You are on page 1of 83

Standards Australia

Information in fields should be filled in if known or left as is.

Bridge design—Commentary
(Supplement to AS 5100.3—2004)
Part Title: Foundations and soil supporting structures
Designation: AS 5100.C3—200X
Part Number: C3
Supersedes Standard No:HB 77.3 Supplement 1—1996
AustralianORJoint: Australian
Creation Date: 2005-01-04
Revision Date: 2007-03-22
Issue Date: XX 2007
Committee Number: BD-090
Committee Title: Bridge Design
Subcommittee Number: BD-090-03
Subcommittee Title: Foundations and Soil Supporting Structures
Project Manager: Eddy Go
PMs Email Address: eddy.go@standards.org.au
WP Operator: Forster/Cotter/Go/Phillips/Forster/Go/Phillips/Forster
Project Number: 198
Combined Procedure?: No
Committee Doc No.: ONE
Stage: 4.1.3
Committee Reps: Australasian Railway Association
Austroads
Bureau of Steel Manufacturers of Australia
Cement Concrete & Aggregates Australia
Engineers Australia
Queensland University of Technology
Steel Reinforcement Institute of Australia
The Association of Consulting Engineers Australia
University of Western Sydney
Additional Interests:
Product Type AS
Document Status Current
Document Availability Private

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 2 DRAFT ONLY

PREFACE
This Commentary was prepared by Standards Australia Committee BD-090, Bridge Design,
to supersede HB 77.3 Supp 1—1996, Australian Bridge Design Code—Foundations—
Commentary (Supplement to SAA HB 77.3—1996).
The objective of this Commentary is to provide users with background information and
guidance to AS 5100.3—2004. This Commentary differs from its predecessor in that many
equations and analytical methods are not presented. The Commentary instead focuses on
design principles and practical considerations and provides references that can be used for
guidance. The Standard and Commentary are intended for use by bridge design
professionals with demonstrated engineering competence in their field and geotechnical
engineers involved with the investigation, analysis and design of bridge foundations and
related soil-supporting structures.
In this Commentary, AS 5100.3 is referred as ‘the Standard’ and the Commentary itself is
referred to as ‘the Commentary’.
The clause numbers and titles used in the Commentary match those in AS 5100.3. To avoid
possible confusion between the Commentary and the Standard, a commentary clause is
referred to as ‘Clause C……’ in accordance with Standards Australia policy.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 3 DRAFT ONLY

CONTENTS

Page

C1 SCOPE............................................................................................................................5
C2 APPLICATION...............................................................................................................5
C3 REFERENCED DOCUMENTS.....................................................................................7
C4 DEFINITIONS...............................................................................................................7
C5 NOTATION.....................................................................................................................7
C6 SITE INVESTIGATION.................................................................................................7
C7 DESIGN REQUIREMENTS........................................................................................11
C8 LOAD AND LOAD COMBINATIONS.......................................................................20
C9 DURABILITY..............................................................................................................26
C10 SHALLOW FOOTINGS..............................................................................................28
C11 PILED FOUNDATIONS..............................................................................................32
C12 ANCHORAGES...........................................................................................................41
C13 RETAINING WALLS AND ABUTMENTS................................................................46
C14 BURIED STRUCTURES.............................................................................................52

APPENDICES
CA SITE INVESTIGATION TECHNIQUES.....................................................................56
CB WORKED EXAMPLES................................................................................................61

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 5 DRAFT ONLY

STANDARDS AUSTRALIA

Australian Standard
Bridge design—Commentary
(Supplement to AS 5100.3—2004)

Part C3: Foundations and soil supporting structures

C1 SCOPE
The Standard provides procedures for the design of foundations and soil-supporting structures
commonly encountered by engineers involved in the detailed design of road and railway bridges
and associated structures. The Standard uses limit state methods. Strength, stability,
serviceability and durability limit state requirements should be satisfied.
The design and construction of reinforced soil structures is not covered by the Standard.
Guidance may be found in BS 8006 (Ref 1) and Elias, Christopher and Berg (2001) (Ref 2) or
State Road and Railway Authority specifications.

C2 APPLICATION
The Standard provides the analytical procedures to be adopted and numerical values for the
required relevant geotechnical strength reduction factors. The loads to be applied to structures
are those specified in AS 5100.2, including earth pressure loadings. However, design engineers
should recognise the following:
(a) AS 5100.2 specifies that soil imposed loads on retaining walls and the like are to be
obtained from AS 4678. The design of foundations and soil-supporting structures for
bridges and other road and rail related structures is to be carried out in accordance with
AS 5100.2 and AS 5100.3, not AS 4678.
The density of soils is factored by the factor ( γge) given in AS 5100.2 where factoring of
loads is specified.
(b) AS 4678 uses factored loads and factored characteristic values of soil and rock parameters
when calculating ground related actions (e.g. active or at-rest ‘disturbing’ forces) and
resistances (e.g. bearing capacity, passive force, sliding resistance and the like). This is
different from the requirements of the Standard.
(c) The Standard requires for the strength and stability design of foundations a similar
approach to that of AS 2159 where the loads and action effects are factored prior to
carrying out the analysis using unfactored characteristic values of soil and rock material
parameters (see Clause 7.3.2(c)) with the calculated resistances subsequently factored and
the design inequality compared to verify that the factored resistances are greater than the
calculated design action effects.
(d) The Standard requires for the strength and stability design of soil-supporting structures
that the geotechnical analysis be carried out using unfactored loads and unfactored
characteristic values of soil and rock material parameters (see Clause 7.3.3(d)) with the
design action effects and resistances subsequently factored and the design inequality
compared to verify that the factored resistances are greater than the calculated design
action effects.
(e) The designer must also demonstrate that all other limit state requirements of the Standard
are satisfied.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 6 DRAFT ONLY

Item (a) should be taken to only mean that the characteristic values of soil and rock material
parameters given in AS 4678 may be used in the absence of more reliable site specific data for
calculating the loads and resistances required by the Standard.
Items (b), (c) and (d) are philosophically different and great care needs to be taken to ensure
consistency in any one analysis or design. Items (b), (c) and (d) should not be mixed or
confused.
The approach taken for the design of soil-supporting structures described in Item (d) is intended
to ensure that the designer is working with realistic values of soil and rock parameters to give
realistic analytical results which can then be compared with limiting strength parameters for the
soils and rocks at the site. Unlike most structural analysis which usually assumes linearly
elastic behaviour, geotechnical analysis must accommodate non-linear soil and rock behaviour
under imposed loads and deformations, with behaviour largely dominated by the soil friction
angle ’.
Any designer who is not clear about the different approaches of Items (b), (c) and (d) should
seek guidance from a more experienced engineer before undertaking design in accordance with
the Standard.
Because the strength and stability limit states are checked using unfactored loads and material
parameters, the serviceability limit state can also be readily checked using the same analytical
model.
Computer analysis has become the norm for geotechnical engineers in recent years, displacing
the earlier reliance on design charts, tables and technical papers. Many of the available
programs are on the face of it easy to use and yet most are complex, involve making specific
assumptions about ground behaviour, and often require input parameters that cannot readily be
determined from normal field and laboratory tests. Programs need to be used with care by
engineers who understand the assumptions made, the parameters being used and the basic
principles relied on by the program. This will normally mean having undergone specific
training, or having had a long familiarity with geotechnical methods of analysis. Comparison of
output values with limiting values should be carried out to assess whether the requirements of
the Standard are met.
Outputs from computer programs are not necessarily in a form that readily indicates whether the
requirements of the Standard have been met. As an example, output from a typical retaining
wall design or finite element program may give—
(i) an indication that equilibrium has been achieved;
(ii) the calculated deflections over the full height of the wall (and surrounding ground in the
case of a finite element program);
(iii) the stresses and deformations in the soil; and
(iv) the calculated stresses, shears and moments in the wall,
without directly providing information that allows the designer to determine either Rug or S*. In
this case, comparisons of the output using hand calculations or spreadsheets with limiting values
of the relevant design action effects may be required to demonstrate that the requirements of the
Standard are satisfied.
Rug and S* being design actions effects are not necessarily forces and could be stresses,
deformations or any other output parameter. If the program does not provide relevant limiting
values for these parameters, then they can usually be readily calculated using basic soil
parameters such as ’, c’, σ z and the like using equations and formulas found in relevant and
reliable texts.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 7 DRAFT ONLY

C3 REFERENCED DOCUMENTS
Standards referred to in the Standard and the Commentary are subject to revision from time to
time and the current edition should always be used. The currency of any Standard may be
checked using the Standards Australia website at www.saiglobal.com/shop.
The predecessors of the Standard and their commentaries generally provided full details of
published methods of analysis, but most of those methods have been superseded following
ongoing research and development. In keeping with a focus on computer methods, most of the
previously published methods have been replaced by references to general texts which contain
many of the original references, but also provide a broader overview of the earlier work, putting
it into context and recommending approaches to geotechnical analysis that have been found by
experience to be the most useful or relevant.

C4 DEFINITIONS
Technical definitions are generally provided in the Standard. Technical definitions that are
applicable only to one clause are usually given in that clause.
In the Commentary, the terms ‘sub-structure’ and ‘sub-structure element’ are used to mean any
footing, piled foundation, abutment or retaining wall which forms part of a bridge or associated
road or railway structure.
The term ‘hand calculation’ is used to mean traditional methods of analysis using design charts,
tables and simple numerical analysis (including spreadsheets), as opposed to commercially
available geotechnical computer programs.

C5 NOTATION
The basis of the notation is generally in accordance with ISO 3898, Bases for design of
structures—Notations—General symbols. Standards Australia’s policy is to use ISO
recommendations for notation wherever practicable in structural design standards. A general
summary can be found in AS/NZS 1170.0.

C6 SITE INVESTIGATION
C6.1 General
As specified in the Standard, site investigation should be carried out as a routine procedure to
provide the information necessary for the design and construction of foundations and soil-
supporting structures. It is desirable to have completed site investigation before design begins,
but investigations can sometimes be in stages as discussed further in Clause C6.2.
The objectives of a site investigation are to—
(a) develop a geological/geotechnical model of the site. This should clearly indicate
stratification and structure of the soils and rocks on the site and identify pre-existing
features (e.g. an old landslide or karst formations) that could adversely impact on the
proposed construction;
(b) determine geotechnical parameters needed for the analysis and design of the sub-structure
elements and their likely variability;
(c) locate the groundwater table, if it is likely to affect foundation construction or
performance; and
(d) assess likely water inflows to excavations where they could impact on construction.
The site investigation needs to be of sufficient extent to be used to evaluate the effect of the
proposed construction on adjacent structures and properties.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 8 DRAFT ONLY

Ground is inherently variable and, as a result, never totally predictable. Unknowns should be
reduced as far as is reasonably possible by investigation. Close liaison between the bridge
designer and the geotechnical engineer is desirable from the start of all but the simplest of
projects. Working together, they should plan the site investigation and testing programme,
taking into account the specific requirements of the proposed structure (e.g. likely foundation
loads, sensitivity of the structure to differential settlement, preferred construction method and
the like) and of the site (e.g. access difficulties, anticipated geology, environmental
considerations and the like). Having understood and agreed what design information is required,
the geotechnical engineer should undertake the investigation and testing, interpret the data and
prepare a geotechnical report.
For foundations bearing on refuse, uncompacted fill or soft, loose or highly compressible soils,
rigorous investigation, testing and analysis will be required unless foundations are to be
founded beneath the problem ground. If problems involving hydrology, vegetation, surface
water, mine subsidence, acid sulfate soils or environmental factors are encountered or
suspected, the investigation should specifically look at these aspects.
Be conscious of what might not be known about a site. For sites with complex or initially
unknown geotechnical conditions, the approach to the site investigation should be flexible, with
a willingness to vary the approach depending on the findings. Where the design investigation
indicates there is still uncertainty, additional investigation and proving of ground conditions
may be required during construction.
C6.2 Design investigations
AS 1726 specifies details of available investigation and testing techniques. In particular,
AS 1726 provides—
(a) a list of available field test methods;
(b) a list of laboratory examination and testing methods; and
(c) an outline brief for a geotechnical investigation which can be used to guide the designer
or the geotechnical engineer, or both, when planning a site investigation.
This Clause and Appendix CA are provided to assist design engineers to evaluate site
investigation requirements and methods. Within reason, site investigation costs should be
considered to be of secondary importance to getting the right information for design. In
deciding what is required of a site investigation, consideration should be given to the
consequences of getting the answers wrong.
As specified in the Standard, a preliminary investigation is often a help on a major project, since
it can provide an initial indication of preferred foundation type and cost during concept design
stage. On major projects, it is sometimes beneficial to undertake site investigation as a
progressive process, with data acquisition keeping pace with the level of detail of the design.
All site investigation data is useful and none should be ignored unless it can be shown to be
erroneous.
The only site investigation requirements specified in the Standard are the minimum number of
boreholes for bridge foundations (Clause 6.2 (a)) and for culverts, retaining walls and the like
(Clause 6.2(b)). It should be noted in relation to what follows that the word borehole should be
taken to include any site investigation method chosen for the particular project. As a further
guide—
Variability
 The number of boreholes should normally be increased when ground conditions are
uncertain or, from a knowledge of the local geology, are expected to be variable.
Borehole depth

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 9 DRAFT ONLY

 Borehole depths should be sufficient to penetrate all strata likely to influence foundation
behaviour. This requires the investigation to extend below the anticipated founding level.
Low strength or highly compressible strata below the founding level should be identified
and their properties determined.
 For spread footings and pile groups, borehole depths should normally extend below the
founding/pile toe level to at least 1.5 times the width of the sub-structure element, for nearly
square sub-structures, and to at least 3 times the width when the length/breadth ratio of the
sub-structure exceeds 3.
 Where individual piles, as opposed to pile groups, are to be used the site investigation
should extend to at least 10 times the pile diameter below the founding level, or at least 5 m,
whichever is the greater. Soft strata encountered below proposed pile toe levels should be
fully investigated to obtain data needed to assess the potential for punching failure or
excessive settlement.
 For embankments, investigation depth should normally extend below the founding level to
at least 1.5 times the embankment width.
 In cuttings, boreholes should extend to a depth of at least one metre beneath the base of the
proposed excavation. Where excavation will leave substantial cut batters, boreholes should
extend sufficiently far into the underlying soils to check slope stability.
Number of Boreholes
 In cuttings the site investigation should involve not less than four boreholes giving general
coverage and a minimum spacing corresponding to one borehole for every 10,000 m 3 to be
excavated. Additional boreholes should be drilled in areas where changes of excavation
method could be required, such as where a soil/rock interface occurs, or the water table will
be intersected. This information is not only needed for design purposes, but is also
important for estimating quantities and the cost of different excavation and disposal
methods.
 Where approach embankments are less than about 2 m in height, a minimum of one borehole
per 100 m of embankment may be adopted. Where poor subsoil conditions (e.g. peat, soft
clay, high water table, acid sulfate soils) are anticipated this number should be increased in
order to define the extent of the problem. Test pits may be substituted for boreholes where
stiff to hard soils, or weathered rock, exists close to the surface.
 Where approach embankments are more than 2 m in height a minimum initial investigation
of 2 boreholes per 100 m of embankment is recommended. Boreholes should be distributed
across the embankment width in order to define differences in the depth of compressible
strata. In situ and laboratory testing should be undertaken so embankment stability and
settlement behaviour can be assessed. Additional boreholes and test pits may also be
required.
Other considerations
 More than the minimum number of boreholes may be required.
 Where ground conditions could vary transversely, boreholes should be located either side of
the centre-line.
 Boreholes should extend sufficiently far into rock to identify its condition, including the
variability of weathering and the nature and spacing of joints. Where relevant, sufficient
information should be obtained to establish the possibility of failure along discontinuities
(e.g. joints, bedding planes, low strength beds) within what is generally a reasonably strong
rock mass.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 10 DRAFT ONLY

 Care should be taken to ensure that ‘rock’ encountered in boreholes is not a boulder
embedded in a soil matrix (‘floater’). This requires consideration of the local geology and
the origin of the ground being investigated. It may be necessary to extend a small proportion
of the boreholes to a minimum depth of 6 m into rock to demonstrate that it continues
downwards.
 Where static cone penetrometer testing is used, it should usually be accompanied by
sampled boreholes, or other investigations, so that the penetrometer test results can be
calibrated. This calibration should only be omitted if the geology of the site is known and
results of the testing can be related to the type and consistency of soils on the basis of well
documented experience.
Groundwater
The presence of groundwater within the depth affected by the proposed engineering work
should be investigated to determine—
 the level of the permanent water table and whether it fluctuates (e.g., tidal response,
seasonal variation);
 likely inflow rates to excavations;
 probable effect of dewatering on the local water table and adjacent structures;
 the presence of and pressures associated with artesian or sub-artesian conditions; and
 the potential aggressiveness of the groundwater to buried concrete, steel and the like.
Investigations should ensure that the most adverse ground water situation is established. Where
fluctuations occur, water levels should be monitored periodically.
During site investigation, boreholes should be left open and ground water levels recorded—
 at the start and finish of each working day during drilling;
 on completion of each borehole; and
 thereafter, on a daily basis until the field work is completed and the boreholes are
backfilled.
It will often be necessary to install standpipes or piezometers in boreholes so variations in the
ground water table can be monitored. Consideration should be given to data loggers or simpler
devices able to indicate the peak water level in a given period of time. It will be appreciated that
water levels are often influenced by rainfall, but a high groundwater level does not necessarily
coincide with a period of high rainfall. In some situations, a significant time lag can occur.
For sites adjacent to tidal waters, water levels in completed boreholes should be recorded
several times daily and compared with tidal heights in order to establish any tidal influence and
time lag.
Where construction will involve excavation below the groundwater table in permeable soils or
fractured rock the investigation should include in place permeability testing of the substrata in
order to estimate inflow rates to the excavation. Dewatering of excavations in permeable sands
and gravels can affect the groundwater table at considerable distances from the excavation and
can cause settlement of nearby structures. This aspect should be carefully investigated on sites
adjacent to existing structures.
In boreholes where artesian conditions (i.e. water rises rapidly up the borehole to above ground
level, causing the borehole to flow continuously) are encountered the artesian pressure should
be recorded. In cased holes this is normally done by adding casing above ground level until the
flow stops and the static head can be recorded. Alternatively, a sealed type of piezometer can be
installed for subsequent reading once the borehole is backfilled. On completion of a borehole in

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 11 DRAFT ONLY

artesian conditions care should be taken to block off and backfill the hole so that the artesian
flow is cut off.
Where sub-artesian conditions (i.e. water rises rapidly up the borehole to stabilise above the
level of the stratum in which it was met, but without flowing out at ground level) the water level
should be recorded and the condition noted on the borehole logs.
Groundwater samples should be chemically analysed to establish whether it is likely to cause
damage to concrete or steel (or other buried materials). Any special precautions to protect sub-
structure elements from this type of damage should be specified accordingly.
Sample storage
For all projects it is advisable to store soil samples and rock cores at least until construction of
the substructure is completed. Where work is to be carried out by contract, all samples should
be made available for inspection by prospective tenderers in order to enable each tenderer to
make its own assessment of ground conditions.
In the event of engineering problems or contractual disputes occurring during construction,
these samples can be valuable in assessing relative conditions at various locations and in
establishing whether the site investigation data represents the conditions as found. For this
reason all unused portions of undisturbed soil samples and representative portions of rock core
should be sealed and stored at in-place moisture conditions.
Reporting
The results of the site investigation should be compiled into a geotechnical report in accordance
with AS 1726. The report should normally—
 state clearly the purpose and objectives of the investigation;
 provide a brief description of the project for which the geotechnical report is being
compiled, giving information about its location, size, layout, anticipated loads, method of
construction and the like;
 provide a description of the investigation, including when it was undertaken, field and
laboratory work performed, and relevant observations such as the weather conditions at the
time;
 present the geotechnical information obtained from the investigation, together with available
data (or reference to data) from previous investigations near the site;
 evaluate the geotechnical information, focussing on the specific purpose and objectives of
the investigation; and
 provide conclusions and recommendations.

C7 DESIGN REQUIREMENTS
C7.1 Aim
(No Commentary)
C7.2 Design
The design process for strength and stability is illustrated in Figure C7.2.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 12 DRAFT ONLY

FIGURE C7.2 DESIGN PROCESS FOR STRENGTH AND STABILITY

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 13 DRAFT ONLY

C7.3 Design for strength


AS 5100.1 specifies that the strength limit state is an elastic, inelastic or buckling state in which
the collapse condition is reached at one or more sections of the structure. Plastic or buckling
redistribution of action and resistance shall only be considered if data on the associated
deformation characteristics of the structure from theory and tests is available.
This definition is different from that normally used by geotechnical engineers and relates to
collapse of the structure, rather than collapse of the sub-structure element itself. By implication,
the geotechnical strength limit state includes any of the following, provided movements are
large enough to lead to a collapse condition in one or more elements of the structure (refer to
Figure C7.3):
(a) Elastic movements including differential movements and rotations of a sub-structure
element, even though the sub-structure element is not failing in geotechnical terms.
(b) Consolidation and creep movements, even though the substructure element is not failing
in geotechnical terms.
(c) Inelastic (or plastic) movements, which might also indicate failure of the sub-structure
element in geotechnical terms.
(d) A more widespread slope stability failure, including the effects of stream bed erosion,
which removes support from a sub-structure element.
(e) More generalised sources of differential movement including such things as settlement,
heave, or lateral movement due to adjacent embankment or cutting construction or to the
effects of dewatering.
The concept of a buckling state has no real equivalent in geotechnical engineering, but the
warning that redistribution of action and resistance should only be considered if data on the
associated deformation characteristics of the structure from theory and tests is available needs
to be heeded. For significant or sensitive structures this implies a close working relationship is
needed between the geotechnical and structural engineer if critical geotechnical strength limit
state situations are not to be overlooked.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 14 DRAFT ONLY

Figure C7.3 GEOTECHNICAL STRENGTH LIMIT STATE DEFINED BY COLLAPSE


OF STRUCTURE

C7.3.1 General

In geotechnical terms, the strength limit state involves loss of static equilibrium or the formation of
a mechanism in the ground, leading to ongoing movement under a constant applied load. This is
referred as “geotechnical failure” (refer to Figure C7.3). In some ground conditions this condition
will be reached at relatively low deflections. In others, the effect of dilation (volume increase) of the
soil may result in very large deflections before the strength limit state is reached. The deformation
of an associated structure may lead to structural collapse before geotechnical failure is reached. In
this case the “geotechnical strength limit state” (e.g. bearing capacity, anchor resistance) is governed
by the structure (refer to Figure C7.3). The geotechnical strength limit state therefore must take into
account the sensitivity to movement of the associated structure.
Hand calculation relied on postulating a failure mechanism and calculating the equivalent
theoretical failure load. Overall factors of safety were then applied to the failure load in order to
determine the permissible load capacity at which movements were limited to acceptable levels.
In the case of foundations, the factor of safety was usually of the order of 2 to 3. Sloan (2005)
(Ref 3) discusses this approach and also presents a more rigorous way of calculating
geotechnical strength for a variety of structural foundation elements (finite element limit
approach). It is possible that for a sensitive structure, the permissible load capacity calculated
this way is in excess of the geotechnical strength limit state, so care should be exercised at all
times.
Some computer programs can calculate the geotechnical strength of a specific sub-structure
element, such as the geotechnical failure load of a spread footing or the pull-out resistance of an
anchor. It is important that the user understands how the particular program calculates its
answer and what limitations apply before assuming that the result is Rug.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 15 DRAFT ONLY

Computer programs that rely on simple elastic theory and do not account for non-linear or
plastic soil behaviour are generally unable to model the geotechnical strength limit state unless
the predicted movements are large enough to cause collapse of part or all of the structure. These
programs can be used to calculate deflections, stresses and moments, but are unlikely to indicate
whether Equations 7.3.2(1) or 7.3.3(1) are satisfied.
Finite difference and finite element methods of analysis, which incorporate non-linear or
elastic-plastic stress-strain behaviour and are intended to model large strains, can be used to
calculate load-deflection performance and the load at which geotechnical failure occurs. These
programs should be used with care. Experience is important to evaluate the input parameters,
many of which cannot readily be determined from normal commercially undertaken site
investigation data. The analytical process used needs to be understood to interpret the results.
This sort of analysis can be used to assess a project specific geotechnical strength limit state.
This is done by comparing the calculated movements with those that are judged large enough to
lead to collapse of part or all of the structure. The same analysis may well show that
geotechnical failure of the sub-structure will not occur.
Calculations using finite element methods can be complicated, particularly where deformations
are significant. The accuracy of the results needs to be compatible with the design
requirements. Generally coarser meshes give less accurate results but they consume less
calculation time and vice versa for finer meshes. It is practical to adopt a coarser mesh for the
initial design to condition the numerical model before attempting finer meshes in the final
design. Depending on user-specified conditions, convergence of a calculation does not
automatically mean an acceptable solution has been reached. As normal design practice, users
must examine the calculated results and be satisfied that all check sums and indicators are
reasonable and internally consistent, within an accuracy compatible with the design. Sensitivity
checks of parameters critical to the analysis should be carried out to verify that the solution
reflects reality and has the required safety.
Although fully 3-dimensional situations can be analysed using finite element and finite
difference methods, the process is normally complex and time consuming. As a result, it is more
common for 2-dimensional representations to be used. It is important to understand the
differences between the 2-dimensional model and the 3-dimensional reality and to select
parameters that take this into account. As an example, the stresses applied to the ground by a
spread footing reduce with depth. If the footing has a low length to breadth ratio, the stresses in
the 3-dimensional ground reduce more rapidly with depth than a 2-dimensional analysis will
predict. In some geological situations, this could lead to an inappropriate assessment of the
geotechnical strength.
C7.3.2 Foundations
The strength limit state of a foundation occurs in relation to the applied vertical loads, lateral
loads and moments as well as combinations of all three. Non-uniform ground conditions and a
sloping ground surface will affect the result and should be considered.
The Standard requires the determination of the ultimate geotechnical strength Rug. Input
parameters for the determination of Rug are—
(a) for foundations in soil:
(i) A geotechnical model of the ground which indicates the foundation geometry and
the depth, dip and strike of distinct strata;
(ii) Undrained shear strength (cu), or drained shear strength (c' and '), of each stratum
(which can include a strength vs. depth profile);
(iii) Bulk density (γ) of the soil in each stratum; and
(iv) Ground water level.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 16 DRAFT ONLY

(b) for foundations in rock:


As for soil, but the unconfined compression strength ( qu) of rock is often used rather than
cu, or c' and ’.
These parameters should be obtained from an interpretation of field and laboratory tests. Values
should be checked against reliable local experience and published values, such as those found in
text books and in AS 4678.
Where the ground is significantly non-uniform, care should be taken to model adverse
situations, such as under one side of a large foundation where the ground has lower strength,
whilst also considering the possibility of other significant mechanisms occurring such as
differential settlement or sliding.
As indicated in Clause C7.3.1, computer programs that are able to model large strain behaviour can
be used to indicate a project specific geotechnical strength limit state. In the absence of this kind of
analysis, or criteria relating to the settlement sensitivity of the specific structure, the geotechnical
strength limit state may be assumed to be the load at which:
(A) in soil, the settlement is 10% of the foundation width; and
(B) in rock, the settlement is 1% of the foundation width.
C7.3.3 Soil-supporting structures
Soil-supporting structures are retaining walls or basement walls but can be abutments which
have to carry structural loads as well as the loads due to the supported soil. Retaining walls can
be gravity, cantilevered, propped or anchored walls.
Although Clause 7.3 is concerned with design for strength, it will be noted that geotechnical
strength is represented on both sides of Equation 7.3.3(1) since for a soil supporting structure S*
is a function of soil strength as is Rug. All load factors should be set to 1.0 at the outset in
accordance with the design process and conservatively estimated but unfactored characteristic
values of the soil parameters used for the analysis..
The geotechnical strength reduction factors applied to the soil resistances ( Rug) reflects that the
loads and design action effects are unfactored (see also Clause C2).
Care should be taken not to over-estimate soil strength, since this will reduce the unfactored
active pressure proportionately more than it increases the factored passive pressure, suggesting
a greater level of safety than is justified. Density, on the other hand, should not be under-
estimated, since it will reduce the unfactored active pressure proportionately more than it
reduces the factored passive pressure.
The Standard requires that S* = 1.0Se where Se is the soil imposed action effect, or the
unfactored load imposed by the soil on the wall.
The design action effects in structural elements are factored by a factor of 1.5 after the analysis
to enable the required structural resistances to be assessed in accordance with the limit state
design principles of AS 5100.5 or AS 5100.6 or the relevant limit state design standard for the
particular material involved.
Gravity walls
Geotechnical failure of a gravity wall is by sliding or overturning.
Sliding
The geotechnical resistance that controls sliding is normally a combination of the passive
resistance in front of the toe and shear resistance generated along the wall/ground interface (or
through the ground beneath the wall if a shear key is incorporated to force the failure surface
away from the foundation).

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 17 DRAFT ONLY

Shear on the base can be calculated using cu multiplied by an appropriate adhesion factor or
frictional resistance of the foundation/ground interface, or the soil/rock strength where a shear
key is used (see Table C10.3.3.4).
Overturning
The geotechnical strength (Rug) that controls overturning failure is the bearing capacity beneath
the toe of the wall (the point about which the wall will rotate if the disturbing moment exceeds
the restoring moment). Overturning failure can occur when the disturbing moment exceeds the
restoring moment, but no bearing capacity failure occurs. This is one of the few situations
where a geotechnical element needs to be designed for stability (Clause 7.4).
Computer programs
Some computer programs can be used to design gravity walls, but may not specifically identify
the ultimate passive and sliding resistance or the way in which a factor of safety is used in the
analysis. Check how the program calculates equilibrium and the assumptions inherent in the
analysis. It may be necessary to undertake additional hand calculations to determine whether
Equation 7.3.3(1) is satisfied.
Cantilevered, propped or anchored walls
Geotechnical failure of a cantilevered or anchored wall is by passive failure at the toe or pull-
out of the anchor.
Passive failure
In soil, passive resistance is determined using the passive earth pressure coefficient ( K p), unless
a finite element or finite difference method of analysis is being used which relies on a failure
criterion (e.g. Mohr-Coulomb) to predict geotechnical failure. When calculating earth pressure
coefficients, the slope of the ground and the slope of the wall should be accounted for, since
they have an impact on the result (see Clause C8.2.1).
In rock, the ultimate passive resistance is equal to the unconfined compressive strength of the
rock (qu), but this should be reduced to take account of discontinuities and lower strength
stratum within the rock mass.
Anchor pull-out
The geotechnical strength (Rug) controlling anchor pull-out is the frictional resistance of the
anchor/ground interface. The active anchor length should be located so that it cannot transfer
load back to the wall and needs to be at sufficient depth to prevent a failure cone forming at the
surface (see Clause C12). Both the short-term and long-term strengths need to examined, as the
latter may often be critical.
Anchor stiffness has a very significant impact on anchor load and the deflection of the soil-
supporting structure. Close cooperation is required between the structural and geotechnical
engineer to ensure that appropriate assumptions are being incorporated in the analysis.
Prop failure
Whether or not a prop will fail is normally a matter determined by the structural designer rather
than the geotechnical engineer. As for anchors, prop stiffness has a very significant impact on
prop load and the deflection of the soil-supporting structure. Close cooperation is required
between the structural and geotechnical engineer to ensure that appropriate assumptions are
being incorporated into the analysis.
C7.3.4 Characteristic values
Characteristic values of geotechnical parameters are required to be conservatively assessed. It
is emphasised, in NOTE 3, that this is not necessarily a low value, but an appropriate value for

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 18 DRAFT ONLY

the particular situation being analysed. This is different from '92 AUSTROADS which took the
5th percentile as the characteristic value.
C7.3.5 Geotechnical strength reduction factors (  g)
When traditional, hand calculation methods are used to design structural elements founded in
the ground, it is apparent that the geotechnical strength reduction factors required by the
Standard can be applied to the resisting forces generated in the ground in order to satisfy
Equations 7.3.2(1) and 7.3.3(1).
With computer methods of analysis, this is not normally possible, since most analyses provide a
design that indicates an equilibrium condition. Additional hand calculations may be required to
demonstrate that a particular computer aided design satisfies the requirements of the Standard.
C7.4 Design for stability
AS5100.1 specifies that the stability limit state is the loss of static equilibrium by sliding,
overturning or uplift of a part or the whole of the structure.
The stability limit state is reached when, in geotechnical terms, the disturbing force exceeds the
resisting (or restoring) force. It could be argued that this applies to all geotechnical failures (see
Clause C7.3.1), but a strict reading of the Standard specifies that it only applies where Rug is not
a relevant criterion. This applies to the toppling failure of a gravity retaining wall, when bearing
capacity at the toe is not a pre-requisite for collapse.
C7.5 Design for serviceability
AS 5100.1 specifies one of four serviceability limit states as the deformation of foundation
material or a major load-carrying element of sufficient magnitude that the structure has
limitation placed on its use, or is of public concern.
From a geotechnical perspective, the serviceability limit state is reached when ground
movements, or more importantly differential movements between sub-structure elements,
become sufficiently large as to cause distress to the structure, or to give the public reason for
concern. Examples of serviceability limit states are—
(a) differential settlements causing unacceptable loads or deflections along a structure;
(b) soil movements rendering expansion joints unserviceable;
(c) poor rideability at bridge approaches due to embankment settlement relative to the
abutment;
(d) unsightly retaining wall movements due to displacement or rotation; and
(e) cracking of the concrete structure due to ground movements.
In practice, experience will often show which type of limit state will govern the design, and
other analyses may be omitted or limited to less detailed control checks.
Computer methods are particularly well suited to analysing the serviceability limit state and can
be used to calculate total and differential movements between different sub-structure elements.
Since the Standard calls for a load factor of 1.0 to be used in relation to loads resulting from soil
movements (see Clause 8.4), the movements to be assessed at the serviceability limit state will
normally be those that come directly from the analysis.
Attention is drawn to the following:
(i) Analytical methods which represent the ground as a series of independent springs or
utilise coefficients of sub-grade reaction can be overly simplistic and should not be used
to assess the serviceability limit state other than for simple structures with widely
separated sub-structure elements. Even then, the coefficient of sub-grade reaction or
spring stiffness should be calculated to be relevant to the foundation being analysed and

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 19 DRAFT ONLY

the geotechnical conditions on the site. Rarely will these be anything like the coefficient
of sub-grade reaction used for pavement design.
(ii) Consolidation settlements in cohesive soils should normally be calculated based only on
structural dead loads (including embankment loads). This approach may need to be
reviewed if vertically applied live loads are present for a large proportion of the total
time. In this case, the mean dead plus live load, averaged over time may be more relevant.
(iii) Settlements in non-cohesive soils and rocks should normally be calculated based on dead
plus live load (without dynamic load allowance).
(iv) Other transient loads (including earthquake effects) should only be considered if they will
have an adverse impact on sub-structure performance.
(v) Where loose sandy or silty soils and a high water table exist, the possibility of
liquefaction during an earthquake should be considered, since total loss of bearing
capacity may be possible.
(vi) Lateral ground movements and dishing often result where vertical loads are applied over
large areas, such as in the vicinity of an embankment or a number of closely spaced,
independent, shallow foundations. The effect on the structure should be considered.
(vii) Load redistribution may occur in the structure, which reduces differential movements
between sub-structure elements. Geotechnical analysis will often not take this into
account. If the effect is likely to be of significance, it will be necessary either to model the
structure as well as the ground in the analysis, or to modify the design actions accordingly
and re-run the geotechnical model on a number of occasions with revised structural loads
until a reasonable representation of the equilibrium condition is reached.
(viii) The ground is rarely uniform. Analytical models normally include a series of discrete
layers to accommodate vertical variations of strength and stress-strain parameters.
Horizontal variations are less often modelled, but can be of significance when considering
the serviceability limit state. Since horizontal variations are more difficult to predict from
site investigation data than vertical variations, consideration should be given to checking
the sensitivity of the structure to assumed possible variations in properties. This can be
done, for example, by assessing reasonable upper and lower bound values of strength,
stiffness and strata thickness, based on the field and laboratory test data, and
incorporating these in the analytical model in order to calculate the worst case of
differential movement. Variations across the site should be applied in a way that produces
the most adverse effects on the structure. The accuracy of calculated settlements of sub-
structure elements even for simpler types of structures depends very much on the
particular site, the data available, and who does the prediction.
(ix) The settlement (immediate, consolidation and creep) of approach embankments and the
underlying ground should be evaluated and added to settlements due to structural loads
when considering the effects on the structure.
(x) The relative timings of embankment, sub-structure and super-structure construction can be
significant where the ground is subject to time dependent settlements. Differential
settlements between abutments and piers should be checked as well as the effects of
negative skin friction on piles.
Ideally, the specific sensitivity of a bridge structure to differential settlements and lateral
movements should be calculated. In the absence of such calculation (i.e. for simple structures),
or for preliminary design purposes, the following differential settlements may be considered as
tolerable:
(A) Bridge with simply supported spans 25 to 50 mm.
(B) Bridge with beams continuous over piers 12 to 25 mm.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 20 DRAFT ONLY

(C) Box girder bridges 12 to 25 mm.


(D) Incrementally launched bridges 6 mm longitudinally, 2 mm transversely.
(E) Pier frames 6 mm.
(F) Culverts and retaining walls 6 mm.
Guidance on the selection of suitable values of differential movements of bridge foundations
may also be found in Baker (1991) (Ref 4).
Where relevant, the following possible causes of additional ground movement and their impact
on serviceability should also be considered:
(1) Changes in ground-water level.
(2) Ground heave due to frost, anticipated moisture content variation or pile driving.
NOTES:
1 In expansive soils, shallow foundations should be located at sufficient depth to ensure that
changes in moisture content are of no importance (see AS 2870). The same should apply to
frost sensitive foundations. Pile caps in either of these situations should either be located at
sufficient depth below the heave affected zone, or incorporate a layer of compressible
material between the underside of the cap and the ground.
2 In soils where ground heave due to pile driving could lift previously driven piles,
construction procedures should be specified to restrike the displaced piles to obtain the
required load capacity or toe levels. Consideration should be given in these circumstances to
the use of low displacement piles or to bored, cast-in-place piles.
(3) Regional long term consolidation of the ground.
(4) Mine subsidence.
C7.6 Design for strength, stability and serviceability by load testing a prototype
Prototype substructure elements may be—
(a) subject to more scrutiny during construction and therefore be better built than typical sub-
structure elements constructed later on the same project; and
(b) tested in isolation from other influences, such as settlement imposed by adjacent
construction, or group action.
Due allowance should be made for these factors when assessing the results.
C7.7 Design for durability
Materials used to construct sub-structure elements should be selected to be durable enough to
have adequate sectional properties at the end of their service life.
If not otherwise specified, the design life of a sub-structure element should not be less than that
of the structure.
C7.8 Design for other relevant design requirements
(No Commentary)

C8 LOAD AND LOAD COMBINATIONS


C8.1 General
(No Commentary)

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 21 DRAFT ONLY

C8.2 Loads
C8.2.1 General
Item C(a)
(See Clause C2)
Item C(b)
Any sub-structure element should be designed for the appropriate earth pressures, which will
range from active through to passive, depending on the ground displacements that take place
during construction and the method used to compact backfill. Some computer programs
calculate earth pressures generated against sub-structure elements from first principles, while
others require the user to input appropriate values of the active and passive earth pressure
coefficients. Guidance on the calculation of earth pressure s can be found in texts such as
Bowles, J.E. (1996) (Ref 17); Fang, H-Y (1991) (Ref 18); Naval Facilities Engineering
Command (1986) (Ref 20). The effects of surcharge, sloping ground and sloping wall faces
should be considered and accounted for if they are likely to be significant. Specific guidance in
relation to earth pressures acting on rigid walls can be found in Lee and Herington (1972) (Ref
5).
Some methods of analysis require the initial lateral stress in the ground, the Ko condition, to be
quantified. Soils can be normally consolidated if they are geologically recent, or over-
consolidated due to stress changes that have occurred in the geological past. Over-consolidation
can lead to horizontal stresses as high as the passive pressure. Ko is not easily measured in
soils, so it is often necessary to make an assessment based on an understanding of the local
geological history. Mayne and Kulhawy (1982) (Ref 6) suggested that—
Ko = (1  sin')OCR sin' . . . C8.2.1(1)
Where
OCR = over-consolidation ratio (the ratio of the highest over-burden
pressure ever experienced at a point in the ground to the current
over-burden pressure at the same point).
NOTE: Ko = (1  sin') only for normally consolidated soils (i.e.
OCR = 1)
Yielding retaining walls are usually designed to withstand active earth pressures, because only
small movements of the order 0.1% of wall height are required for pressures to drop to this
level.
Unyielding retaining walls, including piled structures founded on rock, may be subject to earth
pressures higher than the active values. The following should be considered during design:
(1) Excavation for sub-structure construction allows the horizontal stress to drop well below
the K o value and often to the active value. This can be calculated if a sheet pile, bored pile
or a diaphragm wall is constructed, using finite element, finite difference or specific
programs intended for this kind of analysis.
(2) Similar calculations can be undertaken for walls restrained from movement by ground
anchors, by connecting ties between parallel walls or by abutments which are propped off
the superstructure.
(3) Where excavation sides are left unsupported or are battered it is evident that the
horizontal stress near the excavation drops to zero while the excavation is open. As a
result the stresses applied to the sub-structure element that is subsequently built will
depend on the nature of and manner in which backfill is placed (see Item (g) of the
Clause).

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 22 DRAFT ONLY

(4) Pressures higher than at-rest values can also result from thermal movement of a
superstructure integral with its abutments. These can be calculated from the known
thermal expansion of the superstructure and the elastic properties of the ground. Since the
passive load capacity of the ground is potentially very large the possibility of buckling of
the bridge deck may need to be considered as a stability limit state condition.
When computing earth pressures, the nature and drainage properties of the backfill material are
of primary concern and should always be properly determined. Where practical, free-draining
granular material should be used as backfill and should be properly compacted.
If the backfill is composed of material other than free-draining granular material, the sub-
structure may be subjected to forces from frost heave, swelling, or down-drag caused by
consolidation. The swelling pressure of expansive clay backfill is discussed by Pufahl, Fredlund
and Rahardjo (1993) (Ref 7). Other considerations affecting the choice of backfill are cost,
availability and the possibility of scour, particularly at bridge abutments.
Items C(c) and C(d)
Many computer programs are able to calculate loads imposed on abutments and retaining walls
from surcharges including wheel loads.
Where highway, foundation or other surcharge loads are applied within a zone defined by a
plane rising at 45° to the horizontal from the underside of the heel of a footing or base of a
retaining structure, surcharge loads should be applied on walls in accordance with AS 5100.2.
This surcharge loading provides for live load actions and should be used in the design of all
abutment walls and wing-walls located within the specified distance from the traffic.
For low abutments (h < 3 m) where an adequately designed and suitably proportioned approach
slab supported at one end on the bridge is provided, the effect of live load surcharge will be
small and may be neglected.
For high abutments (h ≥ 3 m), the effect of vertical reactions on the backfill at the approach end
of the slab needs to be considered and may be treated as for dead loads.
The surcharge pressures thus determined are superimposed on those due to earth pressure and
other effects.
AS 5100.2 gives the distribution of live loads on horizontal surfaces of foundations and soil-
supporting structures covered by fill, and applies to individual and overlap conditions and limits
imposed by the boundaries of a structure. The depth of fill includes the pavement and should be
measured to the finished wearing surface.
When surcharges from several loads overlap, the total load is considered as uniformly
distributed over the areas defined by the outside limits of the individual areas. The total width
of distribution should not exceed the total width of the structure supporting the fill.
For single spans, the effect of live load on horizontal surfaces may be neglected when the depth
of fill is more than 2.5 m and exceeds the span length, as the live load acting upon the surface of
the structure is very small compared to the total dead load.
For multiple spans, live load on horizontal surfaces may be neglected when the depth of fill
exceeds the distance between faces of the end abutments.
For the effect of attenuation of dynamic effects through fill, see also AS 5100.2.
Item C(e)
See Clause C8.2.4.
Item C(g)
Some aspects of the stresses induced on sub-structure elements due to the method of
construction are discussed in Item C(b). Backfill is usually placed in layers and compacted with

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 23 DRAFT ONLY

mechanical plant. During construction, lateral stress is induced in the fill by the compaction
equipment which is related to its dynamic load (static plus vibrating). This is normally found to
be in the range 20 to 30 kPa as shown in Figure C8.2(B).
In the top few compacted layers, this exceeds the passive pressure, so the soil fails and the
lateral stress is limited to the passive pressure. At depth, the active pressure exceeds the
compaction induced stress, and so becomes the relevant stress to apply to the wall. This
phenomenon is discussed by Ingold (1979) (Ref 8). If the computer program being used for
analysis does not have the facility to input a stress distribution due to compaction, some way
around it, perhaps by using an artificially increased K a, may have to be found in order to
reproduce something like the appropriate average stress distribution.

FIGURE C8.2(B) EARTH PRESSURE DUE TO COMPACTION BEHIND A WALL

Item C(h)

Wherever possible, the construction sequence should be planned to prevent the possibility of
significant stresses in adjacent sub-structure elements occurring.
Item C(i)
(No Commentary)
C8.2.2 Loads induced by soil movement
Lateral soil movements and settlements can result from a variety of sources, including—
(a) embankment construction;
(b) piling works;
(c) cutting excavation;

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 24 DRAFT ONLY

(d) loading of foundations;


(e) changes in ground water level, and groundwater lowering in particular;
(f) swelling of heavily compacted, or expansive clays; and
(g) slope instability.
Where time dependent ground movements are likely to be significant, the following steps
should be taken to minimise their effects:
(i) Construction can be staged to allow the bulk of the movement to occur before sub-
structure construction commences.
(ii) Preloading of embankments can be considered.
(iii) Vertical drains can be used to speed consolidation.
(iv) Ground improvement techniques can be employed.
(v) Slip layers can be used on piles subjected to negative skin friction.
C8.2.3 Construction loads
Construction loads can be significant, especially when large items of earthmoving plant and
mobile cranes are present on or around the structure. These items of plant will often have quite
different axle and wheel configurations and apply different loads from the design vehicles. The
resulting action effects should be determined and allowed for during the design. Details of the
assumed construction vehicle should be included on the drawings, since the contractor may
choose to use something different.
Where an item of plant is assumed to travel at restricted speeds, or on specific parts of the
structure, or other special precautions need to be taken, these design assumptions and
constraints should also be shown on the drawings.
Loads on geotechnical structures are affected by the construction sequence. Generally speaking,
onerous loading conditions do not necessarily always occur after construction, with some
critical loadings possible during the course of the works.
During the design, any numerical model staging needs to be consistent with a carefully planned
construction sequence for accurate determination of loads and deformations. It is essential to
examine soil/structure and groundwater behaviours during this process to satisfy design
constraints.
C8.2.4 Water pressure
The loads applied by water pressure on a sub-structure element can be significant and should be
taken into account in its design. The effect of buoyancy on structural components and the self-
weight of the soil (submerged density) should be included in the analysis. Water pressure should
always be applied normal to the surface on which it acts.
It is essential that the correct water levels be used taking into account design life and other
environmental and hydrogeological factors. In certain ground conditions piezometric profiles
may differ from initial ground water level, and must be accounted for by sufficient site
investigation and measurements before completing the design.
Anticipated variations in water level should be accounted for in order to derive the worst case
design condition. Where groundwater information is not available it is better to make a
conservative, but rational assumption, since the assumption that the ground is saturated can
more than double the load on a structure.
Water pressure effects can only be ignored in the design of sub-structure elements which are not
subject to immersion and which have free draining granular backfill material and an effective
drainage system. Where backfill does not drain freely (permeability k<104 m/sec), or a thin,

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 25 DRAFT ONLY

man-made, drainage medium is relied on immediately behind a wall, ground water pressures
may be elevated within the active wedge and therefore contribute to the actions applied to the
wall. This should be accounted for in the analysis.
Where tension cracks could occur in the soil within the active wedge (i.e. in soil which is
predominantly clay), lateral hydrostatic water pressure should be included for the full depth of
the crack, or for half the wall height, whichever is the greater. The maximum crack depth ( Zo)
may be theoretically calculated as follows:

2c u
Zo  . . .C8.2.4(1)

Where
cu = undrained shear strength
 = soil bulk density
Many computer programs will ask the user whether tension cracks are to be assumed to be full
of water. The answer should always be yes, unless there are reasons to believe that this situation
could never arise.
Full lateral water pressure should be assumed below the level of weep holes or other drainage
outlets.
Water should preferably be prevented from entering the backfill from the surface, otherwise any
resulting seepage pressures should be allowed for in design.
C8.3 Load combinations for strength and stability design
C8.3.1 General
Different load combinations for strength and stability design apply to foundations and to soil-
supporting structures, as required by Clause 7.
C8.3.2 Foundations
The relevant load combinations for design are detailed in AS 5100.2.
The design procedure for foundations requires the application of the load factors detailed in
AS 5100.2 for each load combination.
The load factors will usually be greater than 1.0 to increase the design action effects in
accordance with limit state design principles.
A default load factor of 1.5 applies where the load is not covered by the Clause or AS 5100.2.
Load factors applicable to the effects of soil movements are detailed in the Clause (see Clauses
8.2.2 and C2).
C8.3.3 Soil-supporting structures
Load combinations for the design of soil-supporting structures requires that the most adverse
combination of loads be used for design. The relevant load combinations are detailed in
AS 5100.2 (see Clauses C2 and C7.3.3).
All load factors are set to 1.0 for this analysis in accordance with the design process for soil-
supporting structures.
C8.4 Load combinations for serviceability design
Design loads and load combinations for serviceability design of foundations and soil-supporting
structures are detailed in the Clause and AS 5100.2.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 26 DRAFT ONLY

All load factors have been set at 1.0, as realistic and conservative estimates of movements or
deformations are required, using conservatively estimated values of material parameters.

C9 DURABILITY
C9.1 General
Geotechnical site investigations should include as a matter of routine the chemical analysis of
soil and water samples to assess the potential for aggressive or corrosive conditions. Only if
reliable local knowledge is already available should this work be omitted.
In natural soils, chemical analysis can normally be limited to the determination of pH and
sulphate and chloride contents to establish the potential for attack on ordinary Portland cement
reinforced concrete.
Where acid sulfate soils are anticipated, appropriate testing should be undertaken based on local
experience and publicly available information (refer to References for guideline publications on
acid sulfate soils).
Soil resistivity tests should be undertaken where unprotected steel will be exposed in natural
ground or in backfill (e.g. sheet piles, reinforced soil walls).
Where fill material is present, particularly if it contains household refuse or industrial wastes,
full chemical analysis should be carried out to identify potentially aggressive substances.
AS 2159 and AS 4678 contain information about the durability of materials used in foundations
and soil-supporting structures in specific environments.
C9.2 Durability of timber
See AS 1604.1 and AS 5605.
The durability of timber used in the construction of foundations and soil-supporting structures
will depend on the species of timber, the type of preservative treatment (if any), and the
environment in which it is placed. A short structural life can be expected where timber will be
subject to—
(a) fluctuating water levels;
(b) damp conditions;
(c) marine conditions where molluscan or crustacean borers may be present; and
(d) where termites or borers are present above ground or above water level.
Treated timber or durable species may be used as permanent components of foundations or soil-
supporting structures where other materials are not economically available. They may also be
suitable in situations where highly chemically aggressive ground water exists and other options
are not feasible. Economic analysis should include the anticipated frequency and cost of
failure/replacement.
Untreated timber permanently located in soil below the ground-water level can be assumed to
have an indefinite life provided it is not subject to attack by borers or other organisms.
Untreated timber may be used for protection fenders in waterways adjacent to bridge
substructures, but regular inspections should be carried out.
C9.3 Durability of concrete
See AS 5100.5 and its Supplement.
Conventional concretes containing Portland or blended cements may be subject to chemical
attack by—
(a) sulphates and sulphuric acid which can occur naturally in soil;

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 27 DRAFT ONLY

(b) corrosive chemicals which may be present in industrial wastes in groundwater and fill;
(c) organic acids and carbon dioxide which may be present in groundwater as a result of
decaying vegetable matter.
As a guide, chemically aggressive conditions exist where—
(i) the concentration of sulphate (expressed as SO 3) exceeds 0.5% in soil or 1 in 200 parts per
million in groundwater; or
(ii) lime-dissolving carbonic acid in groundwater exceeds about 30 milligrams per litre.
Dense, well compacted and adequately cured concrete of low water/cement ratio incorporating
blended cement will generally provide adequate protection against moderate sulphate and acid
attack and provide better durability in marine environments than Portland cement concretes.
Attention to the design of the mix and to proper placement, compaction and curing is required.
Protective coatings such as bitumen, coal-tar epoxy, or vinyl ester can provide protection for
cast-in-place concrete, but may be less reliable on driven piles where the coatings can be
stripped off during driving.
Reduced skin friction will generally result where piles are coated with bitumen. This should be
allowed for in design.
The aggressive chemical effects on Portland cement concrete by the high sulphate contents
typically present in sea water is inhibited by the presence of the sodium chloride in the water.
C9.4 Durability of steel
In the absence of more reliable information, AS 2159 contains information relevant to the
durability of unprotected steel in different types of ground conditions.
The corrosion of steel buried in soil or immersed in water is most commonly electro-chemical in
nature. For electro-chemical corrosion, atmospheric oxygen or oxygen dissolved in the water is
essential. This type of corrosion is stimulated by the presence of dissolved salts in the
electrolyte water or by acidic conditions, which may result from the decay of animal or
vegetable matter, or the presence of natural minerals such as pyrite.
Corrosion may also result from electrolysis caused by stray electrical currents near buried
power cables or electrified rail lines, or in some circumstances may be due to bacterial activity.
Specialist literature should be consulted, or guidance obtained from an expert in these cases.
In marine conditions, the rates of corrosion of unprotected steel surfaces can be expected to be
considerably more than those for land conditions. At any one site the range of corrosion
conditions will vary considerably according to position. The order of increasing severity
according to position is—
(a) submerged and mud zone;
(b) tidal zone;
(c) atmospheric zone; and
(d) splash zone.
Water temperature, the insulating effect of marine growth and rust layers on the surface, the
velocity of the water passing the sub-structure element, protective actions of oil and grease
floating on the water and the type and concentration of water soluble pollutants can all influence
corrosion rates.
For steel surfaces exposed to the atmosphere, the rate of corrosion will depend upon the type of
protective coating and the extent of routine maintenance. See also AS/NZS 2312.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 28 DRAFT ONLY

Consideration should be given to loss of the coating by sand abrasion, damage during
construction and, in particular, pile driving operations. Coatings applied to surfaces exposed to
marine conditions, particularly in the intertidal and splash zones, will normally have a life of
only a few years. The most effective protection for a steel component or structure is to embed it
in concrete.
Careful assessment should be made of the cost of coating steel surfaces, and should be
compared with the cost of providing an additional thickness of metal that will be permitted to
corrode away in time.
Cathodic protection measures will function only where the metal is constantly immersed or is
below the groundwater level. Cathodic protection is an ineffective treatment in the splash zone
and against chemical attack.
C9.5 Durability of slip layers
The durability of slip layer coatings applied to piling should be considered.
There is reported evidence of bitumen coatings on piles being degraded by the action of
sulphate reducing bacteria in soft clays. Caution should be exercised and consideration given to
any reduction in the effectiveness of the slip layer that could result.
C9.6 Durability of other materials
(No Commentary)

C10 SHALLOW FOOTINGS


C10.1 Scope
Shallow footings support applied loads by direct bearing of the underside of the footing on the
soil or rock below. A footing is considered to be shallow when it is founded no more than one
footing width or 3 m below the surrounding ground level, whichever is less.
C10.2 Load and load combinations
See Clause C8.
The loads and load combinations to be considered when designing shallow footings are
specified in Clauses 8.3.2 and 8.4, and AS 5100.2.
C10.3 Design requirements
C10.3.1 General
The weight of the footing and any backfill material should be included in the total load applied
at the founding level.
The combined effect of vertical loads and moments should be taken into account when assessing
the design action effect (S*). On sloping sites, strip footings should normally be founded on a
horizontal bearing surface, stepped where necessary to maintain adequate depth of cover.
Where the geotechnical strength of a footing is calculated for a footing on clay using the
undrained soil strength (cu), groundwater level is not relevant and may be ignored.
Where the geotechnical strength of a footing is calculated using drained strength parameters ( c',
'), the effect of groundwater should be included, unless it will always be at a depth equal to or
greater than the width of the footing below the founding level.
Shallow footings on soil, except those for culverts, should not normally be located in or
immediately adjacent to streams, rivers or flood affected zones, except where a full assessment
of scour has been carried out and proper protection is provided. Where there is sufficient
evidence to show that the soil is scour resistant, having regard to the expected duration of
flooding and the velocity of flood flow, shallow footings should be located at a depth of—

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 29 DRAFT ONLY

(a) not less than 1.8 m below the permanent bed of a defined channel, and
(b) not less than 1.2 m below ground surface on flood affected banks of streams or rivers.
Cut-offs and aprons should normally be provided at the ends of footings for culverts and
associated structures, including wing-walls and drop structures, to prevent erosion undermining
the footings.
Shallow footings for permanent structures should generally not be founded on filled ground.
Only where fill has been properly placed and compacted and comprises selected fill material,
may consideration be given to founding on fill. In this case, the structure should be designed to
cater for the differential settlements which may occur.
Normally consolidated clays, organic soils and peat deposits can settle for long periods under
their own weight. They can also be very sensitive to changing ground water levels. Shallow
foundations should be avoided for permanent structures on these soils unless detailed analysis
shows them to be practicable. Foundations should normally be taken down to less compressible
strata at depth. Negative skin friction or down drag effects should be considered in this case.
An excavation for a footing can reduce the stability of existing adjacent footings, even when the
new excavation is not taken below existing foundation level. In addition, the presence of the
new structure may induce movements in adjacent structures, pipes and services. The magnitude
of all such movements should be calculated and any adverse consequences determined.
It is necessary to comply with local building regulations, by-laws, special regulations and any
other statutory requirements with regard to—
(i) maintenance of support for adjoining structures;
(ii) provision of protection against foreshore erosion; and
(iii) mining subsidence requirements in proclaimed areas.
If compliance creates an unacceptable design constraint, an acceptable resolution to the problem
should be found in discussion with the relevant authority. An example of an unacceptable design
constraint might be a statutory requirement to include an intolerable degree of structural
flexibility in a bridge being built in a mine subsidence area.
C10.3.2 Footing depth and size
(No Commentary)
C10.3.3 Design for geotechnical strength
C10.3.3.1 General
See Clause C7.3.
C10.3.3.2 Overall stability
A footing should not be designed in isolation from its surroundings. Under no circumstances
should design drawings imply a requirement for a subsequent excavation (e.g. for a services
trench) to be dug which will reduce the overall stability of a shallow footing.
C10.3.3.3 Ultimate bearing failure
Footing design should be based on a rational geotechnical model of the ground conditions which
fits the available data.
For conventional, vertically loaded foundations, Terzaghi's bearing capacity factors are often
used to calculate ultimate bearing capacity, although they are recognised today as being
conservative. These will be found in most soil mechanics and foundation engineering texts such
as Bowles (1996) (Ref 17); Fang (1991) (Ref 18); Simons & Menzies (2000) (Ref 19); and
Naval Facilities Engineering Command (1986) (Ref 20).

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 30 DRAFT ONLY

For more complex situations where inclined or eccentric loads, a sloping base, or a sloping
ground surface are involved more generally applicable bearing capacity factors, such as those
proposed by Brinch-Hansen (1970) (Ref 9) or Vesic (1975) (Ref 10) will provide more
economic designs.
The highest values of the geotechnical strength reduction factor  g may only be used when there
is reason to have confidence in the geotechnical model and in the design parameters. This will
normally come from careful site investigation and field and laboratory testing, but could also be
based on well documented, local experience. Lower values of  g should always be used when a
high level of confidence cannot be justified.
The serviceability limit state will often be found to be the controlling design condition,
effectively limiting the actual geotechnical strength reduction factor to a value considerably
lower than those in Table 10.3.3(A).
Where the influence of time dependent effects or transient repeated or vibratory loadings could
be significant, the possible effects on soil strength should be taken into account.
C10.3.3.4 Failure by sliding
Mobilising the expected ultimate sliding resistance Hug of a shallow footing depends on the
following:
(a) The strength of the soil/rock at the founding level. Footings should be inspected to ensure
that the soil/rock strength is as assumed for design. The base of the foundation excavation
should not be allowed to deteriorate before concrete is placed.
(b) The cleanliness of the footing excavation prior to pouring concrete.
(c) The shape/roughness of the underside of the footing which will determine whether sliding
will occur through soil/rock, or along the interface between the concrete and the ground.
In the absence of site specific data, the values of friction between a wall, or foundation, and the
ground given in Table C10.3.3.4 may be assumed.

TABLE C10.3.3.4
SLIDING FRICTION COEFFICIENTS AGAINST POURED CONCRETE
(Source: Naval Facilities Engineering Command (1986) (Ref 20))
Concrete poured against Friction coefficients Friction angle
tan  
Clean sound rock 0.70 35°
Gravel, gravel-sand, coarse sand 0.55 - 0.60 29°-31°
Clean fine to medium sand, silty 0.45 - 0.55 24°-29°
medium to coarse sand, silty or
clayey gravel
Clean fine sand, silty or clayey 0.35 - 0.45 19°-24°
fine to medium sand
Fine sandy silt, non-plastic silt 0.30 - 0.35 17°-19°
Very stiff and hard residual or 0.40 - 0.50 11°-26°
pre-consolidated clay
Firm to stiff clay and silty clay 0.30 - 0.35 17°-19°

Mobilising the expected ultimate passive resistance of the soil in contact with the vertical face
of the footing Epr (see Item (b) of Clause C8.2.1) depends on the following:

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 31 DRAFT ONLY

(i) that the backfilling between the face of the footing and the surrounding excavation has
been compacted or cement stabilised to achieve a strength at least as high as that of the
surrounding soil/rock.
(ii) that the footing can, without danger and without exceeding any serviceability limit state
of the structure, achieve displacements large enough to mobilise the required passive earth
resistance.
(iii) there being no chance that soil in front of the footing will be removed by erosion, or by
subsequent excavation, or other activity.
C10.3.4 Design for structural strength
C10.3.4.1 General
All concrete footings should be reinforced.
When designing reinforced concrete shallow footings for strength, it is generally satisfactory to
assume that the contact pressure is uniformly distributed over the effective foundation area, as
described in this Clause.
Where hand calculation methods are being used, the following simplifications may be applied:
(a) Strip footings When designing the longitudinal reinforcement in a strip footing subject to
column loads, the bearing pressure distribution on the base of the footing may be
simplified by regarding for calculation purposes those parts of the footing beneath the
columns as pad footings concealed within the strip footings. The intermediate sections of
the strip footing between the columns are assumed free of loads. The extent of the footing
sections subjected to bearing pressure can be determined by assuming that the full bearing
resistance is utilised beneath each concealed footing.
(b) Raft footings Simple raft footings may be designed for strength using a similar approach
to that described above for a strip footing. The simplified bearing pressure distribution
beneath the raft may be determined by regarding those parts of the raft beneath columns
and walls as concealed footings. Again the extent of the footing sections subjected to
bearing pressure can be determined by assuming that the full bearing resistance is utilised
beneath each concealed footing.
Raft footings may be partly piled to reduce the settlement of the raft alone. The bearing
capacity of the raft over and above the piles may be calculated using methods such as
contained in Poulos and Davis (1980) (Ref 21 ) or Fleming et al (1994) (Ref 22 ).
(c) Pad footings The critical section for bending for pad footings should be taken at the face
of the supported column for bending and at a distance of d/2 from the face of the
supported column for shear. If the edge of the footing is less than or equal to d from the
face of the supported column, the footing should be designed in accordance with deep
beam theory and should be considered to be rigid for moments in that direction.
The footing of a retaining wall should be designed similarly.
C10.3.4.2 Structural failure as a result of footing movement
Settlement and horizontal displacements of sub-structure elements under the design loads
should be calculated to ensure that they do not lead to a failure of the structure. The effects of
angular distortion on bridges depend on the particular structural articulation and design, so each
case should be individually assessed.
Where a strength limit state occurs, the structure may be beyond the linear elastic range.
Analysis may require progressive iteration between geotechnical software able to calculate
ground movements and structural software able to calculate the structural response. Close

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 32 DRAFT ONLY

cooperation is normally required between the geotechnical engineer and the bridge designer
where major or complex bridge structures are involved.
Footings subjected to inclined loading should be designed so as to minimise and preferably
eliminate horizontal movements.
C10.3.5 Design for serviceability limit states
C10.3.5.1 General
See Clause C7.5.
C10.3.5.2 Tilting
(No Commentary)
C10.3.6 Design for durability
See Clause C9.
C10.4 Structural design and detailing
(No Commentary)
C10.5 Materials and construction requirements
(No Commentary)

C11 PILED FOUNDATIONS


See also AS 2159 and its Supplement.
C11.1 Scope
Note that the design life of piled foundations in AS 2159 is 40 to 60 years whereas in AS 5100 it
is 100 years. This should be taken into account, particularly when considering pile durability.
C11.2 Load and load combinations
See Clause C8.
The actions to be considered when designing piles are specified in AS 2159. However, the
requirements of Clauses 8.3.2 and 8.4, and AS 5100.2 also apply.
Allowance should be made for the effect of axial, eccentric and lateral loads, and their
combinations on individual piles and when applied at pile cap level to pile groups.
C11.3 Design requirements
C11.3.1 General
Scour
In addition to normal design requirements the minimum depth of embedment below anticipated
scour level of any pile should be—
(a) 3 m into a stiff to hard soil;
(b) 6 m into a soft to stiff soil; or
(c) not less than one third of the pile length.
whichever is the greatest. Pile shafts should be checked to ensure buckling will not occur.
Application of loads
The loads imposed on a piled foundation should be applied concentrically with the axis of the
pile or at the centroid of the pile group at pile cap level unless specific provision has been made
for eccentric loading. If permanent eccentric loads are to be supported, due allowance should be

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 33 DRAFT ONLY

made for them in the structural design of the pile(s). Any effect on the supporting soil should be
analysed.
Allowance should be made for piles being out of position and at the limit of tolerance permitted
by the specification. The worst possible situation should always be considered.
Spacing
The spacing of piles should be determined taking into account the—
(i) nature of the ground;
(ii) method of pile installation;
(iii) possibility of soil heave or compaction during installation; and
(iv) impact on pile group behaviour.
The pile spacing should be sufficient to enable the required number of piles to be installed to
the required penetration without damage to previously installed piles or to any adjacent
structures.
The following minimum values for pile spacing are given as a guide:
(A) Friction piles The centre-to-centre spacing should not be less than 2.5 times the diameter
or nominal size of the pile. For steel H-section piles and other less common sections, the
perimeter of the encompassing rectangle should be used instead.
Wider spacings or variable pile batters (rakes) should be used for flexible piles to offset
the possibility of subsurface convergence of piles during driving.
()B Mainly end bearing piles The centre-to-centre spacing should be not less than twice the
least width of the pile cross-section.
()C Enlarged base (under-reamed) bored piles No recommendations are made, but
consideration should be given to the interaction of stresses at the toe and the need to
maintain stability of the under-reams prior to concreting.
()D Bulb base piles The centre-to-centre spacing should be not less than 1.5 times the
expected diameter of the bulb.
Clearances
Consideration should be given to installation tolerances when developing pile layouts.
Where the pile heads are to be incorporated in a reinforced concrete pile cap, the design
clearance from the side of any pile to the nearest edge of the pile cap should be not less than
200 mm.
When the pile heads are to be connected by a crosshead that supports the superstructure, the
design clearance from the side of any pile to the nearest edge of the crosshead should be not less
than 100 mm.
For fully embedded piles, large clearances may be required to avoid interference with the
crosshead or pile cap reinforcement.
Where piles are to be installed adjacent to an existing structure, sufficient room should be left
between the centre of the pile and the adjacent structure to give the piling rig room in which to
work. Appropriate dimensions should be sought from piling contractors prior to finalising the
layout.
Head embedment
For concrete piles, the full cross-section of each pile should project a distance of not less than
50 mm into the pile cap or crosshead after all unsound material or laitance has been removed

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 34 DRAFT ONLY

from the head of the pile. In addition, a sufficient length of the pile reinforcement, or other
means of anchorage, should extend into the pile cap or crosshead to adequately resist all applied
forces.
For steel H-section piles a sufficient length of embedment should be provided for all direct
loads and bending moments, with a minimum embedment of 300 mm. In addition, a sufficient
depth of concrete should be provided above the top of the pile to prevent punching shear failure
occurring.
For timber piles, a minimum length of head embedment of twice the pile diameter should be
used for positive connection if resistance to direct tension or bending is required. For other
conditions, a lesser head embedment length may be used with a minimum embedment of
100 mm.
Effects of installation on adjacent structures
Be aware of the possible impact of piling on adjacent structures.
Driven and vibrated displacement piles can create vibrations that can damage nearby structures
and noise that can disturb people.
Continuous flight auger piles can cause large settlements of the surrounding ground, particularly
where auger progress is likely to slow with depth, for example when the pile will be largely end
bearing on rock.
Design considerations for pile caps
In the design of the pile caps, consideration should be given to the structural actions of the piles
and the members supported on the pile caps, as well as the span to depth ratio of the pile cap.
Guidance for detailing reinforcement of simple compact pile caps may be obtained from
Adeben, Kuchma and Collins (1990) (Ref 11).
Where a pile cap is essentially a beam (for example where a single row of piles is used), beam
theory or deep beam theory may be used to design the cap.
Where a pile cap is essentially a flexible slab supported on numerous piles, the cap may be
designed using slab theory.
Designers should be aware that for normal proportions pile caps are too stiff to allow loads to be
evenly distributed among the piles in a group. Any redistribution should be supported by a
suitable analysis correctly modelling the actions within the pile cap.
Computer programs often use a 2-dimensional (in plan) representation of the pile group and
assume a rigid pile cap to determine the action effects in piles. In most cases, this will produce a
reasonable representation of the actual situation. If there is reason to assume that the locations
of the applied loads combined with pile cap flexibility are likely to make this simplification
unrepresentative of the actual situation, alternative analysis should be carried out using a
flexible pile cap.
C11.3.2 Design for strength
See Clause C7.3.
Computer programs are commercially available to assist with the design of individual piles and
pile groups. Guidance on theory, practice and methods of hand calculation can be found in a
number of texts such as Poulos & Davis (1990) (Ref 21), Fleming et al (1994) (Ref 22),
Tomlinson (1997) (Ref 23).
Static or dynamic pile testing or both is recommended for all pile types, particularly on larger or
more important projects to demonstrate that the design pile performance can be achieved.
Dynamic pile testing together with signal matching of the measured blow to derive a unique

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 35 DRAFT ONLY

solution for the pile resistance using the wave equation can be used to confirm pile resistance
(see also AS 2159 and its Supplement).
In soils (e.g. sands or silts) that could exhibit a significant pore pressure response to vibration or
rapid repeated loading, an apparent increase or decrease in capacity during installation can
occur. A suitable delay should be allowed between pile installation and testing so pore pressures
can stabilise.
C11.3.3 Design for serviceability
See Clause C7.5.
Computer programs are commercially available to calculate the theoretical load-settlement
behaviour of piles and pile groups. Confirmation can and should be provided by load testing
where little relevant local experience is available. Static load testing is likely to give the most
reliable indication of the settlement of a single pile at the design load, but it does not
necessarily indicate how a pile group will behave.
C11.3.4 Design for durability
See Clause C9 and C11.1.
C11.4 Structural design and detailing
C11.4.1 General
AS 2159 contains considerations that should be taken into account when designing piles for
usual types of structures. Additional requirements for bridges are specified in the Standard
because of the longer design life and different loading conditions, which may include horizontal
loads not usually experienced in other types of structures.
Piles that depend largely on shaft adhesion or piles that penetrate at least 3 diameters into a hard
soil or rock stratum should be designed as structural columns with end fixity and lateral support
appropriate to the surrounding soil conditions.
Piles that depend largely on end bearing and may not penetrate 3 diameters into the founding
stratum (e.g. where soil overlies a high strength rock) should be designed as structural columns
with pinned ends and lateral support appropriate to the surrounding soil conditions.
The length of a pile in contact with air, water, or soft material, including the length that may be
exposed as a result of scour, should be considered as laterally unsupported.
Stresses arising from the following sources should be considered in the structural design of a
pile, as appropriate:
(a) Handling (preformed piles only).
(b) Installation (preformed piles only).
(c) Applied axial loads and axial forces caused by negative friction (down drag).
(d) Applied lateral loads and moments.
Handling (preformed piles only)
Bending stresses induced in a pile by handling before installation should be determined having
regard to the number and spacing of the lifting points, the mass of the pile and the length of the
pile. The calculated stresses should be increased by 50% to allow for impact.
Handling seldom affects timber piles, steel H-piles or steel tubular piles, but these effects
should be checked.
Installation (preformed piles only)

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 36 DRAFT ONLY

Provided that piles are installed in accordance with sound construction practice, installation
stresses should not normally be of concern. However, excessive driving stresses can occur in
piles of any length (e.g. short piles driven to rock) if care is not taken. Consideration should be
given to both the compressive and tensile stresses induced during installation of piles driven
harder than to the specified set. Stresses which are high enough to damage a pile can be caused
in several ways during driving, especially under the following circumstances:
(i) Where the surface of the pile head is not normal to the direction of the blow, unwanted
lateral shear and bending stresses will be generated. This can be due to out of square
fabrication of the top of the pile, or when a pile goes off line during driving.
(ii) Where the driving helmet is not in contact with the full cross-sectional area of the pile,
the pile may undergo local pile-top crushing.
(iii) Where the cross-sectional area of the pile which receives the blow is not large enough, the
material may be crushed. This can apply to—
(A) timber or steel piles driven without a helmet; or
(B) where a helmet or packing does not fully cover the head of a concrete pile.
(iv) Where the hammer is oversized (delivers too much energy) for the pile section or in the
case of concrete piles where the pile-top cushioning is insufficient. In the latter case, more
general pile-top damage may result. In such cases, tensile or compressive failures at other
locations in the pile are more likely to occur.
(v) Where the pile is driven to end bearing on a very dense or hard soil layer or rock,
especially where the overlying strata are soft or loose. In this situation compressive
stresses at the pile toe may be as much as twice the compressive stress at the pile top,
leading to toe damage.
(vi) If piles have to be driven through a very stiff or dense upper stratum to reach the founding
stratum at a lower level, excessive driving stresses may be developed while penetrating
the upper stratum. In such a case, consideration should be given to pre-boring through the
upper stratum. If pre-boring is carried out, appropriate measures should be taken to ensure
that the pile is fully supported laterally after installation.
(vii) Where the pile is driven through a dense stratum into underlying weaker soil which offers
little toe support, reflected tensile stresses may be large enough to cause tensile failure of
reinforced or prestressed concrete piles.
(viii) If insufficient head and helmet packing is used during driving, the period of transfer of
energy from hammer to pile will be shortened and the peak magnitude of stress increased
to such an extent that the concrete may shatter in compression adjacent to the head.
Driving stresses may be estimated from a wave equation driveability analysis, which takes
account of reflection of stress waves within a pile during driving and permits evaluation of the
resulting tensile stresses. The use of such analysis requires assumptions of soil response and
driving system efficiency, stiffness and coefficients of restitution, and should be performed by
an engineer experienced in such analyses.
Stresses during driving should be limited to:

Material Stress
MPa
Steel 0.9f y

Concrete in compression 0.8 f c'


Concrete in tension:
Reinforced concrete piles with—

(a) 2% reinforcement 0.8 f c'


389944842.doc - 15/06/2018 7:26:57
(b) 2% reinforcement f c'
Prestressed concrete piles Initial prestress
DRAFT ONLY 37 DRAFT ONLY

where f c' is the compressive strength of the concrete at the age of driving (see also AS 2159).
Axial loads and negative friction
Where downdrag forces are developed in a pile due to negative friction, the maximum load
applied to the pile should include the axial load and the calculated maximum downdrag force. It
should be noted that the maximum load in a pile subject to downdrag occurs at depth below the
head of the pile. This may be significant when considering the design of the pile.
Where piles are installed in expansive soils, they should either be designed to withstand tensile
forces that might develop or be provided with an appropriate slip coat.
Lateral loads and moments
Computer programs are available to calculate stresses induced by lateral loads and moments.
Net areas
The following cross-sectional areas should be used for the calculation of pile strengths:
(A) Timber piles The minimum pile cross-sectional area for adequately protected treated
piles. For untreated piles, the thickness of sapwood should be disregarded for all
naturally round timbers. Even when specifically measured, sapwood should be taken as
not less than 15 mm measured radially.
(B) Steel piles The gross cross-sectional area of steel, less an appropriate allowance for the
loss of section due to corrosion.
(C) Concrete piles Generally the gross cross-sectional area, but in the case of reinforced
concrete piles in pile trestle piers and abutments, use of the cracked section may be
necessary for combined bending and axial load.
(D) Permanently steel-cased concrete piles The gross cross-sectional area of the concrete
shaft plus the remaining cross-sectional area of permanent casing after making due
allowance for corrosion.
(E) Continuous flight auger (concrete and grout-injected) and alpha piles The cross-
sectional area corresponding to the minimum probable shaft diameter.
Splices
Wherever possible, splices should be avoided. If this is not possible (e.g. where there are head
room constraints), splices should be designed to develop the full strength of the pile cross-
section. In addition, the location of any splice should be such that it does not affect the
structural performance of the pile. Splices should be designed so they do not significantly
increase or alter the cross-sectional area of the pile.
If a splice cannot be designed to resist the full bending moment on a pile, it should be located in
a region of the pile where bending is small and can be resisted, and its strength and location
does not affect the performance of the unit during handling, driving or in the finished position.
Pile splices should be located at a depth where corrosion will not be an issue, generally where
replenishment of oxygen cannot take place. If this is not practicable, then durability issues
should be taken into account as specified in Clause 9.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 38 DRAFT ONLY

Driving shoes and toe reinforcement


While not normally necessary when a pile is to be wholly driven in soft soils, for anticipated
hard driving conditions or if driving conditions warrant the toe should be protected from
damage.
Hard driving conditions include driving through weathered rock, gravel, hard clay, dense sand
and the like with standard penetration test (SPT) values greater than 30.
For hard driving conditions, timber, reinforced concrete and prestressed concrete piles should be
fitted with a steel or cast iron driving shoe. The shoe should be concentric with the axis of the
pile and should be designed to be integral with the pile.
In hard driving conditions, consideration should also be given to the need for reinforcement of
the toes of—
(1) steel H-section piles, using plates welded to the flanges and web; and
(2) open-ended tubular piles, using a driving ring welded to the toe.
Proprietary hardened and cast steel tips for steel H-section piles could also be used. If external
stiffening plates are used at the toe of a pile it should be noted that these can reduce the skin
friction along the pile by forming an oversized hole in the ground.
Durability
See Clause C9.
The durability of a pile under service conditions should be considered in all cases.
C11.4.2 Design details relevant to specific types of piles
Timber piles
The following should be considered:
(a) Pile toes The toe of a timber pile should be pointed in the form of a truncated cone or a
pyramid having a square end 50 mm to 200 mm wide and a length one and a half to two
times the pile width.
(b) Driving shoes Unless the pile is to be wholly driven in soft soils, the toe of a timber pile
should be fitted with a protective metal driving shoe. The shoe should be concentric with
the axis of the pile and should be firmly fixed to the end of the pile by bolting or other
suitable means to prevent removal during driving.
(c) Driving rings The head of a timber pile should be fitted with a steel driving ring to
prevent splitting.
C1.1.1.1 Precast reinforced concrete piles
The minimum strength of concrete specified for durability should take contact with the soil into
account.
Item C(a) Size and shape
The minimum cross-sectional area specified for reinforced concrete piles in salt water should
take account of the increased cover requirements for this exposure classification.
Item C(b) Driving straps
(No Commentary)
Item C(c) Reinforcement

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 39 DRAFT ONLY

The requirement for spirals or ligatures takes into account driving stresses as well as those
imposed by service conditions. The closer spacing of this reinforcement at pile ends accounts
for higher driving stress at these locations.
Item C(d) Mechanical joints
Concrete used in mechanically-jointed precast concrete piles should have a 28 day compressive
strength of not less than 40 MPa.
Mechanically-jointed precast concrete piles should comprise precast concrete lengths, with cast-
in-joints, which enable lengths to be rapidly connected together in place to suit any required
length. Precast concrete lengths should be not less than 3 m and not more than 20 m.
Joints should be designed so that they provide a permanent connection between the precast
concrete lengths. The strength of the joint should be not less than that of the precast concrete
lengths of pile being joined, unless specifically designed to resist pile loads at joint level. All
exposed metal joint components should be provided with a suitable protective coating, or
alternatively, a rust-resistant steel should be used in fabrication. Joints should preferably be
located below the permanent low water table.
C11.4.2.2 Prestressed concrete piles
Item C(a) Concrete strength
As the concrete is prestressed, and these piles may be used in aggressive environments, a higher
compressive strength has been specified for prestressed as compared to reinforced concrete
piles. The concrete cover appropriate to the exposure classification should be specified.
Excessive cover should be avoided, as spalling of the concrete may occur during driving.
Item C(b) Size and shape
(No Commentary)
Item C(c) Prestress and reinforcement
The minimum level of prestress specified should take into account driving and handling stresses
as well as structural considerations.
Sufficient non-prestressed longitudinal reinforcement should be provided at each end of the pile
to withstand secondary stresses induced by driving.
The head reinforcement specified should take into account any requirements for splicing and
anchorage to crossheads or pile caps.
The requirements for helical reinforcement may be varied where experience indicates that other
amounts of reinforcement may be appropriate, for example on piles of larger cross-sections.
Item C(d) Mechanical joints
As for precast reinforced concrete piles (see Clause C11.4.2.1(d)).
C11.4.2.3 Cast-in-place reinforced concrete piles
Cast-in-place piles may be constructed with an open bore which can be inspected and proved, to
establish that the specification requirements in relation to soil/rock strength, roughness and
cleanliness have been achieved. However, several types of cast-in-place pile including small
diameter (< 750 mm) piles constructed with an open bore, continuous flight auger piles, Atlas
piles and driven piles with an expanded base provide no such opportunity. Where no local
experience is available to justify the design, pile load testing should be undertaken to
demonstrate that the performance requirements have been met.
The minimum strength of concrete specified for durability should take contact with soil and
permanent water into account.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 40 DRAFT ONLY

Item C(a) Reinforcement


The requirements of the Standard in relation to reinforcement should be applied to all piles
constructed with an open bore. A number of proprietary piling systems do not involve
construction with an open bore and have their own specific reinforcement requirements (e.g.
continuous flight auger piles). In these cases, every effort should be made to adhere to the
intention of the Standard, even if it is not possible to adhere to the specific requirements.
Longitudinal reinforcement should protrude from the top of the pile a minimum of 600 mm, or
whatever greater distance is required for anchorage to the pile cap or crosshead. Any stirrups or
spiral reinforcement should be removed from the protruding longitudinal bars at the top of the
pile.
The specified clear spacing between longitudinal bars is to facilitate concrete placement.
Lapped or fillet welded splices in the longitudinal reinforcement should be long enough to
develop the yield stress of the bars at the splice.
Proprietary mechanical bar splices may be used provided they develop the full strength of the
bars across the joint and concrete cover is maintained.
Item C(b) Casing
Where the toe of a casing is stiffened by a ring protruding on the outside of a casing, skin
friction will be reduced relative to an unstiffened casing or one where the stiffening ring
protrudes on the inside. Wherever possible, casing stiffened on the outside should not be used
but the required inside diameter to accommodate the rock socket cutting tool should be taken
into account.
For cast-in-place reinforced concrete piles with permanent steel casing the residual thickness of
the casing should not be less than 6 mm after allowing for loss of thickness due to corrosion.
Where corrosion is expected, due allowance for corrosion of the casing in accordance with the
provisions of Clause 9 should be made and, if applicable, the remaining casing thickness may be
taken as contributing to the structural strength of the pile.
C11.4.2.4 Steel piles
The 10 mm minimum thickness is retained from previous Australian Bridge Design Codes, with
the additional clarification that the specified minimum thickness is after corrosion losses have
been taken into account. The minimum thickness was probably specified originally to cover
both corrosion and buckling issues.
The following should be taken into consideration:
(a) Shapes Steel piles should generally be made from rolled or welded hollow sections, or
rolled H-sections. Hollow sections can either be open ended or have one end closed.
(b) Toe reinforcement The toe of an open tubular driven steel pile should be reinforced with
a suitable driving ring.
For hard driving conditions, the toe of steel H-section piles should be reinforced with
plates welded to the flanges and the web.
Where steeply sloping bedrock is known to exist, the toe of steel H-section piles should
be bevelled so that a wedging action develops on penetration into rock. In all other cases,
the toe of the pile should be cut square to the pile axis.
(c) Splices Where unavoidable, additional lengths should be spliced to previously driven
lengths by qualified butt welding or by welded or bolted splice plates.
(d) Cap plates In general, cap plates are not required provided an adequate length of the
steel pile is embedded in a concrete pile cap of sufficient depth.

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 41 DRAFT ONLY

(e) Protective coating The most effective means of protecting a steel pile is to use concrete
in the upper section which would otherwise be subject to corrosion. Either a precast
concrete top section can be welded on to a steel pile, or the upper section of the pile can
be encased in concrete. In either case, the concrete section should extend from the pile
cap soffit to a minimum of 3 m below the existing ground level, to the lowest likely scour
level, or to 600 mm below the lowest anticipated water table, whichever level is the lower.
In easy driving conditions, piles can be coated with a bitumastic coal tar, or a tar epoxy
paint may also be effective. Extreme care needs to be taken in handling and driving such
piles to avoid damaging the protective coating.
(f) Sheet piling Sheet piles may be used as cantilevered or anchored walls and for coffer
dam construction. They may be used for temporary works, or left in place as a permanent
part of the structure. In the latter case, the piles should be designed to remain stable under
the worst conditions of scour.
Sheet piling should generally consist of interlocking driven steel sections, which should
be designed as structural members capable of carrying all loads imposed on them. The
stresses produced should not exceed those set out in AS 5100.6.
Allowance should be made for loss of section due to corrosion by reducing the relevant
section properties. For estimating rates of corrosion, see Clause 9.4. and Clause C9.4
C11.5 Materials and construction requirements
C11.5.1 General
(No Commentary)
C11.5.2 Spacing, edge distance and embedment of piles
See Clause C11.3.
C11.6 Testing
See AS 2159 and its Supplement.

C12 ANCHORAGES
C12.1 Scope
The Clause applies to non-prestressed ties and anchors and to prestressed ground anchors.
Design guidance can be obtained from publications such as BS 8081 (1989) (Ref 12), EN1531
(Ref 13) or Littlejohn & Bruce (1975-76) (Refs 14, 15 & 16).
Non-prestressed ties and anchors
Non-prestressed (passive) tie bars may be used to stabilise substructures, wing walls or
retaining walls subjected to earth pressures and/or lateral loads. Horizontal or inclined tie bars
are commonly used to support the tops of sheet pile, bored pile, or diaphragm walls. They can
also be used to prevent over-turning of gravity walls or abutments where there is a restriction on
base width, or a need to raise the retained height above the original design height.
Ties will normally consist of threaded steel rods or reinforcing bars with the possible
incorporation of a turnbuckle for initial tensioning.
When used in compressible soils, sagging of the ties can occur as the soil below the anchors
settles and the weight of the overlying soil is transferred to the ties. This effect can be
minimised by using support piles at 6 to 10 m centres, or by laying the ties in large diameter
pipes to isolate them from the surrounding ground.
For permanent construction, adequate corrosion protection should be provided for the ties.
Generally this is done using bitumen coating and fabric wrapping, or grease filled plastic

389944842.doc - 15/06/2018 7:26:57


DRAFT ONLY 42 DRAFT ONLY

sheathing. Due to the possibility of flexural cracking, the use of concrete encasement as
protection is not recommended, except where the ties are also galvanized.
The reaction of tie bars can be taken by—
(a) Deadmen Deadmen are short concrete blocks or continuous beams deriving their
resistance from passive earth pressure;
(b) Raked piles Resistance is provided by the use of a driven raked pile or piles; and
(c) Sheet piles Sheet piles are driven to form a continuous wall which derives its resistance
from passive earth pressure.
Anchor bars formed by grouting or mechanically anchoring non-prestressed reinforcing bars
into drilled holes in the foundation rock may be used for footings on rock to improve structure
stability, by increasing the foundation's resistance to overturning.
For anchor bars grouted into drilled holes in rock, anchorage depth will be determined by bar-
grout bond, grout-rock bond and resistance of the rock mass to extraction as a block.
Passive ties and anchors will stretch as the load is applied and the soil around the anchor length
will deflect. Movements should be assessed to ensure that deflections do not exceed the
serviceability limit state condition.
Consideration should always be given to physical load testing of ties and anchors to
demonstrate that adequate capacity and load-deflection behaviour is being achieved during
construction.
Ground anchors
The term ground anchor refers specifically to a stressed, high tensile steel tendon anchored in a
hole drilled into soil or rock. In this context, the term tendon can include a bundle of tendons,
each of which is individually anchored and stressed. The most common form of ground anchor
comprises a tendon bonded at its far end in high strength grout and stressed at its active end at a
stressing anchorage. The function of the ground anchor is to provide a large concentrated force
at a precise location on a structure or sub-structure to withstand a combination of applied loads
(see Figure C12.1).

(a) Rock anchor (b) Soil anchor

FIGURE C12.1 GROUND ANCHOR DETAILS

The term horizontal anchor is sometimes loosely used to mean any tendon inclined at a
downward slope not greater than 30° to the horizontal. The term vertical anchor is similarly
used to mean any tendon inclined downwards at more than 30° to the horizontal.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 43 DRAFT ONLY

Ground anchors can be permanent or temporary. Permanent anchors should have adequate
corrosion protection and require careful, often staged, stressing to provide the required load
capacity and design life.
Permanent ground anchors can be used to provide—
(i) lateral support for retaining walls, abutments and piers;
(ii) end fixity (moment capacity) for retaining walls, abutments and piers;
(iii) anchor points for tension structures; and
(iv) long term stability of cut faces in jointed rocks.
Temporary ground anchors can be used to provide—
(A) support to temporary sheet pile walls; and
(B) reaction points for pile load testing.
C12.2 Load and load combinations
(No Commentary)
C12.3 Design requirements
C12.3.1 General
The factored loads applied to the anchors should be less than the total of the factored resistances
of the system resisting these loads. As anchorage systems usually comprise several elements,
analysis of the redundant system is usually complex and the accuracy of the results will depend
on the assumptions made and the capabilities of the analytical method being used.
C12.3.2 Site investigation
See Clause C6.
C12.3.3 Design for strength
Anchors should be designed in accordance with generally accepted principles. Guidance will be
found in publications such as BS8081 (Ref 12) and Littlejohn & Bruce Part 1 (1975) (Ref 18).
Design of deadman anchors
In order to develop the full capacity of a deadman, it should be located in stable ground at
sufficient distance from the structure that the passive sliding wedge of soil in front of the
anchorage does not interfere with the active sliding wedge of backfill behind the structure. The
anchorage should also be located below a line from the bottom of the structure sloping on an
angle  to the horizontal, as shown in Figure C12.3.3(A). Further guidance will be found in
publications such as Naval Facilities Engineering Command (1986) (Ref 20).

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 44 DRAFT ONLY

FIGURE C12.3.3(A) DEADMAN LOCATION

Design of rock anchors


For tendons bonded in grout, the determination of the depth of anchorage and length of the fixed
anchor should take into consideration—
(A) uplift resistance of the rock mass;
(B) rock to grout bond;
(C) grout to tendon bond; and
(D) tendon design.
Design guidance can be found in publications such as Tomlinson (1994) (Ref 23) and Wyllie
(1999) (Ref 24).
Bond between grout and rock
Bond at the rock to grout interface may be considered as uniformly distributed along the fixed
anchor length.
When shear strength tests are carried out on representative samples of the rock mass with due
allowance for in-place joints or other defects, the average effective design bond stress should
not exceed the minimum shear strength of the rock.
In the absence of shear strength data or pull-out tests, typical values of characteristic ultimate
rock-to-grout bond stresses are given in Table C12.3.3.

TABLE C12.3.3
ROCK TO GROUT ULTIMATE BOND STRESSES
Rock type Ultimate bond
(MPa)
Low strength rock 0.7–1.4
Medium strength rock 1.4–2.1
High strength rock 2.1–2.8

A minimum fixed anchor length of 3 metres is recommended to guard against variable rock
quality and constructional imperfections.
Bond between tendon and grout
The characteristic ultimate value of bond resistance may be taken as—

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 45 DRAFT ONLY

(1) 1.0 MPa for clean wire tendons; and


(2) 2.0 MPa for clean strand tendons.
For strand tendons, the effective area for bond should be taken as the surface area of a smooth
cylinder having a diameter equal to the nominal diameter of the strand.
For bar tendons, the bond may be improved by using a nut and washer at the fixed anchor end,
with anchorage lengths determined by actual test.
Tendon
Tendons for rock anchors should be designed generally to the requirements of AS 5100.5.
C12.3.4 Design for other relevant factors
(No Commentary)
C12.3.5 Design for serviceability
Creep can result in ground anchors losing prestress with time and this can be accompanied by
movement of the anchor head. Where such changes could have an impact on the serviceability
of the structure, provision should be made for anchors to be load tested and re-stressed at any
time during the design life of the structure.
C12.3.6 Design for durability
See Clause 9 and Clause C9.
Plastics, greases and grouts used to provide corrosion protection for anchorages should be
selected taking into account the intended anchorage life, its environment and possible
interaction and deterioration of these materials over time.
A multi-level protection system will usually be required to give assurance of durability. Such
systems usually involve specific protection of each anchor strand together with staged grouting.
Short-term accelerated durability testing of materials proposed may be required to verify their
suitability.
C12.4 Materials and construction requirements
The materials for an anchorage should be selected taking into account their ability to protect the
load bearing parts of the anchorage system from corrosion and deterioration over the life of the
anchorage.
The anchorage should be constructed so that the multi-level anchorage protection system is not
compromised as a result of installation practices.
The actual strength of the anchorage compared to design should be verified at the time of
installation by testing, as detailed in Clause 12.6.
C12.5 Anchorage installation plan
On major projects and where considered relevant the anchorage installation plan should be used
as the basis for ‘as built’ and performance monitoring records. Records should include any
problems experienced during construction and stressing, which may be of value later if
performance problems are experienced during the design life of the structure.
C12.6 Anchorage testing
C12.6.1 General
The anchorage testing specified applies mainly to prestressed ground anchorages. It has little
application to other types of anchorage systems.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 46 DRAFT ONLY

C12.6.2 Proof load tests


(No Commentary)
C12.6.3 Acceptance tests
(No Commentary)
C12.7 Monitoring
Monitoring should be specified where the founding material is of doubtful quality or where
there are concerns about the long-term performance of the anchors, or both. Monitoring can be
costly because of the setting up costs involved. Consequently, monitoring should only be
specified when it is essential that long-term anchor performance needs to be verified. Guidance
on monitoring will be found in EN1531:2000 (Ref 13).

C13 RETAINING WALLS AND ABUTMENTS


C13.1 Scope
See Clause C7.3.3.
The Clause refers to the design of retaining walls and abutments which act as soil-supporting
structures, on which the primary loads on the wall and the resistance of the wall are determined
by the properties of the soil being retained.
When the primary load on an abutment is not from the retained soil but is that transferred from
the bridge superstructure or other source and thence to the ground, then the abutment should
also be designed as a footing in accordance with Clause 10.
Where both types of loading exist, then the design of the substructure should be carried out in
accordance with both Clause 10 and Clause 13, and engineering judgement exercised as to the
interpretation of the analytical output and subsequent design of substructure elements.
When locating, proportioning and detailing retaining walls, the requirements of AS 5100.1
regarding location, clearances, protection, hydraulic design and the requirements of the other
Parts of AS 5100 should be considered.
Issues specific to different types of walling systems are discussed in Clause C13.3.
For reinforced soil walls, see Clause C1.
C13.2 Loads and load combinations
See Clause C2.
In the design of retaining walls and abutments, loads and other actions to be applied should be
those—
(a) specified in AS 5100.2, including earth pressure loadings;
(b) resulting from soil movement (see Item (b) of Clause C8.2.1);
(c) from surcharges;
(d) from water pressure;
(e) from compaction during backfilling (see Item (g) of Clause C8.2.1);
(f) resulting from soil displacement during piling; and
(g) from any additional loads or effects.
Where scour or subsequent excavation may remove fill material over a footing projection or
sloping wall, the stabilising effect of both the mass of the fill and any passive resistance should
be disregarded.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 47 DRAFT ONLY

C13.3 Design requirements


During the design of retaining walls, consideration should be given to the following:
(a) Compliance with the various requirements of AS 5100.1 concerning location of the
structure and its individual components, horizontal and vertical clearances at structures,
waterway area and waterway shape.
(b) Satisfying the positional and other site constraints imposed by construction requirements.
(c) The effects of interaction between bridge superstructure and sub-structure and the
consequent effect on overall economy of the structure.
(d) Fixing of shallow foundation levels to satisfy any depth requirements.
(e) Aesthetic requirements at urban sites which may require special treatment or component
shapes.
(f) The ability of the structure type chosen to withstand the expected differential settlements.
Masonry and unreinforced concrete walls, crib walls, gabion walls and reinforced earth walls
should be designed as gravity type structures.
Crib walls
Crib walls may be formed with precast concrete, steel or timber members, and together with the
enclosed soil mass are considered to act as gravity type retaining walls.
Generally, the basic components of crib walls are manufactured as proprietary systems by
manufacturers who issue design information relevant to their product.
The components of crib walls should be selected from materials with durability appropriate to
the location. In a corrosive environment, galvanised steel components may require an additional
protective coating.
Untreated timber members should not be used for crib walls unless there is a specific need to do
so, such as for aesthetic reasons (see Clause C9.2).
When checking individual elements of a crib wall, reference should be made to AS5100.5 or
AS 5100.6, or other appropriate relevant Standards.
The minimum depth of toe foundations for crib walls should be 300 mm for single walls and
600 mm for double or triple walls (see Clause C13.6).
Crib walls should be filled with free draining granular material and should have additional rear
drainage.
Crib walls should not be used for structures subject to wave action, but may be used when
subjected to occasional river flood flows, provided the durability of the wall is ensured by
selecting fill material which is not easily erodible.
For design requirements for the ties of tied crib walls, see Clause 12.
Where multiple cells are used, they should be adequately interconnected to ensure group action,
with individual elements also adequately connected.
Care should be exercised with curved walls, since crib walls with a convex front face are much
more susceptible to damage by transverse deformations than are concave walls.
Sheet pile walls
Sheet pile walls are often used for temporary shoring purposes during construction, but may be
also used as permanent structures under suitable conditions, particularly in marine applications.
Sheet pile walls may be formed from steel, concrete, various plastics, or timber. The use of
untreated timber will generally be restricted to temporary works.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 48 DRAFT ONLY

Closed wall systems which are required to exclude water entry should be designed for the sum
of earth pressure and full ground water pressure.
Sheet pile walls may be designed for cantilever, braced or tied action and, in addition to loads
specified in Clause 8.2, the effects of temporary loads such as those caused by construction
equipment should be considered.
Computer programs are commercially available for the design of cantilevered and
anchored/propped walls. Used carefully these will normally provide detailed bending moment,
shear force and deflection information at different stages of construction.
In the determination of lateral loads due to earth pressure and their distribution, the method and
sequence of construction and the tolerable movements and deformations of the wall and soil
should be taken into account. In the absence of site specific test data, the values of wall friction
given in Table C13.3 may be adopted.
The overall stability of the soil mass surrounding the sheet pile wall system should be
investigated. Unless scour protection is specifically provided, the depth of embedment of a wall
should be measured from the lowest depth of anticipated scour.

TABLE C13.3
SLIDING FRICTION COEFFICIENTS AGAINST STEEL PILES
(Source Naval Facilities Engineering Command (1986) (Ref 20)
Non-cohesive soils
Interface Materials Friction coefficients tan  Friction angle 
Clean gravel, gravel-sand mixtures, well 0.40 22
graded rock fill
Clean gravel, silty sand-gravel mixtures, 0.30 17
single size hard rock fill
Silty sand, gravel or sand mixed with silt 0.25 14
or clay
Fine sandy silt, non-plastic silt 0.20 11

Cohesive soils (terminology changed to suit AS 1726)


Interface Materials Adhesion C w
(kPa)
Very soft soil 0 to 12
Soft soil 12 to 25
Firm soil 25 to 37
Stiff soil 37 to 45
Very stiff soil 45 to 62

C1.1.1 Design for strength and stability


See Clause C7.3.3.
For gravity walls on soil, the resultant action at the founding level should be within the middle
third for static loading and within the middle half for earthquake loading at the serviceability
limit state.
For gravity walls on rock, the resultant action at the founding level should be within the middle
half of the base for both static and earthquake loading.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 49 DRAFT ONLY

Where the foundation is on soil, overturning stability will usually be satisfied if bearing
capacity criteria are satisfied. However, overturning can be critical on strong foundation
materials such as rock, or when the base of the wall is small as with crib walls.
Computer programs are commercially available for the design of retaining walls. Used carefully
these will normally indicate that equilibrium has been achieved under the applied loads, provide
detailed bending moment and shear force diagrams, and indicate deflections at different stages
of construction. Output results should be compared with limiting values and assessed to verify
that the requirements of the Standard have been met by the design.
C13.3.2 Calculation of earth pressures
See Clause C8.2.1.
Propping
When designing propped abutments or rigid frame structures, differential earth loads through
the structure should be investigated by checking the following loading conditions, assuming the
earth pressure coefficients appropriate to the sub-structure deformations:
(a) Full earth pressure and live load surcharge at both abutments;
(b) Full earth pressure plus surcharge at either abutment and half earth pressure (no
surcharge) at the other abutment; and
(c) Full earth pressure (no surcharge) at one abutment only where there is the possibility of
scouring out of the fill at an abutment. For this condition, only the strength and stability
limit states need to be considered.
C13.3.3 Design for eccentric and inclined loads
See Clause C7.3.3, Clause 10.3.3 and Clause C13.3.1.
C13.3.4 Design for serviceability
See Clause C7.5.
C13.3.5 Design for durability
See Clause C9.
C13.4 Structural design and detailing
C13.4.1 General
Base slabs
The projections of reinforced concrete footings for cantilever walls should be designed as
cantilevers supported at the wall face, and loaded with the weight of superimposed loads
together with the forces exerted by the foundations.
For counterfort walls, the heel slab may be designed as a continuous beam spanning between the
counterforts.
The heel slab should be well tied to the base of the counterforts to take the net downward load
from the slab, with due regard being paid to effective anchorage of this reinforcement in both
the slab and the counterfort.
Walls
The vertical stem of a cantilever type retaining wall should be designed as a freestanding
cantilever fixed at the top of the footing, unless tied construction is used.
For counterfort construction, the upper portion of a wall may be assumed to span horizontally
between counterforts, while the lower part will have a two-way bending action with
simultaneous cantilevering from the base slab.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 50 DRAFT ONLY

Counterforts and buttresses


Counterforts may be assumed to act as triangular shaped cantilevers fixed at the base and loaded
by the face slab.
Counterforts may be considered to act as T-beams designed in accordance with AS 5100.5,
while buttresses should be considered to act essentially as rectangular beams.
The main tension reinforcement of counterforts should be effectively anchored to heel and face
slabs by a system of vertical and horizontal bars or stirrups. These stirrups should be anchored
as close as practical to the bottom and outside of the heel and face slabs.
C13.4.2 Joints
Contraction joints should extend the full depth of the wall to the top of the footing. Provision
should be made for the transmission of shear forces by the use of keys or dowels at all joints.
Where dowels are used, they should be hot-dip galvanized or stainless steel and debonded on
one side of the joint. More effective protection measures should be considered in corrosive
environments (see Clause C9.4).
Care should be exercised in detailing keys or dowelled joints between members with
substantially different bending stiffnesses in order to prevent local overstressing or failure of the
joint area due to differential movements under the action of earth loads (e.g. counterfort
abutment/cantilever wing wall construction).
C13.4.3 Shrinkage and temperature reinforcement
All exposed faces of reinforced concrete sub-structures and retaining walls should be reinforced
for shrinkage and temperature effects to the requirements of AS 5100.5. Exposed faces of
nominally mass concrete retaining structures may not need such reinforcement provided other
means of reducing cracking, such as closer spacing of contraction joints (typically at 3 m
centres) are provided.
Where walls thicker than 200 mm are cast on a previously constructed base slab, the heat of
hydration on dissipation and restrained shrinkage may lead to cracking at 2 to 3 m centres along
the wall.
Such cracking may be controlled by—
(a) the inclusion of contraction joints at no greater than 3 m centres;
(b) the use of low heat cement or other measures to reduce heat accumulation; and
(c) large amounts of reinforcement placed horizontally in the lower 2 m of wall (e.g.
equivalent to 20 mm diameter bars at 250 mm centres) in accordance with AS 5100.5.
C13.5 Materials and construction requirements
The exposed face of gravity, cantilever and counterfort retaining walls should preferably be
constructed with a minimum batter towards the supported material of 20 mm per metre, to
counter the visual effect of possible rotation during backfilling.
For crib and gabion walls, a batter of 200 to 250 mm per metre is recommended. Gabion walls
may be terraced and battered.
C13.6 Drainage
See Clause C8.2.4.
The fill behind abutments and retaining walls should be effectively drained by providing a free
draining layer which may consist of a suitable filter cloth or a selected permeable fill material
with a suggested minimum thickness of 600 mm over the full depth of the structure.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 51 DRAFT ONLY

This layer should be adequately drained by weepholes through the walls, by perforated pipes, or
other adequate means. These drainage requirements may be relaxed where there is a reduced
chance of water pressure build-up, such as is likely to occur with spill-through abutments with
shallow headstocks.
For counterfort walls, a minimum drainage requirement consists of at least one weephole for
each pocket formed by the counterforts.
Typical drainage details are shown in Figure C13.6(A).
Crib walls
The drainage system for crib walls should be designed to prevent removal of the filling material
in the crib by scour during draining of the water.
For all heights of crib wall, a continuous 150 mm minimum diameter sub-soil drain should be
provided at the rear of the foundation slab to ensure a dry foundation.
Typical drainage details are shown in Figure C13.6(B).
Adequate drainage of the whole crib structure is essential. Many of the failures in crib walls
have occurred because material of low permeability was used as backfill, allowing the
development of high static or seepage water pressures. Unless effectively drained over the full
height, crib walls should be designed to resist lateral hydrostatic pressures in addition to soil
pressures.

FIGURE C13.6(A) DRAINAGE SYSTEM FOR LOW HEIGHT WALLS

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 52 DRAFT ONLY

FIGURE C13.6(B) DRAINAGE SYSTEM FOR CRIB WALLS

C14 BURIED STRUCTURES


C14.1 Scope
(No Commentary)
C14.2 Loads and load combinations
See Clause C8.2.1.
Consideration should be given to the type of earth pressure (active, passive or at-rest)
appropriate for the structure under consideration. The type of earth pressure chosen will depend
on structure flexibility. Consideration should also be given to the compacting effect of traffic
behind bridge abutments, which may increase earth loads above the active level.
Culverts should be designed for the loads specified in Clause 8.2 of the Standard taking into
account the following:
(a) At-rest earth pressures will generally be required.
(b) Dynamic load allowance should be as specified in AS 5100.2.
(c) Where live loads are distributed through fill in accordance with AS 5100.2, both the axle
loads and vehicle loads should be considered and the most severe loading used for design.
(d) The effect of differential earth pressures on the structure should also be investigated as
required by Clause 8.2 of the Standard.
C14.3 Design requirements
Bending moments in buried structures are sensitive to the relative stiffnesses of the structure
and the surrounding soil. The design should consider variations in these stiffness parameters to
determine structural sensitivity.
The design of base slabs should preferably be carried out using flexible footing analysis.
C14.3.1 Design for strength and stability
See Clauses C7.3 and C7.4.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 53 DRAFT ONLY

When designing rigid frame structures such as concrete culverts, differential earth loads through
the structure should be investigated by checking the following loading conditions, assuming the
earth pressure coefficients appropriate to the sub-structure deformations:
(a) Full earth pressure and live load surcharge on both sides.
(b) Full earth pressure plus surcharge on one side and half earth pressure (no surcharge) on
the other side.
(c) Full earth pressure (no surcharge) on one side only where there is the possibility of
scouring out of the fill on that side. For this condition, only the strength and stability limit
states need to be considered.
C14.3.2 Design for serviceability
See Clause C7.5.
Cracking and deflections of reinforced concrete structures should be limited in accordance with
AS 5100.5.
C14.3.3 Design for durability
See Clause C9.
C14.4 Structural design and detailing
(No Commentary)
C14.5 Materials and construction requirements
(No Commentary)

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 54 DRAFT ONLY

REFERENCES - SPECIFIC
1 BS8006 (1995), Code of Practice for Strengthened/Reinforced Soils and Other Fills,
British Standards Institute (available from www.saiglobal.com/shop/).
2 Elias, V., Christopher, B. R., and Berg, R. R., (2001), Mechanically stabilized earth walls
and reinforced soil slopes design and construction guidelines, U.S. Department of
Transportation, Federal Highway Administration, Report No FHWA-NHI-00-043, March
(2001).
3 Sloan, S., (2005), Geotechnical stability analysis; new methods for an old problem,
Australian Geomechanics, Vol 40, No 3, September 2005, pp 1-28.
4 Baker, R.M. et al., (1991), National Cooperative Highway Research Program Report 343,
Manuals for the Design of Bridge Foundations, National Transportation Research Board,
National Research Council, Washington DC.
5 Lee, I. K., and Herington, J. R., (1972), Theoretical Study of Pressures Acting on Rigid
Wall by Sloping Earth or Rock Fill, Geotechnique 22, No 1, 1-26.
6 Mayne, P. W., and Kulhawy, F. H., (1982), Ko-OCR Relationships in Soil, Journal of
Geotechnical Engineering, Eng Div ASCE, Vol 8, No GT6, June 1982, pp 851-872.
7 Pufahl, D. E., Fredlund, D. G., and Rahardjo, H., (1983), Lateral Earth Pressures in
Expansive Clay Soils, Canadian Geotechnical Journal, Vol 20.
8 Ingold, T. S., (1979), The effects of compaction on retaining walls, Geotechnique 29, No
3, pp 265-283.
9 Brinch Hansen, J., (1970), A revised and extended formula for bearing capacity, Danish
Geotechnical Institute, Copenhagen, pp 5 to 11.
10 Vesic, A. S., (1975), Bearing Capacity of Shallow Foundations, Chapter 3 of
Foundations Engineering Handbook. H. F. Winterkorn and H.-Y. Fang (eds.), Van
Nostrand Reinhold Co., New York, N.Y., U.S.A., pp. 121-147.
11 Adeben, Kuchma, and Collins (1990), Strut and Tie Models for the Design of Pile Caps,
an Experimental Study, ACI Structural Journal, V 87, No. 1, pp. 82-92.
12 BS 8081 (1989), Code of practice for ground anchorages, British Standards Institute
(available from www.saiglobal.com/shop/).
13 EN1531:2000, Execution of Special Geotechnical Work - Ground Anchors, EN Standards
(available from BSI@BSI-Sales.Com).
14 Littlejohn, G. S., and Bruce, D. A., (1975), Rock Anchors - State of the Art Part 1, Design,
Ground Engineering May, July 1975.
15 Littlejohn, G. S., and Bruce, D. A., (1975), Rock Anchors - State of the Art Part 2,
Construction, Ground Engineering September, November 1975.
16 Littlejohn, G. S., and Bruce, D. A., (1976), Rock Anchors - State of the Art Part 3,
Stressing and Testing, Ground Engineering March, April, May 1976.
REFERENCES - GENERAL
17 Bowles, J. E. E., (1996), Foundation Analysis and Design, McGraw-Hill Professional
Books.
18 Fang, H-Y., (1991), Foundation Engineering Handbook, 2nd Edition, Chapman & Hall.
19 Simons, N., Menzies, B. K., (2000), A Short Course in Foundation Engineering, Second
Edition. Thomas Telford.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 55 DRAFT ONLY

20 Naval Facilities Engineering Command (1986), Design Manual 7.02, Foundations &
Earth Structures, (NOTE: Available as a free download from the author's website).
21 Poulos, H. G., Davis, E. H., (1990). Pile Foundation Analysis and Design, Krieger
Publishing Company.
22 Fleming, W. G. K., Weltman, A. J., Randolph, M. E., Elson, W. K., (1994), Piling
Engineering, 2nd Edition, E & FN Spon an imprint of Routledge.
23 Tomlinson, M. J., (1997), Pile Design and Construction Practice, Fourth Edition, E & FN
Spon an imprint of Chapman & Hall.
24 Wyllie, D. C., (1999), Foundations On Rock, Second Edition E & FN Spon an imprint of
Routledge.

ACID SULFATE SOILS


The following websites provide some information about acid sulfate soils in Australia:
Federal: www.deh.gov.au/coasts/cass/links.html
www.nht.gov.au/nht1/programs/cassp/index.html
www.ozestuaries.org/indicators/econ_cons_acid_sulfate_soils.jsp
New South Wales: www.agric.nsw.gov.au/reader/soil-acidss
www.rta.nsw.gov.au/environment/downloads/rta_asm_guideline.pdf
Queensland: www.nrm.qld.gov.au/land/ass
Tasmania www.rpdc.tas.gov.au/soer/lan/2/issue/91/ataglance.php
Victoria www.vcc.vic.gov.au
Western Australia portal.environment.wa.gov.au/pls/portal/url/page/land/acid_sulfate_so
ils

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 56 DRAFT ONLY

APPENDIX CA
SITE INVESTIGATION TECHNIQUES
Table CA1 provides guidance in relation to available site investigation and testing methods,
what they can be expected to achieve, and what their limitations might be.

TABLE CA1
TYPICAL CHECKLIST FOR SITE INVESTIGATIONS
Description of test Information obtained Remarks
Field sampling and testing
Disturbed sample
Cutting removed from Soil/rock description. Inexpensive method of obtaining basic
bottom of auger flight, soil/rock description and properties.
Moisture content and
washings from wash-bored
Atterberg limits (not from Quick.
hole, or bulk samples from
wash-boring).
test pits bagged for
laboratory identification and Suitable for compaction tests
testing. (from test pits).

(b) Pocket (hand) penetrometer


and pocket (hand) shear
vane
Simple tests to assess soil Undrained shear strength. Tests only give a rough indication of the
strength in test pits or consistency of cohesive soils. Several
undisturbed samples. tests recommended at each level, or in
each sample recovered, so an average
result can be obtained. The small area
tested can lead to an over-estimate of
strength in fissured clays.
(b) Undisturbed tube sample
Cohesive soils from very Soil description. In sensitive soils, sample disturbance
soft to hard consistency can occur during sampling,
Moisture content.
obtained by means of an transportation and preparation of
open drive sampler (nominal Density/dry density. laboratory samples.
sample diameter normally Particle size distribution. Fortunately, strength tests on samples
50 mm), or in very soft soils
Samples for strength testing affected by disturbance yield lower
using a piston sampler
(triaxial drained or undrained (conservative) results.
(nominal diameter usually 75
or 100 mm). tests, shear box tests). Standard tube sample relatively quick.
Samples for consolidation Piston sample considerably slower and
testing (oedometer). samples require careful handling.

Undisturbed samples for all


other less frequently used
laboratory tests.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 57 DRAFT ONLY

Description of test Information obtained Remarks


Standard Penetration Test
(SPT)
A record of the number of Disturbed sample available A widely used test with empirical
blows of a 63.6 kg weight for soil description, moisture relationships published for granular
dropped 762 mm to drive a content, Atterberg Limits soils for consistency and bearing
thick walled 50 mm OD (clays only), particle size capacity and settlement of shallow
sampling tube a distance of distribution. footings.
300 mm.
Provides an indication of the Use of test has widened to cohesive
consistency of granular soil soils but the blow count-soil consistency
using correlation curves. relationship is not reliable without
extensive local experience and
calibration against other testing.
Relatively quick.
(d) Vane Shear
Test used in very soft to stiff Undisturbed undrained shear Frequently used in very soft to soft
clays. Vane pushed down strength. clays where sample disturbance affects
into soil and rotated undisturbed tube samples.
Remoulded undrained shear
horizontally until failure
strengths In recent years this test has been under
occurs.
criticism.
Relatively quick.
Pressuremeter test
Carried out in a borehole In place stress-strain Original pressuremeter modified to give
where a membrane is measurement in soils and a self-boring version for avoidance of
expanded laterally against weathered rock. disturbance in very soft or sensitive
the soil in the side of the clays and a high pressure version for
borehole. testing weathered rock.
Gives stress-strain relationships in
weathered rocks where previously
coring did not give samples suitable for
lab testing.
Testing gives stress-strain relationship
essentially in horizontal plane when
most structure loads are essentially in
the vertical plane. Test can therefore
give erroneous results where anisotropic
residual stresses are present in soils due
to over-consolidation and the like.
Slow and time consuming.
Rock coring
Typically NX size (75.7 mm . Continuous rock cores are invaluable
dia. hole, 54.7 mm core), or for assessment purposes for rock
in softer rocks use of NMLC sockets in piles, bearing capacity of
size triple core barrel. rock, excavation techniques in rock and
the like.
Relatively quick-standard borehole
procedure.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 58 DRAFT ONLY

Description of test Information obtained Remarks


(g) Permeability testing
In cased boreholes by means Permeability of subsoils or In soils the test is most suitable for fine
of gravity flow in soils or rock in place. sands and silts but much experience is
under pressure in rock using necessary.
a packer test.
In soils pumping tests from wells are
generally preferred to borehole
permeability tests.
Relatively time consuming.

(h) Screw plate test


A down-the-hole test where Stress-strain relationship in Several tests at various depths enables
a circular plate (or spiral vertical plane at any depth in evaluation of elastic settlement under
auger) is loaded and elastic soil profile. full scale conditions eliminating
settlements are recorded. previous limitations associated with the
plate load test.
Relatively slow and unusual test.
Quasi-static cone
(Electric Friction Cone End bearing side friction data Generally less expensive and quicker
Penetration Test [EFCPT], for full depth of probe. than boreholes and provides continuous
piezocone, seismic cone). data.
Indication of soil type and
Testing provides continuous
consistency. Automated equipment minimises
direct readings of end
operator induced influences (as can
resistance and side friction Good indicator of capacity of
occur with SPT) and provides graphical
of cone and sleeve (33.7 mm driven piles.
printout.
dia. typical). Probe Indicative values of soil
penetrates at constant rate Modern developments using Tee-bar and
strength settlement and soil
(20 mm/sec, typical) which ball ends provide theoretically more
moduli.
approximates static loading precise analysis of data.
of soil. Piezocone also Permeability and
Not suitable for hard/very dense soils or
records pore pressure consolidation parameters
rocks.
dissipation. Seismic cone (piezocone).
determines seismic velocity. Quick test providing more information
Insitu dynamic shear
per location than a borehole.
modulus (seismic cone).
Plate load test
Measurement of stress-strain Elastic modulus of soil or Of limited use due to shallow depth of
relationship of soil or rock rock. soil influenced by the load applied.
beneath loaded plate
Requires careful interpretation.
(typically 762 mm dia).
Modulus is undrained modulus unless
the test is very slow and drainage
assured.
Stress-strain relationship is in the same
plane as most footing loads and results
reflect effects of joints and defects in
rock.
Time consuming unless mobile reaction
system (i.e. truck or bulldozer) can be
used.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 59 DRAFT ONLY

Description of test Information obtained Remarks


(k) Pumping tests
Large scale field test where Permeability data derived on Most reliable form of permeability
water is pumped out from a large scale from soils in measurement when calculation of
well and draw down of water place. inflows to excavations, and the like is
table is monitored in important.
observation holes.
Normally takes several days to set up
and conduct.
Geophysical surveys
(Seismic reflection, seismic Variable depending on Conclusions should be checked using
refraction, gravity, conditions, includes subsoil boreholes, test pits and the like.
resistivity) profile determination
Equipment and interpretive skills
especially soil/rock interface.
generally very specialised.
Capable of providing extensive
geological information about a site and
an indication of engineering properties.
(m) Dynamic cone penetrometer
(DCP) or Perth
Penetrometer
Lightweight, hand operated, Indication of soil density, or Test most useful for preliminary
falling weight penetrometer. stiffness. investigations, or as a tool to prove
construction.
Perth penetrometer
correlated against bearing Quick to use.
capacity in sand.
Laboratory testing
(a) Soil classification
Density, moisture content, General data aiding soil Assists assessment of design parameters
Atterberg limits, particle size description and and the engineering characteristics of
distribution, organic content. classification. the soil.
(b) Triaxial compression test
Stress-strain behaviour of Undrained and effective A wide variety of testing procedures is
cylindrical sample of soil strength parameters of available to obtain various design
with control and monitoring cohesive soils. parameters.
of confining stresses and
Strength results obtained from a given
(sometimes) pore water
sample will vary according to test
pressures.
procedure used.
Quality of results very dependent on the
quality of samples.
(c) Consolidation test
Stress-strain behaviour of Consolidation settlement In layered or fissured soils
confined cylindrical sample properties of cohesive soils. consolidation times predicted from
of soil with controlled consolidation tests may be considerably
Secondary consolidation in
drainage conditions. longer than actually occurs in the field.
soft clays.
This is usually because drainage in the
Degree of over-consolidation test is vertical whilst, on site, higher
in clays. horizontal permeabilities permit
horizontal drainage.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 60 DRAFT ONLY

Description of test Information obtained Remarks


(d) Direct shear test
Stress-strain behaviour of Strength parameters. Results affected by dilatancy, especially
cylindrical or square box in granular soils.
Residual shear strength of
shaped samples sheared
clays. More advanced shear tests (e.g. simple
across a predetermined
shear, ring shear) but are not often
plane. Strength of joints in rock.
undertaken by commercial laboratories.
(e) Point load test
Breaking strength of solid Result (Is (50)) can be Gives upper bound on rock strength.
rock sample under point correlated with unconfined Joints, defects and the like generally
load. compressive strength. significantly reduce allowable bearing
pressures and increase potential
settlement.
(f) Shrink-swell test
Various tests available. Provides and indication of Used to estimate ground movements in
shrink-swell behaviour of expansive soils.
soil with variation in soil
suction and/or moisture
content.
Field testing and instrumentation
(a) Standpipes and piezometers
Water table level and Establish water tables and water
phreatic surface in a confined pressures.
aquifer.
Investigate seasonal or tidal effects on
water table.
(b) Field instrumentation
Piezometers, settlement Used to observe movements Requires extensive experience to plan
plates, sub-surface of a trial embankment, as and interpret the results. Properly
settlement gauges, earth part of the design process, or conducted monitoring can provide the
pressure cells and to monitor performance of most reliable information about
inclinometers. the final construction. engineering parameters on the site.
(c) Full scale test
e.g. Instrumented trial Used to check design Very useful when other data available is
excavation, trial assumptions, suitability of inconclusive or there are uncertainties.
embankment and trial pile equipment, proposed
Good for trial excavations, especially
installation. construction techniques and
where a soil/weathered rock profile is
the like.
present.
(d) Pile load tests to AS 2159
Load-settlement behaviour of Can be used in advance of main project
pile. to check load settlement behaviour and
ultimate capacity if taken to failure.
Used during project to confirm pile
performance is in accordance with the
specification.
(e) Dynamic pile tests
Using pile driving analyser Used to estimate pile Rapid, high strain testing during
(PDA) and CAPWAP. capacity and dynamic load construction to confirm pile driving
settlement behaviour. parameters and capacity.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 61 DRAFT ONLY

APPENDIX CB
WORKED EXAMPLES

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 62 DRAFT ONLY

Example CB1
Analyse the retaining wall shown in Figure CB1. Unit weight of concrete is 24 kN/m3. The wall runs
parallel to a road and the backfill is compacted by a vibratory roller to cater for surcharge loading in
the future. For the compacted backfill the angle of internal friction  = 32° and bulk unit weight
 = 20 kN/m3.
The base of the retaining wall is 1.5 m below road level. Appropriate drainage is provided for the
backfill and drainage along the road ensures that the ground water table does not rise above 2 m below
road level. Properties of the subsoil are determined by testing soil samples in the laboratory, giving
the effective angle of friction ' = 25°, cohesion c' = 10 kPa and bulk unit weight  = 19.0 kN/m3.

400

C o n t r o l le d , c o m p a c t e d f il l
 = 2 0 k N /m 3 ;  = 3 2 °
3000

1500 400
61°

500
3200
D r a in
S u b s o il p r o p e r tie s
G ro u n d w a te r
 = 1 9 k N /m 3 ; c ' = 1 0 k P a ; ' = 2 5 °

FIGURE CB1 RETAINING WALL IN EXAMPLE CB1

Solution
This is an earth retaining structure on which loads are predominantly imposed by the soil.
Accordingly analyse it in accordance with Clause 7.3.3.
Soil Parameters
The wall is free to deflect adequately to develop active pressure on its back face and passive pressure
in front. The active and passive earth pressures are given by:
pa  K ad CB1(A)
p p  K pd  K pc c ' CB1(B)
In this example Rankine formulas have been used to calculate the active and passive earth pressure
coefficients Ka and Kp [refer e.g. Bowles1] and Kpc = 2 K p . The resulting active and passive
pressures are given in Table CB1. Note that the passive pressure given by Equation CB1(B) is the
maximum that can be developed only if the required deformation of soil occurs, otherwise passive
resistance is developed only to the extent necessary to maintain equilibrium.

1 Bowles, J. E. E., (1997), Foundation Analysis and Design, McGraw-Hill Professional Books.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 63 DRAFT ONLY

In accordance with Clause 13.3.1 the effective ground level for passive resistance is lowered by h,
which is the greater of 0.5 m and 10% of the wall height above ground (3.0 m), hence h = 0.5 m.
Vertical forces due to friction are neglected.

TABLE CB1
SOIL PARAMETERS FOR EXAMPLE CB1
Active Passive Passive
  pressure at pressure at pressure at
Soil Ka Kp bottom top bottom
kN/m3 degrees
kPa kPa kPa
Backfill 20 32 0.307 3.255 27.7
Subsoil 19 25 0.406 2.464 31.4 78.2

Due to compaction of the backfill consider an additional permanent lateral pressure (pc) as proposed
by Ingold1 and shown in Figure C8.2(B). Adopt a value of pc = 20 kPa. Accordingly, specify that the
dynamic roller load (Q) must not exceed:
p c2
Q
2 CB1(2)
20 2 
=  31.4 kN/m
2  20
The depths d1 and d2 in Figure C8.2(B) are given by:
pc
d1  CB1(3A)
K p

pc
d2  CB1(3B)
K a
Substituting the values from Table CB1, we get d1 = 0.307 m and d2 = 3.255 m. As d1 is small, the
pressure due to compaction is taken conservatively as a constant value of 20 kPa from top to depth d2.
The specified backfill should extend for a minimum distance beyond the heel of:
4.1 × tan(45-32/2) = 2.27 m.
The forces acting on the wall and their distances from the toe can be readily calculated and are shown
in Figure CB2(a), with values presented to 3 significant digits.

1 Ingold, T. S., (1979), The effects of compaction on retaining walls, Geotechnique 29, No.3, pp 265-283.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 64 DRAFT ONLY

20 kP a 20 kP a

1085
3 2 .5 k N 3 2 .5 k N
230 kN 230 kN
200

3252
3 9 .4 k N
1800 1400
500

5 1 .7 k N

3015
3415
6 2 .2 k N

3 1 .4 k P a 3 0 .7 k N 27 kN

1367
1600

1500
1000

5 4 .8 k N 2 5 .2 k P a
7 8 .2 k P a 2 7 .7 k P a
1531
429

1931 766

155 kP a 155 kP a

237 kN

(a) Forces for stability (b) Forces for structural design

FIGURE CB2 FORCES ACTING ON RETAINING WALL IN EXAMPLE CB1

Overturning
Check the stability of the wall using the above nominal unfactored values of forces in accordance with
Clauses 7.3.3 and 13.3.
The overturning moment is:
M*ov = 32.5×3.415 + 62.2×1.5 = 205 kN.m
The stabilising moment is:
Mst = 39.4×0.2 + 30.7×1.6 + 230×1.8 + 54.8×0.429 = 494 kN.m
From Table 13.3.1(A) select a value of g = 0.55 as soil properties are determined by laboratory
testing. The design resistance to overturning is:
g Mst = 0.55×494 = 272 kN.m
As g Mst > M*ov this check for overturning is satisfied.
Bearing Pressure
The total load on the foundation is:
W*f = 39.4 + 30.7 +230 = 300 kN
The resultant acts at a distance from the toe given by:
a = (Mst – M*ov) / W*f CB1(4)
= (494 – 205) / 300 = 0.965 m
A rectangular pressure distribution is assumed, for comparison with ultimate bearing capacity as
calculated below. As the pressure resultant must be equal and opposite to Wf, the width of the pressure
diagram is given by:
bf =2a CB1(5)
= 1.931 m

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 65 DRAFT ONLY

The reduction in effective contact width complies with Clause 13.3.3.


The pressure on the foundation is:
q*f = W*f / bf CB1(6)
= 300 / 1.931 = 155 kPa
The horizontal load is:
H*f = 62.2 + 32.5 = 94.7 kN
The angle of inclination of the resultant to the vertical is:

 H*f 
  tan 1  *  degrees
 CB1(7)
W f 
= tan-1(94.7 / 300) = 17.5°
Bearing capacity
Calculate the ultimate bearing capacity of the foundation (qu) from the tested soil properties using a
conservative conventional formula, based on depth of foundation of 1.5 m, effective width bf,
inclination of the load to the vertical  and submerged unit weight of soil below the water table. This
gives:
qu = 397 kPa
From Table 13.3.1(A), adopt g = 0.45. The safe bearing capacity is:
g qu = 0.45×397 = 179 kPa
As g qu > q*f this check for bearing capacity of the foundation is satisfied.
Sliding
For resistance to sliding, assess the base friction. As concrete is poured directly over the compacted
base adopt  = tan' = 0.466. This agrees with the value in Table C10.3.3.4 for  = 25°. For base to
soil adhesion adopt a value of ca = 0.7c' = 7 kPa. This is assumed to act over a contact width of 3a,
based on a triangular pressure distribution at unfactored loads. Hence the nominal resistance to sliding
is:
Hr = W*f + cabc + Pp CB1(8)
= 0.466×300 + 7×3×0.965 + 54.8 = 215 kN
From Table 13.3.1(A) adopt g = 0.55. The design resistance to sliding is:
g Hr = 0.55×215 = 118 kN
The horizontal force causing the wall to slide is H*f = 94.7 kN. As g Hr > H*f this check for sliding is
satisfied.
These checks for a soil retaining structure have been satisfied for the dimensions given in Figure CB1.
Structural Design
According to Clause 7.3.3(f), the ultimate loads for structural design of members must be 1.5 times the
soil imposed action effects, determined simply by factoring all loads by 1.5. It can be seen from
Equation CB1(4) that the value of a is not changed, hence the soil pressure distribution remains the
same at ultimate loads and the ultimate soil pressure is simply 1.5 times its nominal value, i.e.:
q*fu = 1.5×155 = 233 kPa
For the wall as well as for the base slab, the critical section for shear is at the face of the support, as
diagonal cracking can extend into the support. The critical section for bending moment is
0.15×400 = 60 mm inside the face of the support, according to Clause 7.2.10 of AS 5100.5. The

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 66 DRAFT ONLY

relevant nominal loads acting on the wall and their locations are shown in Figure CB2(b). As a load
factor of 1.5 is applied to all loads, ultimate shears and bending moments can be obtained by factoring
the nominal shears and bending moments by 1.5.
Reduction due to passive pressure is ignored, as it may not develop fully for small structural
deformations.
The nominal shear force on the wall is:
Vw = 32.5 + 51.7 = 84.2 kN
The ultimate shear force for design of the wall is:
V*uw = 1.5×84.2 = 126 kN
The nominal bending moment for the wall at the top of the base slab is:
Mw = 32.5×3.015 + 51.7×1.367 = 169 kN.m
The design ultimate bending moment is:
M*uw = 1.5×169 + 126×0.060 = 261 kN.m
For the base slab the nominal shear force is:
Vb = 230 + 27 - 237 = 20 kN
The ultimate shear force for design of the base slab is:
V*ub = 1.5×20 = 30 kN
The nominal bending moment on the base slab at the face of the wall is:
Mb = 230×1.4 + 27×1.4 - 237×0.766 = 178 kN.m
The design ultimate bending moment is:
M*ub = 1.5×178 + 30×0.060 = 269 kN.m
The reinforcement can now be designed for the above design action effects. The requirements for
crack control then have to be checked for the serviceability limit state in accordance with AS 5100.5.
The load factors for the serviceability limit state are obtained from Tables 5.2 and 5.4 of AS 5100.2.
As controlled soil fill has been used, the factors for concrete as well as backfill are 1.0. The
serviceability loads are therefore the same as the nominal loads.
The design serviceability bending moment for the wall is:
Ms.1.w = 169 + 84.2×0.060 = 174 kN.m
The design serviceability bending moment for the base slab is:
Ms.1.b = 178 + 20×0.060 = 179 kN.m
Stresses in the reinforcement can be calculated for these bending moments and further checks carried
out in accordance with AS 5100.5.

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 67 DRAFT ONLY

Example CB2
Analyse the bridge abutment shown in Figure CB3. The backfill comprises granular fill with a unit
weight of 20 kN/m3 and an angle of internal friction of 30°. The foundation is on weathered rock, for
which laboratory test gave characteristic values of 20 kN/m3 for the unit weight, 25 for the angle of
internal friction and 25 kPa for cohesion. The unit weight of concrete is 25 kN/m3. The bridge is not
over a watercourse and the backfill as well as the subsoil is adequately drained.
The abutment supports a bridge deck imposing a dead load of 100 kN per metre length of wall. The
vertical live load from the deck is 241 kN per metre length of wall including the appropriate value of
dynamic load allowance to be applied at footing level. The live loads may act either on the deck or the
backfill or on both. The girder is placed after the backfill is placed and compacted. For the purpose of
this worked example other types of loads e.g. braking forces, are not considered.

200 m m
G IR D E R

210
850
1200

C L B E A R IN G
25°

400
1500

1000 700 1500

FIGURE CB3 BRIDGE ABUTMENT IN EXAMPLE CB2

Solution
The abutment functions as a foundation for the bridge deck as well as a soil supporting structure. It
must therefore be analysed in accordance with Clause 7.3.2 as well as Clause 7.3.3. Analysis as a soil
supporting structure is presented first.
The design load (S*) must not exceed the factored geotechnical resistance (gRug) in accordance with
Equation 10.3.3.3. Several load cases with appropriately factored or unfactored loads must be
considered. For brevity only the most critical loading condition for each geotechnical effect is
presented in the calculations below.
Loads
The forces acting on the wall and their locations are shown in Figure CB4. The dead load of the deck
and vertical live load act along the centre-line of the bearing.
Live load on the approach embankment causes a surcharge load on the backfill. In accordance with
Clause 13.2 of AS 5100.2 for depths up to 3 m an equivalent height of fill of 1 m should be
considered. The live load surcharge pressure is therefore 20 kPa. The surcharge can act either as load

389944842.doc - 15/06/2018 7:26:58


DRAFT ONLY 68 DRAFT ONLY

Wq1 above the heel or as load Wq2 behind the soil envelope. Due to load Wq1 both vertical and lateral
forces are considered to act on the structure whereas due to load Wq2 only lateral pressure is considered
and vertical pressure on the heel due to dispersion of the load is conservatively neglected.
The vertical loads and their moments about the toe are calculated and summarised in Table CB2.

TABLE CB2
VERTICAL LOADS AND MOMENTS ON THE ABUTMENT
Item Nominal Nominal
Force Moment
about toe
(kN)
(kN.m)
Self weight of concrete structure W c = 62 M c = 92
Dead load of deck W g = 100 M g = 121
Weight of soil on heel W e1 = 69 M e1 = 169
Weight of soil on toe W e2 = 7 M e2 = 4
Surcharge load W q = 30 M qw = 74
Live load of vehicle W v = 241 M v = 292

Although the design of the girder bearings and transmission of longitudinal forces at bearing level is
not within the scope of this example, allowance is made for propping action of the deck by designing
for at-rest pressure at the back of the wall, using earth pressure coefficient Ko = 0.5.
For bearing capacity of the weathered rock foundation and resistance to sliding the soil parameters
adopted are ' = 25° and c' = 25 kPa. The ground in front of the wall is excavated and backfilled with
the granular fill hence the horizontal component of the passive earth pressure coefficient, computed
using the Coulomb equation with  = 30°, c = 0 and slope of the ground  = -25° is Kp = 1.12. In the
same manner as required in Clause 13.3.1 and in accordance with the intent of Clauses 10.1 and
10.3.3.4, the effective ground level is taken lower than the nominal level by h, where h is the
greater of 10% of 1.2 m or 0.5 m, hence h = 0.5 m.
The lateral pressures calculated as above are shown in Figure CB4. Horizontal force resultants and
their moments about the toe are calculated and summarised in Table CB3.

TABLE CB3
FORCES DUE TO EARTH PRESSURE
Item Nominal Nominal
Force Moment
(kN) (kN.m)
At-rest pressure resultant F o = 36 M o = 33
Surcharge pressure resultant F q = 27 M qf = 37
Passive pressure resultant Fp = 3 Mp = 1

All resultant forces and their distances from the toe are shown in Figure CB4.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 69 DRAFT ONLY

2450
W q1
W q2
1210
W g W v
606 W e1

Fq

W e2 F o
W c
500
h

1350
178

900
Fp 1497
534

12 kP a 27 kP a 10 kP a
b f

*
q *f W f q *f
a a

FIGURE CB4 FORCES ON ABUTMENT IN EXAMPLE CB2

Solution A – Soil-supporting structure


The abutment is designed as a retaining wall in accordance with Clause 13. Therefore the structure
must be analysed with the nominal loads in Tables CB2 and CB3 applied unfactored.
Overturning
Check the stability of the wall using the above nominal unfactored values of forces in accordance with
Clauses 7.3.3 and 13.3. The critical loading condition is with no live load on the deck and surcharge
load Wq2.
The overturning moment is:
M*ov = 33 + 37 = 70 kN.m
The stabilising moment is:
Mst = 92 + 121 +169 + 4 + 1 = 387 kN.m
From Table 13.3.1(A) select a value of g = 0.55 as soil properties are determined by laboratory
testing. The design resistance to overturning is:
g Mst = 0.55×387 = 213 kN.m
As g Mst > M*ov this check for overturning is satisfied.
Sliding
The critical loading condition is with no live load on the deck and surcharge load Wq2. The horizontal
force causing the wall to slide is:
H*f = 36 + 27 = 63 kN
The total load is:
W*f = 62 + 100 + 69 + 7 = 238 kN

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 70 DRAFT ONLY

Distance of the resultant from the toe is:


a = (387 – 70) / 238 = 1.332 m
For resistance to sliding assess the base friction. As the concrete is poured on a rough finished
mudmat over the weathered rock, adopt  = tan = 0.466. For base to soil adhesion adopt a value of
ca = 0.7c' = 17.5 kPa. For a linearly varying soil pressure distribution the contact width is
3a = 3.996 m. The nominal resistance to sliding, given by Equation CB1(8), is:
Hr = 0.466×238 + 17.5×3.996 + 3 = 184 kN
From Table 13.3.1(A) adopt g = 0.55. The design resistance to sliding is:
gHr = 0.55×184 = 101 kN
As gHr > H*f this check for sliding is satisfied.
Bearing pressure
Maximum bearing pressure occurs with live load on the deck and surcharge load Wq1. We have:
M*ov = 33 + 37 = 70 kN.m
Mst = 92 + 121 + 169 + 4 + 74 + 292 + 1 = 753 kN.m
*
Wf = 62 + 100 + 69 + 7 + 30 + 241 = 509 kN.
The soil pressure due to the vertical and horizontal forces and the angle of inclination of the resultant
to the vertical are calculated by means of Equations CB1(4) to CB1(7). We have:
a = (753 – 70) / 509 = 1.342 m
bf = 2 × 1.342 = 2.684 m
*
qf = 509 / 2.684 = 190 kPa
 = tan-1(63 / 509) = 7.1°
Bearing capacity
From Table 13.3.1(A), select g = 0.47. The ultimate bearing capacity is calculated by conservative
conventional formula, with N-factors reduced for downward slope of soil surface [Bowles 1] and other
factors applied in a similar manner to Example CB1, to be :
qu = 822 kPa
From Table 13.3.1(A), adopt g = 0.47. The safe bearing capacity is:
gqu = 0.47×822 = 386 kPa
As gqu > q*f this check for bearing capacity of the foundation is satisfied.
The above checks for a soil retaining structure have been satisfied.
Stability During Construction
The stability of the wall during construction should also be checked. At that stage it acts as a
cantilever wall subjected to earth and compaction pressures in a similar manner to Example CB1.
Relevant g factors for temporary structures can be adopted from Table 13.3.1(A) for this check.
Structural Design
For designing the structural members for strength multiply all soil imposed action effects (Se) by 1.5 in
accordance with Clause 7.3.3. Design each structural member to satisfy Equation 7.3.3(2). As an
illustration the force on the toe outstand is calculated.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 71 DRAFT ONLY

The unfactored soil pressure is q*f = 190 kPa and the weight of soil above the toe is 7 kN, acting at
394 mm from the face of the wall. The factored shear force at the face of the wall is:
V*A = 1.5×(190×1.0 - 7) = 275 kN
The factored bending moment at the face of the wall is:
M*A = 1.5×(190×1.0×0.5× – 7×0.394) = 138 kN.m
For designing the reinforcement, calculate modified values at the critical sections for shear and
bending moment and proceed in accordance with AS 5100.5.
Solution B – Foundation
The abutment is to be designed as a footing in accordance with Clause 10.
The value of g must be selected from Table 10.3.3(A) and the same value is applicable to sliding,
overturning and bearing pressure. Corresponding to the value of 0.47 selected for bearing pressure in
Table 13.3.1(A), g = 0.56 is interpolated from Table 10.3.3(A).
Load factors are applied from the relevant tables in AS 5100.2 as summarised in Table CB4.

TABLE CB4
LOAD FACTORS
Ultimate load factor Serviceability
Load or material
Reduces load Increases load load factor

Abutment self weight 0.85 1.2 1.0


Deck dead load 0.85 1.2 1.0
Foundation subsoil 0.70 1.5 1.2
Soil fill (controlled) 0.85 1.25 1.0
Embankment surcharge 0 1.5 1.0
Vehicle load on deck 0 1.8 1.0

Note that load factors are specified for the unit weights of materials (whether the density is more or
less than the nominal value adopted) and all effects must be calculated consistently for the factored
loads (refer also to Clause J7.1.1 of AS 4678-20021) so that equilibrium is maintained. In some cases
it may not be readily apparent whether a factor which reduces or increases the load should be selected
to obtain the most critical condition e.g. increasing the density of the backfill increases the lateral
pressure and the overturning moment, but the increased load also increases the stabilising moment.
Both alternatives may have to be investigated to achieve a conforming design. For brevity, only the
most critical condition with the lowest value of the ratio gRug/S* is presented below. Note also that
passive pressure and base friction are not loads but resistances that are mobilised only to the extent
necessary to maintain equilibrium.
Overturning
Consider surcharge Wq2 on the embankment with no live load on the deck.. Reduced load of backfill is
found to be more critical than increased load. The overturning and stabilising moments about the toe
therefore are, respectively:
M*ov = 0.85Mo + 1.5Mqf
= 0.85×33 + 1.5×37 = 83 kN.m

1 AS 4678-2002: Earth retaining structures, Standards Australia

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 72 DRAFT ONLY

Mst = 0.85Mc + 0.85Mg + 0.85Me1 + 0.85Me2 + 0.85Mp


= 0.85×92 + 0.85×121 + 0.85×169 + 0.85×4 + 0.85×1 = 329 kN.m
The design resistance to overturning is:
gMst = 0.56×329 = 184 kN.m
As gMst > M*ov this check for overturning is satisfied.
Sliding
The critical condition is with no live load on deck and surcharge load Wq2. However, the soil load
factors are different from the overturning check. The horizontal force causing the wall to slide is:
H*f = 1.25Fo + 1.5Fq
= 1.25×36 + 1.5×27 = 86 kN
The total load is:
W*f = 0.85Wc + 0.85Wg + 1.25We1 + 1.25We2
= 0.85×62 + 0.85×100 + 1.25×69 +1.25×7 = 233 kN
The overturning moment and stabilising moments are, respectively:
M*ov = 1.25Mo + 1.5Mqf
= 1.25×33 + 1.5×37 = 97 kN.m
Mst = 0.85Mc + 0.85Mg + 1.25Me1 + 1.25Me2 + 0.85Mp
= 0.85×92 + 0.85×121 + 1.25×169 + 1.25×4 + 0.85×1 = 398 kN.m
Hence the distance of the resultant from the toe is:
a = (398 – 97) / 233 = 1.292 m
The coefficient of friction is  = 0.466 and the adhesion is ca = 17.5 kPa, as for Solution A. For a
rectangular soil pressure distribution at ultimate load the contact width is:
bc = 2a
= 2×1.292 = 2.584 m
The resistance to sliding is:
Hr = W*f + cabc + 0.85Fp
= 0.466×233 + 17.5×2.584 + 0.85×3 = 155 kN
The design resistance to sliding is:
gHr = 0.56×155 = 87 kN
As gHr > H*f this check for sliding is satisfied.
Bearing pressure
For computing soil pressures on the foundation, the live load from the deck as well as the backfill
(Wq1) is included and all loads are factored up. We have:
M*ov = 1.25Mo + 1.5Mqf
= 1.25×33 + 1.5×37 = 97 kN.m
Mst = 1.2Mc + 1.2Mg + 1.25Me1 + 1.25Me2 + 1.5Mqw + 1.8Mv + 0.85Mp
= 1.2×92 + 1.2×121 + 1.25×169 + 1.25×4 + 1.5×74 + 1.8×292 + 0.85×1 = 1109 kN.m

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 73 DRAFT ONLY

W*f = 1.2Wc + 1.2Wg + 1.25We1 + 1.25We2 + 1.5Wq + 1.8Wv


= 1.2×62 + 1.2×100 + 1.25×69 + 1.25×7 + 1.5×30 + 1.8×241 = 768 kN
The soil pressure due to the vertical and horizontal forces and the angle of inclination of the resultant
to the vertical are calculated by means of Equations CB1(4) to CB1(7) as follows:
a = (1109 – 96) / 768 = 1.319 m
bf = 2 × 1.319 = 2.638 m
*
q fu = 768 / 2.638 = 291 kPa
 = tan-1(86 / 768) = 6.39°
The ultimate bearing capacity is calculated as previously but with the unit weight of foundation
subsoil factored by 0.7. The ultimate bearing capacity is:
qu = 739 kPa
The safe bearing capacity is:
gqu = 0.56×739 = 414 kPa
As gqu > q*fu this check for bearing capacity of the foundation is satisfied.
Structural Design
In accordance with Clause 8.3.2(a) the load factors from AS 5100.2 which are summarised in
Table CB4 are applied for calculating the design ultimate shear force and bending moment. The
maximum soil pressure is q*fu = 291 kPa as above. The nominal load of controlled fill above the toe is
7 kN, acting at 394 mm from the face of the wall, which should be factored by 0.85. A factor of 1.25
was applied to soil loads in the calculation for q*fu; but the reduction in load factor for this small load
does not affect q*fu significantly. The factored shear force at the face of the wall is:
V*B = 291×1.0 – 0.85×7 = 285 kN
The factored bending moment at the face of the wall is:
M*B = 291×1.0×0.5 – 0.85×7×0.394 = 143 kN.m
For designing the reinforcement, calculate modified values at the critical sections for shear and
bending moment and proceed in accordance with AS 5100.5.
Shear force and bending moment at serviceability may be calculated similarly.
Note that the contribution of passive pressure to the overall resistance term gRug is negligible in all
cases. No translation or rotation of the structure for development of passive resistance is expected.
Comparison of Solutions A and B
The ratios of factored resistance to strength (gRug/S*) obtained in Solution A (soil supporting
structure) and Solution B (foundation), as well as the design shear force and bending moment on the
toe slab, are compared in Table CB5.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 74 DRAFT ONLY

TABLE CB5
COMPARISON OF RESULTS
Bearing Moment on
Solution Overturning Sliding Shear on toe
pressure toe
(A) Soil g Mst / M*ov g Hr / H*f g qu / q*f V*A M*A
supporting
structure = 3.04 = 1.60 = 2.03 = 275 kN = 138 kN.m

(B) g Mst / M*ov g Hr / H*f g qu / q*fu V*B M*B


Foundation = 2.22 = 1.01 = 1.42 = 285 kN = 143 kN.m

It is concluded that the dimensions of the structure shown in Figure CB3 are adequate to satisfy the
checks made for safety and stability for the values of g assumed. Solution B is more conservative,
primarily because although the nominal loads are factored by 1.25 to 1.8, the geotechnical reduction
factor is only increased from 0.47 to 0.56 for bearing pressure and from 0.55 to 0.56 for sliding and
overturning. The ultimate shear force and bending moment on the toe outstand for the two solutions
are close, which indicates that the average load factor for Solution B is slightly more than 1.5.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 75 DRAFT ONLY

Example CB3
An anchored sheet pile is provided to retain a 7.9 m deep excavation, as shown in Figure CB5. The
ground water table measured at the time of site investigation was 2 m below ground level but for
design it is conservatively taken at 1 m below ground level. The ground water level inside the
excavation is kept at 1.1 m below the proposed excavation level. A surcharge pressure of 20 kPa is
applied at ground level. Characteristic values of the soil properties are derived from appropriate
advanced laboratory testing, local historic records and engineering experience. The sheet pile wall is
required to function as a temporary structure with a service life of not more than 2 years. Calculate the
required depth of embedment and evaluate the design shear and bending moment for the design of the
sheet pile wall. Determine the design force in the anchor.

2 0 k P a S u rc h a rg e G r o u n d le v e l

G r o u n d w a te r le v e l
2 .0 m

 = 32° ; c = 0
4 .0 m

  = 1 9 k N /m 3

1 .0 m
E x c a v a t io n

20°
7 .9 m

D r e d g e le v e l
A nchor
1 .1 m

 = 26° ; c = 7 kPa
 = 2 0 k N /m 3
Pum ped
w a te r le v e l
S h e e t p ile

FIGURE CB5 ANCHORED SHEET PILE WALL IN EXAMPLE CB3

Solutions
The common methods of analysis for this type of problems can be grouped broadly into three
categories which are, in increasing complexity of analysis:
 Limit equilibrium method (LEM);
 Subgrade reaction and/or pseudo finite element method (PFEM); and
 Finite element (FEM) and/or finite difference method (FDM).
The above problem is solved using a method from each category to demonstrate in general terms the
application of the Standard. It is not the intention to compare the pros and cons of each of these
methods, or to evaluate the commercial software packages used. The first example below uses the
LEM method adopting the conventional CP2 geotechnical design approach 1 for solution. Application
of the Standard's requirements is not possible for the FEM using the conventional method. For this
reason the proposed solutions utilise a new logical method to demonstrate design conformance to the
Standard. The principal differences and benefits of more advanced methods of analysis will become
self evident.

1 Civil Engineering Code of Practice No.2 (1951), Earth Retaining Structures, Institution of
Structural Engineers.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 76 DRAFT ONLY

Only effective stress analysis is undertaken. In some soil conditions total stress analysis may become
critical and will also need to be examined. In the example below only one design scenario is
considered. For detailed design it is the designer’s responsibility to consider all reasonably expected
design load cases and combinations. Optimisation of structural members, which requires iteration of
the calculations, has not been attempted.
For deep excavations with high ground water levels, control of base heave and hydraulic stability
could be critical and designers need to consider these effects separately.
For the LEM and PFEM solutions the water pressure profile is calculated at steady state using a
simplified linear ground water seepage approach. Although the water levels are different on both sides
of the pile the water pressure at the bottom of the sheet pile must have a common value (uo), which is
calculated by means of an approximate formula given by Gaba et al1:
 hw 
uo   w d w 1   CB3(1)
 2d w  hw 
where w is the unit weight of water, dw is the depth of water in the excavation and hw is the difference
in water levels, as shown in Figure CB6. The water pressure on the retained earth side (ue) and on the
dredged earth side (ud) are then assumed to vary linearly from 0 at the water surface to uo at the
bottom.
1 Solution by the Limit Equilibrium Method
1.1 Design Methodology
Analysis of sheet piles by the simplified limit equilibrium method (LEM) does not take into
consideration the interaction between the sheet pile and the soil, the type and flexibility of the anchor
and the construction stages, all of which can have a significant effect on the results. Simplified
distributions of active and passive pressure are assumed, based on laboratory tests and site
measurements conducted by several investigators, some of which have been summarised by
Tschebotarioff in the book edited by Leonards2.
In the "free earth support method" applicable to anchored sheet piles, active pressure is applied on the
retained earth side and passive pressure on the excavated side below the dredge level. In classical
methods explained in textbooks e.g. DAS 3 the depth of penetration below the dredge level (D) is
obtained by solving a polynomial equation derived for piles driven in uniformly cohesive or
cohesionless soil strata. For safety, either the passive pressure coefficient (Kp) is reduced or the
calculated depth of penetration is arbitrarily increased. Neither option is quite satisfactory, as will be
demonstrated later.
The problem can also be formulated in a generalised manner for stratified soils with varying c-
properties (e.g. the "free earth support – general case" solution indicated in the US Navy Manual 4) and
solved by trial and error. Iterative solutions are facilitated by algorithms provided in modern
spreadsheet software. The method presented below complies with Clause 7.3.3 as the sheet pile wall
is a soil supporting structure.
The loads and action effects (Se) comprise the surcharge load (q), the water pressure (u) the active soil
pressure (pa) on the retained earth side and the unknown horizontal component of anchor force (T). In
accordance with Clause 7.3.3(b) and Clause 8.3.3 these loads must be applied with a load factor of 1.0.

1 Gaba, A. R., Simpson, B., Powrie, W., Beadman, D. R., (2003), Embedded Retaining Walls—
Guidance for Economic Design. Construction Industry Research and Information Association , London, CIRIA
C580.
2 Leonards, G. A., Editor, (1962), Foundation Engineering, McGraw-Hill Book Company.
3 Das, B. M., (1990), Principles of Foundation Engineering, PWS-KENT Publishing Company, Boston.
4 Naval Facilities Engineering Command (1986), Design Manual 7.02 – Foundations & Earth
Structures, (NOTE: Available as a free download from the author's website).

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 77 DRAFT ONLY

Due to the presence of ground water the active pressure (p'a) and passive pressure (p'p) must be based
on the effective vertical stress at any level, given by:
' =  – u CB3(2)
where  is the vertical pressure at that level based on bulk unit weight of soil and u is the linearly
interpolated water pressure. All variables are unfactored. Surcharge pressure at ground level, if any,
must also be applied unfactored.
In this example the Rankine equations are used to calculate the active and passive pressures. The
active pressure coefficient (Ka) and passive pressure coefficient (Kp) are calculated from characteristic
values of soil properties determined from laboratory tests.
The total pressure acting on the retained earth side is:
pe  pa  ue CB3(3A)

  ' K a  2c K a  u e CB3(3B)
In accordance with Clause 7.3.3(e), the passive pressure must be factored by the geotechnical strength
reduction factor. Its value is obtained from Table 13.3.1(A) as the sheet pile behaves as a retaining
wall for the purpose of calculating passive pressure. The total pressure acting below the dredge line is:
pd   g p p  u d CB3(4A)


  g  ' K p  2c K p  u d  CB3(4B)
The equations of equilibrium can now be formulated. For a single anchor, it is convenient to take
moments about the anchor point in order to eliminate the anchor force from the moment calculations.
Let the total effective active force be P'a, the total effective passive force be P'p, the total water load on
the retained earth side be Ue and the total water load on the dredged side be Ud. Let these forces act at
depths ya, yp, ye and yd respectively below the anchor. Note that P'p = Rug, the geotechnical resistance.
In accordance with Equation 7.3.3(1) the requirement for moments is:
 g Pp y p  Pa ya  U e ye  U d yd CB3(5)
For any depth of penetration (D) the passive resistance must be adequate to satisfy equilibrium of
moments, hence to satisfy the equality in Equation CB3(5) calculate:
Pa y a  U e ye  U d y d
 g*  CB3(6)
Pp y p
For solution, the pressures are calculated for an assumed depth D, *g is calculated from
Equation CB3(6) and the depth iterated until *g = g, the value required in accordance with the
Standard.
The force in the anchor is given by:
T  Pa  U e  U d   g* P ' p CB3(7)
The shear force and bending moment diagram can now be obtained. Further design is illustrated in the
calculations that follow.
1.2 Calculations
For this worked example the soil stratum is divided into layers such that soil properties and other
physical variables within each layer are constant. The active and passive pressure coefficients within
each layer are calculated from conventional conservative equations.
For application of passive pressure the effective dredge level is lowered in accordance with
Clause 13.3.1 by h, which is the greater of 0.5 m or 10% of 5.9 m i.e. 0.6 m.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 78 DRAFT ONLY

2 0 k P a S u rc h a rg e
6 .1 5 GW L
1

1 1
12
T 2 1 5 .6
 = 2 0 °

2 .0
3 CL a 2 2 .7
nch 2 0 .1
or h w
4

3 .9
0 .6
h 3 9 .6
5 3 .1 5 4 2 .6
W L 6
7 8 .8 4 5 .1
0 .5

D = 1 2 .5 3

d w

Layer 7
ud p 'p p 'a ue

141 303 102 141


uo uo
W a te r U n fa c to re d A c t iv e W a te r
p re s s u re p a s s iv e p r e s s u r e p re s s u re p re s s u re
D im e n s io n s in m ; p r e s s u r e s in k P a

FIGURE CB6 PRESSURES ON SHEET PILE

The geotechnical strength reduction factor ( g) is obtained from Table 13.3.1(A) of AS 5100.3. Based
on the criteria in Table 13.3.1(B), for a temporary structure designed by this approximate method, the
required value of g is 0.55.
Calculation are carried out by assuming a depth of penetration (D) and calculating pressures by means
of Equations CB3(1) to CB3(4). Resultant forces and moments are calculated and the value of *g
required for moment equilibrium is calculated from Equation CB3(6). When *g = 0.55 the minimum
required depth of penetration is found to be 12.53 m. The pressure diagrams for D = 12.53 m are
shown in Figure CB6.
The horizontal force at anchor level is, from Equation CB3(7):
T = 1065 + 1371 – 807 – 0.55×2212 = 412 kN.
A "critical depth" of penetration can be computed with g = 1.0, implying no safety factor. The
"critical depth" is 5.92 m and corresponding T = 252 kN. The substantial increase in depth and anchor
resistance required for the specified value of g indicates the risk associated with increasing the depth
arbitrarily (as advocated in some references) to provide an unquantified safety factor. To provide a
safety factor of 1/0.55 = 1.82 on calculated passive resistance the depth of penetration had to be
increased by a factor of 12.53/5.92 = 2.12. Also note from Equation CB3(4) that dividing Kp by a
factor of safety which is recommended in some references does not give a safety factor equivalent to
1/g.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 79 DRAFT ONLY

The shear force and bending moment on the pile at any section are computed by taking moments of
forces either above or below the section. For this problem a standard structural analysis program was
used, with the sheet pile modelled as a member with one end fixed and the other end free. Loads were
applied in accordance with Equations CB3(3A) and CB3(4A) for the pressures shown in Figure CB6
and with g = 0.55 and T = 412 kN. The result must give zero values of shear force and bending
moment at the fixed end in the model, which provides a useful check on the correctness of the
calculations. Note that this condition will not be satisfied if the depth is altered for any reason, unless
revised pressures and g are calculated by means of Equations CB3(1) to CB3(7). The computed shear
force and bending moment diagrams are plotted in Figure CB7.

1
1 1

T 2 386
-2 1
-2 6
3
2 .0

318 690

4
3 .9

0 .6
62
5 1512
6
0 .5
1 2 .5 3

-1 9 1 715

Layer 7
5 .7

S h e a r fo rc e , k N B e n d in g m o m e n t, k N .m

FIGURE CB7 RESULTS FOR LIMIT EQUILIBRIUM METHOD OF EXAMPLE CB3

The shear force and bending moment for structural strength design must be obtained by applying an
overall load factor of 1.5 to the design action effects in accordance with Clause 7.3.3(f). A moment
reduction factor () recommended in some references to account for the flexibility of the sheet pile
may be applied if relevant. The pile section must therefore be designed for a bending moment of:
M*u = 1.5  1512 ×  = 2268× kN.m per 1 m width of wall
Assume that anchors are provided at 3 m centres. The ultimate horizontal design force at anchor level
is:
T* = 1.5  3.0  412 = 1854 kN
Let the tie rod be inclined at an angle  = 20° to the horizontal. For strength design of the tie rod,
s = 0.9 from AS 5100.6 . From Table 12.3.3(A) of AS 5100.3, the Importance Factor (n) is 0.93.
Hence its area As and yield strength fy must be chosen to satisfy the condition:
s n As fy  1854 × sec20° = 1973 kN
For geotechnical design of the anchor select a value of ga = 0.5 from Table 12.3.3(B). The nominal
force in the tie rod is:
S* = 3.0 × 412 × sec20° = 1315 kN

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 80 DRAFT ONLY

The anchor must develop adequate geotechnical resistance (Ra) such as nominal passive resistance of a
deadman or nominal bond strength of the grout to satisfy the condition:
ga n Ra ≥ 1315 kN
The preliminary design of the anchor and sheet pile sections obtained above may be assessed for the
effects of soil-structure interaction, anchor deformation etc, by using software programs based on
finite difference and finite element methods and can serve as approximate checks on the results of
computer generated solutions.
2 Pseudo Finite Element Method (PFEM)
In the calculations the following assumptions have been made:
 At-rest earth pressure coefficient = 0.7;
 Lateral earth pressure coefficient derived following the Eurocode 7 method;
 Interface wall friction ratio:
 Active side = 0.67;
 Passive side = 0.50;
 Young’s modulus of the top stratum = 20 MPa;
 Young’s modulus of the bottom stratum = 35 MPa ;
 Wall stiffness = 80 360 kN.m2/m;
 Support stiffness = 125 000 kN/m/m;
 Groundwater pressure distribution similar to LEM i.e. linear groundwater seepage, equilibrium
at wall toe;
 Excavation levels:
 Stage 1 to 3.3 m below ground level, targeting 0.8 m below prop level including 10%
over excavation (Cl.13.3.1) for the ultimate limit state;
 Stage 2 to 8.5 m below ground level, targeting 7.9 m below top level including 0.6 m over
excavation (CL.13.3.1) for the ultimate limit state;
 Width of the excavation is 20 m;
Analysis is carried out using a commercial retaining wall analysis program.
The construction staging adopted for this example is as follows:
Stage 1: Apply traffic surcharge on the active side at ground level;
Stage 2: Apply water pressure profiles due to dewatering on either side of the wall;
Stage 3: Excavate to 3.3 m below ground level on the passive side, i.e. 1.3 m below the support level;
Stage 4: Install support at 2 m below ground level;
Stage 5: Excavate to 8.5 m on the passive side below ground level.
The structure is analysed as a soil-supporting structure. All loads and load combinations used in the
design should therefore be unfactored to determine the design action effects in accordance with
Clause 7.3.3. Soil loads are derived from the characteristic soil parameters.
The planned excavation below the support level to permit support construction is arbitrarily taken to
be 0.8 m.
As the stability of the excavation relies on passive soil support, the Standard requires an over-
excavation depth to be incorporated into the design (Clause 13.3.1). Cantilevered walls are considered
differently from supported walls for the over-excavation allowance. The first stage excavation results
in a cantilevered wall with a planned depth of excavation of 2.8 m, and over-excavation allowance h
being the larger of 0.3 m and a minimum 0.5 m. The second stage excavation results in a supported
wall with the distance between the prop and final excavation being 5.9 m, and over-excavation

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 81 DRAFT ONLY

allowance h being the larger of 0.6 m and a minimum 0.5 m. The complying design excavation
depths for the first and second excavation stages are therefore 3.3 m and 8.5 m below ground level.
Having derived the design model, the geotechnical design for lateral toe stability comprises the
following:
Step 1 Determine the minimum toe embedment for critical equilibrium
Step 2 Determine the theoretical geotechnical resistance at minimum toe embedment
Step 3 Determine the geotechnical strength reduction factor g
Step 4 Determine the required geotechnical resistance and revised toe embedment
Step 5 Determine structure loads using the revised toe embedment
The background to this design approach is based on references by Yuen et al 1,2
Step 1 Determine the minimum toe embedment for equilibrium
For a dry excavation or for excavation without dewatering i.e. dredging, determination of the toe
embedment for critical equilibrium is straight-forward. For excavation with dewatering, a few
iterations are required to account for the changes in water pressure distribution as the wall embedment
is varied.
The minimum toe embedment of the sheetpile wall below the second stage excavation level is
calculated to be 4.1 m, resulting in an overall sheetpile length of 12.6 m between ground level and
wall toe.
Step 2 Determine the theoretical geotechnical resistance at minimum toe embedment
Two different types of soil support exist for lateral wall toe stability; fixed earth and free earth. For
simply supported walls at critical equilibrium, toe stability is attained by mobilising limiting passive
soil resistance, which must be analysed for each excavation stage. In the first stage the excavation
depth is shallow, the toe penetration is relatively deep and the problem will not be critical for stability.
Critical stability invariably occurs close to the end of excavation, usually at the last or second last
excavation stage.
For this example, the cantilever wall embedment is 9.3 m at the end of the first stage excavation, well
in excess of that required for stability provided the structure can withstand the soil imposed loads. At
the end of the second stage excavation at a depth of 8.5 m, support is provided by the support at the
top, and by soil resistance at the toe. At critical stability full passive resistance is mobilised.
For c- soils the limiting passive soil pressure is:
p ' p  p ' v K p  c' K pc CB3(8)
where
p'v = vertical effective stress
c' = effective cohesion
Kp, Kpc = passive earth pressure coefficients
Only a single soil layer is penetrated by the toe in this example and one set of earth pressure
coefficients is sufficient to determine the lateral earth pressure profile. Where several soil layers are
penetrated by the toe, the coefficients for each layer should be derived for determining the pressure
distribution.

1 Yuen, C. K. S., (in preparation, 2007), A New Logical Approach to Geotechnical Design of Single-
Propped Cantilever Walls: Background and Theory.
2 Yuen, C. K. S., Forster, G. J. & Bhavnagri, V. S., (in preparation, 2007), A New Logical Approach to
Geotechnical Design of Single-Propped Cantilever Walls: Design Application to Australian Standard AS 5100.3.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 82 DRAFT ONLY

The total lateral resistance P'p is calculated by integrating the pressure distribution over the toe
penetration with the limiting resistance force calculated to be 418 kN/m, which can be calculated using
simple algorithms.
The 418 kN/m force is the action effect (Se) specified in the Standard. For geotechnical design only,
the design action effect (S*) is also the action effect, Se, as a load factor of 1.0 is applicable i.e. for the
geotechnical design only, the following relationship applies:
S* = 1.0 Se = Se CB3(9)
Step 3 Determine the geotechnical strength reduction factor Øg
The geotechnical strength reduction factor for retaining walls is determined based on the method of
assessment with ranges for these values in Table 13.3.1(A) and guidance in Table 13.3.1(B) of
AS 5100.3. Because the soil parameters are determined from advanced laboratory tests, the
geotechnical strength reduction factor for global stability of this temporary sheet pile wall is between
0.55 and 0.65, with the lower value of g = 0.55 selected for the purpose of this worked example.

Step 4 Determine the required geotechnical resistance and revised toe embedment
Clause 7.3.3 requires the ultimate geotechnical strength to be not less than the factored design action
effect. From the calculations in Steps 2 and 3 above, the required minimum design ultimate
geotechnical strength is:
Rug = S* / g CB3(10)
= 418 / 0.55 = 760 kN/m.
The minimum design toe embedment is determined using similar procedures described in Step 2, with
the minimum passive resistance from the passive side of the excavation summed over the revised toe
penetration being more than 760 kN/m.
The revised toe penetration is estimated to be 5.9 m, at which the computed effective passive soil force
is 761 kN/m, satisfying the required minimum passive resistance of 760 kN/m.
The overall sheet pile length required is 14.4 m
At this stage, the other action effects (Se) for the supports and the sheet piles are as follows:
Support force = 385 kN/m
Sheet pile moment = 214 kN.m/m
Sheet pile shear = 200 kN/m
If the calculation involves dewatering below the level of the excavation then the piezometric pressure
distribution on both sides of the wall needs to be updated before the calculation, with a limited number
of iterations usually being sufficient. Overlooking this could lead to under estimation of the required
toe penetration and support load as the increase in toe penetration leads to an increase in pore water
pressure on the active side i.e. higher active load, and a decrease in effective soil resistance on the
passive side.

Step 5 Determine the structure loads using the revised toe embedment
The geotechnical calculation is complete for the load case considered once the final toe penetration
depth is determined at the end of Step 4. The calculated structure loads at final toe penetration
correspond to the action effects imposed through soil (Se). For structural design the design action
effects are given by:
S* = 1.5 Se CB3(11)
The design action effects for the singly propped sheet pile retaining wall are as follows:
Support force = 385×1.5 = 578 kN/m

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 83 DRAFT ONLY

Sheet pile moment = 214×1.5 = 322 kN.m/m


Sheet pile shear = 200×1.5 = 300 kN/m
In this worked example both the effects from ground water and excavation are probably the worst case
scenario. For serviceability limit state deformation estimates, a separate calculation should be
undertaken e.g. the design ground water level may be probably at 1.5 m below ground level instead of
2 m with the design excavation limited to the planned depths only.
3 Finite Element Method (FEM)
Analysis is made using a commercial general purpose 2-D finite element program written for
geotechnical applications. This kind of calculation represents a most sophisticated numerical method
with many complex material models available. Users must understand the complexity of this method
and the underlying assumptions and limitations associated with each individual program.
Most contemporary commercial geotechnical packages recognise the significance of construction
sequence on structure loads and deformations, and are capable of modelling the construction staging.
An identical construction staging as in the PFEM was used.
The design considerations are similar to that of the PFEM discussed above. For simplicity reasons
they are not repeated.
The calculations used a coupled steady state seepage and stress analysis. Unlike PFEM, assumptions
on distribution of pore water pressures are not necessary but the distribution is calculated using
additional knowledge on permeability properties of the materials.
Once the geotechnical design model is assembled and checked, the geotechnical design for lateral toe
stability complying with the Standard could follow the same steps as for the PFEM.
Step 1 Determine the minimum toe embedment for equilibrium
The design process is similar to the PFEM in which the sheet pile penetration in the model is
incrementally adjusted until reaching the critical state to determine the minimum embedment depth.
The minimum toe embedment of the sheetpile wall below the second stage excavation level is
calculated to be 3.5 m, resulting in an overall sheetpile length of 12.0 m.
Step 2 Determine the theoretical geotechnical resistance at minimum toe embedment
All finite element programs calculate stresses at equilibrium. In order to provide a common reference
for measurement of the reserve capacities, the theoretical geotechnical resistance is used in accordance
with Equation CB3(8).
Since the earth pressure coefficients in the Equation CB3(8) are not intrinsic constants, they cannot be
obtained directly from the calculation results and need to be determined separately in a fashion similar
to that by PFEM. Only a single soil layer is penetrated by the toe in this example and the one set of
earth pressure coefficients is sufficient to determine the lateral earth pressure profile. Where several
soil layers are penetrated by the toe, the coefficients for each layer should be derived for determining
the pressure distribution. The vertical effective stresses are obtained directly from the calculation
outputs.
The total theoretical lateral resistance (P'p) is calculated by numerically integrating the pressure
distribution over the toe penetration with the limiting resistance force calculated to be 575 kN/m. The
above calculation can be easily completed using simple algorithms.
The 575 kN/m force is the action effect (Se) specified in the Standard. For geotechnical design only,
the design action effect (S*) is calculated in accordance with Equation CB3(9).
Step 3 Determine the geotechnical strength reduction factor Øg
The geotechnical strength reduction factor is 0.55 as for PFEM.

389944842.doc - 15/06/2018 7:26:59


DRAFT ONLY 84 DRAFT ONLY

Step 4 Determine the required geotechnical resistance and revised toe embedment
In accordance with Clause 7.3.3 and Equation CB3(10)
Rug = 575 / 0.55 = 1045 kN/m.
The minimum design toe embedment is determined using similar procedures described in Step 2, with
the minimum passive resistance from the passive side of the excavation summed over the revised toe
penetration being more than or equal to 1045 kN/m.
It is noted that a limited number of iterations on the toe penetration will have to be carried out in order
to properly account for the pore water pressure effects on the soil mass. The revised toe penetration is
estimated to be 4.6 m, at which the computed effective passive soil force is 1045 kN/m.
The overall sheet pile length required is therefore 13.1 m, 1.3 m shorter than that required via the
PFEM.
At this stage, the soil imposed action effects (Se) for the support and the sheet piles are as follows:
Support force = 285 kN/m
Sheet pile moment = 221 kNm/m
Sheet pile shear = 161 kN/m
Step 5 Determine the structure loads using the revised toe embedment
The geotechnical calculation is complete for the load case considered once the final toe penetration
depth is determined at the end of Step 4. The calculated structure loads at final toe penetration
correspond to the action effects imposed through soil (Se). For structural design the design action
effects (S*) are obtained directly by multiplying the action effects by a constant factor of 1.5. The
design action effects for the singly propped sheet pile retaining wall are as follows:
Support force = 285×1.5 = 428 kN/m
Sheet pile moment = 221×1.5 = 331 kN.m/m
Sheet pile shear = 161×1.5 = 241 kN/m
In this worked example both the effects from ground water and excavation are probably the worst case
scenario. For serviceability limit state deformation estimates, a separate calculation should be
undertaken e.g. the design ground water level may be placed at 1.5 m below ground level instead of
2 m with the excavation limited to the planned depths only.

389944842.doc - 15/06/2018 7:26:59

You might also like