initial stress, i, is equal to the preconsolidation pressure, pc.
For overconsolidated soil, the total final consolidation settlement is calculated by:
Ho
Scf Cr log f
1 eo i
where Cr is the rebound (or recompression or swelling) index of the soil. A soil is
overconsolidated if the initial stress, i, is less than the preconsolidation pressure, pc. The
above equation is valid if f is also less than pc.
For cases where the soil is initially overconsolidated but f > pc, a combination of the above
two equations can be used, such as:
e
H o pc f
Scf Cr log Cc log eo Cr
1 eo i pc
Alternatively, the total final consolidation
settlement of a layer of soil can be calculated
using the coefficient of volume change, mv, as: e
Cc
Scf = mv Ho
where mv can be obtained from the results of ef
oedometer tests as:
1 e
mv log
1 e o i pc f
Note that the coefficient of volume change is
stress dependents. Therefore it is necessary to use a value of mv calculated between the stress
range i and f.
The consolidation settlement at time t, Sct, is calculated as:
Sct = U Scf
where U is the average degree of consolidation for the layer. For one dimensional
CVEN3201 Applied Geotechnics 41 Settlement of Shallow Foundations
0.00
consolidation, U can be obtained
from the chart opposite. Note Average degree of consolidation for one
0.20 dimensional settlement
that Tv is a dimensionless time
factor and defined as:
cv t
0.40
Tv U
DP2
0.60
where cv is the coefficient of
consolidation and DP is the
0.80
maximum length of drainage
path.
1.00
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless time, Tv

Consolidation settlement of footings of limited dimensions


The consolidation settlement of a footing can be calculated assuming that all displacements
are one dimensional and in the vertical direction. This is obviously an approximation to the
real behaviour of the footing in order to simplify the calculations. To estimate the
consolidation settlement of a shallow footing, a soil layer must be divided into a number of
sub-layers and the stress increases due to the imposed load on the footing must be calculated
at the middle of each sub-layers. This is because the change in the effective stresses due to the
footing load varies with depth. The change in the effective stress varies with the distance from
the centre of the footing as well. Therefore it is important to calculate the settlements under
the centre and under the edge of the footing. For rigid footing the settlement of the footing
will be approximately equal to the average settlements calculated under the centre and under
the edge of the footing. Alternatively the average settlement of a rigid footing can be
approximated as 75% of the settlement calculated at the centre of an equivalent flexible
footing. The equations given for calculation of one-dimensional settlement can be used to
calculate the settlement of each sub-layer. Note that in this case Ho is the thickness of each
sub-layer, which is not necessarily similar for all sub-layers.
For the footing of finite dimensions the following methods are used to calculate consolidation
settlements.
1) Conventional one-dimensional analysis:
This method is first proposed by Terzaghi for clays by which the total settlement is assumed
to be the one obtained by the one dimensional consolidation theory explained above. It is
assumed that the foundation loading causes only vertical strains in the soil. It is also assumed
that all the settlements are due to consolidation and that the immediate settlement is negligible
compared to the consolidation settlement. Therefore:
Stf = Scf = Soed
where Soed is the settlement calculated based on one dimensional consolidation theory. The
settlement at time t is calculated as:
Stt = U Soed
2) Modified one-dimensional analysis:
In many cases, especially for soft clay, the immediate settlement is important and cannot be
ignored. The conventional one dimensional method overestimates the time dependent
consolidation settlement for cases where the initial settlement is large. In the modified one-
dimensional method the total settlement is still assumed to be equal to the one dimensional
consolidation settlement. However, the initial settlement, Si, must also be calculated by the
elastic methods. Therefore, settlement at any time can be calculated as:
CVEN3201 Applied Geotechnics 42 Settlement of Shallow Foundations

Stt = Si + U (Soed Si)


where U is the average degree of consolidation.
3) Skempton & Bjerrums method:
In this method the consolidation settlement is calculated by reducing the settlement obtained
by one-dimensional analysis by a factor .
Stf =Si + Soed
Stt = Si + U Soed
Si can be calculated by the elastic methods of analysis and for strip and circular footings is
given in the following chart.

Skempton & Bjerrums settlement coefficient versus pore pressure coefficient


for circular and strip footings

4) Elastic methods
The consolidation settlement can be calculated from elastic methods, as explained before. The
immediate settlement, Si, and the total final settlement, Stf, can be calculated based on elastic
theory using the undrained stiffness parameters, Eu and u=0.5, and the drained stiffness
parameters, E and , respectively. The consolidation settlement can be defined as the
difference between the total final settlement and the immediate settlement:
Scf = Stf - Si
Therefore the time dependent consolidation settlement, Stf, and total settlement, Stt, at any
time can be calculated as:
Sct = U (Stf Si) Stt = Si + U (Stf Si)
where U is the average degree of consolidation.

Rate of settlement and degree of consolidation


The time dependent consolidation settlement can be calculated using the average degree of
consolidation, U, as explained before. The average degree of consolidation has been obtained
for simple cases of footings of limited dimensions and presented in the following figures
(these figures are reproduced from Soil Mechanics data Sheets published by the University
of Sydney). In all the figures, Tv is a non-dimensional time factor, cv is the one-dimensional
coefficient of consolidation obtained from oedometer test, t is the time at which the settlement
CVEN3201 Applied Geotechnics 43 Settlement of Shallow Foundations

is going to be calculated, and h is generally the depth of the layer. Different combinations of
drainage conditions for the top and bottom of the soil layer and the base of the footing are
considered. The following notations are used for different cases:
PB : Permeable base of the soil layer.
IB : Impermeable base of the soil layer.
PT : Permeable upper surface.
IF : Impermeable footing base on permeable upper surface.
The following figures show the consolidation ratio for a homogeneous layer of soil with a
thickness h under a circular footing with a radius a.
CVEN3201 Applied Geotechnics 44 Settlement of Shallow Foundations

The following figures show the consolidation ratio for a homogeneous layer of soil with a
thickness h under a strip footing with a breadth 2b.
CVEN3201 Applied Geotechnics 45 Settlement of Shallow Foundations
CVEN3201 Applied Geotechnics 46 Settlement of Shallow Foundations

For calculation of the average degree of consolidation, U, for a square footing, the footing can
be treated as a circular footing of equal area. The effect of footing shape on the rate of
consolidation settlement for rectangular footings is given in the following charts. The time
required for 50% consolidation is t50. Indices R and S refer to circular and rectangular
footings.
CVEN3201 Applied Geotechnics 47 Settlement of Shallow Foundations

Rate of consolidation of a finite layer beneath a rectangular footing

Secondary Consolidation Settlements (Creep)


The creep settlement is assumed to vary linearly with the logarithm of time and can be
expressed as:
Ho t
Ss C log s
1 eo tp
where Ss is the secondary consolidation settlement or creep, Ho and eo are the initial thickness
and initial void ratio of the soil layer, ts is the time for which the secondary consolidation is
going to be calculated, tp is the time when the primary consolidation is completed, and C is
the secondary consolidation index which can be correlated to the compression index of the
soil. The ratio of C/Cc is found to be approximately constant for soil deposits and can be
obtained from the following table (Mesri et al, 1994).
Soil Types C/Cc
Granular soils 0.02 0.01
Shale and mudstone 0.03 0.01
Inorganic clays and silts 0.04 0.01
Organic clays and silts 0.05 0.01
Peat and muskeg 0.06 0.01
CVEN3201 Applied Geotechnics 48 Settlement of Shallow Foundations

The time at which the creep is assumed to commence is not well defined. Some researchers
assume that the secondary consolidation takes place simultaneously with the primary
consolidation, while others assume that it only commences at the end of the primary
consolidation period (the corresponding time is not well defined either). It is recommended
that the time at which 90-95% of primary consolidation settlement occurs is used as tp.
Note that the use of the simple approach for estimation of creep settlement can only be
regarded as an approximate solution to the problem within a limited time period. This is
because the equation given for calculation of the secondary consolidation settlement is
independent of the magnitude and distribution of the stresses induced in the soil due to
foundation loading, which is clearly an approximation to the problem. It also implies that the
creep settlement does not terminate at any time, but increases indefinitely, which does not
seem to be reasonable.

Remarks on Calculation of Settlements


It is well known that the behaviour of soil under load is not linear elastic. Where the stresses
are large, a portion of soil fails under the applied load and there may be grain crushing. Most
soils are not homogeneous and show various degrees of consistencies. Thin layers of
soft/loose soil may exist within the stratum which cannot be detected by surface exploration
and site investigation. It is therefore recommended that the calculation of settlement is carried
out with lower and upper stiffness characteristic values using various methods, in order to
obtain an understanding of the possible range of settlements.
CVEN3201 Applied Geotechnics 49 Settlement of Shallow Foundations

1750kN/m
Example S4
Find the settlement of a rigid strip footing on an over- 2m 5m
consolidated clay. The properties of a sampling point 2m
below the footing base are as follow:
Cc = 0.2, Cr = 0.05; pc=160 kPa; 4m
t = 20 kN/m ;3
eo = 0.51
Bedrock
The water table is 2m below the ground level.
FOOTINGS ON COHESIONLESS SOILS

DESIGN BASED ON FIELD TEST RESULTS

Introduction
The design of footings on cohesionless soils is usually controlled by the need to limit
settlement. Because it is not possible to take undisturbed samples of such soils they need to be
tested in-situ. The testing aims to determine the relative density which may be related to the
strength and compressibility of the soil. Many empirical relationships between the test results
and settlement have been proposed and in many cases these constitute the simplest design
approach. For small structures, typical bearing pressures range from 50 to 100 kPa, for loose
soils, to 300 kPa for dense soils. However for larger structures where more extensive testing
is carried out higher values are often used. These notes give an outline of some approaches to
the design of footings on cohesionless soils based on tests results. The methods can give
greatly different answers. There is no consensus of which method is the best. It is
recommended to estimate strength and settlement using several methods.

Plate Loading Tests


Plate loading tests are used to assess the stiffness and strength characteristics of soils. In this
test a plate (usually a square or round steel plate having a dimension of the order of 300 mm)
is loaded by jacking against a kentledge or by the use of a dead load. The load is applied
incrementally and measurements of the load deformation behaviour obtained. From these
results an empirical relationship allows for the estimation of footing settlement:
2
2B
SB Sp
Bp
where SB is the settlement of a footing of width B and Sp is the settlement of the plate of
width p. The value of Sp for the plate should be determined at the same pressure which is to
be applied to the footing.
Note that based on the above equation the ratio of settlement of a full size footing to that of
test plate cannot become greater than 4 no matter how large is the footing, and therefore the
equation may underestimate the settlement of large footings. The equation can also be
unreliable particularly where the density of the soil varies with depth. Clearly the zone of
influence of the plate is much smaller than the zone for a full size footing. Where a density
variation is detected by Standard Penetration Testing (SPT) the settlement of actual footing
can be adjusted using the following equation:
SB / Sp = B NP / p NB
where NP is the SPT value for the zone of influence for the plate and NB is the SPT for the
zone of influence of the footing. These N values are computed as the weighted average N
value over a depth equal to twice the footing width. For example if the footing width is B,
three values of N may be obtained at roughly equal B
spacing within the depth 2B beneath the footing as shown
in the figure opposite. A suitable value of NB would be:
N1
NB = (3N1 + 2N2 + N3) / 6
where N1 is the average N from 0 to 2B/3, N2 from 2B/3 N2
2B
to 4B/3 and N3 from 4B/3 to 2B.
If only two values exist within the 2B depth, a suitable N3
weighting would be:
NB = (2N1 + N2) / 3
CVEN3201 Applied Geotechnics 52 Footings on Cohesionless Soils

The results of plate loading tests may also be used to back-figure the Young's modulus, E, of
the soil allowing elastic solutions to be used for calculation of settlements. For example if the
settlement measured for a square plate of dimension B is S, the Youngs modulus of soil
within 2B beneath the plate can be calculated as:
0.7qB(1 2 )
E o 1.0
S 0.9

In the above equation, is the Poissons ratio which 0.8


0
should be assumed for the soil, and o is the depth factor, 0.7

its value can be obtained from the opposite figure for 0.6
plate tested at depth D below the ground level. If the plate 0.5
0.1 0.2 0.5 1 2 5 10 20
is circular, an appropriate equation based on theory of D/B
elasticity should be used.
Note that such estimates of E would also be affected by differences of consistency in the
zones of influence.

Penetration Testing
Penetration testing is the most common means of testing cohesionless soils in-situ. The
Standard Penetration Test has been most widely correlated with engineering properties and
the test has gained wide acceptance. This test is normally carried out down a borehole and
most site investigation drilling rigs have the equipment fitted. In the test a 63.5 kg mass is
dropped through a distance of 750 mm onto drilling rods which have a split-spoon sampler
attached at the bottom of the hole. This sampler has an outer diameter of 50 mm and an inner
diameter of 35 mm and at the completion of the test a disturbed sample can usually be
recovered from the sampler. The number of blows required for each 150 mm of penetration is
counted as the sampler is driven through a total distance of 450 mm. Thus three blow counts
would be obtained and the results recorded as, say, 3, 6 and 4. The standard penetration
number, N, is computed as the blows required for the last 300 mm of penetration which is
found by adding the last two blow counts together. Thus in the example N would be 6+4=10.
Further details on the test procedure may be found in the Australian Standard AS1289-2004.
Typically, SPT N-values are measured at 1.5m intervals which correspond to the standard
lengths of auger drilling rods. SPT tests should also be performed where there is a change in
material type (as the layer may be less than 1.5m thick and may therefore be missed in
testing). For some jobs it may be important to have closer spacing particularly near the
surface. It is recommended that the first SPT test be done near the surface (say 0.5m in natural
soil). Settlements beneath shallow footings will be governed by the properties of the near
surface soils and less so by those at depths of 1.5m or greater.
Different types of hammers are used to drive the SPT sampler into the ground, applying
different amounts of energy and hence give different N values for the same density of soil.
SPT values are often corrected to an energy ratio of 60% (N60).
Studies indicate that the measured value of N, let it be called N', should be corrected for
overburden pressure in order to give a design value N. The so called Peck equation is widely
used to correct N:
N / N = 0.77 log10 (2000 / o)
where o (kPa) is the effective vertical stress at the depth of test. The equation is
recommended to be used only when o is greater than 25 kPa. In very fine sands and silts
below the water table special care is needed and the above correction should not be applied.
These are particularly difficult soils to treat. Terzaghi and Peck suggested the following
equation for correcting N values in these soils if N is greater than 15:
N = 15 + 0.5(N - 15) if N > 15
CVEN3201 Applied Geotechnics 53 Footings on Cohesionless Soils

It is important that drilling mud be used below the water table to minimise soil disturbance.
Rods should also be extracted slowly prior to testing to avoid suction or blow-in effects that
loosen the soil, and hence give incorrect low N values.
Many methods for estimating the material properties from in-situ SPT data have been
proposed by researches in different countries. Those given in this course are only examples
that are most commonly cited. They should be used with caution, because they may not be
applicable locally. Also in many of the correlation equations for material properties and for
bearing capacity of shallow foundations, the correction of the overburden pressure and the
water table, is built in the equations. Such correlation equations require the N value measured
directly from the field as input. A practitioner should compare the local correlations with
them, and as a trend develops, revise the equations.
The value of N has been widely correlated to other properties of soils, an example of which is
given in the following table.
Description V. loose Loose Medium Dense V. dense
N 04 4 10 10 30 30 50 >50
(o) 25 32 28 35 30 40 35 43 >38
t kN/m3 11 16 14 18 17 20 18 22 20 23
Relative density 0 0.15 0.15 0.35 0.35 0.65 0.65 0.85 0.85 1.0

The relative density is defined as:


e e min
Dr
e max e min
where e is the void ratio of the soil in the ground, emin and emax are the minimum and
maximum values of void ratio measured in the lab for the same material in dense and loose
states.
Hatanaka and Uchida (1996) give a correlation between SPT N-value and friction angle, , as:
(deg) 20N 20
The SPT is an example of a dynamic penetration test. There are other forms of dynamic
penetration test, for example the code AS 1289 describes the Perth Sand Penetrometer which
is a portable test where a 9 kg weight is dropped through 600 mm onto a flat ended rod of
10 mm diameter. The blows required for each 100 mm of penetration are measured and N is
given as the blows for 300 mm. This test has been correlated for the so called Perth sand, a
fine uniform grained silica sand. For other sands it should be correlated against the SPT if its
results are to be used in place of the SPT. The test is a very useful means of assessing the
variability of density from place to place. AS 2870 describes a test which may be used for
small residential buildings. A 50 mm square wooden stake is driven with a 5 kg hammer. The
soil is loose if the stake is easily driven, medium dense if some effort is required to drive it
and dense if it is only possible to drive the stake a small distance.
In recent years static penetration tests are being used more often. In these tests the rods are
pushed rather than driven into the soil. Variations of these tests are the Dutch Cone,
Piezocone, and Cone Penetration Test (CPT). Here the pressure required to push a cone with
an apex angle of 60o and a net area of 10 cm2 is measured continuously. The test normally has
two rods which can be pushed, one connected to the cone (which measures the cone
resistance, qc in kg/cm2) and one connected to a friction sleeve. This latter device allows for
measurements to be made of the side friction which is useful in the design of piles.
Correlations are available for the design of footings based on cone testing. One such
correlation for sand is:
Allowable Bearing Pressure (in kg/cm2) = qc / 10 for a settlement of 25 mm.
CVEN3201 Applied Geotechnics 54 Footings on Cohesionless Soils

A CPT test allows nearly continuous testing at many sites which is often valuable. Cone
testing is more established as a means of testing cohesive soils and some care is needed with
the interpretation of results for cohesionless soils. No sample can be recovered during the test.

Prediction of Footing Capacity and Settlement Based on In-situ Test Results


Modified Meyerhof method
Over the years many correlations between N and the likely settlement of a footing have been
proposed. The so called modified Meyerhof approach is widely accepted although it is
recognised that the approach is on the conservative side, i.e., it tends to overestimate
settlement.
Based on the Meyerhof (1956) approach the allowable net increase in the soil pressure, qa in
kPa, can be obtained from the following relationships:
qa = 19.2 N Kd for B < 1.2m
qa = 12 N [( B + 0.3) / B] Kd for B > 1.2m
where N is the weighted average SPT value over the footing zone of influence and Kd is a
depth factor. For a footing with a width B which is embedded to depth D, Kd is given by:
Kd = 1 + 0.33 D / B 1.33
The value qa is the allowable net increase in soil pressure (in kPa) for a settlement of 25 mm.
For settlements other than 25 mm a linear correlation is used, i.e.,:
qs = qa S / 25
where qs is the allowable pressure increase for a settlement of S mm.

Parry method
Another approach which is mainly based on the elastic theory is suggested by Parry (1971). In
this approach the following empirical relationship is used to relate the Youngs modulus of
soil, Es in kPa, to uncorrected SPT value:
Es = 4900 N
Then typical elastic solutions, such as the one given below, are used to find the footing
settlement:
S = q B (1 - 2) 1 / E
A study of Parry's equation for 24 structures showed that the ratio of S (predicted) to S
(observed) averaged 1.2 and ranged from 0.8 to 2.6.
Note that there are many other correlations between the stiffness parameters of sand and SPT
or CPT values, some of them have been presented in a table under Choice of elastic
parameters in the previous section of the
notes. Depth, z
0 z N
Example Fill
2 Fine sand 2.5 37
A large span warehouse building is supported
by columns on concrete spread footings. The 4 WT 4.0 22
columns transfer a load of 14000 kN on the 5.5 28
4 m by 7 m spread footings with their base 6
2 m below ground level. An extensive site Coarse sand 7.0 36
8
investigation involved many readings of SPT 9.0 38
N-values which did not vary greatly from 10
11 44
place to place. As a result the conditions of 12
foundation soil could be idealised as shown in Sandstone
the figure opposite.
CVEN3201 Applied Geotechnics 55 Footings on Cohesionless Soils

The prediction of settlement proceeds as follows. The value of B for the footing is 4 m so its
zone of influence is 2B or 8 m. Thus N values between 2 m (base depth) and 10 m are
important. The N value at 2.5 m (N=37) is ignored because it appears high and out of
sequence with the remaining N values which increase with depth. Ignoring a high value of N
is being on the conservative side as well.
The relevant N values are corrected using Pecks suggestion and t = 18 kN/m3 as in the
following table (note the position of the water table in calculating the effective stress).

z (m) o (kPa) N N
4.0 72 22 24
5.5 84 28 29
7.0 96 36 36
9.0 112 38 36

A single design value of N is required for calculation of bearing capacity of the foundations.
A weighted average will be used with N1 of 24 (N between 2 and 4.7m), N2 of 33 (N between
4.7 and 7.5 m) and N3 of 36 (N between 7.5 and 10 m.). The weighted average N is then
calculated as:
N = ( 3N1 + 2N2 + N3 ) / 6 = (72 + 66 + 36)/6 = 29
Using the Meyerhofs equation the allowable net bearing pressure as:
qa = 1229 ( 4.3 / 4 ) ( 1 + 0.33 (2 /4) ) = 435 kPa for 25 mm settlement
Thus a pressure of 435 kPa would cause 25 mm settlement. For the column load of 14000 kN,
the net pressure at the base of the foundation is 14000/(47) or 500 kPa. Thus the predicted
settlement would be 500/43525 or 29 mm.
Using Parrys equation the uncorrected weighted average N is of the order of 28, giving:
E = 490028 = 137200 kPa
The geometric factors are H/B=10/4=2.5, L/B=7/4=1.75 and D/B=2/4=0.5 giving of 0.88
and of 0.8. Therefore, the settlement of the foundation can be calculated as:
S = q B (1 - 2) / E = 5004 (1 - 0.252) 0.88 0.8 / 137200 = 10 mm
In conclusion it is seen that Meyerhofs equation gives S = 29 mm which is known to be
conservative while Parry's method gives 10 mm. If as an engineer it was necessary to estimate
the settlement of the footings it would be reasonable to say: the settlement should range from
10 to 29 mm with a most probable value of 15 mm. In fact the settlements of 8 footings were
monitored after the building was constructed. They gave values ranging from 8 mm to
16 mm.
Note that other correlations give Es much smaller than that obtained using Parrys equation,
which would results in larger settlements for the foundations.

Strain Influence Factor: Schmertmann method


Schmertmann (1970 and 1978) developed a method to estimate the vertical settlement of
footings on sand when cone penetration test (CPT) data is available. This method is based on
the strain influence approach and elastic theory. In this method it is assumed that the
maximum average vertical strain in the soil beneath a loaded foundation of width B occurs at
a depth of B/2 (for square or circular footings) or B (for strip footings) below the base of the
foundation. The extent of the strain distribution is assumed to be significant up to a depth of
2B (for square or circular footings) or 4B (for strip footings). This depth is called the depth of
influence.
CVEN3201 Applied Geotechnics 56 Footings on Cohesionless Soils

The settlement of a foundation embedded at depth D in sand can be obtained by integrating


the strains over the depth of influence and can be written as:

S C1C2 z dz C1C2 q q o
Layers
Iz
z
0 Es

The definitions of parameters used in the above relationship are given below:
qo
C1 : Foundation embedment correction factor, can be estimated as: C1 1 0.5 .
(q q o )
C2 : Creep correction factor, can be estimated as: C2=1+0.2log(10 t), where t is time in years.
q : Stress at foundation level (allowable bearing capacity when the settlement S is at the
allowable value).
qo : Overburden pressure at the foundation level (effective pressure if water table is above the
footing base)
Iz : Strain influence factor calculated at the middle of a layer
Es : Modulus of elasticity of the soil layers
z : Thickness of each layer
The variation of the strain influence factor, Iz, with depth below the foundation is shown in
the following figure. The influence factor beneath the footing base, at z=0, is 0.1 for square or
circular footings and 0.2 for strip footings.

B L = footing length

Df q o D f qa
Iz
z1 Es1
z1
z2 Es2 Izp
z
z3 Es3 z2

z4 Es4

z
Schmertmann et al. (1978) suggested the peak value of the stress influence factor, Izp, to be:
q qo
I zp 0.5 0.1
v
where (q-qo) is the net bearing pressure and v is the initial vertical effective stress at the
depth of the peak strains.
The values of the variables in the Schmertmann method are specified in the following table
with the peak value of the stress influence factor Izp for square/circular and strip foundations.
CVEN3201 Applied Geotechnics 57 Footings on Cohesionless Soils

Footing shape z* Iz Variation of Iz


0 0.1 z*
I z 0.1 2I zp 0.2
Square or circular B
z1 = 0.5B Izp
2 z*
I z 2 I zp
z2 = 2B 0 3 B

0 0.2 z*
Strip I z 0.2 I zp 0.2
z1 = B Izp B
1 z*
z2 = 4B 0 I z 4 I zp
3 B
z* is the depth below the footing base

Rectangular footings with L/B10 can be regarded as strip footings. For rectangular footings
with L/B<9, the strain influence factor is calculated by linearly interpolating those of the
square and strip footings.

Iz Izs
1
Izc Izs L 1
9 B
where, where L is the length and B is the width of the footing, Izs and Izc are the strain
influence factors for a square footing and a continuous (strip) footing, respectively.
Modulus of elasticity
The modulus of elasticity, Es, can be determined from empirical correlations with the cone
penetration test (CPT) or the standard penetration test (SPT).
Schmertmann et al (1978) further suggested the following values for normally consolidated
soils:
Es = 2.5 qc for square or circular footings
Es = 3.5 qc for strip footings
The recommended value of Es for overconsolidated soils is:
Es = 6.0 qc for overconsolidated soils
Sands that are compressed before can be regarded as overconsolidated soils.
The correlation between SPT N value and the CPT cone resistance can be selected as one
suggested by Meyerhof:
qc (MPa) = 0.4 N
Example
A 3m3m square footing carrying a load of 5400kN is founded 2m below the ground surface
in sand. The water table is at the base of the footing. The total unit weight of the sand can be
taken as t=20kN/m3 above and below the water table. CPT cone resistance, qc, for the sand is
given below.
Depth (m) 1.5 2.5 3.5 4.5 5.5 6.5 7.5 8.5 9.5
qc (MPa) 2.5 2.5 3 3.5 4.5 4 4 6 8
Estimate the settlement of the footing 10 years after the construction using Schmertmann
method.
Note: The cone resistance, qc, is normally measured continuously. For simplicity, an average
is taken over 1m thick layers of soil in this example.
CVEN3201 Applied Geotechnics 58 Footings on Cohesionless Soils

q = 5400/(33)=600kPa
qo = 220 = 40 kPa
q = (q - qo) = 600 - 40 = 560 kPa
v = 3.520 1.59.8 = 55.7 kPa`
Izp 0.5 0.1 560 / 55.7 0.82
C1 = 1 0.5 (40/560) = 0.96
C2 = 1+0.2 log(1010) = 1.4
The depth of influence beneath the footing is divided into 7 layers and different parameters
are calculated and tabulated below.
z qc Es=2.5qc Izz/Es
layer Depth z* (m) z/B Iz
(m) (MPa) (kPa) 1000
1 2-3 0.5 1 2.5 6250 0.167 0.337 0.0540
2 3-3.5 1.25 0.5 3 7500 0.417 0.694 0.0462
3 3.5-4 1.75 0.5 3 7500 0.583 0.767 0.0512
4 4-5 2.5 1 3.5 8750 0.833 0.632 0.0722
5 5-6 3.5 1 4.5 11250 1.167 0.451 0.0401
6 6-7 4.5 1 4 10000 1.500 0.271 0.0271
7 7-8 5.5 1 4 10000 1.833 0.090 0.0090
z* is the depth to the middle of layers below the footing. Total 0.30

The settlement is calculated as:

S C1C2 q q o
Layers
Iz
z = 0.961.4(560)(0.30/1000) = 0.226m = 226mm
0 Es
CVEN3201 Applied Geotechnics 59 Appendices

APPENDICES

Presumptive Bearing Capacity


Allowable bearing
Class Material value, tons/ft2
(approx kPa)3
Massive crystalline bed rocks, such as granite, gneiss, trap rock,
1 100 (10725)
etc.; in sound condition
2 Foliated rocks, such as schist and slate, in sound condition 40 (4290)
Sedimentary rocks, such as hard shales, siltstones, or sandstones, in
3 15 (1610)
sound condition
4 Exceptionally compacted gravels or sands 10 (1070)
5 Gravel; sand-gravel mixtures; compact 6 (645)
6 Gravel, loose; coarse sand, compact 4 (430)
Coarse sand, loose; sand-gravel mixtures, loose; fine sand,
7 3 (320)
compact; coarse sand, wet (confined)
8 Fine sand, loose; fine sand, wet (confined) 2 (215)
9 Stiff clay 4 (430)
10 Medium stiff clay 2 (215)
11 Soft clay 1 (107)
Presumptive bearing capacity from New York State Building Construction
Code, 1977

Notes:
(1) Presumptive bearing values apply to loading at the surface or where permanent lateral support for
the bearing soil is not provided.
(2) Except where, in the opinion of the enforcement officer, the bearing value is adequate for light
frame structures, fill material, organic material, and silt shall be deemed to be without presumptive
bearing value. The bearing value of such material may be fixed on the basis of tests or other
satisfactory evidence.
(3) 1ton/ft2 107kPa
CVEN3201 Applied Geotechnics 60 Appendices

Presumed bearing
Material Remarks
value (kPa)
Hard igneous or gneissic rocks 10000
Only sound
Hard limestone and sandstone 4000
unweathered rocks
Schists and slates 3000
Hard shales and mudstones; soft sandstones 2000 Thin bedded or
Soft shales and mudstones 600-1000 shattered rocks must be
assessed after
Hard sound chalk; soft limestone 600
inspection

Compact gravel or sand/gravel >600


Medium-dense gravel or sand/gravel 200-600 Providing width B
Loose gravel or sand/gravel <200 1m and groundwater
Compact sand >300 level B below base of
Medium-dense sand 100-300 footing
Loose sand <100

Very stiff boulder clays; hard clays 300-600


This group is
Stiff clays 150-300
susceptible to long-
Firm clays 75-150
term settlement
Soft clays and silts <75
Presumptive bearing capacity from British Standard (BS8004, 1986)

Notes:
(1) These are intended as a guide for preliminary design purposes only.
(2) They are gross values, with allowance for embedment.

The following table shows a comparison of the presumptive bearing capacities given in 3
different countries for some soil types. As can be seen there is a reasonable difference
between some of the values given.
Material Ordinance 70 New York (kPa) British (kPa)
Loose sand 150 215 <100
Stiff clay 300 430 150-300
Soft clay 100 107 <75
Comparison of allowable bearing values
CVEN3201 Applied Geotechnics 61 Appendices

References:
Bowles J.E. (1996) Foundation Analysis and Design, 5th Edition, McGraw-Hill
Chen, W.F. (1975) Limit Analysis and Soil Plasticity, Elsevier
Das, B.M. (1999) Principles of Foundation Engineering, 4th Edition, PWS Publishing
Das, B.M. (2000). Fundamentals of Geotechnical Engineering, Thomson-Brooks/Cole
Hansen, J.B. (1970) A revised and extended formula for bearing capacity, Danish
Geotechnical Institute, Bulletin 28, Copenhagen.
Hatanaka, M. and Uchida, A. (1996) Empirical correlation between penetration resistance and
internal friction angle of sandy soils, Soils and Foundations, 36(4), 1-9
Mesri, G., Lo, D.O.K., and Fang, T.W. (1994) Settlement of embankments on soft clays,
Geotechnical Special Publication 40, ASCE, Vol 1, 8-56
Meyerhof, G.G. (1956) Penetration tests and bearing capacity of cohesionless soils, Journal of
the Soil Mechanics and Foundations Division, ASCE, 82(SM1), 1-19
Meyerhof, (1963) Some recent research on the bearing capacity of foundations, Canadian
Geotechnical Journal, Vol. 1(1), 16-26
Meyerhof, G.G. (1965) Shallow foundations, Journal of the Soil Mechanics and Foundations
Division, ASCE, 91(SM2), 21-31
Parry, R. H. G. (1971). A direct method of estimating settlements in sand from SPT values,
Proc. International Symposium in Structure and Foundations, Midlands Soil Mechanics and
Foundation Engineering Society, Birmingham, 29-37.
Schmertmann, J.H. (1970) Static cone to compute static settlement over sand, Journal of the
Soil Mechanics and Foundations Division, ASCE, 96(SM3), 1011-1043
Schmertmann, J.H., Hartman, J.P. and Brown, P.R. (1978) Technical note: improved strain
influence factor diagrams, Journal of the Geotechnical Engineering Division, ASCE,
104(GT8), 1131-1135
Terzaghi, (1943) Theoretical Soil Mechanics, Wiley, New York
Tschebotarioff, G. P. (1951) Soil mechanics, foundations, and earth structures, McGraw-Hill
Vesi, A.S. (1973) Analysis of ultimate loads of shallow foundations, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 99(SM1), 45-73
PILE FOUNDATIONS
Piles are deep foundations which transmit the loads of superstructures through relatively weak
or loose strata deep into the ground and to the stronger underlying soil or rock. Piles replace
shallow foundations when the soil close to the surface does not have sufficient bearing
capacity to withstand loads of the superstructure, or when the settlement of a shallow
foundation is excessive. Piles may also be used to carry uplift loads, to carry loads below
scour level in marine environments, to resist lateral loadings, or to reduce the settlement of
shallow pad or raft foundations. Piles support the applied load by end bearing, adhesion or
friction developed along their shaft, or both. They generally extend to a depth greater than
3 m below the surface. A pile can be installed into the ground by driving, screwing, jacking,
vibrating, drilling or any other methods. A number of piles may be installed in a close
proximity to each other, usually having a common cap, to form a pile group.
Pile foundations are more expensive than shallow foundations. However, they have extra
capabilities and are commonly used in the following situations:

a) Where a stiff layer of soil or rock exists below an upper softer layer. It may therefore be
economical to transfer the structure loads to this lower layer using piles.
b) Where the upper layer is highly compressible and too weak to support the load transmitted
by the super-structure. Gradual transfer of the load to the soil via pile-soil friction and
adhesion is therefore required.
c) To resist uplift. This may be caused by wind loading, machine foundations or may be due
to the use of piles as tension members, e.g., in soil nailing or anchors.
d) & (e) To resist horizontal loads. These loads may be induced by wind, earthquake loads or
may be due to the type of structure itself, e.g., bridge abutments, transmission towers. Piles
can transfer lateral loads much better than shallow foundations due to their depth.
f) Offshore structures where the upper layer is subject to erosion.
g) Adjacent to excavations where excessive ground movements are anticipated if shallow
foundations are used.
h) Where expansive and/or collapsible soils prohibit the use of shallow foundations due to
excessive movement.
Often a group of piles needs to be designed to carry a heavy load. In this case a pile cap is
used to assist in transferring the load to all piles (cases e, f and g).
CVEN3201 Applied Geotechnics 64 Pile Foundations

Classification of Piles
Piles can be broadly classified according to the pile material (timber, steel, concrete), the way
they transmit their loads to the ground, the method of installation (driven, driven and
cast-in-situ, bored or drilled shafts, screwed) and the effects of the installation procedure on
the surrounding soil. The last one is a more useful means of classification for geotechnical
design purposes and two main types can be identified: displacement piles and
non-displacement piles. However, there is a range of other piles installed by various
techniques that can be classified as partial displacement, post grouted, preloaded non-
displacement piles.

Displacement piles
Installation of these piles causes relatively large horizontal displacements and movements of
the ground through which the piles are being installed. These piles may be installed by
hammering, pushing, jacking, vibrating, screwing or other methods to push them into the
ground. Depends on the method of installation, these piles may be subdivided into the
following categories:
Driven cast-in-place piles are formed in-situ by driving a liner, either permanent or
temporary, and filled with concrete. Temporary liners are extracted during concreting or
grouting.
Driven preformed piles are prefabricated piles driven to the ground and left in position.
They may be formed from concrete, steel hollow section or H section, timber, etc.
Installation of steel H piles and open-ended steel tube piles causes minimal lateral ground
movements, and therefore these piles are often classified as partial displacement piles.
Continuous flight auger piles are formed in the ground by drilling with a hollow flight
auger. After drilling the auger is withdrawn while the cavity below the auger tip being
gradually filled with concrete injected from the auger tip under pressure.
Screwed piles are pushed by screwing a threaded tube into the ground. Often screwed piles
are made up of hollow section steel tube with helical plates attached to their tip.

Non-displacement piles
These piles are formed in-situ by removing soil to from a void in the ground which is then
filled with concrete or grout. The soil may be removed using either rotary drilling or
percussion, reverse circulation, grabbing, chiselling, and mechanical or hand excavation
methods. These methods generally induce only small ground movements. During removal of
the soil, the sides of the excavated void may be supported by permanent liners or temporary
liners or drilling mud or continuous flight auger. In strong soil the excavated void may be
stable and is left exposed during excavation. These types of piles are commonly referred to as
bored cast-in-place or cast-in-situ piles.

Classification based on material


Driven piles may be made of steel, pre-cast concrete or timber. Steel piles are usually rolled
sections, e.g., H beams, I beams, circular hollow sections and rectangular hollow sections.
Other types of steel piles include hollow segmented piles and sheet piles. Sheet piles are
almost exclusively used to retain water due to their interlocking nature and strength.
Timber piles are often useful due their low cost and low corrosive potential in saturated
environments. Timber piles are generally not spliced due to the weak nature of their joints.
They are, however, easily cut-off to length.
A summary of the different types of piles and their advantages and disadvantages is shown in
the following table.
Typical Max Usual load Approx
Pile Type Advantages Disadvantages
length (m) length (m) (kN) max load

Easy to handle Relatively costly


See BHP Can stand high driving stresses High level of noise during compaction
Steel 15-60 Unlimited 300-1200
handbook Can penetrate hard layers Subject to corrosion
High load capacity May be damaged or deflected during driving.

Can sustain hard driving


Difficult to achieve proper cut off
Precast Corrosion resistant
10-15 30 300-3000 8000-9000 Difficult to transport
concrete Can be easily combined with concrete
Relatively costly
superstructure

Prestressed
10-35 60 300-3000 7500-8500 As above As above
concrete

Relatively cheap
Cased Difficult to splice after concreting
5-15 15-40 200-500 800 Possibility of inspection before pouring
bored piles Thin casings may be damaged during driving
Easy to extend

Relative cheap
Possibility of inspection before pouring Susceptible to poor compaction
Uncased
5-15 30-40 300-500 700 Fast construction Excavation wall may cave in
bored piles
Quiet Debris at the base may cause excessive settlement
Good uplift resistance

Economical Decay above water table


Easy to handle Can be damaged during driving
Timber 10-15 30 100-200 270
Resistant to decay under water Low load bearing capacity
Easy to cut off Splice no good in tension
Pile types and their typical length, capacity, and advantages/disadvantages
Classification based on load transfer mechanism
Resistance of piles to vertical loads is provided
mainly by either a combination of pile shaft
resistance and pile end bearing resistance, or only
by end bearing resistance. Therefore, piles can
also be classified based on the way they transmit
the applied force to the ground:
a) End bearing piles transmit their forces to the
ground through the reactions developed mainly
at their base. The shaft of the piles has little
contribution to their resistance and therefore
the shaft resistance is often ignored in design.
b) Floating piles transmit their forces to the
ground through the reactions developed along their shaft and on their base. In design of
these piles both shaft resistance and end bearing resistance are considered. For long
floating piles the base resistance has little contribution to the overall pile capacity and is
often ignored.

Design Requirements
The design of a pile shall take into account the ultimate strength, serviceability, and durability
of the pile. Different methods can be used to evaluate the ultimate strength of piles, among
them are: static analysis based on shearing resistance of the soil, static load testing, dynamic
analysis using wave equation or other driving formulae based primarily on the penetration
resistance, and dynamic load testing.
Piles should be designed with the following possibilities in mind:
- Scour by water or air action,
- Uplift by frost heave or expansive soils,
- Negative friction from shrinkage or consolidation,
- Alterations in the ground water table.
The current Australian pile design code (AS 2159-1995) is based on the load and resistance
factor design (LRFD) approach to design. This is a significant change from the previous
code (AS 2159-1978) which was based on the classic global factor of safety approach. While
the new code has been significantly modernised, it contains less design guidelines than its
predecessor. Consequently many engineers are using the previous code for design calculations
while adopting the LRFD in their design approach. Note that AS 2159-1995 refers extensively
to other references for the actual design calculations. It states that: In most cases it is
necessary to assess the sensitivity of a design by use of alternate methods of calculation
In these notes the most basic method of design of piles, the static analysis, will be covered.
Static analysis involves the estimation of ultimate load by material properties and theoretical
equations. It is based on bearing capacity principles and is useful for initial estimation of pile
sizes, even when pile load tests will be subsequently undertaken. The static analysis method
presented here to calculate the ultimate capacity of piles is mostly based on (AS 2159-1978).
Students are encouraged to refer to alternative design methods.
CVEN3201 Applied Geotechnics 67 Pile Foundations

Axial Capacity of Single Piles


The ultimate axial capacity of a pile can be obtained using a static analysis utilising
conventional soil mechanics techniques. In this method the ultimate capacity of a single pile
loaded in compression, Pu, can be determined as:
Pu = Ps + Pb W
where W is the weight of the pile and Ps and Pb are the total shaft resistance and the total base
resistance, respectively. The above equation can be written as:
Pu = fs As + (fb + qo) Ab -W
where fs is the average shaft (or skin) resistance mobilised around the pile shaft, fb is the base
ultimate capacity, qo is the surcharge or overburden pressure at the base level, As is the
perimeter area of the shaft, Ab is the area of the base of the pile, and W is the weight of the
pile. If the soil properties vary below the pile base, the capacity of the base is calculated using
the average properties over two diameters below the base of the pile.
It is usually assumed that WAb qo. Therefore the above equation reduces to:
Pu = fs As + fb Ab
The shaft resistance and the base resistance, fs and fb, can be derived using Mohr-Coulomb
failure criterion and bearing capacity equations, respectively. Their values depend on the soil
type, pile type and the method of installation. The general principles of soil mechanics can be
applied; undrained analyses using total stresses together with undrained strength parameters,
and drained analyses using effective stresses together with drained strength parameters.
Generally two methods are used to determine the capacity of piles, one being used for purely
cohesive soils (often called method) and the other for cohesionless soils. For soils which
have both cohesion and friction, combinations of the both methods may be appropriate.

Cohesive soils
Clay often exhibits its lowest strength when it is saturated and subjected to undrained rapid
loading. Therefore an undrained analysis using total stresses is considered to be more
appropriate for piles embedded in clay. In this type of analysis it is often assumed that the soil
has a zero friction angle, u = 0,
and the shearing strength of the
soil is due only to the undrained
cohesion, cu. Therefore the
ultimate shaft resistance of a
single pile in a uniform soil can be
expressed as:
fs = ca = cu
where ca is the adhesion between
soil and pile which is related to the
undrained shear strength of the
soil by an adhesion factor . The
value of as a function of cu can
be obtained from the figure
opposite (extracted from AS2159-
78). Note that this figure shall be
used in the absence of any other
data for . Great variability has
been experienced with values of
for piles in stiff clays. Adhesion reduction factor
CVEN3201 Applied Geotechnics 68 Pile Foundations

The total shaft resistance for a circular pile with length L and diameter d embedded in a
uniform soil (having uniform shear strength, cu) is:
Ps = fs As = cu ( d) L
Often the shear strength, cu varies with depth. In this case the soil around the pile can be
subdivided into several layers, each having a uniform shear strength, cu. The total shaft
resistance for these soils, or for layered soils in general, can be calculated as:
L
Ps = d c u dL or Ps d c
all layers
u L
0

For slender piles, those with high length-to-diameter rations, the shaft resistance, fs, needs to
be reduced by a reduction factor, Rs. The reduction factor for pile slenderness is given by
(Semple & Rigden, 1984):
1 for L / d 50

R s 1 3(L / d 50) / 700 for 50 L / d 120
0.7 L / d 120
for
where L and d are the length and the diameter of the pile, respectively.
The base of the pile can be treated as a footing embedded deep into the ground and the base
resistance can be calculated using the bearing capacity equation for cohesive soils under
undrained conditions:
qu = cub Nc + qo
The total base resistance for a pile with a base area Ab is:
Qu = (cub Nc + qo) Ab
As described before, the value of qo Ab is approximated to be equal to the weight of the pile,
W, and therefore can be eliminated from the equation. Therefore the base resistance can be
calculated as:
Pb = cub Nc Ab
In the above equation cub is the average undrained shearing strength of the soil within a depth
of two pile base diameter below the pile tip, and Nc is the bearing capacity factor which
includes the depth factor. For L/d 4, Nc will be equal to 9, otherwise Nc = 5.6 + 0.85 L/d9.
The ultimate capacity of a single pile having 4 < L/d < 50 and embedded in a uniform soil can
be written as:
Pu = fs As + fb Ab = cu ( d) L + 9 cub Ab

Cohesionless soils
Granular soils are generally frictional and have zero or very small cohesion. A drained
analysis, using drained strength parameters c & and effective stresses, is considered to be
more appropriate for piles embedded in granular soils. In this type of analysis it is often
assumed that the soil has a zero cohesion, c = 0. Therefore the shaft friction in granular soils
depends on the soil-pile friction angle and the normal effective stress acting to the shaft, n.
For a non-tapered vertical pile in a horizontal soil deposit the normal stress (to the shaft) is
equal to the horizontal stress, n =h = K v. Therefore the shaft resistance can be
presented as:
fs = n tan = K v tan = (tan ) v
where v is the effective vertical stress at the point of interest along the shaft, is the
effective interface friction angle between soil and shaft, and K is the coefficient of lateral
earth pressure.
CVEN3201 Applied Geotechnics 69 Pile Foundations

The friction angle between pile and soil, , depends on the soil type and the method of
installation, and is mostly less than the soil
internal friction angle, . The table opposite Interface materials /
shows typical values of / for various interface Driven Piles
materials. Sand/rough concrete 1.0
The post installation value of lateral earth Sand/smooth concrete 0.8 1.0
pressure, K, also depends on the type of soil Sand/rough steel 0.7 0.9
surrounding the pile and the method of Sand/smooth steel 0.5 0.7
Sand/timber 0.8 0.9
installation of the pile. Some suggested values of
Bored Piles 1.0
K/Ko are given in the following table, where Ko is
the coefficient of lateral earth pressure at rest.
Type of Pile Soil condition K/Ko Comments
Dense 1/2 Soil is removed by water jet
Jetted pile
Soft 2/3 during construction
Stiff clay 1 For cased piles the value is
Bored piles cast in place
Using bentonite slurry 2/3 somewhere in between
Dense 3/4 This can be either an H pile or
Driven pile steel
Soft 5/4 an open pipe pile
Dense 1 Usually suffers large
Driven pile precast concrete
Soft 2 displacements during driving
Typical coefficient of lateral earth pressure

The term K tan will therefore be a function of the soil type and the method of installation
and can be given as a single variable, F = (tan ), known as the coefficient of shaft friction.
Values of the coefficient of shaft resistance, F, for soils with different consistencies and
different methods of installations are given in the following table. Therefore the shaft
resistance can be given simply as:
fs = F v
For a pile in layered soils with different relative densities an appropriate values of F shall be
used for different sections of the pile in each soil layer.

F = K tan Nq
Soil Relative
consistency density Driven Bored, cast-in-situ Driven Bored, cast-in-situ
piles piles piles piles
Loose 0.20 - 0.40 0.8 0.3 60 25
Medium 0.40 - 0.75 1.0 0.5 100 60
Dense 0.75 - 0.90 1.5 0.7 180 100
Shaft friction coefficients and bearing capacity factors

Note that for long slender pile, where L/d>50, the shaft resistance should be reduced by Rs, as
described in the previous section.
The base resistance can be calculated using the bearing capacity equations. The contribution
of soil cohesion to the bearing capacity is equal to zero. The contribution of the self weight of
the soil, the third term in the bearing capacity equations, is assumed to be insignificant
compared to the second term, i.e., the contribution due to the overburden pressure. Therefore
the base resistance may be written as:
fb = qo Nq = vb Nq
where vb is the vertical effective stress at the level of the pile base and Nq is the bearing
capacity factor which includes depth and shape factors. Values of Nq for different soils and
different methods of installations are also given in the above table.
Note that the geostatic conditions in a soil mass indicate that both v and h increase linearly
CVEN3201 Applied Geotechnics 70 Pile Foundations

with depth. However, AS2159-1978 applies a cap on the maximum vertical effective stress,
and therefore on the maximum horizontal stress, around a pile. It assumes that the vertical
effective stress reach its maximum value at a depth equivalent to 10-20 pile diameters (zL),
depend on the relative density of soil. Therefore, for the purpose of calculating the horizontal
soil pressure on piles and vertical soil pressure at the level of the base, a limit is applied on the
maximum vertical stress. However, this limitation has not been appeared in more recent
methods of static analysis of piles, and therefore is not recommended in these notes.

Pile Driving Formulae


The ultimate capacity of driven piles can be calculated based on the actual driving
performance in the field using various formulae. These formulae are mostly semi-empirical,
some include a factor of safety, and therefore care needs to be taken in choosing the right
units for each parameter.
Pile driving formulae are usually functions of energy used to drive a pile and are often
presented in terms of the weight and the fall of hammer, pile movement or set measured
during the last hammer blow or the average movement in the last few hammer blows. The
rational basis of many pile formulae is that the energy input (spent by the equipment) must be
equal to the energy used to drive the pile plus energy lost. The energy input is known or can
be measured. The energy lost can be estimated. The difference between the energy input and
energy lost can be related to the driving resistance, Pu, and pile movement.
An example of pile driving formula, and probably the oldest one, is known as Engineering
News formula:
W .h
Pu r
SC
where Wr is the weight of hammer (same unit as Pu, e.g., kN), h is the hammer drop (mm), S
is the pile movement (mm), and C is a constant equal to 25 mm for drop hammers or 2.5 mm
for steam hammers.
A more comprehensive pile deriving formula has been proposed by Hiley as:
e.Wr . h Wr n 2 Wp
Pu
S 0.5(k1 k 2 k 3 ) Wr Wp
In this equation e is hammer efficiency (typically between 0.75 to 1), Wp is the weight of the
pile, n is coefficient of restitution (typically between 0.2 to 0.5), k1 is elastic compression of
pile head and cap (typically between 0 to 7 mm), k2 is elastic compression of the pile equal to
(Pu L) / (AEp), k3 is elastic compression of soil (typically between 0 to 4 mm), L, A, and Ep
are length, cross section area, and Youngs modulus of the pile, respectively. The Hiley
formula is valid for any consistent set of units.
Pile driving formulae are simple to apply. However, their application is limited because of the
uncertainties in their evaluation of pile capacity. A safety factor between 3 and 6 is usually
considered when using the pile driving formulae in calculation of the bearing capacity of
piles.
A more sophisticated class of pile deriving formulae is based on the travel of an impulse
(shock wave) down a pile as it is being driven. These formulae are generally known as wave
equation. The wave equation is a differential equation of second order and can be solved by
finite difference techniques. The system is modelled by a computer program where the energy
input and the displacement information (the set after each blow) are used to infer the model
parameters and predict the ultimate pile resistance to a reasonable accuracy.
The Australian piling code AS 2159 defines the wave equation methods as dynamic pile
testing and recognises that these methods give a better indication of pile capacity as compared
to theoretical methods.
CVEN3201 Applied Geotechnics 71 Pile Foundations

Vertical Settlement of Single Piles


An estimate of the settlement of a pile can be obtained from a theoretical analysis based on
theory of elasticity or from load tests on the pile. Load tests are the most reliable means of
estimating the settlement of a single pile.
In general about 80% of the total displacement of a pile is considered to be elastic. The other
20% will normally be due to consolidation and thus controlled by the soil consolidation
properties. Creep is usually ignored in deep foundations. The focus here will be on elastic
settlement.

Floating pile
The settlement of a single floating pile in a deep uniform layer of soil can be obtained from
theory of elasticity as:
P
S I
LEs
where S is the vertical
settlement of the pile, P is the
working load applied to the pile
(unfactored load which is less
than the maximum applied load
in some cases), L is the length
of the pile, Es is the modulus of
elasticity of the soil foundation,
and I is an influence factor that
can be obtained from the chart
opposite. In this chart, the pile
stiffness factor, K, is the
relative stiffness of the pile to
that of the foundation soil, and
is defined as:
E
K p R A
Es
where Ep is the Youngs
modulus of the pile material
and RA is the area ratio of the
pile, that is the ratio of area of
the pile section, As, to gross
area bounded by the outer
circumference of the pile, Ag.
RA = As / Ag.
For solid piles RA is equal to 1,
and for piles with hollow
Settlement influence factor for a single floating pile
sections RA is less than 1.

If the floating pile is embedded in a finite layer of soil the settlement calculated based on the
above procedure shall be corrected to include the effect of the finite soil layer. When a rigid
base exists at depth h below the ground level, the settlement can be reduced by a correction
factor, Rh:
CVEN3201 Applied Geotechnics 72 Pile Foundations

P
S I R h
LE s
The correction factor, Rh, can be obtained from
the figure opposite. P
Ep

End bearing pile


L
The tip of an end bearing pile often rests on a h
layer which has a stiffness much larger than the
stiffness of the soil around the pile. The
settlement of an end bearing pile resting on a
Es
rigid stratum can be obtained as: Rigid base
PL
S M
EpAp R

Correction factor for rigid base at depth h
where Ap is the cross section area of the pile (for
hollow sections Ap=As) and MR is a movement
ratio which can be obtained from the figure
opposite.
Note that the settlement calculated by the above
procedure shall be corrected for cases where the
base of the pile is rested on a stratum which is
stiff but not rigid.

In calculation of settlements, (Eu)s and Es can


be used to estimate the immediate and the final
settlements of the pile a long time after
installation. Values of the soil Youngs
modulus, Es, can be best approximated from a
pile loading test and using the elastic theory to 200 500 2000 5000
interpret the results. This modulus can then be
used to predict the settlement of piles of Movement ratio for end bearing pile on rigid base
different dimensions to the test pile. In cases where no data are available, table below can be
used to select Es for cohesive soils and cohesionless soils. Note that no data are available for
bored piles in cohesionless soils, but it is expected that Es will be lower than that for a driven
pile.
For piles in layered soil, a weighted average stiffness, Eav, along the shaft and ~6D below the
base should be used in calculation of settlements of piles. The weighted average stiffness can
n n
be calculated for n layer of soil as: E i d i / d i ,
i 1 i 1
where Ei and di are the stiffness and

thickness of layer i.

Cohesive soils Cohesionless soil


Undrained Youngs modulus, Es (MPa) Youngs modulus,
Soil consistency
cohesion (kPa) Bored piles Driven piles Es, (MPa)
35 3.8 8.5 Loose sand 42
70 8.5 25.0 Medium sand 70
105 22.0 35.0 Dense sand 90
140 70.0 35.0 Medium gravel 200
Typical values of Youngs modulus of drained soil
CVEN3201 Applied Geotechnics 73 Pile Foundations

Example P1
0.5m
Calculate the ultimate geotechnical capacity of an 11m pile driven
into a thick layer of sand. The diameter of the pile is 0.5m. The 1m
relative density of the sand is Dr = 55% and its average unit weight
is t=18kN/m3. The water table is 1m below the ground level.
Sand
10m
t=18kN/m 3

Example P2
A hollow steel pile of circular cross section has a radius of 0.6m. It is driven 15m into a deep
layer of gravel with a relative density of Dr = 60% and unit weight of t =18 kN/m3. There is
no sign of water in the soil layer.
a) Calculate the ultimate capacity of the pile.
b) Calculate the settlement of the pile due to a load 2000 kN. Assume that the Youngs
moduli of the soil and the steel pile are Esoil = 200 MPa and Epile = 200 GPa, respectively,
and the area ratio of the pile section is RA=0.01.
CVEN3201 Applied Geotechnics 74 Pile Foundations

Capacity of Pile Groups


Vertical capacity of pile groups
When two or more piles are constructed close to each
other, they form a pile group. If the piles are close enough
they interact with each other since the stresses each pile
induces in the soil overlap with those of the adjacent piles.
In a closely constructed pile group, loading on one pile not
only induces settlement on the pile itself, but it may also
result in settlement of the adjacent piles and reduction in
their capacity. If the piles are spaced far apart so that the
stress bulbs generated due to loading on each of them do
not overlap, there will not be any interaction between the
piles and each pile tends to act independently. Therefore,
the spacing between the piles in a pile group is an
important factor that must be considered in the evaluation
of the capacity of the group. Typically, a spacing greater
than 2.5 diameters for floating piles is recommended.

Group of floating piles


The ultimate vertical bearing capacity of
a group of floating piles shall be taken as
the lesser of the following cases:
a) The sum of the capacities of the
individual piles in the group.
b) The capacity of a block containing the
piles and the soil between them.
Consideration shall be given to the soil
type and soil profile, pile type, method
of installation, interaction between the
piles and the effects of any eccentric
loading. Where a pile group is founded
on a stratum which overlies a softer stratum, allowance shall be made for the possible
reduction in the ultimate base resistance of the group due to the presence of the softer
underlying stratum. The presence of soft or loose layers of soil below the pile toe may have a
more significant effect on the capacity of a block than on the capacity of a single pile. The
effect of eccentric loading is significant and shall be considered in the design if the
eccentricity of the load is greater than one quarter of the group width in the direction of
eccentricity. If a pile group has a cap cast directly onto a stratum supporting the piles, the sum
of the capacity of the individual piles in the group can be increased by the additional
resistance provided by the cap. The additional resistance is due to the capacity of the net area
of the pile cap, i.e., the gross area of the cap less the sum of the cross section areas of the
piles.
Group of end-bearing piles:
The ultimate vertical bearing capacity of a group of end-bearing piles on rock, or on dense
sand or gravel with equally strong material beneath, can be taken as the sum of the ultimate
capacities of the individual piles in the group.
Typically, the spacing between end bearing piles should not be less than twice the base
diameter of the piles.
CVEN3201 Applied Geotechnics 75 Pile Foundations

Vertical Settlement of Pile Groups


The settlement of a group of piles is the result of the interaction between settlements of all
piles within the group. The settlement of a pile group, Sg, in a single soil layer can be
calculated with reference to the settlement of a single pile as:
Sg = S Rs
where S is the settlement of a single pile calculated under the average load of a pile in the
group, and Rs is the group correction factor which depends on many factors, among them are
the spacing and arrangement of the piles, the pile dimensions, the compressibility of the pile
relative to the soil, and the depth of soil below the pile tips (for friction piles) or the nature of
the bearing stratum (for end-bearing piles). Values of Rs are given in the following tables for
floating pile groups in a deep layer of uniform soil and for end-bearing pile groups on a rigid
stratum. The values on these tables are applicable to square groups of piles in which the
centre-to-centre spacing between adjacent piles is s.
To interpolate between the tables for different values of n, it should be noted that R s varies
with the square root of n.
Number of piles in group, n
Length/dia. Spacing/dia. 4 9 16 25
ratio (L/d) ratio (s/d)
Pile stiffness factor, K
10 100 1000 10 100 1000 10 100 1000 10 100 1000
2 1.83 2.25 2.54 2.62 2.78 3.80 4.42 4.48 3.76 5.49 6.40 6.53 4.75 7.20 8.48 8.68
10 5 1.40 1.73 1.88 1.90 1.83 2.49 2.82 2.85 2.26 3.25 3.74 3.82 2.68 3.98 4.70 4.75
10 1.21 1.39 1.48 1.50 1.42 1.76 1.97 1.99 1.63 2.14 2.46 2.46 1.85 2.53 2.95 2.95
2 1.99 2.14 2.65 2.87 3.01 3.46 4.84 5.29 4.22 5.38 7.44 8.10 5.40 7.25 10.28 11.25
25 5 1.47 1.74 2.09 2.19 1.98 2.61 3.48 3.74 2.46 3.54 4.96 5.34 2.95 4.48 6.50 7.03
10 1.25 1.46 1.74 1.78 1.49 1.95 2.57 2.73 1.74 2.46 3.42 3.63 1.98 2.98 4.28 4.50
2 2.56 2.31 2.26 3.16 4.43 4.05 4.11 6.15 6.42 6.14 6.50 9.92 8.48 8.40 9.25 14.35
100 5 1.88 1.88 2.01 2.64 2.80 2.94 3.38 4.87 3.74 4.05 4.98 7.54 4.68 5.18 6.75 10.55
10 1.47 1.56 1.76 2.28 1.95 2.17 2.73 3.93 2.45 2.80 3.81 5.82 2.95 3.48 5.00 7.88
Settlement ratio Rs for floating pile groups with rigid cap on deep uniform soil mass

Number of piles in group, n


Length/dia. Spacing/dia. 4 9 16 25
ratio (L/d) ratio (s/d)
Pile stiffness factor, K
10 100 1000 10 100 1000 10 100 1000 10 100 1000
2 1.52 1.14 1.00 1.00 2.02 1.31 1.00 1.00 2.39 1.49 1.00 1.00 2.70 1.63 1.00 1.00
10 5 1.15 1.08 1.00 1.00 1.23 1.12 1.02 1.00 1.30 1.14 1.02 1.00 1.33 1.15 1.03 1.00
10 1.02 1.01 1.00 1.00 1.04 1.02 1.00 1.00 1.04 1.02 1.00 1.00 1.03 1.02 1.00 1.00
2 1.88 1.62 1.05 1.00 2.84 2.57 1.16 1.00 3.70 3.28 1.33 1.00 4.48 4.13 1.50 1.00
25 5 1.36 1.36 1.08 1.00 1.67 1.70 1.16 1.00 1.94 2.00 1.23 1.00 2.15 2.23 1.28 1.00
10 1.14 1.15 1.04 1.00 1.23 1.26 1.06 1.00 1.30 1.33 1.07 1.00 1.33 1.38 1.08 1.00
2 2.54 2.26 1.81 1.00 4.40 3.95 3.04 1.00 6.24 5.89 4.61 1.00 8.18 7.93 6.40 1.00
100 5 1.85 1.84 1.67 1.00 2.71 2.77 2.52 1.00 3.54 3.74 3.47 1.00 4.33 4.68 4.45 1.00
10 1.44 1.49 1.46 1.00 1.84 1.99 1.98 1.00 2.21 2.48 2.53 1.00 2.53 2.95 3.10 1.00
Settlement ratio Rs for end-bearing pile groups with rigid cap bearing on a rigid stratum
Example P3
A square group of 4 piles is arranged into 2 rows and 2 columns spaced 2m apart (centre to
centre). Each pile consists of a floating concrete pile having a diameter of 1m and driven 10m
into a deep layer of gravel with medium density having a total unit weight of 19 kN/m3. No
water table exists within the layer. The pile group has a rigid cap and the modulus of concrete
may be taken as 20,000MPa.
a) Determine the allowable load which could be placed on the group assuming that there is a
group efficiency of 80%. The group efficiency is defined as the ratio of the actual group
capacity to the sum of the individual pile capacities. Use an overall factor of safety of 3.
b) Determine the settlement of the pile cap when the load determined in Part (a) is applied to
the pile cap..
CVEN3201 Applied Geotechnics 78 Pile Foundations

Lateral Capacity of Single Piles


Piles are often used due to their capacity to sustain horizontal or lateral loads. The lateral
capacity of a single pile may be limited by the shearing capacity of surrounding soil, or by the
structural capacity of the pile itself, or by excessive lateral deformation. There are different
design procedures depending upon the types of soil, the modes of failure of the pile, and the
rigidity of the pile cap.
If rotation at the head of a pile is not constrained, the pile is called free-head pile. On the
other hand if the rotation at the head is constrained (equal to zero) by the pile cap, the pile is
called fixed-head pile.
For a laterally loaded single pile failure may occur either by a lateral bearing failure of the soil
(mode 1) or by development of a plastic hinge in the pile (mode 2). The third mode of failure
may only occur for fixed head piles where two plastic hinges are developed in the pile. Free
head short piles tend to fail by a rigid movement (fixed head) or by rotation around a point
near their toe (free head), a soil failure, while structural failure of long piles forms a hinge at a
point below the ground level (mode 2 for free head piles) or just under the pile cap (mode 2
for fixed head piles) and sometimes at two pints along the pile (mode 3 for fixed head piles).
The dominant mode of failure depends on the soil stiffness and strength, pile stiffness and
strength, pile length, and the moment and horizontal load applied at the pile head. It is
difficult to establish a relationship between all the above parameters and the dominant mode
of failure, to be able to calculate the horizontal capacity of the pile based on the dominant
mode. A practical method of calculating the horizontal capacity of a pile is to analyse all
conditions of failure and to select the lower lateral load capacity calculated based on all
failure modes. Methods of calculating the lateral capacity of a free head or fixed head single
pile under different failure modes was proposed by Broms in 1964 and expanded by Poulos
and Davis (1980) which are presented here in these notes.

Mmax Mmax Mmax


Hu Hu Hu Hu Hu
e

L L
Plastic hinge
Plastic
hinge

Mode 1 Mode 2 Mode 1 Mode 2 Mode 3

Free head piles Fixed head piles


Modes of failure for free head and fixed head piles

Lateral capacity of single piles in cohesive soils


Free head piles in cohesive soils
The variations of the assumed horizontal stress in front of a free head pile in a cohesive soil
are shown in the following figures for both modes of failure. The soil is assumed to deform
under undrained conditions and has an undrained shear strength (cohesion) of cu. The pile has
a diameter d and length L. The horizontal load, H, is applied on the pile head at a distance e
above the ground level. If a moment M and a horizontal load H are applied to the pile at the
ground level, e can be calculated as M/H.
CVEN3201 Applied Geotechnics 79 Pile Foundations

Distribution of stresses along a free head pile in clay for failure modes 1 & 2
100
Mode 1- Free headed
Mode 1: Hu

Ultimate Lateral Resistance, Hu / cu d2


90 Mode 1- Restrained
Equilibrium of moments and e
horizontal forces results in Hu 80
e/d = 0
and Mmax: 70 1
L
Hu = 9 f cu d = s cu d2 60
2
4
The value of s can be
50 d 8
obtained from the figure
opposite. 40 15
30
30
In this case the maximum 20
bending moment in the pile, 60
10 150
Mmax, which occurs at a point
along the pile where the shear 0
force is equal to zero, can be 0 5 10 15 20 25 30
calculated as: Embedment Length Ratio, L/d

g g g g g 9
M max 9c u d 9c u d c u d g 2
2 2 4 2 4 4
where g = L-f-1.5d, and f = Hu / 9cu d.
Mode 2:
100
As the length of the pile Mode 2- Free headed
Ultimate Lateral Resistance, Hu / cu d2

increases, the maximum Mode 3- Restrained


bending moment calculated in Hu
mode 1 becomes larger than the e
moment capacity of the pile,
My, and the pile yields and a
plastic hinge develops. This 10
L
mode of failure occurs for long
piles. 150
d 60
The ultimate horizontal capacity 30
15
of the pile can be calculated as: 8
4
Hu = L cu d2 1
2
e/d = 0
where L can be obtained from 1
the figure opposite or the 1 10 100 1000
following equation: Yield Moment, M yield / cu d3
e e
2
2M y
L 9 1.5 1.5
d d 9 cu d
3

CVEN3201 Applied Geotechnics 80 Pile Foundations

Lateral capacity of free head piles in cohesive soils


The ultimate lateral capacity of free head piles in cohesive soils will be the lesser value
obtained based on failure in modes 1 and 2:
Hu =s cu d2 (Soil failure, mode 1)
Hu = L cu d2 (Structural failure, mode 2)
If mode 1 is the governing mode, the maximum moment within the pile (Mmax = 9 cu d g2 / 4)
can be checked against the yield moment of the pile structure.
Fixed head piles in cohesive soils
The variations of the assumed horizontal stress in front of a fixed head pile in cohesive soils
are shown in the following figures, for all three modes of failure.
Mode 1
The ultimate horizontal capacity of a fixed
head pile failing in mode 1 can be calculated
from the simple uniform distribution of the
horizontal stress assumed to form in front of
the pile:
Hu = 9 cu d (L - 1.5 d)
Alternatively, Hu can be obtained from the
restrained solution in the mode 1 failure
chart given before for free head piles in clay.
Mode 2
The ultimate horizontal capacity in this
mode of failure can be obtained by solving
the following 3 equations simultaneously:
g = L - 1.5 d - f
Hu = 9 cu f d
My = 9 cu d f (1.5d - f / 2) 9 cu d g2/ 4
This results in a quadratic equation in terms
of g:
9d 2 4M y
g 2 4Lg 2L2 0
2 9cu d
where the appropriate solution to this
equation is:
16M y
g 2L 0.5 8L2 18d 2
9cu d
Mode 3
The distribution of horizontal stress at
mode 3 failure around the pile is complex.
In this case the restrained solution in the Distribution of stresses along a fixed
mode 2 failure chart given before for free head piles in clay for all failure modes
head pile is used to obtain Hu.
Lateral capacity of fixed head piles in cohesive soils
The ultimate lateral capacity of fixed head piles in cohesive soils will be the lesser value
obtained based on failure modes 1, 2 and 3 as explained above.
The maximum moment within the pile is checked using the following relationships:
CVEN3201 Applied Geotechnics 81 Pile Foundations

Mode 1: Mmax = Hu (L + 1.5d)/2.


Mode 2: Mmax = Hu (f + 1.5 d) 4.5 cu d f2 - My = Hu (f + 3d)/2 My.
Mode 3: Mmax = My.

Lateral capacity of single piles in cohesionless soils


The assumed failure mechanisms for piles in granular soils are identical to those for cohesive
soils. However, the distribution of horizontal stresses in front of the piles is different. In the
case of granular soils, a wedge of soil in front of the pile is assumed to displace. This wedge
is assumed to fail under three times the Rankines passive earth pressure, i.e. the soil pressure
in front of the pile is equal to 3 Kpv where Kp = tan2(45 + /2) is the coefficient of passive
earth pressure, v is the vertical effective stress along the pile, and is the friction angle of
the soil. If the diameter of the pile is d, the resisting force per unit length of pile will be
3dKp.
Free head piles
The variations of the assumed horizontal stress in front of a free head pile in cohesionless
soils are shown in the following figures for both modes of failure. Again it is assumed that the
pile has a diameter d, length L, and the horizontal load is applied on the pile at a distance e
from the ground level.

Mode 1:
The ultimate horizontal capacity of a free head pile embedded in a granular soil failing in
mode 1 can be calculated by: 400
Mode 1- Free headed
'v K p d L2
Ultimate Lateral Resistance, Hu / Kp d3

Hu
Hu e
Mode 1- Restrained
2(e L)
300
If the water table is below the pile e/L = 0
tip, the vertical stress at the base of L
the pile will be: v = L, where is 0.2

the total unit weight of the soil, and 200


d
0.4
the above equation is simplified to: 0.7
Kp d L 3
L3 1.0
Hu and f 1.5
2(e L) 3(e L) 100
3.0
Alternatively, Hu can be obtained
7.0
from the chart given in the figure
opposite. 0
0 5 10 15 20 25 30
The maximum moment along the
Embedment Length Ratio, L/d
CVEN3201 Applied Geotechnics 82 Pile Foundations
1000
Mode 2- Free headed
pile, occurs where the shear

Ultimate Lateral Resistance, Hu / Kp d3


Hu Mode 3- Restrained
force is zero, is:
e
Mmax = Hu (e + 2f / 3)
Mode 2
100
The value of Hu for mode 2 L
failure can be found from the
solution to the following
equation: d 150
60
H u 10
H u e 0.54
30
My 15
d K p
8
4
where My is the yield moment 2
of the pile structure. e/d = 0 1
1
Alternatively, the figure
0 1 10 100 1000 10000
opposite can be used to find
the value of Hu. Yield Moment, M yield / Kp d 4

Note that both figures given for horizontal failure loads of a free head pile in modes 1 and 2
have curves that can be used to obtain the horizontal failure loads of a fixed head pile in
modes 1 and 3.
Lateral capacity of free head piles in cohesionless soils
The ultimate lateral capacity of free head piles in granular soils will be the lesser value
obtained based on failure modes 1 and 2. If mode 1 is the governing mode, the maximum
moment within the pile, Mmax = Hu (e + 2f/3), can be checked against the yield moment in the
pile.
Fixed head piles
Mode 1
The assumed distributions of stress and
moment along a fixed head pile in
cohesionless soil failing in mode 1 are shown
in the figure opposite The ultimate horizontal
capacity of the pile can be calculated from the
simple linear distribution of the horizontal
stress assumed to form in front of the pile:
Hu = 1.5 v d L Kp
If the water table is below the pile tip and the
unit weight of soil is , the ultimate horizontal
capacity can be obtained as:
Hu = 1.5 d L2 Kp
Alternatively, Hu can be obtained from the
restrained solution in the mode 1 failure
chart given before for free head piles in sand.
Mode 2
The ultimate horizontal capacity in this mode
of failure can be written as:

Hu
0.5' v
d L2 K p M y
L
If the water table is below the pile tip, the
CVEN3201 Applied Geotechnics 83 Pile Foundations

ultimate horizontal capacity will be:

Hu
0.5d L K M
3
p y
and f
2H u
L 3dK p
Mode 3
The distribution of horizontal stress at mode 3
failure around the pile is complex. In this case the
restrained solution in the mode 2 failure chart
given before for free head piles in sand is used to
obtain Hu.
Lateral capacity of fixed head piles in cohesionless
soils:
The ultimate lateral capacity of fixed head piles in
granular soils will be the lesser value obtained
based on failure modes 1, 2 and 3 as described
above.
The maximum moment within the pile is checked
using the following relationships:
Mode 1: Mmax = 2Hu L/3
Mode 2: Mmax = Hu f - f3 d Kp/2 - My = 2Hu f /3 My.
Mode 3: The maximum moment is My.
CVEN3201 Applied Geotechnics 84 Pile Foundations

Lateral Deflection of Single Piles


The lateral deflection and rotation of a fully embedded single pile can be estimated using the
theory of elasticity. Note that in this way only an estimate of the deflection will be obtained.
A more accurate evaluation of the lateral deflection requires in-situ testing on piles or
comprehensive in-situ and laboratory tests on soil in order to obtain appropriate soil
parameters. AS2159-1978 gives guidelines for calculation of the lateral deflection of piles; a
summary of them will be given here.

Uniform soil modulus


Heavily over-consolidated clays exhibit relatively uniform soil modulus. The ground-line
displacement, , and rotation , for a free head and fixed head piles are given below.
Free head pile
H M H M
IH I , and
2 M
I
2 H
I
3 M
Es L Es L Es L Es L
Fixed head pile
H
I F , and = 0
EsL
The influence factors IH, IM=IH, IM and IF are given in the following table. Note that in
the above equations M is the moment applied to the pile at ground level, that is
M = Mapplied + He.
KR
L/d I 10-6 10-5 10-4 10-3 10-2 10-1 1.0 10
IH 10.84 10.1 8.249 5.951 4.076 3.397 3.304 3.294
IM, IH 124.5 97.11 52.42 22.05 8.348 4.583 4.090 4.039
10
IM 4957.0 2939.0 904.5 199.5 39.67 11.06 7.647 7.296
IF 10.36 8.256 5.630 3.667 2.401 1.532 1.123 1.061
IH 17.44 15.03 11.12 7.468 4.973 4.234 4.141 4.132
IM, IH 233.6 153.5 70.47 26.99 9.871 5.815 5.321 5.270
25
IM 9418.0 4301.0 1101.0 224.7 43.49 13.47 10.050 9.700
IF 15.85 11.11 7.034 4.394 2.831 1.760 1.330 1.271
IH 22.3 18.23 12.92 8.421 5.573 4.808 4.715 4.706
IM, IH 314.3 188.2 80.70 29.82 10.84 6.660 6.165 6.115
50
IM 12593.0 5034.0 1198.0 238.1 45.75 15.12 11.70 11.35
IF 12.88 19.56 7.917 4.862 3.113 1.913 1.473 1.415
IH 26.4 20.91 14.34 9.191 6.076 5.296 5.203 5.194
IM, IH 380.8 214.4 88.39 31.99 11.63 7.376 6.882 6.832
100
IM 15092.0 6541.0 1266.0 247.2 47.53 16.50 13.10 12.76
IF 22.53 14.29 8.632 5.246 3.346 2.043 1.595 1.537
In this table KR is the relative flexibility of the pile which can be defined as:
EpIp
KR
E s L4
In the above equation Ep and Es are the Youngs moduli for the pile and the soil, respectively,
and Ip is the second moment of inertia of the pile section.
Note that the most suitable method to obtain Es for lateral deflection (which is normally
CVEN3201 Applied Geotechnics 85 Pile Foundations

different from Es used for vertical displacement) is to carry out a lateral pile load test and use
a theory to back figure Es. In the absence of any such data, the following relationships can be
used to estimate Es.
Es = 40 cu For normal piles where KR 0.001
Es = 250 cu For large diameter piles or caissons where KR > 1

Linearly increasing soil modulus with depth


Normally consolidated clays and sands exhibit a linearly increasing modulus with depth. The
ground-line displacement, , and rotation , for a free head and fixed head piles are given by:
Free head pile
H M H M
I
2 H
I , and
3 M
I
3 H
I
4 M
Nh L NhL NhL NhL
Fixed head pile
H
I , and = 0
2 F
NhL
The influence factors IH, IM=IH, IM and IF are given in the following table.
KN
L/d I 10-6 10-5 10-4 10-3 10-2 10-1 1.0
IH 355 178.0 81.0 34.7 16.70 12.70 12.30
IM, IH 2,940 1,030 304.0 83.0 24.50 14.10 12.70
5
IM 69,000 14,400 2,460.0 401.0 68.80 22.80 17.60
IF 1303 73.2 33.1 14.2 6.58 2.89 1.70
IH 531 239.0 103.0 43.6 22.70 19.40 19.00
IM, IH 4830 1410.0 384.0 103.0 32.60 22.20 21.50
25
IM 93,500 16,300 2,710.0 437.0 81.80 35.30 30.20
IF 208 98.1 41.3 18.0 8.40 3.58 2.50
IH 653 282.0 120.0 50.4 28.30 25.00 24.70
IM, IH 5920 1640. 437.0 115.0 40.50 30.00 29.0
100
IM 104,000 17,600 2,890.0 465.0 91.90 45.90 41.10
IF 254 115.0 49.2 20.9 9.93 4.18 4.10
Again KN is the pile flexibility factor which is defined as:
EpIp
KN
N h L5
In the above equation Nh is the rate of change of the Youngs modulus of soil with depth:
Nh = Es / z
In the absence of any in-situ soil data, the following table can be used as a guide to estimate
the value of Nh.
Soil Type Nh (MPa/m)
Soft normally consolidated clay 0.2
Normally consolidated organic clay 0.1
Peat 0.04
Dry or moist loose sand* 1.6
Dry or moist medium-dense sand* 4.8
Dry or moist dense sand* 12.6
CVEN3201 Applied Geotechnics 86 Pile Foundations

H 1.2m
Example P4
A 0.6m diameter 12m long free head concrete pile is subjected to a
horizontal loading at a height of 1.2m above the ground level. The soil
is frictionless with a cohesion of cu =40kPa and a unit weight of
12m
20kN/m3. The elastic modulus of the soil increases linearly with depth
as Esoil = 1000z (kPa), where z is the depth below the surface in m. The
pile has a modulus of Ep = 25,000MPa, a yield moment of
My = 750kNm and a moment of inertia of Ip = 0.1m4
a) Determine the maximum horizontal load that the pile can withstand.
b) Determine the maximum moment in the pile under this loading.
c) Estimate the ground line displacement and rotation under 50% of the ultimate horizontal
load obtained in Part (a).
CVEN3201 Applied Geotechnics 87 Pile Foundations

Safety Margin
Design for allowable strength based of overall factor of safety
Traditional methods of pile design have relied on an overall factor of safety against ultimate
failure. Therefore, the design criteria can be written as:
Pa = Pu / F
where Pa is the allowable vertical capacity or the design load applied to the pile, Pu is the
ultimate bearing capacity of the pile, and F is factor of safety. Typical factors of safety are
between 2 to 5, depending on the method of calculation of the capacity, the loading
conditions, the extent of site investigation and the nature and variability of the ground, the
extent of the designers experience, the likely consequence of failure, and the importance of
structure.
In general if theoretical or semi-empirical methods are used for design, but the design is not
verified by loading tests a larger factor of safety, greater than 3, should be used. On the other
hand if the theoretical design is verified by sufficient loading tests, a lower factor of safety,
larger than 2, can be used for the pile.
Note that the movements required to mobilise the ultimate base resistance of driven piles are
generally larger than those required to mobilise shaft resistance. The movements required to
mobilise base resistance of bored piles are even larger due to effects of construction methods,
accumulation of debris at the bottom of the shaft, etc. Therefore some engineers apply
different factors of safety for shaft resistance, fs, and base resistance, fb. A small factor of
safety, in the order of 1.5, is used for shaft resistance while a large factor of safety, in the
order of 3-5, is used for base resistance. In some cases, the contribution of base resistance to
the overall capacity of the pile is ignored altogether, but a factor of safety larger than 1.5 is
used for the calculation of the allowable load based only on shaft resistance.

Design for ultimate strength based on LRFD:


The current Australian Standard 2159-1995 Piling - Design and Installation is based on the
Load and Resistance Factor Design (LRFD) approach. In this method different components of
the applied load are increased by various load factors and the ultimate geotechnical resistance
is reduced by a single reduction factor. The factored loads must be less than or equal the
reduced resistance. Following is a brief extract from the Australian Standard 2159-1995.
Single piles, pile groups and individual piles within a pile group shall be designed for
geotechnical strength. In the design of a single pile or pile group, various factored loads and
other actions shall be applied to the single pile or pile group, and the design effect, S*, shall
be determined for each pile or pile group for each load case. The design loads shall be
determined from AS 1170 Minimum Design Loads on Structures. The design shall take into
account the action effects arising from all loads and actions specified in AS1170.1, the dead
loads of pile and pile cap, and any other additional loads and actions that may be applied to
the pile. The design load shall be the combination of factored loads which produces the most
adverse effect on the pile.
The design geotechnical strength, P*, may be determined based on the methods presented in
the previous sections. The pile shall be proportioned so that the design geotechnical strength
is not less than the design action effect, i.e.
P* S*
The design geotechnical strength P* shall be calculated as the ultimate geotechnical strength
Pu multiplied by a strength reduction factor , based to the following equation:
P* = Pu
In selecting the reduction factor all factors which may influence the reliability of the
CVEN3201 Applied Geotechnics 88 Pile Foundations

ultimate geotechnical strength shall be considered. A range of values for is given in the
following table.

Method of assessment of ultimate strength Range of


Static load testing to failure 0.70 0.90
Dynamic load testing to failure 0.65 0.85
Static analysis using CPT data 0.45 0.65
Static analysis using SPT data in cohesionless soils 0.40 0.55
Static analysis using lab data for cohesive soils 0.45 0.55
Dynamic analysis 0.45 0.65
Range of values for geotechnical strength reduction factor,

In assessing the value to be chosen within the ranges, consideration shall be given to the
factors shown in the following table and appropriate judgement shall be exercised.
In the determination of the area of the shaft, As, allowance should be made for soil near the
surface which may not be effective in providing load transfer. Australian Standard
recommendation is that, in the absence of any data, the surface area from ground level to 1.5
times pile diameter or 1m below ground level (whichever is the greater) shall be assumed to
be ineffective.

Circumstances in which lower end of range Circumstances in which upper end of range
may be appropriate may be appropriate
Limited site investigation Comprehensive site investigation
Simple method of calculation More sophisticated design method
Average geotechnical properties used Geotechnical properties chosen conservatively
Use of published correlations for design Use of site-specific correlation for design
parameters parameters
Limited construction control Careful construction control
Less than 3% piles dynamically tested 15% or more piles dynamically tested
Less than 1% piles statically tested 3% or more piles statically tested
Guide for assessment of strength reduction factor,

References
Broms, B.B. (1964) The lateral resistance of piles in cohesive soils, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol 90, SM2, 27-63.
Broms, B.B. (1964) The lateral resistance of piles in cohesionless soils, Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol. 90, SM3, 123-156.
Poulos, H.G. and Davis, E.H. (1980). Pile Foundation Analysis and Design, John Wiley and
Sons.
Poulos HG (2001) Pile Foundations. Chapter 10 of Geotechnical and Geoenvironmental
Engineering Handbook, Ed. R.K. Rowe, Kluwer Academic Press, Boston, pp. 261-304
LATERAL EARTH PRESSURE
The stresses within the ground and those imposed by the ground on structures, for example
retaining walls, can have a significant influence on the performance of structures supporting
the ground. In most cases, especially for cases involved horizontal ground surface, the vertical
stress can be easily estimated. However, the estimation of horizontal stress is more complex.
The lateral earth pressures imposed by soils on structures can be estimated from the
equilibrium of forces at the onset of failure, or by considering stresses at a point in the ground
that result in shear failure. These are the bases of the two major theories used to estimate the
lateral earth pressure, namely the Coulombs theory and the Rankins theory. While both
theories have their own limitations, they are very much in use today and therefore will be
described in this section.
Note that in the following sections the soil strength parameters are represented by cohesion, c,
and friction angle, . However, these parameters can be exchanged by drained values, c and
, or undrained values, cu and u.

Lateral Earth Pressure at Rest


The vertical total stress at a point located at depth z below a horizontal ground level is simply
the overburden pressure v = z, where is the unit weight of soil. If the soil is at rest, i.e., it
is not subjected to any movement and therefore remains undisturbed, the lateral effective
stress can then be calculated as h = Ko v, where Ko is the coefficient of lateral earth
pressure at rest, v=v u is the vertical effective stress, and u is pore water pressure. Ko is
difficult to determine, its value can be approximated as:
Ko = 1-sin For granular soils and for normally consolidated clays.
Ko = (1-sin )OCR 0.5
For over-consolidated clays, where OCR is the over-
consolidation ratio.

Rankines Theory of Lateral Earth Pressure


Consider a smooth vertical retaining wall wished in place into a soil mass which has a
horizontal ground surface. In this case the soil remains at rest and the horizontal and
vertical effective stresses at depth z are defined, v = z and h = Ko v. Now consider a part
of the soil in front of the wall is excavated and the wall fails and moves toward the
excavation. At the onset of failure, the vertical stress remains unchanged. The shear stress on
any vertical or horizontal plane is assumed to be zero
since the wall is vertical and smooth and the ground
level is horizontal. Therefore, the normal stresses
acting on the wall will be a principal stresses. At the v = 1
onset of failure these horizontal stresses can be
calculated from the Mohr-Coulomb failure criterion.
The Mohr-Coulomb failure criterion can be expressed h = 3
in terms of principal stresses, 1 and 3, as:
v = 3
1 N 3 + 2 c N
where N = tan2(45+/2), and is the friction angle of h = 1
the soil.

Active earth pressure


The horizontal stress at the back of the wall, while the wall is moving away from the soil and
eventually results in failure of the soil, decreases from its initial value of Ko v to a value
which satisfies the failure criterion. After failure any further lateral movement takes place at
CVEN3201 Applied Geotechnics 90 Lateral Earth Pressure

constant horizontal stress. In this


case, the vertical stress at any point at Direction of
failure plane
the back of the wall is greater than
the horizontal stress, and therefore
1 = v and 3 = h. Substituting
these values into the equation for
Mohr-Coulomb failure criterion
results in:
c (45+/2)
v N h + 2 c N h=Ka v v v

or h v 2 c N / N
or h Ka v 2 c Ka
where Ka = tan2 (45 - /2).
Ka is defined as the coefficient of
active earth pressure and the horizontal stress in this case is defined as the active horizontal
stress. This stress is the lowest possible lateral pressure with the whole soil body is on the
point of shear failure. The failure planes make an angle of (45 + /2) with respect to the
horizontal.
At the ground level, where v = 0, the lateral stress is h = -2cKa. The negative stresses in the
soil come mostly from suctions in the pore water,
which are not reliable and therefore usually
zc
ignored. A tension crack is assumed to open to a
depth at which the theoretical pressure becomes
positive, i.e. to depth zc (figure opposite). For a
horizontal ground level where no surcharge is
applied at the ground level, the depth of tension
crack can be calculated as:
2c
zc . 45+/2
Ka
Note that if it is assumed that tensile stresses are
going to develop in the soil above zc, the force
required for stability of the wall will be reduced. Therefore, ignoring the tensile stresses gives
a more conservative solution. Note also that if water is available, it can fill up the tension
crack and provide additional pressures on the wall.

Passive earth pressure Direction of


failure plane
The horizontal stress at a point in the
soil in front of the wall, while the
wall is moving toward the soil and
eventually results in failure of the
soil, increases from its initial value of
Ko v to a value which satisfies the (45-/2)
failure criterion. In this case, the v v h=Ka v
vertical stress in front of the wall
remains constant. The lateral
compression of the soil causes the
horizontal stress to increase gradually
to a value greater than the vertical
stress, to become the new major
CVEN3201 Applied Geotechnics 91 Lateral Earth Pressure

principal stress. Therefore 1 = h and


3 = v. At the onset of failure these
stresses must satisfy the failure criterion.
Substituting these values into the equation
of Mohr-Coulomb failure criterion results
in:
h N v + 2 c N
or h K p v 2 c K p
45-/2
where Kp = tan2 (45 + /2)
Kp is defined as the coefficient of passive
earth pressure and the horizontal stress in this case is defined as the passive horizontal stress.
This stress is the largest possible lateral pressure where the whole soil body is on the point of
shear failure. The failure planes make an angle of (45 - /2) with respect to the horizontal. It
may be shown that Kp = 1 / Ka.
The two Ka and Kp are known as the Rankines coefficients of earth pressure. The Rankines
theory was develop in 1857 for frictional material and later was extended to cohesive-
frictional soil by Bell (1915). This theory is based on a stress state in the soil at failure, and
therefore this approach is referred to as a lower bound method, a method which produces safe
and conservative solutions.
It is worth noting that very small lateral movements are sufficient to mobilise fully the active
state, while large movements are normally required to mobilise fully the passive state.

Active earth pressure on sloping ground


If the ground surface is not horizontal, the vertical stress
in the soil will not be a principal stress but evaluation of
Ka may be carried out in a similar manner to the case of
horizontal ground surface. For cohesionless soil with
ground surface sloping at angle to the horizontal, a
closed form solution was proposed by Rankin as:

h = Ka v
cos (cos 2 cos 2 )
and K a cos
cos (cos 2 cos 2 )
The pressure on the wall, h, is assumed to act parallel to
the surface of the soil, i.e., at angle to the horizontal.

Limitations of Rankines theory of earth pressure


Rankines theory implies that the back of the wall is vertical and perfectly smooth. In practice
most walls are to some extent rough and have an inclined back, and therefore shear stresses
will be developed as the walls move relative to the soil. For cohesive frictional soils, with c
and , the soil/wall interface is characterised by two similar parameters, soil/wall adhesion,
ca (c), and soil/wall friction angle, ( ). For concrete walls, is often taken as 2/3 and ca
is either ignored or taken equal to soil cohesion, c, often with a maximum value of 50 kPa. In
order to take into account the roughness and inclination of the wall back, another method,
based on the Coulombs theory of earth pressure, for the analysis of lateral earth pressure will
be introduced in the next section.
CVEN3201 Applied Geotechnics 92 Lateral Earth Pressure

Example L1
Find the distribution of active earth pressure on the 4m Sand
t = 21 kN/m3
high wall shown opposite. The soil at the back of the wall 2m = 30o
consists of 2m layer of sand underlain by a deep layer of
clay.
Clay
2m t = 20 kN/m3
cu=30 kPa
u =0

Example L2 Clay
t = 20 kN/m3
A sheet pile wall is embedded in a two layered soil stratum; 4m cu=30 kPa, u =0
an upper 5m layer of clay and a lower deep layer of sand, as c=0, =20
shown in the figure opposite. The properties of both layers
are given in the figure. The water table is 4m below the
surface of the clay layer. 1m
Calculate the total force applied on the wall.

3m Sand
t = 18 kN/m3
= 40o
CVEN3201 Applied Geotechnics 93 Lateral Earth Pressure
CVEN3201 Applied Geotechnics 94 Lateral Earth Pressure

Coulombs Theory of Earth Pressure


Coulomb (1776) introduced a method to evaluate the
Direction of
earth pressures, in which a mechanism is assumed for movement
the failure of the soil and wall. At failure, the shear
stress acting on each failure surface is assumed to be
equal to the shear strength of the soil. Equilibrium of Assumed
forces on a failure plane results in the value and the
failure plane
direction of the force applied to the wall.
The shear stress at any point on the failure plane can
be expressed by Mohr-Coulomb failure criterion as:
= c + n tan
The appropriate values for c and will depend on the W
type of analysis. In a total stress (undrained) analysis
cu and u are used whereas in an effective stress ca ntan c ntan
(drained) analysis cand . n n
For the slip failure between soil and the back of a
retaining wall, the failure conditions can also be
described by the Mohr-Coulomb failure criterion. If
the normal stress at a point at the back of the wall is
n, the maximum shear stress which can be
developed would be: W
= ca + n tan Tw
where ca and are the adhesion and the friction angle
between the wall and the soil. Nw
N
The forces acting on the failure plane and on the wall
can be obtained by integration of the stresses over the
area on which they are applied. Considering a unit
thickness for the problem, the forces can be
calculated as:
W
T ds c n tan ds ,
Ca
N n ds , C

and C c ds .
P R
The failure criterion can be expressed in terms of
forces as:
T = C + N tan P
or Tw = Ca + Nw tan
Note that the shear force, T or Tw, consists of two R
components; cohesion or adhesion force, C or Ca, and
R = N tan or P = Nw tan . Force P is the unknown
resultant force that the soil applies to the wall. The W
values of C and Ca are generally known. Their
directions are parallel to the failure surfaces. The Ca
values of R and P are unknown, but these two forces

act at angles and with respect to the normals to C
the failure planes, as shown in the figure opposite.
CVEN3201 Applied Geotechnics 95 Lateral Earth Pressure

The unknown force P can be obtained by drawing the polygon of forces, where the known
forces are drawn first, starting with the self weight of the soil, W, and then C and C a. The
directions of the two unknown forces, R and P, are known and can be drawn last. The size of
these forces can be calculated or measured if the force polygon is drawn in scale.
It is important to note that the directions of the forces on the soil wedge must be consistent
with the assumed failure mechanism, i.e., the forces must be drawn so that the arrows
indicating their directions all point the same way as moving around the polygon. It is possible
that a polygon of forces cannot be constructed, and if this happens it indicates that the
assumed failure mechanism is incorrect.
In a passive failure,
the soil wedge
moves away from Direction of
the wall. The movement W
directions of
P
resisting forces are
R
in the opposite Assumed Ca
C
direction of those in
failure plane
an active failure, as
shown in the figure
opposite.
In an undrained analysis, it is necessary to allow for
the presence of tension cracks and, if water is zc
present, the possibility that these cracks are filled
with water. The depth of tension crack, zc, can be
determined from Rankines theory of earth pressure.
Water in a tension crack adds an extra component
of force to the polygon of forces. This component is
horizontal with a value of 0.5w zc2.
In a drained analysis, the effective normal stresses
and the effective forces must be used in the
analysis. The effective normal force is defined as:
N = N U
where U u.ds is the force due to pore water pressure, u, and acting normal to the failure
plane and to the back of the wall.
In the presence of steady state seepage it may be necessary to draw a flow net to determine
the pore water force, U, acting on the soil wedge.

Limitations of Coulombs theory of earth pressure


The main advantage of Coulombs method, over Rankines method, is that it can be used for
any soil and wall geometry, any soil/wall interface characteristic, and any load on the soil. Its
main disadvantage is that the common layered soil profile cannot be simply accounted for.
Another disadvantage of the Coulombs method is that in this method all possible trial failure
mechanisms, with various slope angles, , must be considered and the mechanism with
greater active force on the wall, or the mechanism with the smaller passive force on the wall,
is taken as the true mechanism of failure.
The Coulomb wedge method is readily extendable to more complex, general problems
involving seeping water, layered backfills, surcharge loadings, ca and , and tension cracks.
A more complex vector polygon must be drawn, otherwise the principles remain the same.
CVEN3201 Applied Geotechnics 96 Lateral Earth Pressure

Drained or Undrained Analysis (Total or Effective Stress Analysis)


For sand and gravel, and for clay a long time after construction, a drained analysis should be
carried out using effective stresses and the effective strength parameters, c and , obtained
by laboratory tests under drained conditions. Whereas for clay immediately after construction
an undrained analysis is carried out using total stresses and the undrained strength parameters,
cu and u.
In an effective stress analysis the state of pore pressure in the soil must be known. Then if
water is present in the soil behind a wall, effective soil stresses must be used for calculating
earth pressures. The lateral earth pressure can be calculated as: h = Ka v. The water also
exerts a full hydrostatic pressure of w zw at any depth. Therefore it is economical to provide
adequate drainage for the soil behind the wall. If the water level is not the same on each side
of the wall, water will flow. The pore water pressures must then be determined from a flow
net before calculating v.
In drained or effective stress analysis, it is more appropriate and safer to use the residual (or
critical state) effective strength parameters, as oppose to the peak strength values.

Effects of Water
Wherever feasible, water pressure caused by
groundwater or seepage should be avoided by the
design of a suitable drainage system and/or use of
granular materials as backfill. If neither of these
conditions is catered for, water pressure behind
the wall needs to be accounted for in the design.
Figure opposite shows a good example of drains
behind retaining walls. The drainage layers may
be omitted if a free-draining granular backfill is
used. However, a drainage pipe should be
provided to discharge water to a suitable outlet.
Note that construction of a vertical drain by itself Drain
does not eliminate the seepage force as part of
the total horizontal pressure and horizontal
and/or sloping drains would therefore be
preferable. z
In calculations of water pressures in front of
quay walls and other retaining walls with a water z w
pressure gradient and water flow around the
wall, the influence of groundwater flow on the
water pressure should be accurately taken into
account. For example for the special case of flow
around a sheet pile wall, shown in the figure
opposite, the flow pressure reduces the steady
state water pressure at the back of the wall but
increases the steady state water pressure in front
of the wall. As a consequence, the active
horizontal pressure at the back of the wall
increases and the passive resistance in front of
the wall decreases due to water flow, as
compared to steady state condition where there is
no flow. This situation may be decisive for the
stability of the wall.
CVEN3201 Applied Geotechnics 97 Lateral Earth Pressure

10o
Example L3
Find the lateral force of the wall shown in the figure cu = 10 kPa
opposite, assuming the failure line given in the figure. u = 10o
ca = 2 kPa
5m 6.4 m = 20o
30 o t = 20kN/m3
RETAINING WALLS
Retaining walls can be broadly classified into three categories: gravity walls, embedded walls,
and anchored walls. Gravity walls rely on their weight for their stability. Embedded walls
mobilise passive earth pressures in the ground to provide resistance. Anchored walls rely on
the force developed in the anchor as well as the earth pressure in front of the wall for their
stability. Within each category there are a variety of wall types. In selecting the appropriate
wall type many factors need to be considered including: soil and groundwater conditions,
height and ground topography, availability of suitable fill material, construction constraints
(space, access, equipment, availability of specialist techniques), environment appearance
and impact during construction, ground movements and
their effects on adjacent structures, underground q
obstructions and services, design life and maintenance
requirements, and cost.
Ww
The forces on a retaining wall may include: active earth
pressure at the back of the wall (Pa), passive earth
W1
pressure in front of the wall (Pp), weight of the wall
Pa
structure (Ww), weight of the soil on top of the
foundation of the wall (W1&2), surcharge (q) due to live W2
load or traffic load, water pressure and uplift, forces of
anchors, forces of construction equipments such as Pp
compaction machines, etc. Rh
Rv

Design Requirements and Safety Margin


The design of a retaining wall shall take into account the stability and serviceability of the
wall. These notes cover only the general design requirement for the stability analysis of walls.
Different types of walls fail in different modes. For different types of walls it is important to
consider the appropriate mode of failure and use it to evaluate the stability of the wall.
Traditionally, the design of retaining walls for strength has been based on the overall factor of
safety method. However, the new Australian standard AS4678-2002 Earth Retaining
Structures is based on the partial factors of safety method, though it recognises the traditional
method of design as well. Both methods of design are covered in these notes in the following
sections.

Design based on overall factor of safety


Traditional methods of retaining wall design are based on an overall factor of safety against
failure. The overall factor of safety, F, can be calculated as:
Re sisting force or moment
F
Destabli sin g force or moment
Typical factors of safety rage from 1.5 to 2, depending on the mode of failure, the loading
conditions, the reliability of the soil strength parameters, the extent of the designers
experience, and the likely consequence of failure.
CVEN3201 Applied Geotechnics 100 Retaining Walls

1.5m
Example R1
Calculate the factor of safety against sliding, overturning, Sand
and bearing failure of the 3m high retaining wall shown = 30o
in the figure opposite. The granular soil at the back of the 2.5m
3m = 18kN/m3
wall has a friction angle of = 30o and a unit weight of
= 18kN/m3. The unit weight of the concrete wall is
c = 25kN/m3. The friction between the base of the wall
and the sand can be taken as = 20o.
= 2/3
CVEN3201 Applied Geotechnics 101 Retaining Walls

Design for ultimate strength based on partial factors of safety


The use of limit state design is relatively new in Geotechnical Engineering and has not been
covered widely by standard text books. The Australian Standard, AS 4678-2002, Earth-
retaining structures, covers the design of retaining walls and other structures which provide
lateral support for soil. The code specifies an ultimate limit state approach to design so the
earth pressures as well as resisting forces and moments are ultimate values. In a design the
ultimate limit state (this is for strength or stability) or serviceability limit state (this is for
control of movement) may be considered. All possible modes of wall failure need also to be
considered, i.e., sliding, rotation, bearing, rupture of components, and global failure.
The approach adopted by AS4678 for design of retaining structures is based on limit state and
partial factors of safety. In this approach, the design criterion for stability is defined as:
n R (ai Pi)
where R is the design resistance, n is structural classification factor, ai are load factors, and
Pi are different components of loading. The design resistance, R, is calculated using the
design strength parameters obtained by reducing the characteristic strength values of the soil
using different partial factors of safety.
Load factors
Retaining walls need to be designed for dead loads due to wall weight and earth pressures,
live load due to any surcharge or traffic load, and hydrostatic water pressure where applicable.
Loads and load factors used in the design are similar to those specified in AS1170 except that:
All structures should be designed to resist a minimum live load of 5 kPa (lower live loads
can only be used on steep backfill slopes).
The requirement in AS 1170.1 for a load factor of 1.5 to be applied to earth pressures is
deemed to be satisfied by applying a factor of 1.25 to the soil unit weight.
Lower load factors shall be used where the lower factors lead to a more critical limit state
design.
Various combinations of partial load factors are prescribed by the code. For normal design the
following load combinations are usually adopted:

Bearing Rotation
Load factor Serviceability
capacity & sliding
Dead load of wall and contained soil 1.25 0.8 1
Dead load of earth pressure behind wall 1.25 1.25 1
Dead load of fill in front of wall 0.8 0.8 1
Water pressures on either side of wall 1 1 1
Live load on top of wall and contained soil 1.5 0 0.7 or 0.4*
Live load on backfill behind wall 1.5 1.5 0.7 or 0.4*
Live load on fill in front of wall 0 0 0
* 0.7 for long term case and 0.4 for short term case
Soil strength reduction factors
In order to take into account the uncertainties in selected material properties, the soil shear
strength parameters, c and , are reduced by applying strength reduction factors. The lateral
soil pressure is then calculated based on the reduced strength parameters, c* and *, obtained
from the normally interpreted values of cohesion, c, and friction angle, , obtained form
in-situ tests or laboratory tests:
c* = uc c * = tan-1 (u (tan ))
CVEN3201 Applied Geotechnics 102 Retaining Walls

The strength reduction factors uc and u are defined in the following table. Note that in an
undrained analysis u should be taken as zero for design purposes.

Soil or fill properties: c and


Strength factor
Fill Class I (98%) Fill Class II (95%) Uncontrolled Fill In-situ Soil
u 0.95 0.90 0.75 0.85
Strength
uc 0.90 0.75 0.50 0.70
u 1.0 0.95 0.90 1.00
Serviceability
uc 1.0 0.85 0.65 0.85
Soil or fill property: cu
Strength uc 0.60 0.50 0.30 0.50
Serviceability uc 0.90 0.80 0.50 0.75

AS4678 recommends different reduction factors for soil strength parameters, depending on
the quality of the fill material. For this purpose different classes of fills are defined:
Fill Class I: Material that has been placed at a site in a controlled fashion to ensure the
resultant material is consistent in character, placed and compacted to an average density
equivalent to 98% of the maximum dry density for the material. For cohesionless soils,
material compacted to at least 75% Density Index is satisfactory.
Fill Class II: Material that has been placed at a site in specified layers in controlled fashion
to ensure that the resultant material is consistent in character placed and compacted to an
average density equivalent to 95% of the maximum dry density for the material. For
cohesionless soils, material compacted to at least 65% Density Index is satisfactory.
Uncontrolled fill: Material that has been placed at a site and that does not fall under the
definitions for materials for Class I and Class II.
In situ soil: Natural soil, weathered rock and rock materials.
Structural classification factor
The Australian standard AS4678 takes into account the consequence of failure of the retaining
wall in design. The design resistance, R, needs to be factored by a structural classification
factor, n. The value of n for the strength limit state is given in the table below. For the
serviceability limit state n is 1.0 for all structure types. Structure types are classified
according to the perceived consequences of failure as defined in the following table.

Structure Type Consequence of failure n


1 Where failure would result in significant damage or loss of life 0.9
2 Failure results in moderate damage and loss of services 1.0
3 Failure results in minimal damage and loss of access 1.1

Appendix D of AS 4678 gives some of information on soil properties. Designers would


normally be expected to use actual measured values of soil properties. However values are
given for typical soils and these are given in the following table.
CVEN3201 Applied Geotechnics 103 Retaining Walls

Soil group Typical soils in group c (kPa)


Soft to firm clay of medium to high plasticity, silty clays,
Poor 05 17 25
variable clayey fill, loose sandy silts
Stiff sandy clays, gravelly clays, compact clayey sands and
Average 0 10 26 32
sandy silts, compacted clay fill (Class II)
Gravelly sands, compacted sands, controlled crushed
Good 05 32 37
sandstone and gravel fills (Class I), dense well graded sands
Weak weathered rock, controlled fills (Class I) of road base,
Very good 0 25 36 43
gravel and recycled concrete

Design of Gravity Retaining Walls


The most common retaining walls are massive gravity walls, cantilever walls and counterfort
wall. Massive gravity walls rely only on the mass of the wall for stability. Examples are
gravity masonry walls, crib and bin walls, and gabions. Cantilever walls, usually in the shape
of or , rely on the weight of the wall and the soil on top of the wall foundation for stability.
Counterfort walls are similar to cantilever walls strengthened by vertical brackets at the back
connecting the wall slab to the wall foundation and therefore provide greater moments
capacity for wall structure.
In the stability analysis of a gravity wall four modes of failure are usually considered. These
are sliding (or translation), overturning, bearing capacity, and overall failure of the soil and
wall. To satisfy serviceability criterion, it is generally necessary to limit the wall deformations
within acceptable limits.

Modes of failure
Sliding failure
Sliding failure occurs when the horizontal frictional force
developed at the base of the wall foundation, Rh, is less
than the resultant horizontal force due to the active and
passive pressures. The active and passive pressures can be
determined from either Rankines theory of earth pressure
or graphical Coulomb method. Because the resultant
horizontal force applied to the wall is dependent on any
passive pressures developed in front of the wall, it is
normal to ignore the upper 0.5 to 1 m of soil contributing
to the passive pressures. This reduces the possibility of
inadvertent excavation leading to failure.
Overturning failure A
If the wall height becomes large then there will be a
significant moment due to the active earth pressures. The
wall may rotate about the toe, point A in the figure
opposite. At this limiting state the overturning moment due
to the earth pressures must be balanced by the restoring
moment due to the weight of the wall.
Bearing capacity failure
If the stress due to the weight of the wall and the weight of
the soil above the wall footing is large there is a possibility
that the underlying soil will not be able to support it. This
CVEN3201 Applied Geotechnics 104 Retaining Walls

is known as bearing capacity failure. It should be noted that due to the earth pressures acting
on the wall there will be a moment (eccentricity of the normal load on the foundation) and
horizontal force acting on the base of the wall. The moment and the horizontal load
significantly reduce the vertical bearing capacity of the
wall foundation.
Global failure
A check is required on the overall stability of the wall
combined with the surrounding the wall. This may be
analysed by a method used for assessing slope stability.

Design procedure
Having the design strength parameters, c and for a design based on an overall factor of
safety or c* and * for a design based on partial factors of safety, the distribution of lateral
earth pressure on both sides of the wall is determined. Lateral earth pressures can be evaluated
by the Rankines theory or the Coulomb method, using the appropriate values for soil unit
weight and surcharge or any other load applied on the ground level. AS 4678 recommends a
minimum live load of 5 kPa to be applied to the soil at the back of the wall.
For a design based on an overall factor of safety, the resisting forces or moment must be
greater than the de-stabilising forces or moment multiply by a factor of safety.
For a design based on partial factors of safety, the design resistance must be greater than or
equal to the design action forces or moment.

0.4m
Example R2
Use AS4678-2002 recommendations and evaluate the = 35o
stability of the cantilever wall shown in the figure opposite = 18kN/m3
4.4m
against sliding and overturning. The concrete wall has a unit
weight of c = 25kN/m3. The uncontrolled backfill soil has a
friction angle of = 35o and a unit weight of = 18kN/m3. 0.4m
The friction between the wall base and the soil is = .
0.6m
2m 1.5m
CVEN3201 Applied Geotechnics 105 Retaining Walls
CVEN3201 Applied Geotechnics 106 Retaining Walls

Design of Embedded Retaining Walls


Embedded walls are generally steel sheet piles, soldier piles with sheeting placed in between
them; and contiguous bored pile walls. Each wall type may act as a cantilever or be supported
by one or more rows of anchors or props. They can be used either as temporary supports
during construction, or for permanent structures such as quay or basement walls. The walls
range from relatively flexible steel sheet piles to relatively stiff diaphragm walls. Several
structural sections are made especially for sheet pile walls, some with watertight connections.
Construction commences with the installation of the piles. Then either the soil in front of the
wall is excavated or fill is added to the back side of the wall. Cantilever embedded walls, used
usually to support low heights or temporary works, rely on the passive pressures developing
on a section of pile embedded below the ground. Anchored sheet pile walls are usually used
as the height of walls increases. Application
of anchors reduces the moment developed in Concrete pile Wale
the structure of a high cantilever wall. Strut
Braced walls are temporary walls used
during excavations and behave similar to
anchored walls. Anchored walls and braced
walls rely on anchors and struts respectively
for lateral stability, as well as the passive
pressures developed on the embedded
section of the wall. The figure opposite
shows a contiguous pile wall made up of Anchor Embedded
cast-in-situ concrete piles and its various section of pile
components.
Embedded walls are generally more expensive than gravity walls but their cost is balanced by
the speed of construction and lack of temporary support. Cantilever walls are only suitable for
moderate retained heights, typically less than 5 m, and at most 10m if a stiff reinforced
concrete wall is formed. Significant ground movements can occur behind cantilever walls,
and they are generally unsuitable if services or foundations of adjacent buildings are close.
The use of anchors or props can reduce the required
penetration length, the ground deformation, and the
bending moments in the walls.

Embedded cantilever walls Deformed wall


Active
The failure mechanism for the embedded cantilever walls
is due to the rotation of the wall caused by lateral yielding
of the ground, as shown in the figure opposite. The dotted Passive O
line shows the wall after movement. The wall rotates
about a point below the toe, point O. Above point O there Active Passive
will be active pressures at the back of the wall and
passive pressures in front of the wall. But below point O
the wall moves in an opposite sense relative to the soil so
that the active pressures will develop in front of the wall
and passive pressures at the back.
H
In assessing the stability of the wall it is usually assumed
that the active and passive earth pressures vary linearly Active
with depth, which is a realistic assumption if the wall is
rigid. This assumption is an approximation to the real
distribution of earth pressures for flexible walls, however, x Passive O
even for these walls it is usually assumed that the D
pressures vary linearly with depth. The effect of this Active Passive
CVEN3201 Applied Geotechnics 107 Retaining Walls

approximation is small in the final design.


The figure opposite shows the distribution
of lateral earth pressures acting on a
H
cantilever embedded wall due to a
cohesionless soil assuming a linear
variation for earth pressures. In this figure
H is the height of the wall, D is the
embedment depth and x is the depth of the
x Kp x Ka x)
point of rotation.
Active
To determine the depth of embedment, D, D
Passive Kp x)
O
for a given wall height H, equilibrium
equations for both moment and horizontal Ka x Active Passive
force must be formulated and solved, and Ka D Kp D)
the two unknown, D and x, are calculated.
The forces applied on the wall from a cohesionless soil are shown in the figure opposite. The
forces are:
PA1 = 0.5Ka (x+H)2
PP1 = 0.5Kp (x)2
PA2 = Ka x(D-x) +0.5Ka (D-x)2 H
PP2 = Kp (x+H)(D-x) +0.5Kp (D-x)2
The sum of the active forces, PA, is:
PA = PA1 PA2 PA1
= 0.5 Ka (H + 2xH + 2x D )
2 2 2
x PP1
The sum of the passive forces, PP, is: D
PP = PP2 PP1 O
PA2 PP2
= 0.5 Kp (2DH - 2xH - 2x2 + D2)
Equilibrium of forces results in:
Ka (H2 + 2xH + 2x2 D2) - Kp (2DH - 2xH - 2x2 + D2) = 0
or
Ka (H2 + 2xH + 2x2 D2) = Kp (-2DH + 2xH + 2x2 - D2) (A)
The left hand side of the above equation is due to the active forces which cause failure,
whereas the right hand side is due to the passive forces which resist failure.
Taking moment about the point of rotation, point O, results in the following components of
moment:
MA1 = Ka (x+H)3/6
MA2 = Ka x(D-x)2/2 +Ka (D-x)3/3
MP1 = Kp (x)3/6
MP2 = Kp (x+H)(D-x)2/2 +Kp (D-x)3/3
The resultant moments due to the active and the passive earth pressures are:
MA = MA1 + MA2 , MP = - (MP1 + MP2)
Therefore equilibrium of moments results in:
MA1 + MA2 = MP1 + MP2 (B)
Equations A and B can be solved simultaneously for the unknown x and D. These two
equations are complex and can best be solved numerically, graphically or by a trial and error
method. Note that these equations are obtained for a simple case where the soil is uniform and
cohesionless. For other cases the equations become more complex.
CVEN3201 Applied Geotechnics 108 Retaining Walls

The right-hand-sides of equations A and B represent contributions from the forces or


moments which resists failure. These contributions are all to the passive earth pressures.
Therefore, in design of walls based on an overall factor of safety, it is convenient to first
reduce the passive earth pressure by a factor of safety and then formulate the equations for
equilibrium of forces and moments and solve them for the unknowns. In a design based on the
partial factors of safety method and AS 4678, the loads and unit weight of the soil are factored
depending on the state of the soil. For example, must be multiplied by a load factor of 1.25
when computing active pressures and by 0.8 when computing passive pressures. The shear
strength parameters, c and , are also reduced by different strength reduction factors as
indicated in AS 4678. The equations for equilibrium of forces and moments are formulated,
similar to equations A and B. The right-hand-sides of the equations are the resisting moments
and forces and, after applying the structural classification factor n to them, must be equal or
greater than the left-hand-sides of the equations which are the disturbing forces and moments.
To find the maximum moment in the wall it is first noted that the maximum moment occurs
where the shear force is zero. The point of zero shear
can be found by formulating equilibrium of forces on
the section of the wall above the point of zero shear.
The form of the earth pressure diagram suggests a
simplified approach which is often referred to as the
Free Earth method of analysis. The segment of wall H
below the point of rotation is always relatively small.
The moment on this segment can be ignored in the
analysis and the effects of forces applied on this
segment can be replaced by a single lateral force, Rd, PA1
acting at the point of rotation. Moment equilibrium
about point O may be used to determine the depth to
x PP1
the point of rotation, x. The depth of embedment is D
then approximated as being 10% to 20% greater than x, O Rd
i.e., D 1.1x 1.2x. Note that it is not necessary to
determine the value of Rd in the simplified method.

Example R3 Sand
Use a factor of safety of 1.5 and determine the depth of = 30o
2.0m
embedment, D, required for a 2m high retaining wall embedded in a = 18kN/m3
soil which has a friction angle of = 30o and a unit weight of
= 18kN/m3.
What would be the required embedment depth based on AS4678?

D
CVEN3201 Applied Geotechnics 109 Retaining Walls
CVEN3201 Applied Geotechnics 110 Retaining Walls

Embedded anchored walls Anchor


An anchored or tied back wall is analysed by assuming
a mode of failure as shown in the figure opposite. The Deformed
depth of embedment is calculated assuming that the wall
Active
wall fails by rotating outwards below the anchor point.
The anchor force is then calculated using equilibrium of
forces on both sides of the wall.
It is assumed that a full passive earth pressure will be Passive
developed in front of the wall. The lateral passive and
active earth pressures are assumed to vary linearly with
depth. The variation of earth pressures on an anchored
wall due to a cohesionless soil is shown in the figure
opposite. The effect of the anchor is represented as a a T
force, T, acting at the anchoring point.
The active and passive forces on the wall, PA and PP, H
developed by a cohesionless soil are:
PA = 0.5Ka (D+H)2 PA
PP = 0.5Kp D2
Equilibrium of moments about the anchor point gives D PP Active
the following cubic equation in terms of the unknown
depth of embedment, D. Passive
0.5Kp D2 (2D/3+H-a) = 0.5Ka (D+H)2(2D/3+2H/3-a) (C)
where a is the depth of anchor point below the ground level behind the wall.
The anchor force per unit distance along the wall, T, can be obtained from equilibrium of
forces on both sides of the wall in the horizontal direction.
T = PA PP = 0.5Ka (D+H)2 - 0.5Kp D2 (D)
Note that if the depth of the embedment is increased for practical purposes, the passive force,
PP, in front of the wall will also increase. However, in calculation of the anchor force, T, the
minimum design requirement for the depth of embedment, D, obtained from the cubic
equation (C) should be considered for the equilibrium of forces. Alternatively, the anchor
force can be obtained as the maximum force required to maintain both equilibrium of forces
and equilibrium of moments.
In design of anchored walls based on an overall factor of safety, it is convenient to first reduce the
passive earth pressure by a factor of safety and then formulate the equation for equilibrium of
moments, equation (C), and solve it for the depth of embedment. The design anchor force obtained
from the equilibrium of forces, equation (D), also needs to be multiplied by a factor of safety.
In a design based on partial factors of safety method and AS 4678, the loads and the soil unit
weight are factored depending on the state of the soil. The shear strength parameters, c and ,
are also reduced by different strength reduction factors. The equation for equilibrium of
moments, equation (C), is formulated. The left-hand-side of this equation is the resisting
moment and, after applying the structural classification factor n, must be equal or greater
than the right-hand-side of the equation which is the disturbing moment. The ultimate anchor
force, T, can be obtained from equation (D) after multiplying both T and the passive
resistance by the structural classification factor, n.
In the design of anchored walls for structural integrity, the maximum bending moment and shear
force developed during construction, when the anchor has not been installed, and under permanent
loads, when the anchor is installed, need to be calculated. Typically the excavation is taken down
0.5m below the level of anchor to allow some vertical space for installation of the anchor.
CVEN3201 Applied Geotechnics 111 Retaining Walls

Effects of nonlinear distribution of earth pressure


The distribution of earth pressures acting on a flexible wall
tend to be more uniform rather than linearly varying with
depth. Part of the reason is that more flexible walls may
yield under the earth pressures which results in reduction of Actual earth
earth pressures. The actual earth pressure distribution due
pressure
to a cohesionless soil on an anchored wall is similar to that
shown in the figure opposite. Note that a design based on
linear variation of earth pressure underestimates the force
Assumed
that would be developed in the anchor, T. However, a earth pressure
design based on a linear variation of earth pressure results
in a conservative estimate of the depth of penetration and
overestimates the design moment for the wall. For design
purposes, the moment calculated based on a linear distribution of pressures, as explained in
these notes, can be reduced. The reduction takes account of yielding of the wall and is a
function of the relative stiffness of the wall and the type of soils behind the wall. Details of
the reduction factors are given in many text books on geotechnical engineering.

Multiple anchor levels and braced walls


Where there are relatively deep temporary/permanent
excavations, it is common to support the walls during
construction by a system of bracing. This procedure is also used
for permanent structures with struts forming the floors or the
basement. Alternatively, the walls can be supported by multiple
rows of anchors at different levels.
A wall with several layers of struts or anchors will have increased
restraint as each layer of anchors is added. Consequently the
lateral deformations are limited and the restrained soil is unlikely
to attain failure in a normal active state. The distribution of earth
pressure tends to be more uniform with depth.
A system of multilevel anchors and wall is statically
indeterminate and analysis is complex. The earth pressure acting
on the wall will depend on the relative stiffness of wall and the
soil behind the wall, as well as the spacing between different
levels of anchors or struts and their load deformation response.
In practice empirical methods are used to estimate the pressures
on the wall and the forces developed in the anchors or struts.
Details of such a method and a distribution of the active earth
pressure were presented by Terzaghi and Peck (1948) for sand,
soft clay and stiff clay, as shown in the following figure.

0.25H 0.25H

H H H 0.50H
0.75H
0.25H

0.65KaH
The greater of
0.2H to 0.4H
H-4cu) and 0.3H
Sand
Soft to firm clay with H/cu>4 Stiff clay with H/cu<4
CVEN3201 Applied Geotechnics 112 Retaining Walls

The structure of braced walls is often designed assuming that there is a hinged at each
strut/anchor level. This simplified assumption makes it possible to easily calculate the forces
in the struts or anchors and the bending moment in the wall.

Anchors
The top anchor level is normally 1.5m below the ground level and subsequent levels of
anchors are then spaced at vertical intervals of around 2.5m to 3m. Anchors are usually
spaced 2-3m apart (in the
Fixed anchor length 45-/2
horizontal direction along
the wall) and a wale is used
to provide the necessary
Deadman
stiffness between anchor
anchor
points. Anchors are usually
inclined downwards at 15o Critical Critical
o failure line failure line
to 20 or more in order to
increase the soil stresses
around the anchor thereby 45+/2 45+/2
increase the capacity of the
anchors.
Anchors must be able to provide a resistance equal to the required anchor force, T, without
excessive displacement of the anchorage toward the wall. The design of anchors is beyond the
scope of these notes. There are many anchoring systems used in practice. They can generally
be classified as deadman anchors or tie-back anchors, as shown in the figures opposite. In
each case the actual anchoring action must be generated outside the failure zone, drawn as
dashed lines in the figures, behind the wall.

Design for serviceability limit states


Gravity walls
Most of the deformations of gravity walls occur during construction. Any additional
deformations after backfilling are normally due to changes in groundwater conditions and
surcharge loadings. Evaluation of deformation is normally not necessary for gravity retaining
walls unless the consequence of deformations is severe. For example, walls which support
sensitive services or heavy structural loading require special consideration. The effects of
consolidation settlement of foundation soil, especially under eccentric loading of wall footing,
may results in tilting of walls.
If deformation calculations are required, the at-rest earth pressure and compaction-induced
lateral pressures should be considered unless movement of the wall is sufficient to mobilize
active earth pressure. AS4678 makes no recommendations regarding at rest design as it is a
limit state code and the at rest state is a working state. As Ko is a very difficult property to
measure, many engineers use an empirical relationship as Ko=1sin. In partial factors of
safety design approach a reduced value of friction angle, , shall be used in calculation of Ko.
Embedded walls
Some ground movements must be expected around sheet pile walls. Where movements can be
tolerated then design is normally based on active earth pressures, i.e., Ka is used. Where wall
movements need to be minimised, for example where there is a building near the wall, design
should be based on the at rest earth pressures using Ko.
There are many examples in the literature where the deformations around tie-back walls have
been measured. Typically they show vertical and horizontal movements in the soil at the top
of the wall ranging from zero to 0.005H for walls in sand and gravel and very stiff to hard
CVEN3201 Applied Geotechnics 113 Retaining Walls

clay. For walls in soft to stiff clay the deformation is much greater, 0.005H to 0.015H.
Movements extend back into the soil behind the wall for a distance of around 2H however the
bulk of the movement occurs in the first H back from the wall.

References
Terzaghi, K. and Peck, R.B (1948), extended and reprinted as Terzaghi, K., Peck, R.B. and
Mesri, G (1996) Soil Mechanics in Engineering Practice, John Wiley & Sons.
CVEN3201 Applied Geotechnics 114 Retaining Walls

1.5m
Example R4
Use the requirements of AS4678-2002 and determine the depth of
embedment for the 4m high retaining wall shown in the figure
4.0m
opposite. The soil has a friction angle of = 30o and a unit weight Sand
of = 18kN/m3. = 30o
Determine the tension in the anchors if they are going to be = 18kN/m3
installed at 3m intervals.
D
PAST EXAM QUESTIONS

2008
1- (Now a tutorial question) An
eccentrically loaded strip footing
with a width of 2.5 m is embedded in Proposed excavation 1.0m
soil. The base of the footing is 1.0 m Pu
below the soil surface. The ground
water level is well below the footing. 1m
The soil has the following properties: 2.5m
Unit weight t = 18 kN/m 3

Friction angle = 30o 25o


Section
Cohesion c=10kPa
An excavation is going to be made close to the strip footing as shown in Figure 1 in dotted
lines.
a) Determine the ultimate load bearing capacity of the foundation, Pu, in kN per metre run of
the footing before excavation. (2024 kN pmr)
b) Determine the ultimate load bearing capacity of the foundation, Pu, in kN per metre run of
the footing after the excavation. (946 kN pmr)
Use the Hansens theory of bearing capacity and the effective width concept in all your
calculations.

2- A soil profile at a site is given below where a 4m4m square footing is to be constructed at
a depth of 2 m below the ground level. The footing carries a load of 8000 kN. The properties
of different soil layers are as follows:
P
Sand t = 18 kN/m3
E = 100MPa 2m
Sand
= 0.3 4m
Clay t = 19.8 kN/m 3 4m
Cc = 0.4
Cr = 0.1 Clay 2m
pc=260 kPa
Rock
eo = 0.5
The water table is 6 m below the ground level.
a) Calculate the settlement of the footing due to deformation of the sand layer only. (6.9mm)
b) Calculate the long term settlement of the footing due to the consolidation of the clay layer
only. Settlement calculation for the clay layer may be based on a one-point estimation
using the soil conditions at the middle of the layer under the centre of the footing. (30mm)

3- A square group of 4 circular concrete piles is arranged with 2 rows and 2 columns and a
centre to centre spacing of 2 m. The piles have a diameter 1.0 m and are to be cast-in-situ to a
depth of 10 m in a deep layer of stiff clay. The clay has an undrained cohesion of cu=140 kPa
and a Youngs modulus of 70 MPa.
a) Calculate the allowable vertical bearing capacity of the each individual pile assuming a
factor of safety of 3. (916.3 kN)
b) Calculate the ultimate vertical bearing capacity of the pile group, taking into account the
CVEN3201 Applied Geotechnics 116 Past Exam Questions

block failure of the entire group as well as the capacity of individual piles. (10996 kN)
c) Calculate the vertical settlement of a single pile when it is loaded to 1000 kN. You may
assume that the Youngs modulus of the concrete pile is 21GPa. (2.3 mm)
d) Determine the settlement of the pile group, assuming that the group has a rigid pile cap
which carries a total load of 6000 kN. (8mm)

4- Figure opposite shows a section of a 4m


4 m high gravity wall. The wall is
inclined toward the backfill at an angle Uniform loading = 10kPa
equal to 80o measured with respect to the
horizontal. The backfill soil is
cohesionless and has a unit weight of
80o
t=18 kN/m3 and a friction angle of
=30o. The backfill supports a building
4m
which applies an average pressure of Failure line
10 kPa to the ground.
Calculate the lateral force on the wall due
to the active earth pressure using the 40o
Coulombs method and the trial failure
line shown in Figure 3. The wall friction
angle may be taken as = 20o. (32 kN @ 10o wrt horizontal)

5- Figure 4 shows a section of a 4 m high retaining wall embedded 2 m into the ground. The
soil profile consists of 4 m of sand over deep layer of clay. The material properties of the sand
and clay are:
Sand: t = 16 kN/m3 = 34o c = 0
Clay: t = 16 kN/m 3
u = 0 o
cu = 150 kPa
There water table is at the surface of the soil in front of the wall.
a) Use the Rankins theory of lateral earth 5kPa
pressure and determine the active and
passive total horizontal stresses applied to
points 1 to 6, as shown in Figure 4, shortly 1
after installation of the wall. The rotation
point for the wall is shown in the figure. You
should apply the requirements of AS4678 4m Sand
and the partial factors of safety method in
estimation of the soil pressures and assume
the soil is in-situ. You should also consider a 2
minimum live load of 5kPa applicable to the 5 3
soil at the back of the wall. If tension cracks 1.8m Clay
occur in the soil below the water table, you
should assume that the cracks are filled with 6 4
0.2m
water. Rotation point
Clay
b) Evaluate the stability of the wall based of the
partial factors of safety method and the
requirements of AS4678. Use a structural classification factor of n=0.9. Note that you
need to calculate both R* and S* for the wall. You may use the simplified procedure which
ignore the moments generated below the rotation point.
(S*=214.6 kN.m pmr, n R*=230 kN.m pmr)
CVEN3201 Applied Geotechnics 117 Past Exam Questions

2009 3m
P
1- A circular footing with a diameter of 3m is
founded 1m below the surface of a deep layer 1m
of uniform clay. The ground water level is 1m Clay
below the base of the footing. The soil 1m
consists of a homogeneous clay having the
following properties:
cu= 60 kPa u= 0o
c' = 5 kPa ' = 25o
t= 20 kN/m3 E = 7MPa = 0.3
a) Use the Hansens theory of bearing capacity to determine the allowable net bearing load (in
kN) which could be placed on the footing using a factor of safety of 3. Consider both the
following cases: (1013 kN)
(i) Immediately after construction of the footing
(ii) A long time after construction of the footing
b) Estimate the settlement of the footing immediately and a long time after application of the
allowable load, using the elastic theory. (29 mm, 39 mm)

2- A load test was carried out on a cast-in-situ reinforced concrete pile embedded 10m in a
deep layer of uniform clay. The pile diameter is 400mm.
a) Failure occurred at a load of 670 kN. Deduce the average value of soil cohesion, cu.(60kPa)
b) Settlement of the pile under a load of 250 kN was 9.5 mm. Deduce a representative value
for Youngs modulus of the soil. Assume that the Youngs modulus of the concrete pile is
Econcrete= 20,000MPa. (5MPa)
c) Estimate the settlement of a group of 9 piles of this type with a spacing equivalent to 5
diameters, carrying a total load of 1800kN. (28.1mm)
Note that an iterative solution may be required in this question.

3- Figure opposite shows a section 1.2m


75o
of a 4 m high gravity wall. The wall
is inclined toward the backfill at an
angle equal to 75o measured with
respect to the horizontal. The
backfill soil is cohesionless and has
a unit weight of t=20 kN/m3 and a 4m Failure line
friction angle of =30 . The soil-
o

wall friction angle is =25o. There


is no sign of water in the soil.
a) Calculate the lateral force on the 50o
wall due to active earth pressure
using the Coulombs theory of
lateral earth pressure and the trial failure line shown in the figure. (31.8 kN pmr)
b) Calculate the overall factor of safety against overturning of the wall assuming that the wall
has a uniform thickness of 1.2 m and a unit weight of 20 kN/m3. (2.83)
CVEN3201 Applied Geotechnics 118 Past Exam Questions

4- Figure opposite shows a section of 10kPa


a 4 m high anchored retaining wall
embedded 0.8 m into the ground. A 1
uniform surcharge of 10kPa is 1.2m
applied on the surface of the soil at Td
the back of the wall.
The soil profile consists of 4m of 4.0m Sand
sand over deep layer of clay. The
material properties of the sand and
clay are:
2
Sand: t = 18 kN/m3 5 3
Clay Clay
= 34 o
c = 0 0.8m
Clay: t = 20 kN/m3
6 4
u = 0 cu = 80kPa
The water table is at the surface of the clay layer.
a) Use the Rankins theory of lateral earth pressure and determine the active and passive
horizontal stresses, shortly after installation of the wall, at points 1 to 6, as shown in the
figure. You should apply the requirements of AS4678-2002 and the partial factors of safety
method in estimating the earth pressures. Assume in-situ conditions for the soil.
b) Determine the ultimate tension force for the anchor, Td, per meter run of the wall. You
may assume a structural classification factor of n=1.0. (38.9 kN pmr)
CVEN3201 Applied Geotechnics 119 Past Exam Questions

2010 3m
1- A square footing of 3m 3m is founded 1.2m
below the surface of a deep layer of soil. The soil
has the following properties:
M
c' = 10 kPa ' = 35o 3m
P
t= 20 kN/m3
The foundation is subjected to a centrally applied
vertical load of P = 7000 kN and a moment of PLAN
M = 2100 kN.m, applied parallel to one side of
the footing as shown in Figure 1. The ground 3m
water level is 5m below the base of the footing. P
a) Use the Hansens theory of bearing capacity M
and determine the factor of safety of the
1.2m
footing against bearing failure. (2.91)
b) To consider the effects of flooding on the
factor of safety, assume the water table is at
the surface of the soil and re-calculate the
SECTION
factor of safety for this case. (1.92)
Figure 1

2- (Now a tutorial question) A 0.5m diameter circular concrete pile is driven to a depth of
10 m in a deep layer of homogeneous dry dense sand. The sand has a friction angle of = 38o,
a unit weight of t = 18 kN/m3, and a Youngs modulus of Es = 100 MPa. The Youngs
modulus of concrete pile can be assumed as Ep = 25GPa. The pile is subjected to horizontal
and axial forces as shown in Figure 2.
a) Determine the allowable axial bearing capacity of the pile
assuming a factor of safety of 3. (2827kN) P
b) Evaluate the settlement of the pile under an axial load of
H
P = 2000 kN. (5.4mm)
c) Determine the ultimate horizontal capacity of the pile, Hu.
Assume that the concrete pile has a yield moment capacity
of My = 2500 kN.m. (945kN)
d) Estimate the horizontal displacement the pile head if a 10m
horizontal load of H = 500 kN is applied to the pile head.
Assume that the pile section has a second moment of
inertia of Ip = 0.004m4. (5.1mm, 0.185o)
e) What would be the influence of water on the vertical and
horizontal capacity of the pile? Assume that the water =0.5m
table is at ground level and evaluate the ultimate vertical
and horizontal capacity of the pile based on the values Figure 2
obtained in parts (a) and (c) of the question. No detailed
calculation is required.
CVEN3201 Applied Geotechnics 120 Past Exam Questions

3- Figure 3 shows a section of a 4 m high gravity wall. The backfill soil is inclined with a
slope angle of 20o with respect to horizontal. The backfill soil is cohesionless and has a unit
weight of t=20 kN/m3 and a friction angle of =30o. The soil-wall friction angle is =25o.
There is no sign of water in the
soil. 2.9m
a) Calculate the horizontal force
on the wall due to the active
earth pressure using the 1.05m
20o
Coulombs theory of lateral
earth pressure and the trial
failure line shown in Figure 3.
(52.76kN)
b) Calculate the horizontal force 4m Failure line
on the wall due to the active
earth pressure using the closed
form solution proposed by
Rankine. (62.3kN)
60o
c) Briefly give a reason for the
different values you obtained
from part (a) and part (b) of Figure 3
the question.

4- Figure 4 shows a section of a 4 m high anchored retaining wall embedded 1 m into the
ground. A uniform surcharge of 5kPa is applied on the surface of the soil at the back of the
wall.
The soil profile consists of a deep layer of sand. The sand has a unit weight of t = 18 kN/m3,
and a friction angle of = 34o. There is no sign of water in the soil
a) Use the Rankins theory of lateral earth pressure and determine the active and passive
horizontal stresses at points 5kPa
1 to 4, as shown in Figure 4.
You should apply the
requirements of AS4678- 1
2002 and the partial factors
of safety method in
estimating the earth Anchor
pressures. Assume in-situ 4.0m
conditions for the soil.
b) Evaluate the stability of the
wall against rotation around
point 1, assuming that the
anchor has sufficient Sand
3
horizontal capacity to
prevent pullout failure. You 1.0m
may assume a structural
Sand
classification factor of 4 2
n=0.9.
(Ma=343.75kN.m, Figure 4
Mp=100.8kN.m)
CVEN3201 Applied Geotechnics 121 Past Exam Questions

2011
1- A square footing is going to be designed to carry a vertical load of 5900 kN. The footing
will be constructed at the surface of a deep soil stratum which has a cohesion of c = 45 kPa, a
friction angle of = 25, and a unit weight of t = 17 kN/m3. There is no sign of water up to a
depth of 5 m.
a) Use the Hansens theory of bearing capacity and determine the size of the footing,
assuming a factor of safety of 3. (3.4m)
b) What would be the factor of safety of the foundation if the water table rises to the ground
level due to flooding? (2.87)

2-(Now a tutorial question) A strip footing with a width of 2m is founded 2m below the
ground level and subjected to a vertical central load of 1500 kN per metre run of the footing.
The soil is a well graded dense sand (previously under large compressive pressure) with a unit
weight of t = 20kN/m3. The water table is 4m below the ground level. The following table
shows the results of a cone penetration test, given as average values over given depth ranges
below the ground level.

Depth below the ground (m) 0-2 2-4 4-6 6-10 10-14 14-20
qc (kPa) 2500 3000 3500 4000 5000 6000

Use the Schmertmanns method and evaluate the settlement of the footing 5 years after the
construction. (151mm)

P
3- A 15m long concrete pile with a diameter of H
0.5m is driven 12m into the ground. The pile is 2.5m 3m
2m water
subjected to a horizontal load, H, applied at a
height of 2.5m above the ground level, and a
vertical load, P, as shown in Figure 1.
The soil, which is submerged under 2m of water,
can be regarded as to behave under undrained 12m
conditions with an undrained cohesion of
cu = 60 kPa. The undrained Youngs modulus of
the soil is assumed to be Esu = 46 MPa.
The Youngs modulus of the concrete pile can be
taken as Ep = 24 GPa. The pile has a yield =0.5m
moment capacity of My = 3000 kN.m and a
second moment of inertia of Ip = 0.004 m4. Figure 1

a) Determine the ultimate axial capacity of the pile, Pu. (1011kN)


b) Determine the ultimate horizontal capacity of the pile, Hu. (668kN)
c) Estimate the ground line horizontal displacement the pile if a horizontal load of
H = 460 kN is applied to the pile. (21.3mm)
CVEN3201 Applied Geotechnics 122 Past Exam Questions

4- Figure 2 shows a section of a 4 m high gravity retaining wall. The wall is inclined toward
the backfill at an angle equal to 76 measured with respect to the horizontal. The surface of
the backfill soil is inclined with a slope angle of approximately 20 with respect to the
horizontal. The backfill soil is cohesionless and has a unit weight of t=20 kN/m3 and a
friction angle of =35. The soil-wall
1.5m 3.06m
friction angle is =24. There is no sign
of water in the soil.
a) Calculate the lateral force on the wall 1.2m
due to active earth pressure using the
Coulombs theory of lateral earth 1
pressure and the trial failure line 4
shown in Figure 2. (53.55 kN)
76o Trial Failure line
b) Using your answer from part (a), 4
calculate the overall factor of safety
against overturning of the wall
assuming that the wall has a uniform
thickness of 1.5 m and a unit weight 60o
of 25 kN/m3. (3.52)
Figure 2

5- A concrete retaining wall supports a sandy clay soil as shown in Figure 3. The soil has the
following properties:
t = 17 kN/m3, c=20 kPa, =25.
There is no sign of water in the soil. However, 0.5m
rain water may fill any tension crack that may
exist on the surface of the backfill before being
drained out. The unit weight of concrete can
be taken as concrete = 25kN/m3.
Check the stability of the wall against rotation
based on the partial factors of safety and the 5m Sandy clay
requirements of AS4678. Use a structural
classification factor of n = 0.9 and assume
uncontrolled backfill. You may ignore the
effects of soil in front of the wall.
(R*=146.25kN pmr, S*=136.2 kN pmr) 2m
Figure 3
CVEN3201 Applied Geotechnics 123 Past Exam Questions

2012
1- A large footing, 6m long and 4m wide, is founded at 1.5m below the ground level in a deep
soil layer. The footing carries 3000kN vertical load from column A and 1000kN vertical load
from column B. The locations of the columns are shown in Figure 1.
The soil has the following properties:
Undrained cohesion, cu 60 kPa
Undrained angle of friction, u 0
Effective cohesion, c 0
Effective angle of friction, 25
Dry unit weight, d 16kN/m3
Saturated unit weight, sat 17.5kN/m3
Use the Hansens theory of bearing capacity and determine the minimum factor of safety of
the footing against bearing failure. For design the water table can be considered to be at a
depth of 1.5m and the soil above the water table to be dry. (2.33)
6m

A B
4m

3m 1m 2m
PLAN

A B

3000kN 1000kN
1.5m

SECTION

Figure 1

2- A strip footing with a width of 3m is founded 1m below the ground level and subjected to a
vertical central load of 2400kN per metre run of the footing. The soil is a well graded dense
sand (compressed previously under a higher surcharge) with a unit weight of t = 20kN/m3.
The water table is 4m below the ground level. The following table shows the results of cone
penetration test, given as average values over given depth ranges below the ground level.

Depth below the ground (m) 0-2 2-4 4-8 8-15 15-20
qc (kPa) 2500 2500 3500 4000 5000

Use the Schmertmanns method and evaluate the settlement of the footing 10 years after the
construction. (290mm)
CVEN3201 Applied Geotechnics 124 Past Exam Questions

3- A 10m long concrete pile with a diameter of 0.5m is driven into a deep layer of dense sand.
The sand has a friction angle of = 40, a unit weight of t = 19kN/m3, and a Youngs
modulus of Es = 100MPa.
The pile is subjected to a vertical load, P, and a
horizontal load, H, applied 2.5m above the P
ground level as shown in Figure 2. There is no
sign of water in the soil. H
The Youngs modulus of the concrete pile can be 2.5m
taken as Ep = 25 GPa. The pile has a yield
moment capacity of My = 5700 kN.m and a
second moment of inertia of Ip = 0.004 m4.
a) Determine the ultimate axial capacity of the
pile, Pu. (8950 kN) 10m
b) Determine the ultimate horizontal capacity of
the pile, Hu. The pile may be assumed to be a
free head pile. (1092 kN)
c) Estimate the ground line horizontal
displacement of the pile if a horizontal load of =0.5m
H = 500 kN is applied to the pile. (13.2mm)

4- Figure below shows the geometry of a soil behind a crib wall. A strip footing applies a
uniform pressure of 40 kPa on the soil at the back of the wall, as shown in the figure. The soil
behind the wall is sand which has a unit weight of t = 20 kN/m3 and a friction angle of
= 40o. The friction angle between the soil and the wall can be taken as = 30o. There is no
sign of water in the soil.
a) Use the Coulomb theory of lateral earth pressure and determine the magnitude of the
maximum resultant force acting on the crib wall shown in Figure 3. The force should be
found based on the two trial wedges defined by the dotted lines in the figure. Sketch the
direction and point of action of the force on the wall. (60 kN pmr)
b) Evaluate the factor of safety of the wall against overturning based on the maximum lateral
force obtained in part (a). The crib wall has a thickness of 1.4 m and a unit weight of
21 kN/m3. (1.67 or 2.31)

1.4m 3m 1m 1m
Strip footing

1m

Sand
5m

Trial line 2

Trial line 1

Crib wall
CVEN3201 Applied Geotechnics 125 Tutorial Questions

BASIC DESIGN
(Self Assessment Tutorial)
You should be able to solve the following problems based on your knowledge from Soil
Mechanics

1- Consider the semi-circular mechanism shown in the figure below for the failure of a long
strip footing on a soil deforming under undrained conditions. Based on this mechanism
failure causes a rotation about point O.
The following soil properties are known: B B
t = 20kN/m 3
q
cu = 100kPa,
u = 0.
a) Calculate the ultimate bearing pressure O
for the footing when the soil deforms
under undrained conditions. (2cu)
b) Evaluate the effects of any surcharge,
qo, which may be applied on the
surface of the soil, on the ultimate
bearing capacity of the footing. (2cu+qo)

2- A long vertical drain is going to be made in a soil. Assume the failure plane shown in the
figure below and evaluate the safety of the cut assuming the following properties for the soil:
t = 22kN/m3
cu = 30kPa, u = 0
c = 0, = 27o.
Consider the following cases:
1) No water in the drain: 5m
2) Drained full of water.
For both cases consider:
a) Drained conditions 45o
b) Undrained conditions.
You should assume that the lining of the
drain does not provide any resistance to failure.
(1a: Safe, 1b: Fail, 2a: Safe, 2b: Fail)

3- Write an expression for the undrained


B O
bearing capacity of a strip footing shown in
the figure opposite. Assume the slip q R
mechanism shown, which is an arc centred
above one edge of the footing (point O).
You may use the following parameters
t , cu , u = 0.
Take as a variable and calculate the
minimum bearing pressure that causes failure
of the soil? (5.52cu)
CVEN3201 Applied Geotechnics 126 Tutorial Questions

BEARING CAPACITY OF SHALLOW FOUNDATION

1- A section through a long strip footing


is shown in the figure opposite. On one Surcharge 20 kPa
side of the footing a slab exerts a
1m
pressure of 20 kPa on the foundation P
0.4m
soil while on the other side the soil is
unexcavated. The foundation soil 1.5m
consists of a layer of clay having the
following properties:
t = 18 kN/m3, cu = 30 kPa, u = 0, c = 5 kPa, = 25o.
The ground water level is 0.4m above at the base of the footing. Determine the maximum
allowable force (pmr) which could be placed on the footing assuming a factor of safety of 3.
(94.5kN pmr)
2- An eccentrically loaded strip footing 1.5m
carries a line load of P (kN per meter run).
The footing is 4m wide and the load is applied
1.5m from one side of the foundation. The 1.5m P
footing base and the water table are located 3m
1.5m and 3m below the ground level, 4.0m
respectively. The properties of the sandy soil
obtained from SPT results are as follow:
t = 20 kN/m3, = 40o.
Determine the maximum allowable capacity of the footing, Pa in kN per meter run of the
footing, assuming a factor of safety of F = 2.5. (4715kN pmr )

3- A 3m diameter circular footing is going to be constructed on the surface of a loose dry


sand with a friction angle of = 30o and a unit weight of = 16 kN/m3.
a) Calculate the ultimate bearing capacity of the foundation. (53.3 kPa)
b) One method for increasing the capacity is to increase the depth of embedment of the
footing, by constructing the footing below the ground level. What would be the ultimate
bearing capacity of the foundation if it is built 1 m below the ground level. (231.1kPa)

4- A 3m3m footing supports an inclined column P


o
of a whare house. The column makes an angle of 15
o
15 with the vertical and carries an axial load of P
(kN). The footing is founded 1m below the surface
0.5m
of a level site. The foundation consists of a deep 1m
layer of sand having a unit weight of t = 19kN/m3
and a friction angle of =30o. The water table is 3m3m
0.5m below the surface of the soil.
a) Determine the allowable axial load, P, which the column could transmit to the footing
with a factor of safety of 4 against bearing failure. (496 kN)
b) Assuming a friction angle =0.6 between the soil and the footing, is there a possibility
that the footing could slide laterally under the inclined load? Is there any resistance
against sliding failure from the soil adjacent to the footing? (No sliding)
CVEN3201 Applied Geotechnics 127 Tutorial Questions

5- (Exam 2008) An eccentrically loaded strip footing with a width of 2.5 m is embedded in
soil. The base of the footing is 1.0 m below the soil surface. The ground water level is well
below the footing. The soil has
the following properties:
Unit weight t = 18 kN/m3 Proposed excavation 1.0m
Friction angle = 30o Pu
Cohesion c=10kPa
1m
An excavation is going to be
made close to the strip footing as 2.5m
shown in the figure opposite in
dotted lines. 25o
Section
a) Determine the ultimate load
bearing capacity of the foundation, Pu, in kN per metre run of the footing before
excavation. (2024 kN pmr)
b) Determine the ultimate load bearing capacity of the foundation, Pu, in kN per metre run of
the footing after the excavation. (946 kN pmr)
Use the Hansens theory of bearing capacity and the effective width concept in all your
calculations.
CVEN3201 Applied Geotechnics 128 Tutorial Questions

SETTLEMENT OF FOUNDATIONS

1- It is proposed to construct a 1m wide strip footing adjacent to an existing raft footing. The
raft footing has plan dimensions of 10m 10m. A uniform pressure of 54kPa is applied to
the surface of the underlying soil by the raft. The raft has been in existence for a very long
time. The proposed strip footing is to exert a uniform pressure of 150kPa on the soil. The
Plan and Section below illustrate the situation.

10m Existing raft Point X Proposed


strip footing
Point X
0.5m
10m
0.5m 1m
2.5m
Clay
Rock
Existing raft
Proposed
PLAN strip footing SECTION

The soil profile at the site consists of a 2.5m deep layer of clay overlying rock. The water
table is at the surface of the soil. A sample of clay is taken from a depth of 1.25m directly
below Point X. The following soil properties are obtained by testing the sample:
Initial moisture content: m= 31% t = 20 kN/m3
Cc = 0.4 Cr = 0.1 pc = 50 kPa
In the design of the project concern has been expressed that the new strip footing may cause
excessive settlements of the soil beneath the existing raft. Determine the total final
settlement of the clay layer under Point X due to the strip footing. Point X is located mid-
way along the side of the raft nearest to the footing. Use the 1-D consolidation method
employing a one-point settlement computation based on conditions at the sampling point.
(52mm)

2-The soil conditions at a site consist of a 2m thick layer of clay over a 3m thick layer of
sand over rock. A 2m wide strip footing is to be constructed at the site. The footing will
carry a uniform pressure of 250 kPa and the base of the footing will found 1m below the
surface. Use the elastic settlement theory to determine the mean settlement of the footing due
to the settlement of the sand layer alone. The Youngs modulus and Poissons ratio of the
sand may be taken as E=80000 kN/m2 and =0.3, respectively. (3mm!)

3- The soil at a level site consists of a 2.4m deep homogeneous sand layer overlying rock
and there is no water table. A plate loading test is conducted at the site using a 0.3 by 0.3m
plate founded at a depth of 0.3m below the surface. When a pressure of 150kPa is applied to
the plate it settles 0.84mm. A full sized footing, 2m by 2m, is to be built at a depth of 0.4m
below the surface. Determine the mean settlement of the full sized footing when it carries a
pressure of 250kPa. Use elastic settlement theory based on a Youngs modulus back figured
from the results of the plate loading test. Assume a Poissons ratio of 0.3 for the soil. (8mm)
CVEN3201 Applied Geotechnics 129 Tutorial Questions

4- (Exam 2011) A strip footing with a width of 2m is founded 2m below the ground level
and subjected to a vertical central load of 1500 kN per metre run of the footing. The soil is a
well graded dense sand (previously under large compressive pressure) with a unit weight of
t = 20kN/m3. The water table is 4m below the ground level. The following table shows the
results of a cone penetration test, given as average values over given depth ranges below the
ground level.

Depth below the ground (m) 0-2 2-4 4-6 6-10 10-14 14-20
qc (kPa) 2500 3000 3500 4000 5000 6000

Use the Schmertmanns method and evaluate the settlement of the footing 5 years after the
construction. (151mm)
CVEN3201 Applied Geotechnics 130 Tutorial Questions

PILE FOUNDATIONS
1- A 6m long concrete pile is driven through a deep layer of loose sand. The pile has a
diameter of 500mm. The unit weight of the sand may be taken as 18kN/m3.
a) Determine the allowable vertical capacity of the pile using a factor of safety of 3. (560kN)
b) Determine the settlement of the pile if a load of 400kN is applied to the pile. The Youngs
modulus of the sand and the concrete pile may be taken as 40MPa 0.6m0.6mand 20,000MPa,
respectively. (2.7mm)
Sand
2- A concrete pile is driven through 5m layer of loose sand 5m
and penetrates 5m into a medium density gravel. The pile t=15kN/m3
has a square section of 600mm by 600mm. There is no
sign of water in the sand or gravel layer. The unit weight of Gravel
sand and gravel may be taken as 15kN/m3 and 20kN/m3, 5m
respectively. t=20kN/m3
Determine the allowable vertical capacity of the pile using
a factor of safety of 3. (2720kN)

3- What load would cause failure immediately after


2m
construction of the cast-in-situ reinforced concrete pile
Brown clay 4m
shown in the figure opposite? The pile has a diameter of
600mm and a length of 8m and the following soil
properties are given: Dense sand 2m
Brown clay: t = 16kN/m3, cu = 30 kPa, u = 0
Dense sand: t = 21kN/m3, = 40o Grey clay 2m
Grey clay: t = 19kN/m3, cu = 120 kPa, u = 0
(880kN)

4- A square group of 9 piles is arranged with 3 rows and 3


columns spaced 3m c/c apart. The cast-in-situ piles have a
diameter of 600mm embedded 6m in a deep layer of stiff 3m
to very stiff clay having an undrained cohesion of
100kPa, a unit weight of t=20kN/m3, and a Youngs 3m
modulus of 20MPa. The pile group has a rigid cap. The
Youngs modulus of the pile material can be taken as
2000MPa. 3m 3m
a) Determine the ultimate capacity of the pile group
assuming that there is no interaction between the piles,
i.e, the capacity of the group is the sum of capacity of
all piles. (7578kN)
b) Calculate the load required to cause block failure of the 6m
entire group. (~43718kN)
c) Calculate the allowable bearing capacity of the group
piles based on the results of parts (a) and (b), using a
Clay
factor of safety of 2. (3789kN)
d) Determine the pile cap settlement assuming that a total t=20kN/m3
load of 3600kN is applied to the pile cap. (15.7mm)
CVEN3201 Applied Geotechnics 131 Tutorial Questions

5- (Exam 2010) A 0.5m diameter circular concrete pile is driven to a depth of 10 m in a deep
layer of homogeneous dry dense sand. The sand has a friction angle of = 38o, a unit weight
of t = 18 kN/m3, and a Youngs modulus of Es = 100 MPa.
The Youngs modulus of concrete pile can be assumed as P
Ep = 25GPa. The pile is subjected to horizontal and axial
forces as shown in the figure opposite. H
a) Determine the allowable axial bearing capacity of the pile
assuming a factor of safety of 3. (2827kN)
b) Evaluate the settlement of the pile under an axial load of
P = 2000 kN. (5.4mm) 10m
c) Determine the ultimate horizontal capacity of the pile, Hu.
Assume that the concrete pile has a yield moment capacity
of My = 2500 kN.m. (945kN)
d) Estimate the horizontal displacement the pile head if a
horizontal load of H = 500 kN is applied to the pile head. =0.5m
Assume that the pile section has a second moment of
inertia of Ip = 0.004m4. (5.1mm, 0.185o)
CVEN3201 Applied Geotechnics 132 Tutorial Questions

EARTH PRESSURE
0.6m
1- Consider the wall in the figure opposite which has
been built in front of a stable rock. Gravel backfill has
been placed in the space between the wall and the rock
with a drainage pipe. The gravel has the following 0.5m
properties: 3m Stable
t = 19kN/m3, c=0, =35o. rock
a) The worst case design scenario is when the drainage
pipe is blocked and the water table rises to the top Pipe
surface of the backfill. Determine the resultant lateral
force on the wall for this case using Rankines theory
of active earth pressure. (55.3kN)
b) Determine the resultant force when the drainage pipe is working properly. (23.2kN)
c) Determine the resultant force as in part (a) using Coulombs earth pressure theory.
Consider only one possible mode of failure as shown in the figure above in dotted line.
The friction angle between the wall and the gravel may be taken as =25o. (50kN)

10kPa
2- A section through a long retaining wall embedded
Clay
into rock is shown in the figure opposite. The wall 2m
retains 4m of clay and 3m of sand and supports a WT
surcharge of 10kPa as shown. The following properties
2m
of the soils are given:
Clay: t = 18kN/m3, c=5kPa, =20o. Sand
Sand: t = 20kN/m3, c=0, =30o.
3m
Use Rankines theory of lateral earth pressure to
determine the bending moment (per meter run) in the
wall section at point A, a long time after construction of Point A Rock
the wall. (544.5kN.m pmr)

3m
3- The figure opposite shows a section of a
4 m high gravity wall. The wall is inclined
toward the backfill at an angle equal to 80o 1m
measured with respect to the horizontal. The
backfill soil is cohesionless and has a unit
weight of t=18 kN/m3 and a friction angle of 80o
=30o. The wall friction angle is =20o. There
4m
is no sign of water in the soil. Failure line
a) Calculate the lateral force on the wall due to
active earth pressure using the Coulombs
method and the trial failure line shown in 52o
the figure. (41.1 kN)
b) Sketch the direction of the lateral force
applied to the wall, clearly mark the inclination of the force with respect to the horizontal.
CVEN3201 Applied Geotechnics 133 Tutorial Questions

RETAINING WALLS

1- Gabions consist of wire baskets filled with rocks. They are 1m


often used as retaining walls. Consider the gabion wall
shown in the figure opposite. The design of this wall is to be
checked according to AS4678-2002. The wall retains a Uncontrolled
natural soil near a minor road. There are no buildings in the 2m backfill sand
vicinity. A live load surcharge of 5kPa should be applied to
No groundwater
the soil behind the wall. The following soil properties have
been determined from testing:
Soil: = 18 kN/m3 = 35o c=0
Gabion: = 22 kN/m3 = 25o
You are to apply code factors as appropriate and use Rankines theory for calculation of Ka.
Evaluate the stability of the wall against sliding and overturning only.
(Sliding: S*=16.6 kN.m, nR*=19.4 kN.m Overturning: S*=22.2 kN, nR*=13.5 kN)

2- A concrete retaining wall supports a clay soil as shown in 10kPa


the figure opposite. Behind the wall the clay is sealed with
Asphalt seal
asphalt pavement so that no water can enter the backfill. A
surcharge of 10kPa is applied to the surface of the backfill
0.4m
due to traffic loading. The following properties of clay are
Clay
known: 4m
t = 18kN/m3, c=5 kPa, =30o.
a) Use Rankines theory of earth pressure to determine the
factor of safety against overturning for the wall. You
should ignore any passive soil resistance in front of the 0.4m
wall. Use concrete = 25kN/m .3
(1.82) 1.4m
b) Assuming class I for clay backfill, use AS4678-2002 recommendations and check the
stability of the wall against overturning. (S*=83.9 kN.m, nR*=66.3 kN.m)

3- A small precast concrete retaining wall is shown in 50kPa


the figure opposite (concrete = 25kN/m3). The area
behind the top of the wall is used as a storage area
which is subjected to a surcharge of 50kPa. The soil
Gravel
behind the wall is gravel with the following
2.5m
properties:
t= 18kN/m3 = 30o 0.3m
The friction angle between the wall footing and the
underlying soil can be assumed as = 25o. Use
Rankines theory of active earth pressure to 0.3m
determine: 0.6m
a) The overall factor of safety of the wall against
sliding, taking the full passive force into account. 0.6m 1.4m
(1.24)
b) The factor of safety of the wall against overturning. (1.97)
c) The safety factor for the bearing capacity of the wall foundation. (0.67)
CVEN3201 Applied Geotechnics 134 Tutorial Questions

4- An anchored wall is embedded 1.5 m into a 5kPa


saturated stiff clay and retaining a sandy soil. A
1m 1
5 kPa surcharge is applicable at the top of the Td
wall. The following properties of the soils are
known:
Sand: t = 15 kN/m3, c= 0, = 30o, 4m Sand
Clay: t = 20 kN/m3, cu = 25kPa, u = 0o
The water table is 0.65m below the surface of the
clay layer. The short term stability of the wall is
going to be considered.
2
a) Use Rankins theory of lateral earth pressure 5 3
and determine the active and passive total 0.65m Water table

horizontal pressures at different points on the 1.5m Clay


wall (points 1 to 6).
(1.67, 21.67, 15, 45, 50, 80 kPa) 6 4
b) Use a global factor of safety of F = 1.7 and determine the design tension force for the
anchor, Td, per meter run of the wall. (58.33 kN pmr)

5- Use the requirements of AS4678 and calculate the design tension force for the anchor of
example 1. Use a structural classification factor of 1. (128.7kN pmr)

Water tank

60kPa

6- A 2m deep excavation has to be performed in


front of a water tank. The water tank is applied a
pressure of 60kPa on a sandy soil with a unit
weight of 20kN/m3 and a friction angle of 34o.
There is no sign of water in the soil. 2m
Use the requirements of AS4678 and calculate the
minimum total length of the sheet pile retaining
wall required to be installed before the excavation
of the soil. Assume a structural importance factor Sand
of n=0.9 (8.36m)

Das könnte Ihnen auch gefallen