You are on page 1of 568

Introduction

1
The purpose of the book is to present, in a selective way, design and construc-
tion for deep excavations made for civil engineering purposes. Emphasis is
placed on descriptions of work constructed, in some instances within the
Author’s personal responsibility, but for the most part as observed by the
Author and as described and reported by others.
Both temporary construction and permanent works are described. The design
of temporary construction to support soil and rock at the excavation periphery,
and to exclude groundwater, frequently becomes the responsibility of the
contractor, and there may be incentives to devise methods that economize in
construction time and cost. The term ‘temporary’ may mislead; on major
works, measures of peripheral soil support and groundwater exclusion may
require sufficient strength and durability to last several years. The difficult
task of the temporary works designer, to provide an adequate but time-related
solution without waste, should not be underestimated. The final cost of tempor-
ary works may also depend upon the ease of their eventual removal. In some
cases, their incorporation into the permanent works, as with diaphragm walls,
can mitigate against the cost of soil support during the construction period.
Permanent construction is divided in the text, for the sake of convenience,
into work in shafts and caissons, basements and cut-and-cover construction,
with some obvious overlap with temporary works. Traditionally, the design of
permanent work and responsibility for its adequacy lies with the consulting
engineer. The exceptions are those projects designed by the owner’s organiza-
tion or which are the subject of a turnkey or design-and-construct arrange-
ment where a contracting firm may assume professional responsibility for
permanent construction. The designers of temporary and permanent works
are therefore often not the same persons on a particular job, and in many
instances they are not employed by the same organization.
The choice of which deep excavation works to include in the book has been
guided by a definition used in the CIRIA Report on trenching practice1 . This
report covered trench excavations to a depth of 6 m. As an approximate
division between shallow and deep excavations, this 6 m depth has been
adopted by the Author: for the most part, this book features work greater
than this depth. Mention should nevertheless be made of the high risk of
excavations less than 6 m deep. Of those accidents involving fatalities or
major injuries in relatively shallow excavations from 1996 to 2001 where
details are available, analysis shows the following causes:
. unsupported excavation: 54% of cases
. working ahead of support: 12% of cases
. inadequate support: 16% of cases
. unstable slopes of open cut: 6% of cases
. other causes (principally unsafe machine operation): 12% of cases.
Within these statistics an underlying cause in accidents involving inadequate
support and working ahead of support involves misuse of drag boxes and
2 Deep excavations

trench boxes. There is insufficient evidence within this period to make a corre-
lation with particular soil conditions but there is some evidence that the proxi-
mity of existing services and backfilled trenchworks and the effects of wet
weather are not adequately considered as stability risks. Throughout, a
prime reason for these accidents is the absence of capable supervision on site.

Safety and avoidance of The design of soil support to deep excavations on land must ensure an
damage adequate factor of safety against collapse in the short term, that is, during
the construction period, and an adequate factor of safety against collapse
for the design life of the permanent substructure. In addition, in both the
short and long term the design of the works must be such as to contain defor-
mation of the soil or rock adjacent to the excavation to limits which do not
cause distress to existing structures or services. Standards of construction at
each stage must be such that the work complies with the strength assumptions
used in the design and is sufficiently durable to avoid deterioration and move-
ment or collapse. Additionally, construction standards must be such as to
avoid loss of ground into the excavation which might cause an unacceptable
risk of subsidence or collapse. In the United Kingdom, legislation extends to
the rights of neighbours sharing a common party wall between their properties
and any damage that occurs due to construction work, including excavations,
on one side of the wall (see the section on Party walls at the end of this
chapter).
Deep excavations below river or sea beds require specific design considera-
tion. For the most part, deformation or subsidence will be less important than
excavations on land unless existing works are nearby. The risk of scour effects
to the sea or river bed, which are possible as a result of the new works them-
selves, may prove to be an additional hazard which could cause structural
collapse, and must be guarded against. Variations in sea and river states in
terms of tide, storm swells, current and wave conditions all require assessment
in terms of risk to safety and to structure; given the likely consequences of
collapse during construction or thereafter to foundation works at sea or on
major rivers, strenuous care is required in the assessment of such factors.
The risk to construction personnel and users of the permanent structure
must be defined separately to the risk of subsidence damage to property
and services. In the former case, the awesome consequences of inadequate
standards of design and construction in deep excavation works should be
self-evident, and particularly so on the site of the works. Frequent reference
is made to modes of failure throughout the text; the experience of previous
failures is seldom reported, and the advantage that should be gained for
future works is lost.
Design and construction works for the support of deep excavations require
investigations of the site topography, the subsoil and groundwater conditions,
the states of sea and river water and the stability of sea and river beds (where
applicable), the risk of seismic loading, the extent of superimposed loads, the
state of existing structures and services, and the availability and quality of
available structural materials. None of these matters is treated specifically,
although without such information of adequate quality prior to design of
both the temporary and permanent works, all reference to avoidance of risk
to life and reduction of damage to property becomes meaningless.

Construction In many developed countries, the design and site works of deep excavations
regulations: safety are subject to statutory regulations that are devised to maintain minimum
standards of site construction safety.
Introduction 3

In the UK the principal legislation is the Health and Safety at Work etc.
Act 19742 , which specifies the responsibility of employers, suppliers and
employees for all work together with the Management of Health and Safety
at Work Regulations 19993 . This legislation is complementary to the Con-
struction (Health, Safety and Welfare) Regulations 19964 , which apply to any
type of building works and most types of civil engineering construction. These
UK regulations list specific work operations, section by section, for example:
section 9 covers stability of structures; section 10, demolition or dismantling;
section 11, explosives; section 12, excavations; section 13, cofferdams and cais-
sons. Other legislation refers to operations and specialist activities, for example,
the Lifting Operations and Lifting Equipment Regulations 19985 , the Provision
and Use of Work Equipment Regulations 19986 , the Work in Compressed Air
Regulations 19967 , the Confined Space Regulations 19978 , and the Control of
Explosives Regulations 19919 . An explanatory manual to the Construction
Regulations, published by CIP Ltd10 , is continuously updated.
In the UK, the Health and Safety Executive is responsible for the implemen-
tation of the Health and Safety at Work Act and the Construction Regulations.
These pieces of legislation place duties on both employer and employee: while
the employer must provide safe access, a safe place of work and a safe system of
work for employees, every employee must take reasonable care for the safety of
others and must cooperate with the employer in such matters. The employer
must not intentionally or recklessly interfere with or misuse anything provided
in pursuance of the requirements for health, safety and welfare. All cofferdams
and caissons must be properly constructed, altered or dismantled under compe-
tent supervision and, wherever possible, by experienced operatives. Every
cofferdam or caisson must be provided, so far as is reasonably practical, with
ladders or other means of escape in case of flooding. Inspections of cofferdams,
caissons and trenches must be made when work is in progress and, in addition,
they must be thoroughly examined and records made whenever explosive
charges have been fired, whenever any damage has occurred, or, in any case,
every seven days. Other regulations require the inspection of lifting plant and
excavators, and the management is held responsible for the competence of
designers, supervisors and operatives.
In 1992 an extension to the Management of Health and Safety at Work Act
required all employers in the UK, not only those in construction, to carry out
assessments of risks to safety. A further widening of responsibility was made
in 1995 when the Construction (Design and Management) Regulations 199411
placed a duty on designers to avoid, so far as is practicable, risks to safety and
health during the demolition, construction and maintenance of construction
works. A recent amendment to these regulations in October 2000 placed
responsibility not only on designers but also on staff under their control.
The above in no way gives a complete explanation of safety legislation
applying to deep excavation sites in the UK but shows the change, particularly
at site, made first in 1961 and thereafter with the introduction of statutory
regulations for site safety.
Permanent works for building construction such as basements, are also the
subject of building standard control legislation either nationally or by city in
all developed countries.

Contractual In addition to the responsibility spelt out in statutory regulations for works
responsibility: client, on site and the design of temporary and permanent works, the contracts
engineer and between employer and contractor and between client and consultant will
contractor define responsibility for the adequacy of both the temporary and permanent
soil support. Contract conditions will vary between countries and from job
4 Deep excavations

Fig. 1.1. Conditions of contract


currently used in construction
in the UK12

to job. Standard forms of contract between employer and contractor for deep
excavation work include recent editions of the ICE contract, FIDIC inter-
national conditions, the New Engineering Contract and other design-and-
build contracts and management contracts. The range of conditions of
contract in use in the UK was reviewed by Clayton12 in 2001 for general con-
struction works and is summarized in Fig. 1.1. The ICE contract still remains
in use, however, in particular the sixth edition first published in 1991.
In the UK, an earlier standard form of contract between contractor and
client for civil engineering works, the ICE form (fifth edition), was commented
on by Abrahamson13 who concluded that the responsibility for temporary
works was complex. He examined four issues.
(a) Responsibility of the contractor to the employer. The contractor’s
responsibility to the employer is clear by virtue of the clause wording:
‘The contractor shall take full responsibility for the adequacy, stability
and safety of all site operations and methods of construction’. So, if any
temporary works design by the contractor, or subcontractors (whether
nominated or not), is inadequate, the deficiency must be remedied by
the contractor, and if any damage to either permanent or temporary
works is caused by the deficiency the contractor becomes liable to
rectify this also. Temporary works designed by the engineer do not
become the design responsibility of the contractor under this clause.
(b) Responsibility of the engineer to the client. The engineer has a plain duty
to the client to ensure that the permanent works are not distressed by
loads induced from the temporary works and that the temporary
works are built in accordance with the design whether by the contractor
or the engineer. In addition, the engineer carries a duty to the client to
design the temporary works where it would not be satisfactory to allow
the contractor to make the design.
(c) Responsibility (or perhaps lack of responsibility) of the employer, via the
engineer, to the contractor. The engineer has no obligation to the con-
tractor to detect or prevent faults in the temporary works. While the
engineer has rights of control under the contract, the contractor
Introduction 5

cannot excuse bad workmanship in temporary works on the basis that


the engineer made no objection.
(d) Responsibility of the employer, engineer and contractor to employees and
members of the public. The contractor is liable to employees and other
third parties if a duty of care is not discharged in designing or construct-
ing temporary works, and the contractor is probably also liable for a
defective design by the engineer when an experienced contractor
would have known it to be defective. Abrahamson concluded that the
engineer’s liability to third parties as a result of temporary works failure
was most difficult to define. Without doubt the contractor, under this
particular form of contract, does hold much of the responsibility for
the safe design and performance of temporary works, such as tempor-
ary soil support. In particular, he holds responsibility towards the client
for the adequacy in design and construction of such temporary support
works by subcontractors and even nominated subcontractors.
It is evident that works such as piling or diaphragm walls, which at different
stages serve functions both of temporary and permanent soil support, require
specific reference in the conditions of contract for such work. The case law
on such matters remains sparse and the division of responsibility between
contractor and engineer may be without legal precedent. In the UK it is not
unusual for a subcontractor to provide the design of a specialist soil support
system, say a diaphragm wall, to act both during construction and as part of
the permanent structure. Presumably, under ICE conditions, the engineer’s
approval of the subcontractor’s design for the permanent performance of
the wall element would to some extent relieve the contractor’s responsibility
in that direction, whereas the performance of the same element during con-
struction would remain solely the contractor’s responsibility. It is interesting
to note that severe distress of such a wall panel, should it occur during
construction, say below formation level, would not necessarily become
apparent at that time and may only be revealed by the non-performance of
the permanent works. The assessment of fact and the legal position of
works designed to act in temporary and permanent stages may prove to be
complicated.

Causes of failure in The failure of a soil support system does not necessarily occur by structural
deep excavations collapse; other types of failure include excessive deformation of the soil and
soil support structure, inadequate groundwater exclusion, and insufficient
durability of the soil support structure resulting in failure over time.
In the Author’s experience the causes of failure may be summarized as
follows.
(a) Open excavations
(i) inadequate site investigation resulting in optimistic design
assumptions of soil, rock strength and groundwater conditions
(ii) inadequate appreciation by the designer of susceptibility to settle-
ment of adjacent structures and services
(iii) lack of appreciation by the designer and constructor of the effects
of weathering and time on soil strength.
(b) Braced excavations
(i) inadequate site investigations resulting in optimistic design
assumptions of soil and rock homogeneity, strength of soil and
rock fabric, and groundwater conditions
(ii) inadequate quality of structural detailing
(iii) inadequate coordination between designer and constructor
6 Deep excavations

(iv) lack of appreciation by the designer of the limitations of specialist


techniques such as diaphragm walling and anchoring
(v) lack of appreciation by the designer of the influence of deflections
in the soil support structure and retained soil deformations
(vi) changes in loading from natural conditions – groundwater, tidal
states, waves, temperature – and lack of appreciation by the con-
structor of the possible consequences of these changes
(vii) changes in soil and rock conditions and the lack of appreciation
by the constructor of the possible consequences
(viii) overloading of soil support structure by temporary plant loads
(ix) bad workmanship in site temporary works.

Sowers and Sowers14 stated that, within their experience, failures of anchored
sheet pile walls and braced excavations seldom occur as the result of inadequa-
cies of modern earth pressure theories. Instead, they are caused by the more
obvious neglect of backfill loads, construction operations that produce
excessive earth pressures, poorly designed support systems and inadequate
allowances for deflections, deterioration and corrosion, and poor design in
construction details.
In the Author’s experience, structural failure of braced and anchored walls
has usually occurred within the strutting or anchorage, or by passive soil
failure below formation level caused by inadequate sheeting penetration. In
other instances, fewer in number, very bad standards have caused gaps
within walls allowing cofferdams to blow with extensive loss of ground
from behind the walls.
Deflections caused by loads applied to soil support systems observed by the
Author have generally been less than the tolerances allowed for in con-
struction of the systems and frequent actual settlements and deformation to
adjacent existing structures have been less than those predicted by calculation
unless workmanship standards have been poor.
The extent of soil deformations around large excavations is referred to in
published work more readily than records of structural failure; Peck15 and
Clough and Davidson16 reviewed the likely range of horizontal and vertical
movements. Clough and Schmidt17 , in considering the design and performance
of excavations in soft clay, used data from Peck15 , D’Appolania18 and
Goldberg et al.19 to show that settlements associated with excavations where
basal stability is a problem exceed those where no such stability problem exists.
Records of soil deformations caused by particular excavations in London
were referred to by Cole and Burland20 and by Wood and Perrin21 . Burland
et al.22 examined movements near excavations into London clay and stated
that while the magnitude of ground movement will depend on methods of con-
struction and day-to-day sequences of work made on site, it should be possible
to make reasonable estimates of upper and lower limits of movement, espe-
cially when field measurements add to knowledge of the conditions. Calcu-
lated predictions of deformation using numerical methods rely on accurate
assessment of soil deformation parameters; back analysis of field measure-
ments from nearby excavations may provide these values.
Clough and Davidson16 concluded that for a given depth of excavation the
amount of ground movement depends on the properties of the retained soil
and not on the stiffness of the temporary supporting wall, but Goldberg
et al.19 stated that wall stiffness is an important factor in such soil deforma-
tion. Defining wall stiffness by a parameter given by EI=h4 , where EI is the
flexural stiffness of the wall and h is the vertical distance between supports,
and plotting this against the stability number for excavation in clays H=cu
(where H is the total depth of excavation and cu is the undrained shear
Introduction 7

strength of the clay), boundary lines were given to show orders of expected
lateral wall movement.
The use of temporary berms and other construction procedures to reduce
movements were referred to by Clough and Schmidt17 and will be treated in
detail in the following chapters.
Subsidence may occur near excavations as a result of the installation pro-
cess of walling, sheeting and anchorages in addition to deformations caused
by loading. Deformations which occur during installation of unlined borings
for piles and diaphragm walls and movements caused by pre-loading of
ground anchors are rarely appreciated at the design stage. White23 referred
to several cases where the declination of rock anchors used to support tem-
porary sheeting caused settlement to occur as a result of the vertical load
component overstressing rock below the tip of soldier piles.
More recently, the response of buildings to induced settlements caused by
nearby deep excavations and tunnelling works has been the subject of both
analytical and observational research (see Boscardin and Cording24 and the
Proceedings of the Conference on Building Response to Tunnelling 200125 ).
This aspect is further discussed in Chapter 11.

Risk evaluation Casagrande26 emphasized that risks were inherent in any project, their exis-
tence should be recognized and, using steps representing a balance between
economy and safety, these risks should be treated systematically. Casagrande
defined ‘calculated risk’ in two parts:
(a) the use of imperfect knowledge, guided judgement and experience, to
estimate the possible ranges for all pertinent quantities that enter into
the solution of the problem
(b) decisions on an appropriate margin of safety, or degree of risk, taking
into consideration economic factors and the magnitude of losses that
would result from failure.
Casagrande did not quantify risk. Later, Whitman27 reported considerable
advances in reliability and probabilistic theory, but stressed that the use of
such methods was no substitute for physical measurements and sound
engineering interpretation. Concluding, Whitman said that the satisfactory
evaluation of risk could be answered in two ways.
(a) If a relatively large probability of failure (0.05 or more) under design
loading were tolerable, then this risk could be evaluated (by reliability
theory) with sufficiency accuracy for decision-making purposes. This
situation applies only when economic loss and not safety are of concern.
(b) If a very small probability of failure (say less than 0.001) under design
loading conditions is required, the actual risk cannot be evaluated by
analysis. However, conducting a formal evaluation of the probability
of failure can help greatly in understanding the risk and what might
best be done to reduce it.
The design of many deep excavation schemes must certainly lie within the
second category where the acceptance of failure probability must be very
low indeed because of the risk to life.
Whitman illustrated his paper with applications of reliability theory to
examine systematic and random errors when evaluating risk in slope stability,
factors of safety in risk analysis of liquefaction, and the use of system analysis
techniques for quantifying risk on a project basis. Examples of risk evaluation
were given for an industrial plant built on potentially liquefiable sands and for
earth dam construction.
8 Deep excavations

Hoeg and Muraka28 considered the conventional design of a simple gravity


retaining wall and carried out a statistical analysis. For given soil properties
and backfill height, this design used factors of safety of 1.9, 3.7 and 1.6
against overturning, bearing failure and sliding respectively, yet despite
these apparently conservative values the statistical analysis showed that the
corresponding failure probabilities were 1/10 000, 13/1000 and 3/1000. The
probability of bearing failure is particularly high, and large differences were
also indicated in failure probabilities between each failure. This example
shows the ease with which conventional factors of safety are able to mislead.
Hoeg and Muraka then redesigned the gravity wall using probabilistic
methods, evaluating initial costs, construction costs, costs of failure and the
probabilities of failure by overturning, bearing and sliding, to determine the
expected total cost. The optimum design was the system with minimum
expected total cost.
The principle of risk assessment is within the scope of this book, but prob-
ability is not (for this see Whitman27 and Hoeg and Muraka28 and the refer-
ences therein).
The primary intent of Hoeg and Muraka was to provide a model for the
probabilistic design, by similar methods, of more complicated structures
such as braced and anchored walls, but despite their intrinsic logic there is
little indication that such methods have gained acceptance by designers.

Risk management The management of geotechnical risk, including that associated with deep
excavations, was reviewed by Clayton12 in 2001. Clayton also comments
that procurement methods have an important influence on geotechnical risk
management and refers to the extent to which geotechnical risk is shared
between client and contractor as the type of contract is varied. Figure 1.2
(Flanagan and Norman29 ) suggests a broad risk division for forms of contract
which have gained increasing acceptance and do not conform to the tradi-
tional fully designed, remeasured type of contract such as the ICE Conditions
of Contract.
Examples of risk registers compiled by both designers and contractors are
given by Clayton together with a summary of software for risk management
from a 1998 survey. A recent edition of the Project Risk Management Software
Directory30 lists over 40 different software titles although their use in risk
management in geotechnical work is limited in the Author’s experience.
Generally, the Prima Vera programme appears to be the most popular
software for this purpose at present.

Fig. 1.2. Risk management:


division by forms of contract12
Introduction 9

Party walls Although outside the scope of this book, mention should be made of the
statutory rights of building owners and their tenants regarding construction
works, such as excavation and underpinning affecting party walls. In the
UK the Party Wall Act applies to this matter; Ensom et al.31 and Anstey32
should be consulted.

References 1. Irvine D.J. and Smith R.J. Trenching practice. CIRIA, London, 1992, Report 97.
2. Health and Safety at Work Act 1974. HMSO, London.
3. Management of Health and Safety at Work Regulations 1999. HMSO, London.
4. The Construction (Health, Safety and Welfare) Regulations 1996. HMSO,
London.
5. The Lifting Operations and Lifting Equipment Regulations 1998. HMSO, London.
6. The Provision and Use of Work Equipment Regulations 1998. HMSO, London.
7. Work in Compressed Air Special Regulations 1996. HMSO, London.
8. The Confined Space Regulations 1997. HMSO, London.
9. Control of Explosives Regulations 1991. HMSO, London.
10. Construction: Health and Safety Manual. CIP Ltd, Birmingham, 2000.
11. Construction (Design and Management) Regulations 1994. HMSO, London.
12. Clayton C.R. Managing geotechnical risk. Thomas Telford Ltd., London, 2001.
13. Abrahamson M.W. Engineering law and the ICE contracts. Applied Science,
London, 1983.
14. Sowers G.B. and Sowers G.F. Failures of bulkhead and excavation bracing. Civ.
Engng, 1967, 107, No. 1, 72–77.
15. Peck R.B. Deep excavations and tunnelling in soft ground. Proc. 7th Int. Conf.
S.M.F.E., Mexico City, 1969, State of the art volume, Vol. 2, 225–290.
Sociedad de Mexicana de Mecanica de Suelos, Mexico City, 1969.
16. Clough G.W. and Davidson R.R. Effects of construction on geotechnical
performance. Proc. 9th Int. Conf. S.M.F.E., Tokyo, 1977, 15–53. Japanese
Society of Soil Mechanics and Foundation Engineering, Tokyo.
17. Clough G.W. and Schmidt B. Design and performance of excavations and
tunnels in soft clay. Soft clay engineering. Ed. E.W. Brand and R.P. Brenner,
Elsevier, London, 1981, 567–634.
18. D’Appolonia D.J. Effects of foundation construction on nearby structures. Proc.
4th Pan American Conf. S.M.F.E., San Juan, 1971, Vol. 1, 189–236 (discussion
Vol. 3, 171–178).
19. Goldberg D.T. et al. Federal Highway Administration Reports. National
Technical Information Service, Washington, DC, 1976.
20. Cole K.W. and Burland J.B. Observations on retaining wall movements
associated with a large excavation. Proc. 5th Euro. Conf. S.M.F.E., Madrid,
1972, Vol. 1, 445–453.
21. Wood L.A. and Perrin A.J. The performance of a deep foundation in London
clay. Proc. 11th Int. Conf. S.M.F.E., San Francisco, 1985, 2277–2280. Balkema,
Rotterdam, 1985.
22. Burland J.B. et al. Movements around excavations in London clay. Proc. 7th
Euro. Conf. S.M.F.E., Brighton, 1979, Vol. 1, 13–29. British Geotechnical
Society, London, 1979.
23. White R.E. Anchored walls adjacent to vertical rock cuts. Proc. Conf. Diaphragm
Walls. Institution of Civil Engineers, London, 1974, 181–188.
24. Boscardin M. and Cording E. Building settlement to excavation induced
settlement. ASCE J. Geotech. Engng, 115, No. 1, 1–21. Jan. 1989.
25. Proc. of Conference on Building Response to Tunnelling, London, 2001. Thomas
Telford Ltd., London, 2001.
26. Casagrande A. Role of the calculated risk in earthwork and foundation
engineering. ASCE J.S.M., 1965, 91, No. 4, July 1.
27. Whitman R.V. Evaluating calculated risk in geotechnical engineering (17th
Terzaghi Lecture). ASCE J. Geotech. Engng, 1984, 110, No. 2, 145–188.
28. Hoeg K. and Muraka P.P. Probabilistic analysis and design of a retaining wall.
ASCE J. Geotech. Engng, 1974, 100, No. 3, 349–366.
10 Deep excavations

29. Flanagan R. and Norman G. Risk management and construction. Blackwell


Science, London, 1993.
30. Project risk management software directory. Association of Project Management.
Euro Log Ltd., 2000.
31. Ensom D., Roe E. and Anstey J. The party wall act explained. Pyramus and
Thisbe, Parrot House Press, Weedon 1996.
32. Anstey J. Party walls. Royal Institution of Chartered Surveyors, 5th edition, 1998.

Bibliography Benjamin J.R. and Cornell G.R. Probability statistics and decisions for civil engineers.
McGraw-Hill, New York, 1970.
Health and Safety Commission. Management of health and safety at work. HMSO,
London, 1992.
Health and Safety Executive. Successful health and safety management. HMSO,
London, 1991.
Lamb P. Applications of statistics in soil mechanics. Newnes-Butterworth, London,
1974.
Newmark N.M. Rankine lecture: Effects of earthquakes on dams and foundations.
Ge´otechnique, 1965, 15, 139–160.
Peck R.B. 9th Rankine lecture: Advantages and limitations of the observational
method in applied soil mechanics. Ge´otechnique, 1969, 19, No. 2, 171–187.
The control of groundwater
2
The occurrence of groundwater on site directly influences construction
methods, permanent works design and, thereby, construction time and cost;
possibly, in the longer term, the durability of the structure and its maintenance
costs will also be affected. This chapter considers only the first of these, con-
struction methods of dewatering deep excavations. It addresses groundwater
control in three ways: the problem presented to constructors; the techniques
available to them; and the calculation methods to assist in the design of
groundwater control. The chapter draws on data from four references,
CIRIA Report C5151 , BS 80042 and textbooks by Cashman and Preene3
and Powers4 .

Groundwater problems The sources of groundwater on a particular site may be threefold: rainfall,
run-off or groundwater flow through pervious soils or rock from streams,
rivers or the sea. Variations in soil and rock conditions, in particular
permeability, horizontally and vertically, cause variations in groundwater
flow both on the ground surface and below it. The degree to which ground-
water is contained by relatively impermeable soil above or below a permeable
stratum will in turn influence any excess or artesian pressure within stored
groundwater.
Reducing the quantity of groundwater within subsoil adjacent to an
excavation by a dewatering process such as pumping will increase the strength
of the soil as the groundwater pressure, or pore pressure, is reduced. Reduc-
tion in groundwater head therefore reduces the load, say on the bracing to a
deep excavation, and provides a method of improving soil strength.
The effective stress is the difference between the applied, total stress and the
pressure induced by loading to groundwater within the pores of saturated soil.
As the soil is loaded, say from a building foundation, the increase in load is
shared between the soil structure and the pore-water within the soil. The
stress carried by the soil structure, known as the effective stress 0 , is therefore
equal to total stress  less the pore-water pressure u. Since water possesses no
strength, the soil reduces in volume as the water is displaced. The time-
dependent rate of this change in volume and change in pore pressure depends
on the permeability of the soil fabric and physical drainage conditions. The
dissipation of pore pressure and the volume change stops when equilibrium
is reached with external forces applied to the soil mass.
So, the shear strength of the soil  0 depends on the effective stress and at
failure is
 0 ¼ c0 þ 0 tan 0 ð1Þ
0 0
where c is the cohesion of the soil in effective stress terms, and  is the angle
of shearing resistance in effective stress terms. The total stress, therefore, is
a stress state which applies only at one instant, whereas the fully drained
equilibrium condition when effective stress is maximized occurs after some
12 Deep excavations

Fig. 2.1. Groundwater control applied to a deep sheeted excavation in various ground conditions

time. This time is relatively short with a granular, permeable soil but longer
with a cohesive, impermeable soil.
Reduction of groundwater within an excavation may be necessary for
access by workers and machines. However, removal of groundwater from
below the excavated level and externally to the excavation may be required
to improve soil stability below and around the excavation itself. Examples
of groundwater control applied to a deep sheeted excavation in various
ground conditions are shown in Fig. 2.1. Reduction in groundwater levels
and piezometric head allows progressive excavation in the dry, reduces
pressure on the sheeting, reduces the risk of base uplift at formation level of
the completed excavation, and allows the strength of the soil to increase
progressively as effective stress conditions apply to a fully drained, dewatered
soil condition. Figure 2.2 shows a similar improvement to working conditions
and soil strength in a battered, open excavation.

Fig. 2.2. Groundwater control applied to a battered excavation


The control of groundwater 13

The quantity of groundwater that may be abstracted from individual sites


may be restricted by local legislation. In Berlin, for example, a maximum dis-
charge from any one site into watercourses has for some years been limited to
1.5 litres/s for every 1000 m2 of wetted wall and base slab areas. Any discharge
into watercourses or public sewers may be controlled by legislation regarding
the quality of the discharge water. In the United Kingdom such discharge may
be considered to be trade waste.

Available methods of The four main methods used to exclude groundwater from deep excavations
groundwater control are1 :
(a) stopping surface water from entering the excavation by using, for
example, cut-off ditches, low walls and embankments
(b) allowing water to flow into the excavation and subsequently pumping it
from drainage sumps, grips, ditches or French drains
(c) pre-draining the soil by lowering the groundwater level ahead of the
excavation, for example, by use of wellpoints or deep wells
(d) stopping the groundwater from entering the excavation by a cut-off
wall within the soil, such as a cement–bentonite slurry wall.
Table 2.1 shows the wide range of available techniques, which fall broadly
into these four categories. Selection of the most effective method at minimum
cost will depend on a number of factors, such as the dimensions of the excava-
tion (in particular its depth), the thickness and type of soil strata, the depth of
the excavation relative to soil types, the magnitude of groundwater pressure in
each stratum, the prevention of damage to nearby structures and services, and
the length of time the excavation is to remain open.
Preliminary guidance on choosing the best method may be gained from
Table 2.2, which shows the influence of the width and depth of the excavation.
The range of application of dewatering techniques (related to permeability
and drawdown) is shown in Fig. 2.3, and similarly the range of groundwater
exclusion methods varying by soil particle size is shown in Fig. 2.4.
In very broad terms, open excavations are frequently dewatered using
single- or multi-stage wellpointing systems and sheeted excavations for base-
ments and cut-and-cover construction often use sump pumping where the
sheeting can be economically driven or excavated into an impermeable
stratum to get a natural seal. Where a natural cut-off cannot be obtained,
a horizontal grout plug can be injected to obtain a cut-off. Where artesian
pressure heads need to be relieved in deep strata below excavations, this
may be done by relief wells or deep pumped wells. For both open and
sheeted excavations, deep wells with submersible pumps at the well screens
are often used to obtain a drawdown which would not have been possible
by wellpointing or sump pumping. The use of vertical cut-off walls to
isolate construction areas from inundated surrounding areas of subsoil
can be applied to achieve economies in pumping resources and construction
time.
A site investigation is necessary before choosing a dewatering method. This
will accurately define the depths and types of strata, from which permeabilities
may be estimated and groundwater levels assessed. The site investigation must
disclose any tidal influences. Where groundwater is to be removed from the
site, its method of disposal must be investigated and any risk of contaminants
within the groundwater carefully evaluated.
The contractor is therefore faced with a choice of dewatering or ground-
water exclusion methods which will each have advantages and disadvantages
in terms of cost, overall efficiency, time and convenience. Some indication of
14 Deep excavations

Table 2.1 Methods of groundwater control38

Method Soils suitable for Uses Advantages Disadvantages


treatment

Group 1: surface water control


1. Ditches All soils if used in Open excavations Simple methods of May be an obstruction
2. Training walls conjunction with diverting surface water to construction traffic
3. Embankments polythene sheeting

Group 2: temporary groundwater control


Internal pumping
4. Sump pumping Clean gravels and Open, shallow Simple pumping from Fines easily removed.
coarse sands excavations ground Encourages instability
of formation
5. Gravity drainage Impermeable soils Open excavation Simple pumping
especially on sloping equipment
sites
Groundwater lowering
6. Wellpoint systems Sandy gravels down Open excavation Quick and easy to Difficult to install in
with suction pumps to fine sands (with including progressive install in suitable soils. open gravels or ground
(including the proper control can trench excavations. Economical for short containing cobbles and
machine-laid also be used in silty Horizontal drain pumping periods of a boulders. Pumping must
horizontal system) sands) system particularly few weeks be continuous and noise
pertinent for pipe of pump may be a
trench excavations problem in a built-up
outside urban areas area. Suction lift is
limited to about 4:0–
5:5 m, depending on
soils. If greater lowering
is needed, multi-stage
installation is necessary
7. Eductor system using Silty sands and sandy Deep excavations in No limitation on Initial installation is
high-pressure water silts space so confined that amount of drawdown. fairly costly. Risk of
to create vacuum as multi-stage Raking holes are flooding excavation if
well as to lift the wellpointing cannot possible high-pressure water
water be used. More main is ruptured
appropriate to low-
permeability soils
8. Shallow bored wells Sandy gravels to silty More appropriate for Generally costs less to Initial installation is
with suction pumps fine sands and water installations to be run than a comparable fairly costly. Pumping
bearing rocks but pumped for several wellpoint installation, so must be continuous and
particularly suitable months or for use in if pumping is required noise of pump may be a
for high-permeability silty soils where for several months costs problem in a built-up
soils correct filtering is should be compared. area. Suction is limited
important Correct filtering can be to about 4:0–5:5 m,
controlled better than depending on soils, If
with wellpoints to greater lowering is
prevent removal of fines needed, multi-stage
from silty soils installation is necessary
9. Deep-bored filter Gravels to silty fine Deep excavations in, No limitation on High installation cost
wells, i.e. those with sand and water- through or above amount of drawdown as
submersible pumps bearing rocks water-bearing there is for suction
(line-shaft pumps formations pumping. A well can be
with motor mounted constructed to draw
at well head used in water from several layers
some countries) throughout its depth.
Vacuum can be applied
to assist drainage of fine
soils. Wells can be sited
clear of working area.
No noise problem if
mains electricity supply
is available
The control of groundwater 15

Table 2.1 continued

Method Soils suitable for Uses Advantages Disadvantages


treatment

10. Electro-osmosis Silts, silty clays and Open excavations in In appropriate soils can Installation and running
some peats appropriate soils or be used when no other costs are usually high
to speed dissipation water-lowering method
of water during is applicable
construction
11. Drainage galleries Any water-bearing Removal of large Very large quantities of Very expensive, galleries
strata underlain by quantities of water water can be drained may need to be
low permeability for dam abutment, into gallery and concreted and grouted
strata suitable for cut-offs, etc. disposed of by later
tunnelling conventional large-scale
pumps
12. Collector well Clean sands and Dewatering deep Minimizes number of Only suitable for large
gravel confined aquifers pumping points excavations

Group 3: exclusion methods


Temporary methods
13. Ammonium/brine All types of saturated Formation of ice in Imparts temporary Treatment takes time to
refrigeration soils and rock the voids stops water mechanical strength to develop. Installation
flow soils. Treatment costs are high and
effective from working refrigeration plant is
surface outwards. Better expensive. Requires
for large applications of strict site control. Some
long duration ground heave
14. Liquid nitrogen As for 13 As for 13 As for 13, but better for Liquid nitrogen is
refrigeration small applications of expensive. Requires
short duration or where strict site control. Some
quick freezing is ground heave
required
15. Compressed air All types of saturated Confined chambers Gives stability to sides High set-up costs;
soils and rock such as tunnels, of chamber by limiting possible health hazards
shafts and caissons ingress of water.
Reduces pumping to a
minimum
16. Slurry trench cut-off Silts, sands, gravels Practically A rapidly installed, Must be adequately
with bentonite or and cobbles unrestricted. cheaper form of supported. Cost
native clay Extensive curtain diaphragm wall. Can be increases greatly with
walls round open keyed into impermeable depth. Costly to
excavation strata such as clays or attempt to key into
soft shales hard or irregular
bedrock surfaces.
Not effective in soils of
greater permeability
than 5  103 m/s
17. Impervious soil Silts, sands, gravels As for 16 Relatively cheap. Local Must be placed some
barrier and cobbles materials may be used distance from
excavation. Restricted
depth of installation
18. Sheet piling (can be All types of soil Practically Well understood Difficult to drive and
permanent) (except boulder beds unrestricted method using readily maintain seal in
and where natural or available plant. Rapid boulders. Vibration and
unnatural installation. Steel can be noise of driving may not
obstructions exist — incorporated in be acceptable. Capital
particularly timber permanent works or investment in piles can
baulks) recovered be high if re-usage is
restricted. Seal may not
be perfect; proprietary
seals may be expensive
16 Deep excavations

Table 2.1 continued

Method Soils suitable for Uses Advantages Disadvantages


treatment

Permanent methods — diaphragms


19. Diaphragm walls All soil types Deep basements. Can be designed to High cost may prove
(structural concrete) including those Underground car form part of a uneconomical unless it
containing boulders parks. Underground permanent foundation. can be incorporated
(rotary percussion pumping stations. Particularly efficient for into permanent
drilling suitable for Shafts. Dry docks. circular excavations. structure. There is an
penetrating rocks and Cut-and-cover Can be keyed into rock. upper limit to the
boulders by reverse construction, etc. Minimum vibration and density of steel
circulation using noise. Treatment is reinforcement that can
bentonite slurry) permanent. Can be used be accepted
in restricted space.
Can be put down very
close to existing
foundations
20. Secant (interlocking) All soil types, but As for 19. Can be used on small Ensuring complete
and contiguous bored penetration through Underpasses in stiff and confined sites. Can contact of all piles over
piles boulders may be clay soils be put down very close their full length may be
difficult and costly to existing foundations. difficult in practice.
Minimum noise and Joints may be sealed by
vibration. Treatment is grouting externally.
permanent Efficiency of reinforcing
steel not as high as
for 19
Permanent methods — grouted cut-offs
21. Thin, grouted Silts and sands As for 16 As for 16 The driving and
membrane extracting of the H pile
or sheet pile element
used to form the
membrane limits the
depth achievable and
the type of soil. Also as
for 16
22. Jet grouting All types of soil and Practically As for 16 Expensive
weak rocks unrestricted
23. Cementitious Fissured and jointed Filling fissures to Equipment is simple Treatment needs to be
rocks stop water flow (filler and can be used in extensive to be effective
added for major confined spaces.
voids) Treatment is permanent
24. Clay/cementitious Sands and gravels Filling voids to Equipment is simple A comparatively thick
grouts exclude water. To and can be used in barrier is needed to
form relatively confined spaces. ensure continuity. At
impermeable barriers Treatment is least 4 m of natural
(vertical or permanent. Grout is cover needed (or
horizontal). Suitable introduced by means equivalent)
for conditions where of a sleeved grout
long-term flexibility is pipe which limits its
desirable, e.g. cores spread. Can be sealed
of dams to an irregular or hard
stratum
25. Silicates, Joosten, Medium and coarse As for 24, but non- Comparatively high Comparatively high
Guttman and other sands and gravels flexible mechanical strength. cost of chemicals.
processes High degree of control Requires at least 2 m of
of grout spread. natural cover or
Simple means of equivalent. Treatment
injection by lances. can be incomplete in
Indefinite life. Favoured silty material or in
for underpinning works presence of silt or clay
below water level lenses
The control of groundwater 17

Table 2.1 continued

Method Soils suitable for Uses Advantages Disadvantages


treatment

26. Resin grouts Silty fine sands As for 24, but only Can be used in High cost, so usually
some flexibility conjunction with clay/ economical only on
cementitious grouts for larger civil engineering
treating finer strata works. Requires strict
site control
Permanent methods — soil strengthening
27. Electrochemical Soft clays Improving shear See ‘Uses’ Installation and running
consolidation strength of soft clay costs are usually high
without causing
settlement

Table 2.2 Depth and width


restrictions for excavations that Depth limits Width limits Other limits
use groundwater control
methods38 Groundwater control by pumping
1. Sump pumping Limits of excavation: Increasing width Flatter slopes may be
Up to 8 m below increases required required for
pump installation sump and ditch unsupported
level capacity excavations in silts
and fine sands
2. Single system Maximum limit of Limited by soil cone Space required for
wellpoints drawdown: 3–4 m in of depression (R0 ) unsupported side
silty fine sands, slopes
5–6 m generally
3. Multi-stage Unlimited Limited by soil cone Requires increasingly
wellpoints of depression (R0 ) larger land-take for
side slopes
4. Horizontal Limits in installation As for 2 Segmental
wellpointing below ground level: installation lengths
4 m normally, 6 m usually 100 m long.
maximum Space required for a
machine 13 m by 3 m
5. Eductor Unlimited but for As for 7 As for 7
wellpoint type
drawdown usually
restricted to 25 m
6. Shallow wells Limit of drawdown: Not usually critical,
6–8 m below pump but the wider the
installation level excavation the more
wells are required.
Limited then by
cone of depression
(R0 )
7. Deep bored Unrestricted using Not usually critical, Extremely large
wells submersible pumps but the wider the excavation may
excavation the more require ancillary wells
wells are required within the excavation
8. Electro-osmosis Limits of excavation: Not critical Available power
8 m below pump supply
installation level
9. Drainage Can be installed at Unlimited May require large
galleries any depth where working space at
access is available installation level
10. Collector well As for 7 As for 7 As for 7
18 Deep excavations

Table 2.2 continued


Depth limits Width limits Other limits

Groundwater control by exclusion


11. Freezing Unlimited (cases Not critical, Circular construction
recorded to >900 m excavation base can highly desirable for
below ground level). be frozen. However, stability. Long time
Depends on depth to because of required for
which receiving holes economics usually installation and
can be drilled. Liquid confined to narrow freezing
nitrogen required for excavation
deeper projects
12. Compressed air See statutory Depends on depth Must be used in an
regulations below ground level enclosed
environment, as in
tunnels and shafts
13. Slurry trenching 25 m below ground None
level or as restricted
by reach of digging
plant employed
14. Impervious soil Usually 5 m or less None, since cut-off Must be placed some
barrier achieved distance from
excavation. Space is
required for
construction
15. Sheet piling Recommended None, providing Overhead space for
maximum below adequate driving required.
ground level 20 m. penetration When used as double
Have been used to achieved. Wide wall cofferdam, ratio
>30 m, but piles may excavation may of width to retained
not then be require ancillary height >0:8. Noise
recoverable central dewatering problem
16. Diaphragm wall Installation below None, but minimum Space required for a
ground level to 40 m diameter of a stabilizing bund if
normal. Up to 100 m circular cut-off wall is not tied or
can be achieved about 4.5 m propped
17. Secant Maximum depth of Overhead space for
(interlocking) installation 30 m boring rig required
and contiguous below ground level
bored piles or to hard strata
18. Thin grouted Limits of installation None
membrane below ground level:
15 m if driven (usual)
25 m if vibrated
19. Jet grouting Cannot be used
through hard rock
20. Grouting Determined by depth Unlimited, but more
processes to which receiving efficient in confined
hole can be drilled areas rather than as
and presence of strata curtains
which cannot be
penetrated by chosen
grout. 12 m below
ground level for
driven lance methods
(e.g. Joosten).
>250 m for tube-à-
manchette methods
in soft deposits
21. Electrochemical Not critical, but Not critical Available power
consolidation preferably <8 m supply
The control of groundwater 19

Fig. 2.3. Tentative range of


application of dewatering
techniques related to soil
permeability and drawdown37

Fig. 2.4. Tentative ranges of


soils for groundwater exclusion
methods38

the cost of each method can be seen from Tables 2.3–2.5, originally included in
CIRIA Report C5151 but here augmented and corrected to 2003 prices.
In terms of overall efficiency, BS 80042 recommends that the following
conditions should be fulfilled when dewatering excavations.
(a) The lowered groundwater level should be kept under full control at all
times, to avoid fluctuations that could affect the stability of the excava-
tions (and presumably the continuity of construction work within it).
(b) The method adopted should be chosen so that the excavation remains
stable at all times, that is, slips do not occur in the sides of the excava-
tion and excessive heaving of the base does not arise.
(c) When the aquifer to be drained consists of a fairly uniformly graded
granular material, it can establish itself as a material filter to prevent
loss of ground as a result of the pumping. (If this is not the case,
filter material needs to be placed around the well screen to avoid the
transport of fines, particularly in sandy silts and fine sands. The
20 Deep excavations

Table 2.3. Approximate costs of groundwater control38

Method Cost approximation

Sump pumping Cheapest. Cost of excavating sumps, pump hire, attendance and operation. Note
high relative cost of fuel: for 150 mm pump, fuel costs approx. £30 per 24 h pump
hire £190 per week
Wellpointing Very competitive for reasonable length excavations to moderate depth over short
period. Approximate cost for 100 m nominal length of wellpoint equipment with
wellpoints at 1.5 m centres: mobilization, installation and demobilization allow
£3000 (sum). Hire of equipment including two 150 mm pumps £600–£700 per
week. Cost of operating system and fuel costs: £650–750 per week. Disposable
wellpoints are cheaper for long-term projects
Shallow wells More expensive than wellpoints of same depth, but competitive on confined sites
with medium- to high-permeability soils
Eductors Expensive but may be the best engineering solution
Deep bored wells Relatively costly depending on number installed, depth and strata. Usually only
economic for large projects. Installation only costs approx. £250 per linear metre
of well, with operating costs extra
Horizontal wellpoints Dependent on plant availability. Expensive mobilization but can be competitive for
the right job, e.g. pipelines or drainage. Installation only £15 to £20 per linear
metre installed plus pump hire/operation
Electro-osmosis Very high energy costs
Electrochemical consolidation Very high energy costs
Note. 2003 UK prices; costs are only approximate and vary from job to job, influenced by job size and location, access conditions,
ground conditions and period of dewatering. Mobilization and demobilization costs are not included.

Table 2.4. Relative costs of permanent methods of excluding groundwater38

Method Relative cost

Impervious soil barrier Inexpensive form of cut-off for shallow depths but dependent on availability of
local clay soil. May require dewatering during construction
Shallow concrete cut-off If local soils are permeable, mass concrete walls of limited depth may be feasible
Slurry trench with soil backfill Relatively inexpensive cut-off. Performance not necessarily adequate. Often
undertaken with extended backhoe without concrete guide walls
Thin grouted membrane An inexpensive cut-off particularly in granular subsoils where H pile former can be
inserted by vibration economically
Jet grouting and intrusion Secant walls in jet grouted soils and intrusion grouting may be cost feasible
grouting depending on soil conditions
Soil mix walls Use of overlapping columns in soil mix may be feasible depending on type of soil
and availability of local expertise
Slurry trenching Competitive at moderate depths where extended backhoe excavation is feasible.
Plastic sheets may be inserted to improve performance
Diaphragm walling Expensive but feasible to depths in excess of 50 m by use of cutter rigs where high
quality is required. May be integrated into permanent structural walls to improve
overall expense
Piled wall cut-offs Contiguous piled walls with possible later jet grouting or hard/soft secant pile
construction are feasible in most subsoils to moderate depths. Expensive
The control of groundwater 21

Table 2.5. Approximate costs of


geotechnical processes for Method Relative cost per m3 of grouted soil
excluding groundwater38
Grouting
Clay Cheapest
Cement/fly ash £30–£40
Cement £50–£100
Cement/bentonite £50–£120
Bentonite gel £50–£120
Jet grouting £150–£200
Silicates Joosten £50
Aluminates £100
Esters £120
Resins £250–£350
Chemicals Up to £300–£400
Compressed air Very expensive
Freezing Extremely expensive. Usually regarded as a last resort. More
expensive than electro-osmosis. Only economic at great depth
Note. Installation methods, periods, soil and rock types, job size and location all influence site-
specific prices. The above prices (2003 UK prices) are given only for broad comparison
purposes.

quantity of fines pumped should be checked by using a silt trap within


the discharge pipework.)
(d) There should be an adequate margin of pumping capacity and standby
power plant.
(e) Water removed by pumps should be discharged clear of the excavation
areas, avoiding erosion, silting or contamination of existing drains or
watercourses.
(f) The pumping methods adopted for groundwater lowering should not
lead to damage of existing structures (in particular, the risk of removal
of fines by pumping and the long-term consolidation of soils, especially
fine-grained soils, should be carefully evaluated; care in filter and well
screen design will determine the extent of the transport of fines by
pumping). The risk of settlement caused by removal of groundwater
will be determined by the extent of drawdown and the consolidation
characteristics of the soil, and both should be examined carefully,
especially when nearby buildings are sensitive to settlement.
(g) At least for economy in pumping, the water level should not be lowered
further than necessary to keep the excavation clear of water at all times.
(h) The method applied should avoid excessive loss of ground by seepage
from the sides of the excavation.
The methods available to implement these conditions are described in more
detail below.

Surface drainage
It is prudent to avoid surface run-off entering the excavation by the use of a
cut-off ditch or French drain (using a porous pipe with a granular surround).
This applies particularly on sloping ground where the cut-off drain is placed
on the side of the excavation with greater elevation.

Gravity drainage
In relatively impermeable soils water can be conducted to sumps by open-
jointed pipework surrounded by gravel with very little fall. Some advantageous
22 Deep excavations

drawdown will occur if the trench excavations are lowered when groundwater
levels are close to the main formation level.
In deep cofferdams with a clay formation, gravity drainage is frequently
used to conduct groundwater to a sump sited outside the plan area of the
permanent works. These temporary drains should be grouted on completion
of the permanent works construction to avoid introducing a weak bearing
area below the permanent works and a source of free water which could
cause leakage through the permanent structure. Where possible, cofferdam
widths should be increased to accommodate these drains outside the perma-
nent works and to accommodate any tolerance required in the installation
of the sheeting to the cofferdam.

Sump pumping
Sump pumping is the traditional method of removing groundwater from
within a sheeted deep excavation. The sump, ideally formed within the sheeted
enclosure outside the plan area of the permanent construction, is equipped
with a suction head to a lift pump, or with a filter head to a submersible
pump where the lift would be more than 6 m or so below a suction pump.
The suction or filter head can often be made with a perforated 45 gallon oil
drum surrounded by a gravel filter medium. Frequently, a greater pump
capacity is needed to pull the groundwater down in the sump, either prior
to bulk excavation or during the course of excavation, than is required there-
after to maintain the depressed level of groundwater. It may be convenient to
install the pumps in multiple units so that the pumping capacity can be
reduced in the steady-state pumping condition.

Pumping from wellpoints or wells


Wellpoints and deep well systems installed prior to excavation and outside the
excavation area can be used to cause groundwater to flow away from the
excavation to improve stability to its side batters and base, and to allow con-
struction works to proceed in the dry.
Traditionally, wellpoints have been an easily installed system in loose to
medium-dense and well-graded sandy gravels. With a drawdown limit for
each wellpoint of 5 or 6 m from the level of the pump suction, either single-
or multi-tier systems can be used in battered excavations. Multi-tier systems
are installed progressively as drawdown is achieved at each level of wellpoint-
ing. The wellpoints consist of small-diameter (50 mm) wellscreens and are
generally spaced at regular intervals around the perimeter of the excavation.
When wellpoints are spaced relatively close to one another, the completed
overlapping cones of depression of the installation are such that the depressed
water level is brought below the final formation level (Fig. 2.5). The wellpoints
are normally jetted in, although soils such as open gravels may lead to high
jetting-water loss, and other soils such as compact sands, stiff clays and
boulder clays may be difficult to penetrate by jetting, in which case pre-
boring may be needed. The wellpoints, once installed, are connected to a
header pipe, typically 150 mm in diameter, by swing-joint connections
equipped with a gate valve to each wellpoint. The header pipe is joined to
the suction side of a vacuum-assisted centrifugal pump with delivery to a con-
venient point of discharge. Figure 2.6 gives an indication of the soil grading in
which wellpoints are economical. Wellpoints are extracted after use by jetting
out, although this may prove difficult if the dewatering system has been in
place for a long time. Where soil contamination constitutes a serious risk of
corrosion to the wellpoint screen, a disposal wellpoint with a plastic slotted
tube and nylon filter fabric may prevent the screen from blocking during
long-term pumping.
The control of groundwater 23

Fig. 2.5. Overlapping cones of


depression in a wellpoint
installation to produce a
depressed water level below
the formation level

Fig. 2.6. Approximate soil


grading limits for effective
wellpoint dewatering38
24 Deep excavations

Fig. 2.7. Cross-section of a


large open excavation showing
typical wellpoint installation
applied in two stages

Typical wellpoint installations for large open excavations of increasing


depth are shown in Fig. 2.7. After drawdown has been obtained, it may be
possible to remove the highest tier of wellpoints for reuse at a lower level. It
should be noted that multi-stage systems often impose large construction
widths to accommodate the batters and the berm widths at each wellpoint
tier. Heavy construction plant may damage header pipework and it is usual
to install valves within the header pipework to allow speedy isolation and
replacement of any damaged sections.
Wellpoints are spaced 1 to 4 m apart, depending on soil conditions and
drawdown requirements. Nomograms for spacing wellpoint installations in
clean, uniform sands and gravels and stratified clean sand and gravel are
shown in Fig. 2.8. The spacing should be based on the most permeable of
the strata, and it should be noted that the lower the permeability of the
ground, the steeper the drawdown curve. Variations between horizontal and
vertical soil permeability should be noted; in stratified sands the horizontal

Fig. 2.8. Nomographs for wellpoint spacing: (a) uniform sands and gravels; (b) stratified sand and gravel38
The control of groundwater 25

permeability (expressed in cm/s) may be up to one order less than the vertical
soil permeability.
The design of wellpoint systems is considered in more detail by Cashman
and Preene3 and by Powers4 , who gives methods for estimating yield and
drawdown for confined and unconfined aquifers for various well layouts.
Drawdown by wellpoints becomes progressively slower as soil grain size
reduces, and becomes very slow in silts, silty sands and soils with a D10
grain size less than about 0.05 mm. Friction losses within pipework become
critical in such soils and efficient drawdown is less likely. In these circum-
stances, vacuum wellpointing can be used to advantage. A bentonite clay
seal is used to maintain a high vacuum at the well screen and although a con-
ventional drawdown of groundwater is not achieved, continuous pumping
and close spacing of the vacuum wellpoints prevent groundwater flow from
the soil towards the excavation. A typical sealed vacuum wellpoint is shown
in Fig. 2.9.
Wellpointing is often advantageous for dewatering trenchworks and can be
applied to either one or both trench sides. The system can be installed progres-
sively for long trenchworks by leapfrogging in increments of 100 to 120 m.
Within this length successive operations of wellpoint installation, operation,
Fig. 2.9. Vertical cross-section trench excavation, trench backfilling and wellpoint extraction are carried
of a sealed vacuum wellpoint out prior to moving forward to the next length. Although systems installed
to one trench side only allow favourable access for construction activities,
the extent of drawdown over the trench width may not be sufficient other
than in homogeneous permeable soils. The effective depth of trench
dewatering is limited of course by the 5 to 6 m maximum depth of the
wellpoint itself. Figure 2.10 shows a typical installation for a single-sided
wellpoint system.
Some improvement to drawdown in a layered depth can be achieved by
vertical sand drains to improve vertical permeability. The use of a granular
soil surround to the wellpoint, known as ‘sanding-in’, achieves the same
purpose. The use of a double-sided wellpoint system for trenchworks is
shown in Fig. 2.11. This system provides a more effective drawdown and
allows a greater trench width than the single-sided system for similar soil
conditions. The improvement in drawdown reduces the risk of instability at
the bottom of the trench. This also results from groundwater flow towards
the wellpoints and away from the bottom of the trench.

Fig. 2.10. Typical installation of


a single-sided wellpoint system
26 Deep excavations

Fig. 2.11. Typical installation of


a double-sided wellpoint
system

Where fine-grained soils predominate, a vacuum can be applied to the well-


points to improve system effectiveness, although the time taken to achieve full
drawdown may prove to be longer than desired. Some reduction in efficiency
may also occur with time as air is drawn into the system through ineffective
seals and joints. The effective depth of drawdown may be increased in all
soils by a multi-stage system, as shown in Fig. 2.7.
The use of a dewatering system outside a sheeted excavation such as trench-
works reduces the groundwater pressure from the sheeting and trench bracing,
but means an increased drawdown of the water table from the area outside the
excavation which is limited in extent only by the permeability of the soil itself
and the volume of the aquifer. Similarly, the yield of groundwater to achieve
this drawdown and the pump capacity needed is controlled by the permeability
of the subsoil and the aquifer volume. Where the drawdown curve is flat, in
relatively permeable strata, the removal of groundwater, sometimes un-
intended, from beneath existing structures and services a considerable distance
from the excavation may cause settlement and subsidence damage.
In addition to vertical drainage, horizontal drains and wellpoints are also
used, particularly for trenchworks. In layered soils the difference between
horizontal and vertical permeability of the soil fabric may be as large as one
order of magnitude when permeability is measured in cm/s. It is therefore
illogical to install dewatering that relies on the vertical flow of groundwater
to the wellpoint or well screen if the means are available to lay a horizontal
drain at the required depth to achieve drawdown. Trenches for drains are
dug by backhoe or, alternatively, specially built machines for laying a
perforated pipe within a gravel surround may be used. The maximum depth
for the operation of these excavator/pipelayer machines is only of the order
of 6 m, but they find application for shallow trenchworks in homogeneous
sandy soils and have attractive installation outputs of up to 1000 linear
metres of drain per day. Although currently available in Holland and
North America these machines are now less commonly available in the
United Kingdom than previously.
Ejector systems, also known as eductor systems, may provide an alternative
to both conventional and vacuum wellpointing, although their adoption
probably requires a more reliable and detailed knowledge of subsoil and
groundwater conditions. Overall, the plant and installation costs are likely
to be less than those for deep wells, and although individual ejector efficiency
is low, the operational depths are not limited in the same way as wellpointing
and therefore multi-tier systems become unnecessary.
Both twin- and single-pipe ejectors are used; typical pipe layouts are shown
in Fig. 2.12. In the twin-pipe ejector, high-pressure water through the supply
pipe and the body of the ejector is fed through the tapered nozzle. A partial
The control of groundwater 27

Fig. 2.12. Single- and twin-pipe


wellpoint ejector systems4

vacuum is caused in the suction chamber, causing groundwater to be sucked


through the foot valve. Incoming supplies from the downward supply pipe
and through the foot valve mix in the suction chamber and pass into the
Venturi where the velocity decreases and the pressure increase is sufficient
to send the combined supplies through the second pipe to ground level.
This twin-pipe system is usually installed in a casing and well screen, unlike
the single-pipe system, which is shown in section detail in Fig. 2.13. The
single-pipe ejector can be installed within a 50 mm diameter wellpoint and,
like the twin-pipe model, is self-priming and will evacuate air from its own
well screen. The single-pipe system relies on downward flow of water (this
time in the annular space between the casing and the inner return pipe) to a
suction chamber and then, with groundwater sucked in, the water passes
into a Venturi chamber where increased pressure causes the supply and the
groundwater to flow upward to ground level for disposal.
Ejectors are best suited for removal of groundwater from finer-grained,
low-permeability soils and are used in deep excavations and shafts beyond
the depths economically operated by multi-tier wellpointing. Their operation
depends on submergence below groundwater level, otherwise cavitation
occurs. Deep well systems are likely to be used in preference to ejectors
unless this submergence condition is reasonably certain to be maintained.
Case studies of the use of ejectors and vacuum wellpoints in fine-grained
Eocene soils in the UK were described by Preene5 , and methods of estimating
steady-state flow rate discussed.
28 Deep excavations

Fig. 2.13. Cross-section of


single-pipe ejector4

The most convenient subsoil conditions for excavation through inundated


soils are where an impermeable stratum can be economically reached by
sheeting driven, or a cut-off installed to penetrate below excavation level.
Some penetration of the impermeable stratum will be needed to form a seal,
and this may be achieved with difficulty in hard soils or rock strata. Grout
injection may improve this seal. In these circumstances, groundwater flow
below the sheeting or cut-off will be small, and providing there are no serious
openings or ‘windows’ within the sheeting or cut-off below excavation level,
there will be no risk of base failure and heave of the completed excavation
at formation level. It is only necessary to pump out the volume of ground-
water within the sheeting near formation level and to maintain the dewatered
level. Where such a cut-off does not occur at convenient depth, a seal can be
obtained with a horizontal cut-off by jet grouting or by intrusion grouting
where soil permeability makes this feasible. In these circumstances, again,
only limited dewatering is needed within the sheeting to keep the formation
dry, providing the sheeting or cut-off is effective below formation level and
leakage through the grout plug is small.
Where an impermeable stratum does not occur at a convenient depth or the
use of a horizontal grout curtain is not feasible, the flow of water through
permeable strata occurs downwards, under the sheeting, and then upwards
to formation level. As the penetration depth of the sheeting is increased, the
flow path is increased and the quantity of exit water and its exit velocity are
reduced proportionally. This ingress of groundwater upwards to formation
level may have serious effects because of the induced seepage forces on the
passive wedge of soil below excavation level supporting the external sheeting.
Marsland6 showed the relevance of this seepage force in model tests and, later,
Soubra and Kestner7 considered the effects of this seepage force on alternative
passive failure mechanisms. The risk of instability below formation level
within a sheeted excavation is described later in this chapter. In finer-grained
The control of groundwater 29

Fig. 2.14. Effect of seepage


pressure on base stability
within an excavation in sand:
calculation of exit gradient and
factor of safety against piping

soils where the differential head of groundwater between the outside and
inside of the sheeting to the excavation is high, soil liquefaction can occur
when the vertical effective stress within the soil reaches zero, a condition
caused by the seepage pressure. This condition can be examined by consider-
ing the critical exit velocity with the use of a flow net, as shown in Fig. 2.14,
and will give an early indication of risk. Figure 2.15 shows the reduction in
factor of safety against piping as the width of a trench in fine sand is progres-
sively reduced. The possibility of base instability within excavations in soft
clays should be examined, as described in Chapter 7 on cofferdam design.
Where the ingress of groundwater below sheeting is such that seepage pres-
sures are excessive or the quantity of water is too great to remove by horizontal
drains and sump pumping, the use of deep wells near the toe of the sheeting
becomes necessary. Screens for these deep wells may be sited below the toe
of the sheeting to reduce flow and upward seepage pressures, but at the expense
of greater yield of groundwater and increased radius of drawdown.
A cross-section through a deep well is shown in Fig. 2.168 . Methods of
boring for the installation of deep wells are summarized in Table 2.6 from
the CIRIA report1 . The diameter of the bore is dependent largely on the
diameter of the submersible pump required. The initial equipment and
installation costs for a deep well scheme are higher than those for wellpoint
systems, although pumping efficiency is generally better and similar to that
in water well schemes. The wells typically consist of a plastic well casing
surrounded by granular filter media over the screen length with a bentonite
seal above the filter. A geotextile or nylon sock is usually placed over the
well screen to avoid any movement of the filter media towards the pump.
More durable and stronger screens from galvanized or stainless steel are
employed where long-term performance in fine-grained soils is required. An
electric submersible pump of the required power and capacity is suspended
within the well screen depth and connected to riser and header pipes.
BS 80042 provides a useful guide to the design of filters to surround deep
well installations, pumping sumps or blankets to control seepage:
30 Deep excavations

Fig. 2.15. Reduction in the factor of safety against piping as cofferdam width is reduced, shown by flow net calculations

(a) The 15% size of the filter should not be greater than four times the 85%
size of the natural soil of the protected material surrounding the filter
(point A, Fig. 2.17).
(b) The 15% size of the filter should not be less than four times the 15% size
of the protected material (point B, Fig. 2.17).
(c) Filters should not contain more than 5% material passing through a
75 mm sieve (point C, Fig. 2.17), and such material should be cohesion-
less. Where the size given by (b) is less than that given by (c), the latter
should apply.
(d) The grading curve of the filter material should roughly follow the same
shape as the grading curve of the protected material.
(e) The 50% size of the filter should not exceed 25 times the 50% size of the
protected material (point E, Fig. 2.17).
( f ) Where the protected material contains a large proportion of gravel or
coarser material, the filter should be designed on the basis of the
The control of groundwater 31

Fig. 2.16. Cross-section of a


typical deep well with
observation wells8

grading of that proportion of the protected material finer than the


19 mm sieve.
(g) Where the soil is gap graded (e.g. a silty fine sand with some gravel) the
coarser particles cannot prevent the finer particles from migrating
through the large pore spaces in the filter, if the latter is designed on
the complete grading curve of the soil. Therefore, the coarse particles
should be ignored and the grading limits for the filter should be selected
on the grading curve of the finer soil.
(h) Where a filter is placed against a variable soil, the filter has to be
designed to protect the finest soil. Generally, rule (a) should be applied
to the finest soil and rule (b) to the coarsest.
(i) Filter material should be well graded within the limits permitted by the
range of particle sizes to avoid segregation when placing.
(j) The maximum size of the filter material should not be more than about
80 mm.
BS 8004 adds that in practice slightly coarser grading may be needed to ensure
reasonable flow of water to the well without initial loss of an unreasonable
quantity of fines.
Deep well systems are often applied in deep excavations within temporary
cofferdams or basement excavations to remove groundwater from sands and
gravels or to intercept flow from water-bearing rocks. In particular, they can
be used to relieve artesian groundwater pressure at depth from contained
aquifers. In order that pumping prior to excavation may be effective, the
period required to achieve reduction of head should not be estimated too
32 Deep excavations

Table 2.6 Summary of principal drilling techniques used for dewatering well installations38

Method Resources Typical diameter and depth of Notes


bore

Jetting with hammer- Supervisor 300 mm cased to 20 m depth Not widely used. Not
action placing tube Labourers usually cost-effective for
Crane operator installation of just a few
Hammer-action tube wells. Can be difficult to
Large jetting pump monitor and control.
Large compressor Special safety measures may
Crane, twin roped, be necessary
free fall
Cable percussion Drill rig operator 150–600 mm cased to about Widely available. Effective
Assistant drillers 50 m depth in unstable ground at penetrating granular and
Cable percussion drill with casing telescoped 100 m cohesive soils. Slow
and casing depth or more in stable penetration if rock or
formations uncased cobbles and boulders
present
Rotary open-hole with Drill rig operator 150–600 mm uncased to 100 m Rapid installation rates
mud, direct circulation Assistant driller depth or more with achievable in granular and
Rotary drill and rods appropriate rig cohesive soils. Cobbles and
Mud boulders can cause difficulty
Mud handling system
Rotary open-hole with Drill rig operator 400 mm plus uncased to 100 m Similar to direct circulation
mud, reverse circulation Assistant driller depth or more with system, but usually used for
Rotary drill and rods appropriate rig larger holes
Mud
Mud handling system
Rotary cased hole with Drill rig operator 100–250 mm cased to 30 m An appropriate rig can
water flush Assistant driller depth or more with penetrate virtually any
Jetting pump appropriate rig ground from hard rock to
Rotary drill and casing soft clay
Rotary down the hole Drill rig operator 76–600 mm to 100 m depth or Requires the use of duplex
hammer Assistant driller more with appropriate rig systems in unstable
Large compressor formations
Rotary drill and rods
Down the hole hammer
(Foam)

optimistically. Failure to achieve sufficient drawdown may limit excavation


progress or, worse, induce base failure of the formation where reduction of
the artesian head is not measured or taken into account.
Design of a deep well system for a sheeted or enclosed excavation should be
made with reliable data regarding groundwater level, soil stratification and
soil fabric permeability. Due to the importance of the installation and the
capital costs related to the number of wells required, soil permeability
values should be assessed from full-scale pumping tests wherever possible.
Where buildings or services are nearby, it is essential that settlement risk
due to dewatering is assessed and, where necessary, the cost and construction
time of local recharge schemes allowed for.
An example of the successful application of a deep well dewatering scheme
is shown in Fig. 2.18. The site for an underpass some 600 m long and maxi-
mum depth 12 m had been reclaimed using hydraulic fill from the coastline
at Jeddah. The subsoil conditions are also shown in Fig. 2.18: 1 m thick
loose sand and gravel fill overlaid the reclamation material consisting of
The control of groundwater 33

Fig. 2.17. Design rules for filters2

very loose sand and a medium dense to dense clayey sand to a depth of 7 m.
Below this reclamation a layer of soft coral limestone interbedded with dense
coralline sand 7 m thick overlaid dense clayey sand. The existing groundwater
level was generally 1 m deep over the whole of a fairly flat site. The underpass
walls were designed to be built in structural diaphragm walls and the whole
structure was anchored by tension piles. The structural walls, spanned by a
reinforced concrete floor slab, resembled a floating structure anchored by
the piles. The structural walls did not achieve either a temporary or permanent
seal into impermeable soil or rock at depth. Cross-diaphragms in self-
hardening slurry works subdivided the excavation into three compartments,
but none of these walls achieved a cut-off with depth. The three compartments
were successively excavated and dewatered by a total of 29 deep wells. The
specification for the wells and the pumps is shown in Fig. 2.19. The cost of
the four cross-diaphragm slurry walls was significantly less than savings
made from reducing the programme time; early commencement of excavation
within the first compartment was possible prior to completion of the struc-
tural diaphragm walls further along the underpass. Compartment working
also reduced drawdown of groundwater at the shallower excavation levels
at each end of the underpass. The dewatering scheme was particularly
successful, bearing in mind the limited opportunity for pumping prior to
bulk excavation within a very short construction programme worked on a
24-hour, seven-day basis. At the deepest section of the underpass, the deep
wells reduced groundwater to a metre or so below formation level with two
to three weeks of pumping prior to excavation.
In such conditions, the penetration depth of the external diaphragm walls is
selected to meet the most severe of four design criteria: first, the minimum
depth to avoid piping of loose sand below formation level; second, the mini-
mum depth to give sufficient passive resistance to support the external walls
prior to casting the floor slab to the underpass; third, the optimum depth
to extend the drainage path to formation level for external groundwater;
34 Deep excavations

Fig. 2.18. Diwan Underpass, Jeddah, Saudi Arabia, longitudinal and transverse section

last, to site the well heads within the walls at the most economical depth. The
optimum depth for the well heads minimizes the number of wells required to
achieve drawdown, at the same time reducing the quantities to be pumped in
terms of available pump capacity and energy used. Wall penetration and the
design of the dewatering scheme in sheeted deep excavations is referred to
later in this chapter.

Use of relief wells


The use of vertical relief wells each consisting of a granular filter backfill
placed within and surrounding a slotted well screen can be effective, without
pumping, for transferring perched water tables to lower, partly-filled aquifers
below deep excavations. Ward9 also referred to the use of this simple
The control of groundwater 35

Fig. 2.19. Location of cut-offs and deep wells for the Diwan Underpass, Jeddah, Saudi Arabia (courtesy of Bauer)

procedure to relieve the artesian pressure which had caused uplift of the
formation and damage to pipework in a trench excavation penetrating
laminated sand and silty clay with entrapped artesian groundwater.

Groundwater control by cut-offs


Relatively impermeable walls or cut-offs can be effective in temporarily
excluding groundwater from deep excavations, reducing the need for removal
of groundwater by pumping. The use of cut-offs can therefore materially
improve construction progress by dividing a basement or cut-and-cover
excavation into compartments, allowing excavation to proceed below the
water table in construction areas where perimeter soil-retaining walls have
been completed, the cut-off achieving a seal into an impermeable stratum
below the aquifer prior to the completion of the whole perimeter wall. Cut-
offs can also be effective in restricting the flow of groundwater from aquifers
exposed by battered, open excavations.
The cut-off may possess both flexural strength and relative impermeability,
such as walls built from sheet piling, reinforced concrete piling and diaphragm
walling, or the cut-off may have low permeability, relatively high flexibility
and low flexural strength. Cut-offs in the latter category include unreinforced
diaphragm walls in plastic concrete, thin diaphragm cut-offs made by grout
36 Deep excavations

Table 2.7 Summary of cut-off methods8

Method Soils suitable for treatment Uses

1. Sheet piling All types of soil (except boulder beds and Practically unrestricted
rock)
2. Diaphragm walls (structural All soil types including those containing Deep basements
concrete) boulders (rotary percussion drilling Underground car parks
suitable for penetrating rocks and Underground pumping stations
boulders by reverse circulation using Shafts
bentonite slurry) Dry docks
3. Slurry trench cut-off Silts, sands, gravels and cobbles Practically unrestricted
Extensive curtain walls round
open excavations
4. Thin grouted membrane Silts and sands As for 3
5. Contiguous and secant bored All soil types but penetration through As for 2
pile walls (and mix-in-place boulders may be difficult and costly
pile walls)
6. Cement grouts Fissured and jointed rocks Filling fissures to stop water flow
(filler added for major voids)
Grouted cut-offs
7. Clay/cement grouts Sands and gravels Filling voids to exclude water.
To form relatively impermeable
barriers — vertical or horizontal.
Suitable for conditions where
long-term flexibility is desirable,
e.g. cores of dams
8. Silicates: Joosten, Guttman Medium and coarse sand and gravels As for 7, but non-flexible
and other processes
9. Resin grouts Silty fine sands As for 7, but only some flexibility
Freezing
10. Ammonium/brine refrigeration All types of saturated soils and rocks Formation of ice in the voids
stops water flow
11. Liquid nitrogen refrigerant As for 10 As for 10

injection, mix-in-place pile walls, unreinforced bentonite slurry walls, cement–


bentonite slurry walls and membranes. These low strength cut-offs rely for
their support above formation level on soil berms between the cut-off and
the excavation, and find economic application in temporary works where
only the property of relative impermeability is needed. Specialist methods
such as installation of freeze soil barriers using ammonia/brine or liquid
nitrogen find application in particular locations in fine-grained saturated
soils which justify relatively high installation and running costs. Table 2.7
summarizes cut-off methods8 .
The effectiveness of cut-offs constructed from grout curtains and slurry
trenches was discussed by Ambrasseys10 and later by Brauns11 and others.
The effectiveness of cut-off walls using jointed methods such as sheet piling,
concrete piling and concrete diaphragm walls was reviewed by Telling
et al.12 , referring to the use of the cut-offs below dam structures. Analytical
appraisal of cut-off efficiency, while important for permanent cut-offs below
dams, is not of prime importance for temporary works in deep excavations,
although the practical aspects of avoiding local windows caused by bad
workmanship and leading to leakages in temporary cut-offs may prove to
be vital.
The control of groundwater 37

Fig. 2.20. Triple auger rig used


in mix-in-place piling (courtesy
of Raito Inc.)

Mix-in-place pile walls


The use of mix-in-place methods, combining cement additives, water and soil,
was first applied by the Japanese Ministry of Transportation in 1975 using
multi-shaft augers. The method, now popular in Japan and the United
States is finding progressive application in Europe, as plant manufacturers
develop equipment to efficiently mix in place either powdered additives to
the soil or wet grout pre-mixed and achieved to the requisite depth through
the hollow stem of a deep auger. Figure 2.20 shows a typical triple auger rig.

Cut-off construction using slurry wall techniques


The development of slurry cut-off walls began in the mid 1940s when the US
Corps of Engineers used permanent cut-off construction below levees on the
Mississippi river. Wide cut-offs excavated by dragline were put down under
bentonite slurry, and the excavated soils were later backfilled by dozer. The
slurry stabilizes the trench during excavation and is left in place, with the
backfilled soil used to form the cut-off. A filter cake slurry deposit forms on
the trench sides during excavation and acts compositely with the backfilled
soil and slurry to resist the flow of water across the trench. Many cut-offs
of this type were subsequently made in the United States to depths of 30 m
or more and widths varying up to 3 m for temporary and permanent control
38 Deep excavations

of seepage into excavations, as foundation and embankment cut-offs for


water-retaining structures, and to prevent seepage of various pollutants
from contaminated groundwater. D’Appolonia13 described construction
methods by dragline and backhoe and provided values of wall permeability
as a function of the bentonite content of completed backfill and the permeabil-
ity of soil–bentonite backfill related to fines content. Overall, however, the
application of such wide, deep soil–bentonite slurry cut-offs has been replaced
in deep excavation works by bentonite–cement slurry membranes which are
installed by backhoe with long booms to relatively shallow depths up to
10 m, to greater depths using conventional diaphragm wall equipment, or a
combination of both. These cut-offs are narrower than the earlier soil–
bentonite walls and vary between 600 mm and 1 m, and may also be backfilled
with excavated soil, as described by Hetherington et al.14 where imperme-
ability requirements are less rigorous and a cheaper membrane material is
needed. This innovation was first used in 1975 at Alton Water Dam, Ipswich,
UK, for the construction of a permanent seepage cut-off to an impounded
area. Initially, the 600 mm wide trench was excavated by backhoe and then
by conventional diaphragm wall equipment, a hydraulic grab mounted on
kelly bar equipment. Conventional, temporary tubular stop ends were used
at the end of each working day, but guide trench construction was limited
to the use of timber baulks. An initial site trial used six panels with varying
cement contents for the slurry. With a constant bentonite content of 4% by
weight, the cement was varied in increments from a minimum of 70 kg/m3
to a maximum of 150 kg/m3 : permeability test results varied from 1  108
to 1  109 m/s at the lower and higher cement contents, respectively. The
water bleed did not appear to vary with cement content; a final cement content
of 125 kg/m3 was used.
The properties of bentonite–cement slurry cut-off walls, known as self-
hardening slurry walls, were described by Caron15 . He examined the character-
istics of the slurry in its two separate phases, initially as a means of trench
support and subsequently as the hardened trench fill mixture. During the
excavation phase the initial properties of the self-hardening slurry are similar
to those of a bentonite slurry, but over time the cement content causes stiffen-
ing and, as excavation proceeds, the bentonite–cement slurry progressively
combines with soil particles. Later, water bleed from the slurry and filtration
through the sides of the trench causes further stiffening. The change in slurry
stiffness and the increase in viscosity with time in the initial excavation phase
were described by Caron for a range of cement types, retarder additives and
soil addition to the slurry. Hardened slurry properties were also examined,
but as a first approximation Caron concluded that slurry strength was
governed by the cement/water ratio, viscosity by the bentonite/water ratio,
and the setting time by the retarder/water ratio. More detailed analysis
showed the relationships were a little more complicated, as shown in Table 2.8.
The principal characteristics to be addressed by the designer are cut-off
strength, minimum strain at failure and permeability. For a typical self-
hardening slurry for groundwater cut-off, these properties would be specified
as follows:

Table 2.8 Hardened slurry


properties15 Characteristic Principal factor Secondary factors

Strength Cement/water ratio Bentonite/water, retarder/water ratios


Viscosity Bentonite/water ratio Cement/water, retarder/water ratios
Settling time Retarder/water ratio Cement/water, bentonite/water ratios
The control of groundwater 39

. unconfined compressive strength: at 28 days, samples from panels, 90%


of all results in the range 100–500 kN/m2
. permeability: target value less than 1  109 m/s; test results from
control samples from panel: all samples less than 1  108 m/s and at
least 50% of all samples less than 1  109 m/s
. consolidated drained triaxial testing: all results of samples tested from
panels to demonstrate strain at failure in excess of 5%.
On a cut-off contract (Kielder Water Dam, Icos, 1976) the mix proportions
to achieve a similar specification were, in kilograms per cubic metre of
slurry: bentonite, 54 kg/m3 ; water, 942 kg/m3 ; Portland cement, 28.8 kg/m3 ;
ground blast furnace slag, 67.2 kg/m3 ; retarder additive, nil. The parameters
specified were:
. permeability: 500 h after placing under a hydraulic gradient of 1:450, the
sample first saturated under a back pressure of 150 kN/m2 :108 m/s
. deformation: minimum deformation under a deviator stress of 125 kN/
m2 at 2% strain per hour with cell pressure of 500 kN/m2 , 90 days after
mixing: 5%.
The cut-off at Kielder was 19 m deep and 600 mm wide.
Overall, it is rare that the minimum compressive strength of the slurry
proves difficult to achieve with economical cement contents. Caron proposed
an empirical relationship using test results with bentonite from various
sources, and concluded that a 28-day compressive strength R ¼ 104 
ðcement/water ratio)2 kPa was dependable for cement/water ratios between
0.1 and 0.7. Caron commented that deviations from this relationship could
be attributed to the nature of the bentonite, but considered that this con-
tradicted the more reasoned assumption that the cement should have a
more important influence on slurry strength than the colloidal agent that
keeps it in suspension. Tornaghi16 in a review of self-hardening slurries
using Italian cements, commented that the value of the multiplier to the
square of the cement/water ratio could vary considerably, depending on the
nature of the cement, from a value of 103 to more than 104 kPa for cement/
water ratios between 0.15 and 1.
The specified values of minimum deformation at failure, commonly of the
order of 5% axial strain, imply a plastic failure and preclude brittleness from
the failure mechanism, presumably to accommodate lateral ground movement
during the performance of the cut-off. In practice, this value is not difficult to
achieve with normal bentonite and cement contents in drained tests, although
as Tornaghi reminded us, the rate of strain in drained tests on cement–
bentonite mixes is critical in its effect on test values of axial strain at failure.
The permeability of the cut-off material is generally of the order of 108 m/s
after one month, according to Tornaghi, a somewhat higher figure than that
Fig. 2.21. Development over shown by Caron (Fig. 2.21).
time of the permeability of a The effects of mixing methods on the properties of slurries were reported by
bentonite–cement slurry15
Jefferis17 . He concluded that poorly mixed slurries never develop cut-off
properties that are as good as those of efficiently mixed slurries and, for
cement–bentonite slurries in particular, good mixing reduces bleeding of the
fluid material and the permeability of the set material, although some increase
in strength and brittleness also occurs.
The design of cement–bentonite slurry mixes can be varied (as at Kielder)
by the inclusion of cement-replacement materials such as Pulverized Fuel
Ash and ground blast furnace slag for economy. The use of ground clays
other than bentonite can also reduce material cost. Apart from material
costs, the largest cost variables are the provision and type of guide walls
and the means of slurry trench excavation. The risk of ‘windows’ occurring
40 Deep excavations

Fig. 2.22. Plan and cross-


section of cut-off construction
for deep excavation at Caorso,
Italy16

within cut-offs excavated by backhoe should be noted as it can affect overall


cut-off performance detrimentally and lead to wash-out through scour.
Tornaghi described two examples of the use of cement–bentonite cut-offs to
exclude groundwater in deep excavations. The first, installed by Rodio in
1970, was used to cut off groundwater flow through fill, silty sands and
sands and gravels in a deep excavation at a nuclear power complex at
Caorso, near Milan. A plan and cross-section through the excavation is
shown in Fig. 2.22. The cut-off is 450 mm thick and has a mean depth of
23.5 m. The second example, installed in 1982 for a deep excavation at the
nuclear power complex at Montalto di Castro, near Rome, is shown in plan
and cross-section in Fig. 2.23. The cut-off, 800 mm wide, was taken through
silty sands containing silty clay layers and through a conglomerate rock to
obtain a cut-off into clay. The total depth of the wall was 36 m and the
depth of cut-off into the clay was 2 m. The required drawdown of the
groundwater was approximately 20 m, which was maintained by ten deep
wells shown in plan position in Fig. 2.23.
The control of groundwater 41

Fig. 2.23. Plan and cross-


section of deep excavation with
cut-off and ten deep wells at
Montalto di Castro, Italy16

Cut-off construction using grout curtains installed by intrusion


grouting or jet grouting
Where peripheral walls to a deep excavation are required to resist soil
pressures and provide a barrier to the ingress of groundwater, structural
walls are necessary. Where such walls need to be extended in depth to prevent
ingress of groundwater or where the waterproofing of structural wall elements
requires improvement, it may be economical to use grout injections or jet
grouting techniques.
The process of grout injection into soils is used both to strengthen the
soil mass and to reduce its overall permeability. The acceptance of grout by
the soil structure and the travel of the grout within it is governed by the
permeability of the soil fabric. Where soil permeability is insufficient for
grout take and grout pressure is adequate, the grout enters the soil mass as
a grout body and does not permeate between soil grains. Such an effect,
known as claquage or hydrofracture, is the basis of compaction and squeeze
grouting. In contrast, jet grouting is used to destroy the soil structure deliber-
ately and replace it with a soil/grout mixture. Jet grouting can be used in most
soil types but is particularly effective in sands and cohesionless silts. Although
very high jet pressures are used, hydrofracture pressures are not likely to be
reached in jet grouting, the grouted void being formed by the erosive action
of the jets and the high pore-water pressures set up at shallow depth in the
soil near the jet.
The jet grouting process deserves more explanation. The process was origin-
ally introduced in Pakistan by the UK contractor Cementation, but was later
42 Deep excavations

Fig. 2.24. Jet grouting systems: (a) mix-in-place; (b) replacement method

developed by Japanese firms. In the early 1980s the process was imported to
Europe from Japan.
The two systems most frequently used are mix-in-place and replacement
grouting, as shown in Fig. 2.24. Both systems use water or air/water jets
with diameters in the range 2 to 3 mm. The first is the simpler method in
which a single grout jet at a pressure of 300 to 500 bar is rotated during extrac-
tion from a pre-drilled hole. Rotation speed is 10 to 20 rpm and the rate of
extraction is low (about 100 to 150 mm/min). Since grout loss has to be
low, mix-in-place is carried out in a tight hole. Eroded soil brought to the
surface is recirculated with additional grout. To improve the cutting action
a shroud of air (at low pressure, 5 to 7 bar) can be used to wrap the cutting
jet. Columns between 400 and 800 mm in diameter are formed in sands –
more in coarse sands and less in cohesive soils.
The second method, replacement grouting (known originally as the
Kajima–Keller process), uses a triple-tube drill stem to grout from an open
hole. The tube delivers a jet of water at very high pressure (400 to 500 bar)
which breaks up the soil at the periphery of the bore. The cuttings are returned
to the ground surface by this water by direct circulation. Air encapsulation
can also be used in the second method. This air jet improves soil cutting per-
formance and the efficiency of soil cuttings returning to the ground surface.
Grout is then pumped down the third tube at a pressure of 5 to 7 bar to
enter below the eroding jet. The system, cutting and grouting simultaneously,
is rotated and lifted simultaneously at about 5 rpm and 50 mm/min, respec-
tively. Column and panel construction by the triple-jet process is shown in
Fig. 2.25.
The grout holes can be located within 300 mm of existing foundations so
that the jet can undercut them by as much again and bores can be inclined
below existing foundations. Drilling can often be made through existing foot-
ings with inclined bores to treat below them. The process has been used to
depths in excess of 40 m. A wide range of soils above and below groundwater
level can be treated, from gravels to clays, although stones tend to fall to the
base of the column in open gravels.
Typical dimensions and properties of the soil/cement grout mass are shown
in Table 2.9.
Quality control may be applied on site by pre-contract trials of column
construction. Site measurements are made of lift speed and rotation, depth,
pressures and flow rates of grout, water and air. Grout mix quality tests
and core sampling are used where appropriate, and the specific gravity of
The control of groundwater 43

Fig. 2.25. Jet grouting: column and panel construction by the triple-jet process (courtesy of Soletanche–Bachy)

Table 2.9 Typical dimensions


and properties of the Granular soils Cohesive soils
soil/cement grout mass
Diameter (m) 0.8 to 1.8 0.5 to 1.5
Unconfined compressive strength (N/mm2 ) 1 to 10 0.5 to 5
Shear strength (kN/m2 ) 500 250
Permeability (cm/s) 104 to 107 104 to 107

the grout waste slurry is measured. Both cement and cement/Pulverized Fuel
Ash grouts may be used, the latter weaker but costing less than the cement
grout alternative.
Two examples are presented to show the use of jet grouting to improve
grout water exclusion from deep excavations. Figure 2.26 shows the use of
a jet-grouted cut-off to allow a new underground railway tunnel construction
in Rotterdam to pass under the existing Blaak station built on a wall in H-
section steel profiles (the ‘D’ wall) which terminated in the middle of the Pleis-
tocene bed of fine sands at level 24. To allow excavation under the station
without considerably reducing the water level, the H pile wall was extended
downwards by a cut-off in secant jet grout columns to a cut-off in clay at
level 30 using the triple-jet method. The grouting was made from a working
platform at level 8 under Blaak station with a headroom of only 4.3 m. The
groundwater was reduced temporarily to level 0.5 during the jet grouting
works. To avoid risk of instability of the existing station during installation
of the cut-off, the row of jet-grouted columns was offset a small distance
from the line of the H pile wall and the gap between the H piles and the jet
grout secant wall was plugged with a double row of short jet grout columns.
The second example shown in Fig. 2.27 is the use of jet-grouted columns at
the rear of a hand-dug caisson for a peripheral basement wall to a commercial
44 Deep excavations

Fig. 2.26. Example of jet


grouting for cut-off construction
at the Blaak metro station,
Rotterdam: (a) plan of cut-off
walls; (b) jet mix column
installation; (c) cross-section of
deep and shallow jet mix walls
(courtesy of Soletanche–Bachy)

development in Hong Kong. The soil on the west side of the basement was
retained by conventional diaphragm wall construction to depths greater
than 20 m, while hand-dug caissons were used at shallower depths on the
side where bedrock was at depths between 5 and 20 m. Prior to the installation
of the caisson wall a cut-off was necessary to avoid groundwater drawdown
outside the site and the consequent settlement to neighbouring structures.
The cut-off was made by a grout curtain where soils were groutable using
bentonite–cement grout in the first phase followed by silicate gel injection
using tubes à manchette. Where the fine silty sands overlying bedrock were
not easily grouted, the cut-off was made from secant jet mix columns. The
columns, shown in Fig. 2.27, were generally 0.8 m in diameter and a pressure
of 200 bar was used to install them with a cement grout proportion of
350 kg/m3 of soil cement grout.

Horizontal grout curtains


Within a sheeted or walled deep excavation a horizontal grout curtain may
be needed to reduce the vertical inflow of groundwater from permeable
strata below formation level where the peripheral sheeting or walls do not
achieve a cut-off into bedrock or a relatively impermeable stratum. Such a
situation arises where a deep permeable stratum or layered permeable strata
exist. The depth of a peripheral vertical cut-off wall may prove uneconomic
in these soil conditions and the alternative of a box-like cut-off using a
horizontal grout curtain to seal the base of the excavation enclosure may be
attractive.
The control of groundwater 45

Fig. 2.27. Example of jet grouting for cut-off treatment at the rear of hand-dug caissons at Kowloon, Hong Kong (courtesy of
Soletanche–Bachy)

Two options are available. A shallow horizontal curtain may be formed by


jet grouting or, where soil conditions allow, a deeper horizontal curtain may
be formed by intrusion grouting, using either clay–cement (ultrafine grain size
cement can be used) or silicate grouts, depending on soil permeability. It is
often economical to anchor shallow grout curtains with mini piles or deeper
jet-grouted tensile anchors, which reduces the effective span of the grout
curtains and resists the upward groundwater pressure upon the removal of
the significant dead weight of the overburden pressure of the soil above it.
The installation of jet-grouted columns improves the subsoil strength below
46 Deep excavations

formation level and increases the subsoil passive resistance, giving support to
the peripheral walls. The shallow jet-grouted, anchored cut-off therefore has
the prime objective of excluding groundwater, but it can also materially
assist in supporting the peripheral walls. Deeper silicate grout curtains,
which in turn require deeper peripheral walls, may be formed more economic-
ally by intrusion grouting rather than jet grouting where soil permeability
allows. These deeper cut-offs benefit from the dead weight of the retained
soil above the grout curtain to resist the upward groundwater pressure and
probably avoid the need for intermediate anchoring from jet-grouted
columns. Specified maximum leakage rates in the range 2 to 5 litres/s per
1000 m2 of plan area would be typical of this type of construction in cohesion-
less soils. More traditional methods of underwater concreting may prove
economical as horizontal cut-offs, particularly in long, narrow excavations.
Additives are vital with underwater concrete to prevent the cement from
being washed out and to improve flowability, waterproofness and control
setting time. Most recent innovations include the incorporation of steel
fibre reinforcement in tremie concrete plugs.
Mention should also be made of the installation of horizontal curtains by
horizontal drilling and grouting from shafts for special projects.

Design of dewatering Design objectives


systems The objectives of the methods described to control groundwater in the
previous sections, by exclusion (with surface drains or by cut-off walls) or
by removal of groundwater (by drainage or by pumping), are to:
(a) lower the water table and intercept seepage which would otherwise
enter the excavation and interfere with the work
(b) improve the stability of slopes and prevent slippages
(c) prevent heave in the bottom of excavations
(d ) reduce lateral pressures on temporary sheeting and bracing.
Design methods to remove groundwater by pumping, either by various well-
pointing methods or deep wells, will be considered in the remainder of this
chapter. It is assumed that the means of dewatering has been selected, the
soil properties, including permeability, are known or can be estimated, and
the depths of each stratum and the head of groundwater have been deter-
mined. The design will determine the number and spacing of wellpoints or
wells to achieve the required drawdown and the required capacity of pumping
to remove the groundwater yield at the well heads. The design yield, or
quantity of water to be pumped, depends on various factors including the
permeability of the soil fabric, the groundwater source (whether flow is
radial or from a line source such as a river or shoreline), whether the wells
penetrate fully or partly to the base of the aquifer, the plan shape of the
excavation and the plan layout of wells.
The design methods make fundamental assumptions regarding soil and
groundwater flow which are only partly fulfilled on site; therefore, a balance
has to be struck between the results of such calculations and experience, pre-
ferably local experience, of the selected dewatering method. The assumptions
made include the following.
(a) The aquifer extends horizontally with uniform thickness in all direc-
tions without encountering recharge or barrier boundaries.
(b) The aquifer is isotropic, that is, the permeability is the same in all
directions.
(c) The aquifer releases water from storage instantly when the head is
reduced.
The control of groundwater 47

(d ) The pumping well is frictionless and has a very small diameter.


(e) Under steady-state seepage conditions, Darcy’s law applies and flow
is laminar (Darcy’s law states that for a given soil the velocity of
flow is directly proportional to the hydraulic gradient and the soil
permeability, v ¼ ki, where v is velocity of flow, i is the hydraulic
gradient and k is the soil permeability).

Numerical modelling of groundwater flow may be used either with spread-


sheet programs to allow speedy calculation of established design equations or
by groundwater modelling packages which provide design solutions too
complex for traditional computation. These latter methods are described by
Anderson and Woessner18 . The use of these packages is likely to increase as
reliance on computer design increases; their principal application at present
is in the initial design of large complex schemes and the sensitivity analyses
applied to layouts, pump depths and capacities in such schemes.
Standard well formulae still remain a practical, albeit approximate, basis of
design for typical dewatering systems as described here. Relationships for
discharge into a slot from a line source or sources are used with empirical
expressions for radius of influence to model flow quantities to wellpoints or
wells for trenchworks. Formulae for radial flow to a single well are used to
model flow to a whole excavation, often square or rectangular in plan. The
application of two-dimensional flow nets is convenient, although not neces-
sarily accurate, to obtain estimates of quantities of groundwater flow into
sheeted excavations and to verify the stability of the formation soils due to
seepage pressures as a result of the dewatering. The design is therefore a
process of approximation and does not replace either previous experience
or common sense.
Powrie and Preene19 compared flow rate/drawdown relationships com-
puted by equivalent well and infinite slot techniques with finite element
methods, and advised the range of validity of the equivalent well method
for various excavation plan geometries. The same authors compared case
records of 30 dewatering schemes in fine-grained soils20 to assess the validity
of analysis methods in steady-state conditions. They concluded that equiva-
lent well/slot methods could be effectively used to estimate flow ratio for
dewatering systems in these soils providing the method of analysis is
appropriate to the boundary conditions in the field. Close sources of recharge
were found to increase flow ratio considerably, and it was recommended their
effects be modelled by flow net techniques.
Both two-dimensional steady-state flow programs and equivalent well
formulae therefore provide means of estimating flow rates, each with advan-
tages and disadvantages. The importance of accurate selection of a value for
soil fabric permeability nevertheless remains common to both. The CIRIA
report on groundwater control1 provides a tentative guide to the reliability
of permeability estimates from various methods (see Table 2.10), but con-
cludes that some uncertainty of permeability values is inevitable. Powrie
and Preene19 confirmed the widely held practical opinion that the most
reliable permeability estimates are obtained from field pumping tests with
drawdown measured by piezometers. The method of analysis should be
varied according to the aquifer boundary conditions, using the well formula
where the recharge aquifer is not close to the well, and finite element analysis
otherwise.
Where no pumping test is available and the soil has isotropic permeability
and contains less than approximately 20% of silt and clay size particles,
Hazen’s rule may be used to estimate permeability from grading curves,
providing representative particle size test results are available. (Hazen’s rule
48 Deep excavations

Table 2.10 Tentative guide to reliability of permeability estimates by various methods38

Method Notes Reliability

Groundwater control trials Appropriate for large-scale works or Good if appropriately analysed
where observational method is being
used. Costs high but can be offset
against main works

Pumping tests Test a large volume of soil. Provide Good if appropriately analysed. May
information on well yields, water be difficult to carry out and analyse in
chemistry and distance of influence. fine-grained soils (e.g. silt) or if there
Costs high to moderate depending on are different strata with significant
complexity (simple tests can sometimes variations in permeability (e.g. gravel
be very useful) over fine sand). Good instrumentation
is essential, adequate number of
piezometers, etc.
Inverse numerical modelling Uses groundwater monitoring data (e.g. Good if adequate groundwater data
piezometer readings) to back analyse are available and are appropriately
permeability. Costs low to moderate analysed

Tests in boreholes Only a small volume of soil tested; affected by soil disturbance from drilling
Falling head Very prone to clogging of borehole. Very poor
Costs very low
Rising head Prone to clogging or loosening of Poor to moderate. Better results in
borehole base. Costs very low coarser, less silty soils
Constant head Inflow tests prone to clogging. Costs low Poor to moderate. Better results in
coarser, less silty soils
Packer test Normally carried out in rock. Results Poor to good. Can confirm presence of
are hugely influenced by fissure network. fissures depending on fissure spacing
Costs low to moderate

Tests in piezometers and Only a small volume of soil tested; dependent on the design and quality of
standpipes piezometer installation and on any soil disturbance
Falling head Prone to clogging. Costs very low Very poor
Rising head Costs very low Poor to moderate. Better results in
coarser, less silty soils
Constant head Inflow tests prone to clogging. Costs low Poor to moderate. Better results in
coarser, less silty soils
Specialist in situ tests, Can identify small stratigraphic changes Can be good in fine-grained soils (silts
e.g. piezocone, in situ and provide permeability profile with and clays). Can be difficult to use in
parameter depth. Costs moderate coarser soils or weak rocks

Laboratory tests Only a small volume of soil tested; sample disturbance can affect results
Particle size distribution Loss of fines during sampling may lead Very poor, especially in laminated or
(PSD) analysis of bulk to overestimates of permeability. Costs structured soils
samples very low
Particle size analysis of tube Not representative in structured soils or Moderate to good in uniform sands
samples if silt and clay content is more than with low silt and clay content; poor in
about 10 to 20 per cent. Costs very low laminated or structured soils
to low
Permeameter testing, e.g. Soil fabric and structure means sample Good in clays and some silts where
triaxial cell, Rowe size affects results: smaller samples tend minimally disturbed samples, large
consolidation cell, to underestimate in situ permeability. enough to be representative, can be
oedometer consolidation cell Costs low to moderate obtained
The control of groundwater 49

is k ¼ CD210 , where k is permeability (in m/s), D10 is the 10% particle size (in
mm) and the constant C is usually taken in the range 0.01 to 0.0125.) From
tube samples the mean value of permeability from all curves should be
used, but from bulk samples the minimum value is likely to be more reliable.
If the soil contains more than 20% of silt and clay particles or, less
significantly, anisotropic permeability, in situ rising or falling head tests
should be used in boreholes. Providing that an adequate number of tests
are made, the maximum result should be used. The use of permeability results
from laboratory testing is not recommended.
In the next section, well formulae are described as a series of cases depend-
ing upon the source of water, confinement of the aquifer and penetration of
the well. Design methods are then proposed which use either well formulae
or flow net methods to determine the quantity of water flowing towards the
pumps, and hence the number, and size, of pumps required.

Well formulae
Mansur and Kaufman21 estimated discharge and drawdown from various well
configurations for the cases presented below.

Dewatering for trenchworks


Case 1. Partial penetration by a single row of wellpoints of an unconfined
aquifer with gravity flow fed from a single line source (Fig. 2.28(a)). Applica-
tion: narrow trench work, wellpoints to single side, unconfined aquifer, river
or similar line source.
The total discharge Q (in m3 /s) from wellpoints is
  
: : ðH  h0 Þ kx 2 2
Q¼ 0 73 þ 0 27 ðH  h0 Þ ð2Þ
H 2R0
and the maximum residual head hD downstream from the slot is
 : 
1 48
hD ¼ h0 ðH  h0 Þ þ 1 ð3Þ
R0
where x is the length of the trench (m), H is the height of the static water table
(m), h0 is the height of the water table in wells (m), hs is the difference in head
between the outside and inside of the well (it is small, approximately 0.001H),
k is the soil permeability (m/s), and R0 is the distance of the line source, taken
as equal to the radius of influence R0 (m). Mansur and Kaufman stated that
these expressions for Q and hD are valid for the ratio of radius of influence to
the height of the static water table ðR0 =h0 Þ equal to or greater than 3.0. This
ratio covers most site conditions.
Case 2. For artesian conditions, partial penetration of a single row of
wellpoints fed from a single line source. Application: narrow trench work,
wellpoints to single side, artesian conditions, river or similar line source
(Fig. 2.28(b)).
The total discharge (in m3 /s) is
 
kDxðH  he Þ
Q¼ ð4Þ
R0 þ E A
where EA is an extra length factor which depends upon the ratio of slot
penetration to the thickness of pervious stratum, and is obtained from Fig.
2.28(b); he is the head of water at the well above the base of the aquifer,
and D is the thickness of the aquifer. The maximum residual head (in m)
50 Deep excavations

Fig. 2.28. Partial penetration by a single row of wellpoints from a single line source: (a) gravity conditions; (b) artesian
conditions, plot of W=D against EA =D 38

downstream from the slot is


EA ðH  he Þ
hD ¼ þ he : ð5Þ
R0 þ EA
Case 3. Partial penetration by a single row of wellpoints of an unconfined
aquifer with gravity flow midway between parallel line sources (Fig. 2.29(a)).
Application: narrow trench work, wellpoints to single side, unconfined aquifer,
two line sources, say two rivers, a trench midway between them.
The discharge is
  
: : ðH  h0 Þ kx 2 2
Q¼ 0 73 þ 0 27 ðH  h0 Þ : ð6Þ
H R0
Mansur and Kaufman reported that this expression, as in cases 1 and 2, is
based on model studies by Chapman22 for gravity flow from a line source to a
single partially penetrating slot. The model test showed slight irregularities
The control of groundwater 51

Fig. 2.29. Partial penetration by a single row of wellpoints midway between parallel line sources: (a) gravity conditions;
(b) artesian conditions, plot of W=D against  38

and the equation should therefore be regarded only as an estimate of the flow
required to provide the given head reduction.
Case 4. For artesian conditions, partial penetration by a line of wellpoints
midway between two parallel line sources (Fig. 2.29(b)). Application:
narrow trench work, wellpoints to single side, artesian conditions, two line
sources, say two rivers, a trench midway between them.
The total discharge is
2kDxðH  he Þ
Q¼ : ð7Þ
R0 þ D
At distance y from the slot, when y exceeds 1.3D, the head h increases
linearly as y increases and can be expressed as
 
y þ D
h ¼ he þ ðH  he Þ ð8Þ
R0 þ D
where  is a factor dependent upon the ratio of slot penetration to aquifer
thickness (see Fig. 2.29(b) in which W is the depth of the base of the well
below the upper horizon of the aquifer).

Dewatering for a wide trench or narrow rectangular excavation


Case 5. Partial penetration by a double row of wellpoints of an unconfined
aquifer with gravity flow midway between two parallel line sources (Fig.
2.30(a)). Application: trench works with double row of wellpoints, unconfined
aquifer, two line sources, a trench midway between them.
52 Deep excavations

Fig. 2.30. Partial penetration


by a double row of wellpoints
midway between parallel line
sources: (a) gravity
conditions, plots of C1 against
l=h0 and C2 against b=H;
(b) artesian conditions38

Q is the total combined flow from both slots and is twice that for a single
line source (see Eq. (2) in Case 1). The head is
 
C1 C2
hD ¼ h0 ðH  h0 Þ þ 1 ð9Þ
R0
and C1 and C2 are obtained from Fig. 2.30(a).
Note that for large or square excavations, wellpoints will be required on all
four sides of the excavation. A conservative approximation of the pumping
capacity needed can be made by calculating the values of Q separately for
opposite sides of the excavation.
Case 6. For artesian conditions, partial penetration by a double row of well-
points midway between two parallel line sources (Fig. 2.30(b)). Application:
The control of groundwater 53

trench works with double row of wellpoints, artesian conditions, two line
sources, a trench midway between them.
Again, Q is the total combined flow from both slots and is twice that for
single line sources (Eq. (4)). Values of EA from Fig. 2.28(b) are as for case
2. The head hD midway between the slots can be calculated as before from
Eq. (5) (except where the slots are very close; in this case, a conservative
estimate results from the calculation).

Dewatering for square or rectangular plan shape unsheeted


excavations
Forchheimer23 derived a formula from a system of perfect gravity flow wells of
equal length and capacity. This work forms the basis of the design of
dewatering systems based on radial flow to a number of wells.
Case 7. Full penetration by a single well of unconfined aquifer with gravity
flow fed by circular source. Application: square and rectangular plan shape
excavations, unconfined aquifer (Fig. 2.31(a)).
From Darcy’s law it can be shown that
kðH 2  h2w Þ
Q¼ ð10Þ
loge ðR0 =rw Þ
and drawdown ðH  hÞ at a distance r from the well can be obtained from
 
2 2 Q R0
ðH  h Þ ¼ loge : ð11Þ
k r
Case 8. For artesian conditions, full penetration by single well fed by
circular source (Fig. 2.31(b)). Application: square or rectangular plan shape
excavations, artesian conditions.
2kDðH  hw Þ
Q¼ ð12Þ
loge ðR0 =rw Þ

Fig. 2.31. Full penetration by a


single well fed by a circular
source: (a) gravity conditions;
(b) artesian conditions38
54 Deep excavations

and drawdown ðH  hÞ at distance r from the well can be obtained from


Q R
Hh¼ loge 0 : ð13Þ
2kD r
Case 9. Full penetration of circular arrangement of wells in an unconfined
aquifer. Application: square or rectangular plan shape excavations, uncon-
fined aquifer. From Forchheimer’s work it can be shown that for a circular
arrangement of wells
kðH 2  h2e Þ
Q¼ ð14Þ
loge R0  loge a
where Q is the total flow to the circular well array (m3 /s), a is the radius of the
wells from the centre of the circular well array (m), and he is the height of water
above the impermeable stratum at the centre of the circular well array (m).

Design methods
Formulae which estimate flow and drawdown from line sources to slots,
radial flow to a single or circular array of wells, and flow nets used to
model two-dimensional flow conditions, are explained and illustrated in the
remainder of this chapter to solve design problems for dewatering excavations
of different sizes and construction. The design work of estimating pump
resources, pump depths and locations should all be considered as much art
as science, with adequate allowance for in situ conditions which may differ
from those presumed in the calculations.

Dewatering of trench using progressive wellpoint system: design


procedure
(a) Determine the geometry of the trench cross-section, the extent of
drawdown required and the permeability of strata.
(b) Determine the extent of the dewatered length of the trench works.
Assume the dewatered length will need to be twice the excavated
length at any one stage.
(c) Determine the side slopes to the excavation and whether single- or
double-sided wellpoint installation is required.
(d) Check that the wellpoint will partially penetrate the aquifer.

Fig. 2.32. Plot of h0 =H against


hD =H for various R0 =H 38
The control of groundwater 55

(e) Consider the radius of influence R0 for the required drawdown


R0 ¼ 1500ðH  hD Þk1=2 ð15Þ
Check that R0 =H  3:
( f ) If a double line of wellpoints is required, check the operating level at the
line of wellpoints from the relationships h0 =H and hD =H for different
R0 =H (Fig. 2.32). Check the required drawdown at the line of well-
points from the header pipe level to the top of the wellpoint does not
exceed the maximum wellpoint drawdown of 5 to 6 m.
(g) Check the value of hD from the value of h0 chosen for wellpoint tip
elevation from
 
C1 C2
hD ¼ h0 ðH  h0 Þ þ 1 (see below). ð16Þ
R0
(h) Compute the total flow Q (in m3 /s) from
  
: : ðH  h0 Þ kx 2 2
Q¼ 0 73 þ 0 27 ðH  h0 Þ ð17Þ
H R0
for double-sided wellpointing (divide the total flow by two for the flow
from a single row of wellpoints).
(i) Estimate spacing, and hence the number of wellpoints, from nomo-
graph for uniform clean sand and gravel or stratified clean sand or
gravel (Fig. 2.8).
( j ) Calculate flow per wellpoint and check the capacity of the wellpoint
from Fig. 2.33.
(k) Check approximate size of header pipe required for flow calculated per
side (Fig. 2.34).
(l ) Check head losses in the system and calculate loss in the header pipe
from Table 2.11. Allow equivalent pipe length for fittings (valves,
bends and tees). Calculate the total head loss.
(m) Determine the pump size from total head (required drawdown and head
loss) and required capacity (allow at least 50% above calculated capa-
city per side to achieve initial drawdown). Refer to the pump manufac-
turer’s data for vacuum-assisted pumps (see Fig. 2.35 for typical head
capacity and priming time curves). Note that the estimated radius of
influence R0 calculated in this method is the value when equilibrium

Fig. 2.33. Maximum yield of


wellpoints as a function of soil
permeability38
56 Deep excavations

Fig. 2.34. Wellpoints: header pipe capacity and friction losses38

drawdown has been established. Before this, during the initial


drawdown period, the radius of influence will be reduced and the
yield will be greater; both pump and pipework sizes must allow for this.
(n) For single- and multi-stage installations to relieve artesian pressure
from a contained aquifer, calculate R0 using appropriate well formulae:
hD from Eq. (5) and Q is twice that for single line sources (Eq. (4)), and
follow the design procedure for installation with the gravity flow
method above.

Table 2.11 Friction losses in valves and fittings, expressed as a length of straight pipe (in m)38

Type of fitting Diameter (mm)

Open 1.1 1. 4 1. 7 2.0 2.8 4. 3 5. 2 6.4 7. 6


1 . 7 9 10 1 12 2 18 3 24 4 30 5 41 2 48.8
. . . . . . .
4 closed 61
1. Gate valve 1
2 closed 30.5 39.6 51.8 59.5 91.5 122.0 152.0 213.0 244.0
3
4 closed 122.0 159.0 213.0 244.0 366.0 488.0 610.0 854.0 976.0
2. Standard tee Flow in line 2.9 4. 3 5. 0 5.9 9.1 11.9 15.1 22.0 24.7
Flow to/from branch 9 8 12 8 16 8 19 8 30.5 39.6 50.3 73.2 82.3
. . . .
3. Standard 908 elbow 4.9 6. 4 8. 4 9.9 15.2 19.8 25.2 36.6 41.2
4. Medium sweep 908 elbow .
43 .
55 .
67 7.9 12.2 15.9 21.3 28.0 32.0
5. Long sweep 908 elbow 3.2 4. 3 5. 3 6.1 9.1 12.2 15.2 21.3 24.4
6. Square (908) elbow 9 8 12 8 16 8 19 8 30.5 39.6 50.3 73.2 82.3
. . . .
7. 458 elbow 2.3 3. 1 3. 7 4.6 6.4 8.5 10.7 15.2 18.3
8. Sudden enlargement d=D ¼ 4 1 .
49 .
64 .
84 9 9 15 2 19.8 25.2 36.6 41.2
. .
d=D ¼ 2 1
3.2 4. 3 5. 3 6.1 9.1 12.2 15.2 21.3 24.4
d=D ¼ 4 3 .
29 .
37 .
49 .
56 8.4 11.0 13.7 19.8 22.9
9. Sudden contraction d=D ¼ 4 1
2.3 3. 1 3. 7 4.6 6.4 8.5 10.7 15.2 18.3
d=D ¼ 2 1 .
17 .
23 .
29 .
34 .
49 6. 4 8.2 11.3 12.8
d=D ¼ 4 3
1.1 1. 4 1. 7 2.0 2.8 4. 3 5. 2 6.4 7. 6
The control of groundwater 57

1720 r.p.m.

Fig. 2.35. Head/capacity and


priming time curves for Sykes
Univac pump, 150 mm suction
and discharge
58 Deep excavations

Dewatering of square or rectangular plan shape excavation by


single- or multi-level wellpoints
(a) Use well formulae for case 5 for opposite sides of the excavation,
estimating the radius of influence, drawdown and total flow, as detailed
in the section above. Estimate the spacing of wellpoints from the total
flow on each side of the excavation using nomographs as before, and
determine the header pipe dimensions, head loss and size of pump or
pumps, as above.
(b) For multi-stage wellpointing, make successive calculations of draw-
down, calculating the radius of influence, drawdown and total flow
for each stage.
(c) For artesian conditions use well formulae for case 6, calculating in turn
the radius of influence, drawdown, total flow, number of wellpoints,
header pipe size, head losses and pump size.

Dewatering of deep square or rectangular plan shape excavation


with battered side slopes by deep wells
The following method was described by Hausmann24 .
(a) Make an initial estimate of the total quantity of groundwater to be
pumped by replacing the actual excavation with a circular plan shape
of approximately equal area. Use the well formula from case 9 (Eq.
(14)). If the actual excavation is rectangular with length X and width
Y, then
 
XY 1=2
a¼ : ð18Þ

Use the radius of influence R0 from
R0 ¼ 3000ðH  he Þk1=2 : ð19Þ
(b) Estimate the number of wells needed. For an individual well of radius
rw (in m) the discharge quantity is
Qi ¼ 2rw hw kie ð20Þ
where hw is the height of well screen (m), and ie is the average entry
gradient. According to empirical findings ie should not exceed 1=15k1=2
to avoid turbulence and filter instability for spacings larger than 15 well
diameters. Thus, the capacity of an individual well should be limited to
 1=2 
k
Qi ¼ 2rw hw ð21Þ
15
and if h0 , the height of the top of the well screen above the bottom of the
aquifer, is set equal to hw , the height of the well screen, then
 1=2 
k
Qi ¼ 2rw h0 : ð22Þ
15
If the number of wells is n, then n ¼ Q=Qi .
(c) Check the original estimate of he using
kðh2  h2e Þ
Qtotal ¼ ð23Þ
loge R0  ð1=nÞ loge ðx1 ; x2 ; . . . ; xn Þ
where ðx1 ; x2 ; . . . ; xn Þ are the radial distances of the wells from the
centre of the excavation. Solving for he results in a new, improved
value for he . Using this value, a new R0 and Q are computed. Steps
The control of groundwater 59

(a) to (c) are repeated until he assumed is sufficiently close to he calcu-


lated in step (c).
(d ) Return to the original excavation. Distribute the n wells around the
perimeter. Check the water level at critical points below the excavation
(at the centre and at the corners) using Eq. (22). If the water level is too
high, the overall pumping rate Qtotal has to be increased. This in turn
results in a reduced value for he and may result in an entry gradient ie
in excess of the maximum recommended of 1=15k1=2 . If so, increase
the number of wells and repeat the calculation.
(e) Submersible, centrifugal pumps with one or more impellers driven by
an electric motor are used for most deep well installations. The required
pump capacity is
Qhw 3
N¼ m =s ð24Þ

where w is the density of water to be pumped, and  is the system effi-
ciency. Taking into account friction loss in the delivery pipework,  is
usually in the range 0.3 to 0.5. Assuming  ¼ 0:3 and w ¼ 10 kN/m3 ,
Qh
N¼ ðkWÞ ð25Þ
40
where Q is in litres/s and h is in metres.

Dewatering of sheeted excavations where the sheeting does not


achieve a seal within an impermeable stratum or horizontal grout
cut-off
The design of dewatering systems for sheeted excavations is most conveniently
undertaken with the use of two-dimensional flow nets, making provision for
the length of the excavation in the calculation to take into account three-
dimensional groundwater flow. Where the soil type or depth of excavation
vary around the perimeter of the excavation, separate flow net computations
can be made for each length and summated.

Flow net construction


The three-dimensional flow of water through a porous media can be repre-
sented by the Laplace equation which, when simplified into two dimensions,
can be modelled by flow net construction. The flow net consists of flow
lines, or stream lines, which represent an almost infinite number of ground-
water flow paths and intersecting equipotential lines. The family of curves
therefore consists of curvilinear squares. Flow nets can be constructed in
four ways: by computer program, by graphical means, by electrical analogy
or by physical model. The last two alternatives are both expensive and
time-consuming and in practice nowadays can be deleted from the list. On
the other hand, computer programs based on the finite element method or
finite differences provide the most convenient and fastest solution (standard
spreadsheet programs can be used employing finite differences as described
by Williams et al.25 ).
The graphical construction of flow nets was described by Taylor26 and these
methods still provide a practical means of determining the base stability of an
excavation and the groundwater flow rate into the excavation. Whereas com-
puter programs may indicate improved accuracy, care must be exercised by
choice of an appropriate program, understanding its basis, and by correct
input of reliable parameters and model geometry.
In earlier times, the electrical analogy method was described by Montague
and Thomas27 and until the widespread introduction of computer programs
60 Deep excavations

was used in preference to physical modelling, especially for flow net construc-
tion for flow through dams and embankments. The basis of the analogy is that
with electricity the current is proportional to the voltage drop, whereas with
groundwater flow through soils seepage is proportional to head dissipated;
in this analogy, conductivity corresponds to the permeability of the soil.
A flow net may represent flow in plan or in vertical section. In summary, the
rules which apply to the graphical solution are as follows.
(a) Flow lines and equipotential lines intersect at right angles and form
curvilinear squares.
(b) Where the entire section cannot be divided conveniently into squares, a
row of rectangles will remain and the ratio of the lengths of the sides of
each rectangle will remain constant.
(c) A discharge face under atmospheric pressure is neither an equipotential
nor a flow line; therefore, squares are incomplete and flow lines need
not intersect such a boundary at right angles.
(d) In gravity flow systems, equipotential lines intersect the phreatic surface
at equal intervals of elevation, each interval being a constant fraction of
the total net head.
(e) In sedimentary strata it is not unusual for the horizontal permeability to
be greater than the vertical permeability. It is necessary to model this
difference in seepage calculations by the graphical construction of flow
nets with a deliberate scale distortion (this was described by Taylor26 ).
The scale change to the geometry of the vertical section is made prior
to drawing the flow net as follows. If the horizontal permeability is kh ,
the vertical permeability is kv and the transformed section of the natural
horizontal scale is x, the distorted scale x1 required to take into account
the difference in vertical and horizontal permeability is
ðkv =kh Þ1=2
x1 ¼ : ð26Þ
x
( f ) To determine the quantity of seepage, the discharge q per unit width
and the head h at any point can be determined by
Nf N ne n
q ¼ kH1 ¼ kðH  he Þ f ; h¼ H1 ¼ e ðH  he Þ ð27Þ
Ne Ne Ne Ne
where k is the coefficient of permeability of the soil, H is the total head
at entry, he is the head at the flow exit, H1 is the overall net head
ðH  he Þ, Nf is the number of flow channels in the net, Ne is the total
number of equipotential drops between the full head H and head he
at the point of flow exit, and ne is the number of equipotential drops
from the exit to the point at which head h is desired.
A range of flow net constructions is shown in Fig. 2.36. Flow nets can also be
used to model flow through strata having two or more different permeabilities;
the solution to these composite sections was described by Cedergen28 .

Stability of the base of dewatered sheeted excavations


The risk of failure at the base of a sheeted excavation due to seepage below the
sheeting was referred to by Terzaghi and Peck29 and McNamee30 , and is
summarized in BS 8004 and the German Recommendations of the Committee
for Waterfront Structures31 .
McNamee referred to local failure by piping or boiling and a general failure
of the soil below formation level by heave. BS 8004 concluded that, in order to
avoid risk of boiling at formation level in a sheeted excavation in a cohesionless
The control of groundwater 61

Fig. 2.36. Typical flow net


construction

stratum, with pumping from sumps at formation level, the minimum penetra-
tion depths should be as shown in Fig. 2.37. The German Waterfront
Code31 , however, provided more detailed guidance for a similar excavation.
It referred to risk of base failure in a cohesionless stratum in two ways, as
did McNamee. The following summarizes its recommendations, referring to
the specific problem shown in Fig. 2.38. First, the risk of local failure by
piping is regarded as a risk which cannot be accurately assessed by analytical

Fig. 2.37. Minimum values for


cut-off depth for cofferdam
sheeting in cohesionless soils2
62 Deep excavations

Fig. 2.38. Factors of safety


against failure by heave at
formation level calculated
using the exit gradient from the
flow net with a curved failure
surface and a prism-shaped
failure soil body29

methods because of the diversity of conditions, including variations in local soil


conditions. Where these variations are not excessive, the Waterfront Code
concluded that the danger of failure increases in proportion to the overall
head difference inside and outside the excavation, and increases in the presence
of loose, fine-grained non-cohesive or weakly cohesive material in the subsoil,
especially where loose sand lenses occur. The failure risk does not generally
occur in strong cohesive soils. Such a failure is first indicated at the formation
level by ground swelling and ejection of soil particles in the progressive failure
manner shown in Fig. 2.39. An effective preventative measure at an early stage
would be to place several thick layers of granular materials, as a graded filter, to
retain the subsoil but avoiding excessive restriction of water flow. Only if such
measures failed, or were applied too late, would it be necessary to flood the
excavated space to equalize the water pressure.
Although the Waterfront Code31 advised care and observation to counter
risk of piping, that is, the local reduction of effective stress to produce
quick conditions, it is prudent to check the value of hydraulic gradient at
the exit point with the minimum seepage path length. The maximum exit
gradient can be calculated by the use of a flow net, from
ðh=Nd Þ
iexit ¼ ð28Þ
a
where a is the length of the flow element at the formation level, and Nd is the
number of equipotential drops. The factor of safety against piping is icr =iexit
where icr is the critical hydraulic gradient. It can be shown that for zero
effective stress
Gs  1
icr ¼ ð29Þ
eþ1
The control of groundwater 63

Fig. 2.39. Development of


foundation failure by
subsurface erosion31

where Gs is the specific gravity of the soil grains and e is the voids ratio of the
soil. For most soils, icr lies in the range 0.9 to 1.1, with an average value of 1.0.
A minimum factor of safety against piping, icr =iexit , of 1.5 is advisable.
The Waterfront Code31 considered failure risk by heave at the base of the
excavation and (see Fig. 2.38) recommended examination of the stability in
the vertical direction of volumes of soil contained between the inside face of
the sheeting and alternative failure planes. One trial failure plane would be
that recommended by Terzaghi and Peck, enclosing a rectangular block of
soil of depth equal to the sheeting penetration and width equal to half the
sheeting penetration. Other failure planes, including those with curved sur-
faces, would be examined. The vertical stability of the block of soil is assessed
as the ratio of the downward force due to the mass of soil wea above the failure
plane to the vertical force cfl caused by the seepage water. A ratio of at least
1.5 is recommended for safety.
The vertical component cfl of the total seepage pressure may be conveni-
ently calculated by flow net. In this method, the residual hydrostatic pressure
acting normal to the failure surface is plotted at each intersection of the
equipotential lines and the failure surface, the area of the plot representing
the total seepage pressure acting on the trial failure surface. The pressure
at each intersection point may be computed from nhw , where n is the
64 Deep excavations

number of equipotentials at the intersection, h is the total head difference


divided by the number of equipotentials, and w is the density of water.
The risk of failure of the formation to sheeted excavations in soft clays,
which is not related specifically to seepage pressures, is described in Chapter
7 on cofferdam design.

The risk of formation heave or local failure by piping changes


when deep wells are used to remove groundwater below
formation level
Where deep wells are used outside the sheeting, external drawdown is
increased and the quantity of water pumped is increased as compared with
drawdown and flow quantity where the wells are sited inside the sheeting. It
should be noted, however, that where deep wells are installed outside the
sheeting, seepage pressures tend to be reduced on the subsoil between the
sheeters below formation level, whereas with wells that are sited between
the sheeters, seepage pressures tend to increase, and the risk of base instability
and heave increases. In the latter case, where the well screen is installed
between the sheeters, the closer this is to formation level the higher the seepage
pressure and the instability risk, but the drawdown outside the sheeters is less
and the quantity of water pumped is smaller. Where deep wells are used
between sheeters which do not find a seal (with the toe of the sheeters in an
impermeable stratum), the optimum sheeter depth will be defined by five
factors: the risk of instability of the soil below formation level by seepage pres-
sure; the risk of drawdown outside the sheeters causing settlement damage to
existing structures; the pump resources required (and the pump energy con-
sumed) to obtain and maintain adequate drawdown below formation level;
the maximum values of strut or anchor loads, particularly at the lowest
frame level; and lastly, the horizontal stability of soil below formation level
before the blinding strut or slab is constructed.
Where a cut-off cannot be obtained, the designer frequently defines sheeter
depth by the stability requirements of the subsoil below formation level and
the lowest bracing loads. The dewatering scheme design is then applied to
ensure the necessary drawdown below formation level to obtain dry working
conditions with the minimum number of pumps. However, a more refined
approach is advisable, with coordinated sheeting depth, formation stability,
bracing design and required pumping capacity, and would improve economy.

Development of drawdown with time


Dewatering systems are designed on the basis of a steady-state condition being
achieved at the required drawdown. Invariably, construction programmes on
site are tight and the period of pre-drainage to achieve this steady state may
become critical; indeed, extra pump capacity may be needed to avoid delay.
Assessment of the period to achieve drawdown may therefore become impor-
tant. Hausmann24 referred to work by Theis32 , later modified by Jacob33 ,
which gives a solution. Theis provided a non-equilibrium well formula for
gravity flow to a single well based on the following assumptions:
(a) the well penetrates a homogeneous, isotropic, horizontal aquifer over-
lying an impermeable stratum
(b) the Dupuit–Thiem approximation holds; the hydraulic gradient below
any point of the drawdown curve is equal to the slope of the drawdown
curve at this point
(c) the aquifer is not recharged
(d ) water flows out of the pores of the soil simultaneously with the
drawdown of the phreatic surface
The control of groundwater 65

Fig. 2.40. Lowering of phreatic


surface with time24

(e) the drawdown s is small relative to the aquifer thickness h so that ðh  sÞ


is approximately equal to h. This is equivalent to assuming a non-
artesian aquifer of thickness m ¼ h.
Theis, as reported by Hausmann, equated the water flow through a cylindrical
area around the well at a distance x to the volume of water removed from the
soil beyond x due to lowering of the phreatic surface, as shown in Fig. 2.40.
Theis showed that the drawdown is
Q
s¼ WðuÞ ð30Þ
4km
where Q is the quantity of flow, k is the permeability of the soil, m is the thick-
ness of the aquifer, and WðuÞ is the well function and is defined by
u2 u3 u4
WðuÞ ¼ 0:5772  loge u þ u  þ    ð31Þ
2  2! 3  3! 4  4!
and
x2 S
u¼ : ð32Þ
4tkm
The function S represents the storage coefficient of the aquifer and is the ratio of
the volume of drainable water to total soil volume for an unconfined aquifer,
and may be as low as 105 for confined aquifers which remain saturated
during pumping. For unconfined aquifers, it ranges from 0.01 to 0.03. t is the
pumping time. WðuÞ is tabulated in hydrology textbooks (e.g. Driscoll34 ).
Hausmann showed, from Jacob’s modification of the work by Theis, that if
t is sufficiently large, u is small and for u < 0:05 the well function may be
approximated by
2:25 2:25kmt
WðuÞ ¼ 0:577  loge u ¼ loge ¼ loge : ð33Þ
4u x2 S
The drawdown as a function of time is then
Q 2:25kht
s¼ loge : ð34Þ
4km x2 S
To apply Jacob’s formula (for artesian flow) to unconfined flow, make m ¼ h
and use the analogy 2ms ðartesianÞ ¼ h2  y2 (unconfined), which holds for
fully penetrating wells. Thus
h2  y2
s¼ ð35Þ
2h
66 Deep excavations

and

h2  y2 Q 2:25kht
¼ loge : ð36Þ
2h 4km x2 S
Drawdown s can therefore be plotted against t and x, using the geometry of
the section to obtain h and y and using soil property values for k and S.

The use of the observational method


The foregoing methods of design for dewatering systems are known to be
approximate and dependent on soil fabric parameters which themselves vary
spatially within the soil and upon their accuracy of measurement. Whilst this
may be accepted by the dewatering scheme designer, the subsequent risk at
the earliest stages of a construction contract are often mis-judged both in
terms of cost of resources and construction delay. The use of the observational
method permits a rational approach to these uncertainties particularly in
design-and-build contracts where the risks generally are taken by the contrac-
tor (and in turn by the specialist dewatering subcontractor). Examples of the
observational method applied to dewatering schemes are given in Roberts
and Preene35 , Nicholson et al.36 and Preene et al.1
Nicholson explained that the observational method constitutes a way of
minimizing costs by modifying design during construction in two basic
applications:
(a) ab initio from inception of the project
(b) best way out – during construction when unexpected site problems
occur.
In the event, it is likely that the method will find application in the second of
these two options. When site drawdown fails to be achieved the plant
resources can generally be readily increased and the parameters that deter-
mine extent of dewatering are easily determined (such as drawdown and
discharge flow rate). The matter then becomes the timely application of
good engineering sense based on performance.

References 1. Preene M. et al. Groundwater control. CIRIA, London, 2000. Publication C515.
2. BS 8004. Code of practice on foundations. British Standards Institution, London,
1994.
3. Cashman P.M. and Preene M. Groundwater lowering in construction. Spon,
London, New York, 2001.
4. Powers J.P. Construction dewatering. Wiley, New York, 2nd edition, 1992.
5. Preene M. Case studies of construction dewatering in fine grained Eocene soils.
Ground Engng, 1993, 26, Sept., 23–27.
6. Marsland A. Model experiments to study the influence of seepage on the stability
of a sheeted excavation in sand. Ge´otechnique, 1953, 3, No. 6, 223–241.
7. Soubra A.H. and Kestner R. Influence of seepage flow on passive earth pressures.
Proc. Conf. Retaining Structures. Thomas Telford Ltd., London, 1993, 67–76.
8. Cashman P.M. Groundwater control. Ground Engng, 1973, 6, No. 5, Sept.
9. Ward, W.H. The use of simple relief walls in reducing water pressure beneath a
trench excavation. Ge´otechnique, 1957, 7, 134–146.
10. Ambrasseys N. Cut-off efficiency of grout curtains and slurry trenches. Proc.
Symp. on Grouts and Drilling Muds in Engineering Practice. Butterworths,
London, 1963, 43–46.
11. Brauns J. The effectiveness of imperfect cut-offs beneath hydraulic structures.
National German Conf. on S.M.F.E., Nurnburg, 1976 (in German).
12. Telling R.M. et al. The effectiveness of jointed cut-off walls beneath dams on
pervious soil foundations. Ground Engng, 1978, 11, No. 4, May, 27–37.
The control of groundwater 67

13. D’Appolonia D.J. Soil–bentonite slurry trench cut-offs. ASCE J. Geotech. Engng,
1980, 106, Apr., 399–417.
14. Hetherington J. et al. A development in slurry trench technique for cut-off
construction. Ground Engng, 1975, 8, Nov., 16–20.
15. Caron C. Un nouveau style de perforation; la boue autodurcissable. Ann. Inst.
Tech. Bat Trav. Publ., 1973, 311, Nov., 1–39.
16. Tornaghi R. L’experience italienne dans la domaine des parois souples
d’etancheite. Symp. technologie et organisation de l’execution des parois moule´es,
Sofia, 1984.
17. Jefferis S. Effects of mixing on bentonite slurries and grouts. Proc. ASCE Conf.
Grouting in Geotech. Engng, New Orleans, 1982, 62–76. ASCE, New York, 1982.
18. Anderson M.P. and Woessner W.W. Applied groundwater modelling. Academic
Press, New York, 1992.
19. Powrie W. and Preene M. Equivalent well analysis of construction dewatering
systems. Ge´otechnique, 1992, 42, No. 4, 635–638.
20. Preene M. and Powrie W. Steady-state performance of construction dewatering
systems in fine soils. Ge´otechnique, 1993, 43, No. 2, 191–205.
21. Mansur C. and Kaufman R. Dewatering. Foundation engineering. Ed. G. Leonards,
McGraw-Hill, New York, 1962, ch. 3, 241–350.
22. Chapman T.G. Groundwater flow to trenches and wellpoints. J. Inst. Engrs,
Australia, 1956, Oct.-Nov., 275–280.
23. Forchheimer P. Uber die Ergiebigkeit von Brunnen-Anlagen. Des Arch-und-Ing.-
Verein zu, Hannover, 1886, 541–547.
24. Hausmann M.R. Engineering principles of ground modification. McGraw-Hill,
New York, 1990.
25. Williams B. et al. Flow net diagrams – the use of finite differences and a spread-
sheet to determine potential heads. Ground Engng, 1993, 26, June, 32–36.
26. Taylor D. W. Fundamentals of soil mechanics. Wiley, New York, 1948.
27. Montague A. and Thomas A. The use of the Servomex potentiometer to
determine pressures, pressure gradients and exit velocities by the method of
electrical analogy. Works on permeable foundations. Sir M. Macdonald and
Partners, London, 1964.
28. Cedergen H. Seepage, drainage and flow nets. Wiley, New York, 1977.
29. Terzaghi K. and Peck R.B. Soil mechanics in engineering practice. Wiley, New
York, 1967.
30. McNamee J. Seepage into a sheeted excavation. Ge´otechnique, 1949, 1, No. 4,
229–241.
31. EAU 90. Recommendations of the Committee for Waterfront Structures, Harbours
and Waterways. Ernst and Son, Berlin, 6th English edn, 1993.
32. Theis C.V. The relation between the lowering of the piezometric surface and the
rate of discharge of a well using groundwater storage. Trans. Am. Geophysical
Union, 1935, 16, 519–524.
33. Jacob C.E. On the flow of water in an elastic artesian aquifer. Trans. Am. Geophy-
sical Union, 1940, 21, 574–586.
34. Driscoll F.G. (ed.) Groundwater and wells. Johnson Div., St Paul, Minnesota, 2nd
edition, 1986.
35. Roberts T. and Preene M. The design of groundwater control systems using the
observational method. Ge´otechnique, 44, No. 4, 727–734.
36. Nicholson D.P., Tse C.M., and Penny C. The observational method in ground
engineering. CIRIA, London, 1999, Report C185.
37. Roberts T. and Preene M. Range of application of construction dewatering
systems. Groundwater problems in urban areas. Thomas Telford Ltd., London,
1994, 415–423.
38. CIRIA. Control of groundwater for temporary works. CIRIA, London, 1986,
Report 113.

Bibliography Barron R.A. Consolidation of fine grained soils by drain wells. Trans. ASCE, 1948,
73, 811–835.
Bouwer H. Groundwater hydrology. McGraw-Hill, New York, 1962.
68 Deep excavations

Casagrande L. et al. Electro-osmatic stabilisation of a high slope in loose saturated


silt. Proc. 5th Int. Conf. S.M.F.E., Paris, 1961, Vol. 2, 555–562.
Casagrande A. and Poulos S. On the effectiveness of sand drains. Canadian Geotech.
J., 1969, 6, No. 3.
Darcy H. Les Fontaines Publiques de la Ville de Dijon. Dalmont, Paris, 1856.
Dupuit J. Etudes The´oriques et Pratiques sur le Movement des Eaux. Dunod, Paris,
1863.
Forchheimer P. Hydraulik. Teubner Verlag, Leipzig, 1930.
Gray D. and Mitchell J. Fundamental aspects of electro-osmosis in soils. ASCE J.
S.M., 1967, 93, No. 6, 209–236.
Harr M.E. Groundwater and seepage. McGraw-Hill, New York, 1962.
Howden C. and Crawley J.D. Design and construction of the diaphragm wall. Proc.
Instn Civ. Engrs, 1995, 108, Feb., 48–62.
Margason E. and Arango L. Sand drain performance on a San Francisco Bay mud
site. ASCE Speciality Conf. Performance of Earth and Earth Supported Structures,
Lafayette, 1972, 181. ASCE, New York 1972.
Powers J. Notes on the design and selection of wellpoint systems. Short course and
seminar on groundwater analysis and design of dewatering systems, University of
Missouri, Rolla, 1976.
Preene M. and Powrie W. Construction dewatering in low permeability soils: some
problems and solutions. Proc. Instn Civ. Engrs, Geotech. Engng, 1994, 107, Jan.,
17–26.
Roberts T. O. and Preene M. Case studies of construction dewatering in chalk. Chalk.
Thomas Telford Ltd., London, 1990.
Sanger F. and Golder H.Q. Ground freezing in construction. S.M.F. Division, ASCE,
1968, 94, Jan., 131–157.
US Dept. of the Interior Bureau of Reclamation. Groundwater manual. US Govern-
ment Printing Office, Washington, 1977.
Worth R. and Trenter H. Gravity dock at Nigg Bay for offshore structures. Proc.
Instn Civ. Engrs, Part 1, 1975, 58, Aug., 361–376.
Open excavation: side slopes and
3 soil retention

This chapter examines the problems of supporting side slopes or cut faces to
the periphery of wide, deep excavations and to smaller sites where the soil
batters can be accommodated in the available land area. For large cuts
associated with heavy earth moving, the scale and size of these cuts means
cross-bracing would be quite impractical, whereas for smaller sites the
choice between battered or braced excavations is dependent on available
space and cost considerations. The wide, open excavations considered may
be for permanent use or as part of temporary works for construction. The
only criterion for inclusion is that the work is related to civil engineering
and not to an industrial process such as mining or quarrying.

Battered excavations Temporary battered excavations are more economical in direct cost and
construction time compared with retention schemes for soil support, and
find application where the slopes can be accommodated within the site area
and groundwater discharge on to the slopes is small or can be controlled.
The decision to omit a support or retention system can only be made on the
basis of adequate subsoil knowledge and appropriate stability analyses,
however, and should not be made solely on the basis of a cost calculation
and an assumed gradient to the batter.
The stability of a cut slope in granular soils can be determined simply from
knowledge of the angle of shearing resistance for the soil or the soil layers. In
normally consolidated cohesive soils, the stability can be assessed by a repeti-
tive calculation of disturbing and restoring moments for trial potential failure
surfaces, the restoring force being based on quick undrained shear strengths of
the soil on the failure surface. Where possible, it is usual to avoid the effects of
disturbance on the measurement of clay strengths by measuring in situ shear
strengths with a vane apparatus rather than relying on laboratory test values.
For cut slopes within over-consolidated clays, it is necessary to relate the
method of analysis to the time the cut is to remain exposed, to take account
of drainage and the equalization of excess pore pressure generated by the
relief of vertical overburden pressure. While an effective stress analysis
using drained shear strength parameters would apply to the fully drained con-
dition for permanent works, the designer of temporary batters for construc-
tion periods of, say, six months has some difficulty in assessing short-term
slope stability without knowing the rate of pore-water pressure dissipation.
The choice of slope angle in these clays will depend on the consequences of
slope failure. If safety is not endangered by soil slippage, the financial cost
of repairs to the slope can be calculated and compared with the actual cost
of reducing the slope angle to achieve greater stability.
Vaughan and Chandler1 calculated soil failure strengths from short-term
slope failures in various stiff, over-consolidated clays and concluded that
average clay strengths on the failure surface lie in the range 50 to 100% of
the undrained shear strength of 38 mm dia. samples. Short-term slope stability
70 Deep excavations

in over-consolidated clays in practice depends on the precautions taken to


avoid groundwater discharging down the slope (say from permeable strata
above the clay) and rainwater falling on to the slope. Harmful effects on
stability of the swelling of such clays, due to relief of overburden pressure
and the further softening and swelling as groundwater enters fissures, are
good reasons to adopt measures to protect the slope. Simple measures such
as polythene sheet covers or sprayed concrete blinding can prove beneficial
and cost-effective. Slope stability analysis methods were reviewed by
Bromhead2 .

Improving the stability There are six methods for improving the stability of a cut slope:
of slopes
(a) regrading the profile of the slope and, for example, weighting the toe of
the slope locally with a soil berm to counter the disturbing moment
(b) using tensioned ground or rock anchors to increase the effective stress
on the potential failure surface, thereby improving soil strength
(c) intercepting potential failure surfaces with sheet piles, mix-in-place
piles, or jet-grouted columns installed from the face or the top of the
slope
(d ) increasing the effective vertical stress on the potential failure surfaces by
reduction of pore-water pressure by drainage
(e) improving composite soil strength by regrading the slope and the
inclusion of reinforcement to intercept the potential failure surface,
using reinforced soil
( f ) driving soil nails through potential failure surfaces.
These methods are well established, enable the engineer to minimize land take
at the batters to the periphery of a wide excavation, and may produce a steeper
slope with more economy by virtue of reduced excavation quantities. Methods
(a)–(d ) are described by Bromhead2 , Mitchell3 and Barley4 . Methods (e) and
( f ), reinforced soil and soil nailing, deserve fuller explanation within the
context of deep excavations.
In the UK the Department of Transport’s manual for reinforced soil and
soil nailing5 defines these techniques as follows.
. Reinforced soil is the technique whereby fill material (frictional or
cohesive) is compacted in successive layers on to horizontal placed
sheets or strips of geosynthetic or metallic reinforcement.
. Soil nailing is the technique whereby in situ ground (virgin soil or exist-
ing fill material) is reinforced by the insertion of tension-carrying soil
nails. Soil nails may be of metallic or polymeric material, grouted into
a pre-drilled hole or inserted using a displacement technique. They will
normally be installed at a slight downward inclination to the horizontal.
In summary, reinforced soil is built-in fill from the base upwards, whereas
soil nailing is used for cuts in virgin soils working from the ground surface
downwards.

Reinforced soil Although historically several soil reinforcement methods have been exploited
for both civil and military use, the modern form of earth reinforcement was
developed and introduced commercially by Henri Vidal in the 1960s. His
concept was a composite material of frictional soil and a reinforcing strip
which enabled the gravity forces on a wall or slope to be resisted by tensile
forces generated in the strip and thence transferred by friction to the soil.
The first major structure to be built by this method was a retaining wall
Open excavation: side slopes and soil retention 71

near Menton, France, in 1968. By the end of the 1960s, walls were reinforced
with galvanized steel strips placed horizontally in layers and connected to a
shaped sheet metal member which formed the facing unit. In 1970 a cruci-
form-shaped precast reinforced concrete member was introduced to replace
the sheet metal facing, and this form of wall facing is now widespread and
has become the means of identifying Reinforced Earth walls. Over 15 000
structures have been built throughout the world by Vidal’s original company,
the Reinforced Earth Company Ltd, and many more have been built by com-
petitors using similar methods and, in some cases, alternative reinforcement
materials. At the end of 1989 the Reinforced Earth Company listed 9000
retaining wall structures and 2000 bridge abutments in service. Recently,
research into the Reinforced Earth wall has centred on the durability of
the galvanized steel strips, organic coatings for steel strips in aggressive
environments and the effects of seismicity on Reinforced Earth walls. The
development of polymer materials in either fabric or grid (geogrid form) in
the 1970s allowed the introduction of alternative reinforcement materials in
both walls and slopes. Use of these materials, which are visco-elastic, brought
the need for further investigation of reinforcement properties, this time into
short and long-term creep in addition to durability. A typical reinforced soil
wall using polymeric geogrid reinforcement is shown in Fig. 3.1.
A Code of Practice for design and construction of reinforced soil walls was
published by the British Standards Institution6 in 1995 and a draft European
Code of Practice for reinforced fill7 was circulated for comment in mid 2002.
Four types of reinforcement are used in current designs:
(a) a steel strip, typically a galvanized, high-adherence steel strip (by
Reinforced Earth Company); other uses of steel reinforcement consist
of rods, ladders and rods with anchor plates

(a)

Fig. 3.1. Modular block


retaining wall (courtesy Tensar
(b) (c)
International Ltd)
72 Deep excavations

(b) polymer geogrids (by Tensar International Ltd and others); cells and
meshes are also used
(c) polyester bands known as Parastrip (by Terram Ltd)
(d) synthetic fibres mixed in place with soil, known as Texsol.
The strips are listed in order of use in the UK.
The reinforcement strips must have the following characteristics:
(a) a high resistance to tension, a failure mode which is not brittle and only
a very low, limited creep
(b) a high friction coefficient with the backfill material
(c) low deformability under working loads (not exceeding a few percent)
(d) they must be flexible enough not to limit the deformability of the
reinforced soil material and to make construction easy
(e) high durability
( f ) they must be economical.
The cost of reinforcement in a reinforced soil wall is a relatively high propor-
tion, between 20 and 30% of the total plant labour and material costs, and is
therefore a prime issue in terms of overall wall costs.

Types of reinforcement
Metal strip reinforcement
The metallic reinforcing strip currently recommended is galvanized mild steel,
generally 5 mm thick and in standard widths of 40 or 60 mm. The surface of
the strip is ribbed to improve soil reinforcement friction. The strips are
called high-adherence strips. In permanent walls it is usual to allow a sacrifi-
cial thickness of the metal to be lost due to corrosion (from 0.5 to 2 mm
depending on design life and exposure). It is also usual to place dummy
strips through the wall face on aggressive sites in order that strips can be
jacked out occasionally to monitor the rate of corrosion. The galvanized
steel strip and precast concrete facing panel are competitive for vertically
faced walls of medium to large height, due to the high elastic modulus of
mild steel and the low creep properties.

Fig. 3.2. Reinforced Earth wall


during construction showing
reinforcing strips and precast
concrete facing units (courtesy
of Reinforced Earth Company
Ltd)
Open excavation: side slopes and soil retention 73

Fig. 3.3. Terratrel panel for use


in steep slope construction
(courtesy of Reinforced Earth
Company Ltd.)

Figure 3.2 shows a typical cross-section of a Reinforced Earth structure


using metal strip reinforcement and precast concrete panel facing. The
economics in construction time and cost are determined by a simple and repe-
titious construction process. After placing the base course of the facing panels
on a prepared foundation, each additional panel mechanically interlocks with
the previous course. The reinforcing strip and backfill are placed, and compact
backfill is then placed in successive layers. Varying shapes of facing panel are
available to suit aesthetic requirements. The Reinforced Earth Company
developed a technique which has been used in many steep and near-vertical
slopes, using steel strip reinforcement and a soft facing of steel mesh backed
with geotextile, and suits the requirements of an economic, quick-to-build,
soft-faced steep slope able to accept vegetative cover. This technique is
called Terratrel8 . The use of wraparound geotextiles to steepen soil slopes
has often been associated with relatively slow construction methods, with
excessive amounts of plant and labour. Terratrel uses low-cost steel mesh
facing panels in discrete heights between reinforcement levels, allowing a
repetitive erection technique without the need for propping of formwork;
Fig. 3.3 shows a typical panel. The technique allows high-adherence galva-
nized steel reinforcing strips to be fixed directly to the steel mesh facing
unit, which is sufficiently rigid to support itself as a vertical cantilever, fixed
in advance of earth filling. The geotextile used as a backing material behind
the steel mesh facing retains the soil and guards against soil wash-out
before vegetation can become established. Plant growth can be started by
hydroseeding, or turf can be placed behind the mesh as construction proceeds
to give an instant green effect. Steep slopes have been built using a coarse
stone fill behind the mesh to give a deliberate gabion-type finish. In the
UK, an 11 m high, 808 slope of 2000 m2 face area has been built at Leeds,
and lower slopes have been built at Nottingham (the A52 road improvement),
the Elan aqueduct at Birmingham, and the A74 Bogbain bypass in Scotland.

Polymer geogrids
Tensar geogrids are high-strength polymer grids specifically made as tension-
resistant inclusions in soils. The manufacturing process begins with an
extended sheet of polyethylene or polypropylene which is punched with a
regular pattern of circular holes. This sheet is then stretched under controlled
heat conditions so that randomly long chain molecules are drawn into an
aligned state. This process of stretching increases the tensile strength and stiff-
ness of the polymer. The resulting geogrid structure of ribs and bars produces
an effective means of transferring load from soil to geogrid. The prime dis-
advantage of using a polymer grid is the low elastic modulus and high creep
value compared with steel strip. The influence of long-term strength has
74 Deep excavations

been the subject of vigorous research9 , and the influence of load capacity on
the design life of the geogrid can be predicted by extrapolation from labora-
tory curves of strain against load with increasing time at various temperatures.
Details of characteristic strength derived in this way are shown in Fig. 3.4. The
overall effect of this long-term creep feature of the polymer is, however, to

Fig. 3.4. Derivation of


characteristic strength of
Tensar geogrid SR80 from
short-term stress–strain tests
(courtesy of Netlon)
Open excavation: side slopes and soil retention 75

Fig. 3.5. Techniques of wall


construction using geogrid
reinforcement (courtesy of
(a) (b)
Tensar International Ltd)

limit the use of geogrids to relatively modest vertical wall heights (typically 6
to 8 m) in order to remain competitive with metal strip Reinforced Earth
walls. The prime advantages of these geogrids are their resistance to attack
from all aqueous solutions of acids, alkalis and salts encountered in soils,
and high resistance to biological attack. The geogrids are made from polymers
formulated to resist ultraviolet light degradation, and when buried or covered
by vegetation their life is indefinite. Resistance to corrosion is therefore
considerably greater than that of steel strip. In recent years a number of
flexible geogrids consisting of a polyester core rib coated in PVC have entered
the market. Care should be exercised to ensure that adequate junction and
interlock strength and resistance to alkaline environments are provided by
such geogrids.
Geogrid reinforced soil walls can be faced with modular concrete block units
or can allow a façade of masonry or brickwork to be connected to blockwork
with stainless steel ties for aesthetic effect. Full-height or incremental-height
reinforced concrete facing units may also be used with a short starter length
of the appropriate geogrid cast in at the rear of each unit. A further innovation
uses steel stanchions with precast concrete planks secured to a reinforced soil
backfill. These techniques of wall construction are shown in Fig. 3.5, courtesy
of Tensar International Ltd. Wraparound faces using geogrids may also be
used in the same way as metal strip reinforcement to form the face of the struc-
ture, using either temporary external formwork or internal bagwork (see Fig.
3.6). In addition, steel mesh panel facings lined with vegetation matting are
increasingly used for speed and economy with geogrid reinforcement on
steep slopes. A recent example of this technique involved the reinforcement
of a 708 slope, 9 m high and 2 km in length for the Channel Tunnel Rail
Link. Typical uniaxial geogrid details by Tensar are given in Fig. 3.7.

(a) (b)

Fig. 3.6. Geogrid reinforcement used as wraparound reinforcement (courtesy of Tensar International Ltd)
76 Deep excavations

Fig. 3.7. Typical uniaxial geogrid details (courtesy of Tensar International Ltd)
Open excavation: side slopes and soil retention 77

Polyester-reinforced thermoplastic webbing


High-strength polyester-reinforced thermoplastic flat webbing also provides a
tensile reinforcement material for reinforced soil walls, steepened slopes and
mattresses for settlement control. The webbing, known as Paraweb, is manu-
factured in two forms, either bonded in strips to Terram melded geotextile or
woven into mats.
The composite Paraweb and Terram is supplied in rolls 4.5 m wide and up
to 150 m long; Paraweb strip is typically 50 to 80 mm wide with centres
240 mm apart within the Terram. Using variable material properties, a range
of ultimate tensile strengths are available within the range 95 to 125 kN per
metre width.
Polyester and polyaramid bands with a ribbed surface for improved soil-to-
band adhesion are also produced in widths of 90 mm and thicknesses of 6, 7
and 8 mm, with breaking loads per strip of 50, 100 and 150 kN, respectively.
These strips are used as reinforcement in a similar manner to steel strips in
Reinforced Earth walls with precast concrete or metal sheet facing panels.
The plastic strips overcome the problems of durability associated with steel
strips and the disadvantages of long-term creep associated with polyethylene
geogrids. Meshes of these strips are also manufactured under the trade names
Paragrid and Paralink.

Synthetic fibre mixed with soil in situ


The use of a fine diameter synthetic fibre mixed in place within the soil to form
tensile reinforcement has been patented by the French Bridges and Roads
Research Laboratory (LCPC). Contracts have been confined to France,
Italy, Holland and Japan, but since 1985 a number of retaining walls have
been built using Texsol, with typical heights from 4 m to more than 20 m
and front slopes between 458 and 808. Texsol can only be used to enhance
the slope angle to embankments in granular fill materials. Multiple continu-
ous threads are used in the ratio 0.1 to 0.2% by weight of natural soil. The
growth of vegetation to walls built in this way is a natural process, although
problems of erosion to the wall face may arise at times of heavy rainfall prior
to the establishment of the plants.

Application of reinforced soil in excavation works


Retaining walls
Initial projects in the early 1970s for the Reinforced Earth Company were a
23 m high wall on the Nice–Menton highway at Peyronnet, France, the
walls to the coal and ore handling facility at Port Dunkirk, major retaining
walls to highways in California, Quebec and Bilbao, Spain, and an 11 km
long wall built on the Reunion Island in the Indian Ocean.
After transportation works, industrial and commercial sectors are the next
largest area for reinforced soil application. These include crushing and screen-
ing facilities, sloped wall bunkers for coal and ore storage, safety containment
dykes for liquid petroleum gas (LPG) and crude oil tanks, civil and military
projects providing protection against explosions, and hydraulic structures
such as dam spillways, canal and river walls, coastal defence structures,
marine walls and reservoirs.
The use of retaining walls in the UK has been inhibited by patents registered
by Vidal in the late 1960s (for the process of Reinforced Earth) and in the
early 1970s (for the facings). The Department of the Environment obtained
a settlement with the Reinforced Earth Company to build Reinforced Earth
walls on public works in the mid 1980s and, more recently, the patent for
the wall facing has expired. Prior to this, competitor companies were unable
to offer schemes for vertical walls in the UK.
78 Deep excavations

Steep slopes
The economic advantages of steepening soil slopes are self-evident both in
terms of construction cost and land take. Geotextile reinforcement is widely
used to reinforce fill slopes or regraded slopes in cut. These geotextile layers
are placed successively within the slope as the filling is placed and compacted.
As deformation within a steep slope is less critical, it is possible to use poorer
grades of fill material within steep slopes than vertical wall structures.
For slopes between 458 and 908 a face support will generally be required;
if geogrids or Paragrid are being used, the material can be wrapped around
successive lifts of fill, thereby making rigid facings unnecessary. Alternatively,
soil reinforcement by Texsol synthetic fibres may be considered. For slopes of
less than 458 it will not usually be necessary to use wraparound geogrid or
Paragrid, and in such cases the use of a Tensar mat pinned to the slope may
be necessary to protect seeded slopes from erosion before substantial growth.

Design in reinforced soil


In the UK two principal standards apply to design: BS 80066 and the High-
ways Agency document HA 68/945 for highway works. BS 8006 considers
both reinforced walls and slopes separately whereas HA 68/94 considers
only slopes with a maximum angle to the horizontal of 708. Both documents
use limit state design methods.
The design of walls in reinforced soil is based on one of two available
methods known as the two-part wedge (or tie-back wedge) or the coherent
gravity system. The two-part wedge method follows design principles used
for classical or anchored retaining walls. The coherent gravity method on
the other hand is based on monitored behaviour of reinforced soil structures
using inextensible reinforcement over a number of walls corroborated by
theoretical analysis.
BS 8006 refers to both design methods for internal stability but advises that
for inextensible reinforcement (metal strip) the preferred method of design is
the coherent gravity method whereas for extensible reinforcement (polymeric)
the two-part wedge should be used. The different uses are based on the
development of earth pressures within the soil structure; for inextensible
reinforcement it is likely that pressure near the top of the wall will remain
‘at rest’, whereas for extensible reinforcement, active pressure will develop
as strain increases. The Highways Agency code HA 68/94 on the other hand
prefers the use of the two-part wedge method for all reinforced soil structures.
(The French Ministère des Transports specifies the use of the coherent gravity
method of design.)
The design (as recommended in BS 8006) requires consideration of both
external and internal stability for the reinforced mass of soil.
External stability checks are required for:
. bearing and tilt failure, using a bearing pressure based upon a Meyerhof
distribution (see Fig. 3.8)
. sliding along the bases, and sliding forward at the base of the reinforced
soil, either soil to soil or soil to reinforcement contact
. external slip surfaces, to the rear and below the reinforced soil structure
or through and below it.
Settlement must also be checked of the whole at serviceability state.
Internal stability is concerned with the statical stability of the reinforced
block of soil – to ensure that the tensile resistance of each successive layer
of reinforced soil has sufficient tensile resistance to counter internal collapse
due to its own weight and surcharge loading. Consideration is needed for
the local stability of reinforcing elements, sliding on horizontal planes and
Open excavation: side slopes and soil retention 79

Fig. 3.8. Bearing pressure


based on a Meyerhof
distribution6

the stability of wedges of reinforced soil. With cohesive frictional fills it will be
necessary to consider the effect of pore water pressure on stability.
At ultimate limit state, an expression for the tensile force to be resisted by
any element below the top of the structure is used with appropriate safety
factors to check the resistance of that element by rupture or failure of
adherence.
Wedge stability is considered by examining arbitrary trial wedges at succes-
sive depths of elements a, b, c, etc. as shown in Fig. 3.9. For each wedge, the
gross tensile force T required to maintain stability is calculated and then
compared to the frictional/tensile capacity of the elements anchoring that
wedge. The resistance of each reinforcement element layer used is the lesser
of its tensile strength or the adherence of that element.
BS 8006 then examines the use of the coherent gravity method as an
alternative method of internal stability analysis. The method is based on the
following assumptions.
. The reinforced soil block consists of an active zone and a resisting zone
separated by a line of maximum tension in the reinforcement, shown in
Fig. 3.10. The failure mode is similar to that of a mass of cohesionless
soil supported by a rigid wall rotating about the top. BS 8006 assumes
the change from at rest pressure to active values occurs at a depth of
6 m from the top of the wall.
. The reinforcement interacts and interlocks with the fill to provide an
effective overall resistance against pull-out.
As before, in the coherent gravity method the calculations of the vertical
stresses are based on the Meyerhof distribution of pressure. Calculations for
the tensile force are made assuming a logarithmic spiral for the locus of
maximum tension as shown in Fig. 3.11, with a simplification to the geometry
shown in Fig. 3.12 where superimposed strip loads do not apply; consideration
of a further maximum tension line is required where strip loads apply. Tensile
80 Deep excavations

Fig. 3.9. (a) Two part wedge


method: wedge stability
examined for arbitrary trial
wedges6

loads are also calculated at the facings. Reinforcement is then provided, with
appropriate partial factors applied, for the lesser capacity of the reinforcement
element in either tension or adherence. Deformation of the wall face and
settlement is calculated at serviceability state.

Design of reinforced slopes


As the angle of the face of the reinforced soil structure declines from the
vertical or near-vertical face of walls to those of reinforced slopes, the
influence of the retained soil reduces and the proportion of the stability
provided by the reinforcement decreases.
The design methods at limit state are based on limit equilibrium methods.
In a similar way to wall design both external and internal stability are consid-
ered in reinforced slope design. The external stability, as before, is considered
for bearing and tilt failure, forwards sliding and slip failure around the
reinforced block.
Internal stability may be verified at ultimate limit state either by a two-part
wedge analysis (as described in HA 68/94) assuming a bilinear failure surface
as shown in Fig. 3.13, circular or non-circular analyses, log spiral failure
analyses or the coherent gravity method.
The use of the log spiral analysis and the assessment of moment equilibrium
has been described by Leshchinsky and Boedecker10 . A computer program,
ReSlope, has been developed using the method. The use of the log spiral
simplifies the analysis procedure as the de-stabilizing moment is determined
directly. The restoring moment due to the presence of the reinforcement
should equal or be greater than the de-stabilizing moment.
A further failure mode needs to be checked for potential failure surfaces
that intersect reinforcement elements but also pass outside the reinforced
soil mass (Fig. 3.14). An extension of the analysis method, log spiral or
other assumed failure surfaces, is used, and is referred to as ‘compound
stability’ assessment.
Open excavation: side slopes and soil retention 81

Fig. 3.9. (continued) (b) Internal


wedge stability analysis of
simple problem6

Fig. 3.11. Coherent gravity


Fig. 3.10. Active and resisting method: logarithmic spirals
zones separated by a line of used for the basis of calculation
maximum tension in the of maximum tension in the
reinforcements; coherent reinforcement, for strip
gravity method6 loading6
82 Deep excavations

Soil nailing
Soil nailing is the strengthening of existing ground by the introduction of steel
reinforcement or more recently carbon fibre rods into the exposed face. There
are two particular applications: the retention of a vertical or near-vertical
cut soil face, and the stabilization of a soil slope. The soil nailing process is
typically a top-downwards operation, whereas reinforced soil construction
is usually bottom-upwards.
Steel ‘nails’ are usually rods 20 to 40 mm in diameter or small steel angle
sections typically 50  50 mm (the latter are used in particular in a patented
method known as an Hurpinoise wall). The nails are grouted into pre-drilled
holes or driven using a percussion drilling device or high-energy firing device.
The nails are not prestressed and their density is relatively high (1 bar per 0.5
to 0.6 m2 ). The nail length is dependent on slope geometry and soil conditions,
but would typically be 0.4 to 0.5 times the vertical height of the cut soil face.
Fig. 3.12. Definition of
maximum tension line 2,
The soil is progressively excavated in depths of about 1 to 1.5 m and the steep
simplified geometry where exposed soil face is protected by a layer of sprayed concrete reinforced with
superimposed loads do not high tensile steel mesh. The nails are driven or drilled into the face before
apply – coherent gravity or after guniting, depending on risk of soil instability. The process is shown
method6 in Fig. 3.15.

Fig. 3.13. (a) Reinforced slopes: bilinear failure surface at ultimate limit state in two-part wedge analysis.
Open excavation: side slopes and soil retention 83

Fig. 3.13. (continued) (b)


Reinforced slopes: alternative
methods of internal stability
analysis, by circular slip
analysis, log spiral analysis
and coherent gravity method6

For nailed slope protection a covering of topsoil secured between galva-


nized steel mesh layers gives an environmentally friendly finish. Details of a
patented system by the Phi Group are shown in Fig. 3.16. Nailed slopes
with similar ‘green’ finishes have recently been used extensively in the UK
for steepened slopes, for highway widening schemes.

Fig. 3.14. Failure mode with


failure surface intersecting
reinforcement but passing
outside the reinforced soil
mass: two-part wedge analysis
of compound stability6
84 Deep excavations

Fig. 3.15. Execution phases of


soil nailing11

The first soil-nailed wall was completed in 1973 in Versailles, France,


to retain a cut for a railway (Fig. 3.17). The technique is applied most
economically in granular soils and relies on relatively dry ground conditions
for practical installation. The growth of the method in Western Europe has
centred on France. Figure 3.18 shows the extent of the increasing use of soil
nailing, which is explained by its low cost compared with traditional soil
support systems.
In the mid 1980s, the use of soil nailing was advancing so quickly in France
that the state of site and design work was exceeding theoretical knowledge and
proven methods. In part, specialist geotechnical firms found that compared
with general contractors there was insufficient expertise in installation and
little sophistication in equipment to keep the market to themselves. The use
of the process therefore expanded rapidly in the hands of general contractors.
In Germany and the UK the method found less favour, possibly due to less
favourable ground conditions.

Fig. 3.16. Details of nailed slope protection, patented process (courtesy of Phi Group)
Open excavation: side slopes and soil retention 85

Fig. 3.17. First soil-nailed wall at Versailles, Fig. 3.18. Increase in use of soil-nailing in
France in 1972/7311 France, 1972–198912

Following the growth of this method, a four-year national research


programme called Clouterre was commissioned by the French Ministry of
Transport. Twenty-one firms and research centres collaborated under the
technical direction of Professor F. Schlosser. The report incorporates the
results of full-scale trials of nailed cuts and includes directives in both
design and specification11 .

Design methods
For design purposes, a general stability analysis is applicable with potential
failure surfaces either falling outside the volume of reinforced soil, or partly
within it. Figure 3.19 shows the location of the potential failure surfaces.
The potential failure surface which intersects the reinforcement will cause
bending and shear stresses within the reinforcement in addition to the tensile,
pull-out forces. The ability of a typical nail to withstand bending and shear is
dependent upon its stiffness: the stiffer the nail the greater its resistance to
moment and shear in the same way as a pile with an applied horizontal
force and moment at its head. Figure 3.20 shows bending and shear forces
applied to a stiff nail at the failure surface and the analogy thereto. The failure
surface is not the classical Coulomb wedge but the locus of maximum tensile
forces in the nails. This potential failure surface is approximately parallel to
the wall facing in the upper part of the retained height and separates the
soil mass into two zones: the active zone, where shear stresses between soil
and nails are directed towards the facing; and the passive zone, where shear
stresses are directed towards the inside of the soil mass beyond the potential
failure surface.
The stiffness of the nails also modifies the classical Rankine earth pressure
distribution against the wall facing. The soil pressures in the upper part are
near K0 earth pressure at rest, but reduce to less than Ka values lower
down. For retaining walls, two principal methods have been applied, particu-
larly in France:
(a) ‘micropiles’, where the nails are relatively long and have considerable
tensile strength (200 to 500 kN), grouted in pre-drilled boreholes with
relatively wide spacing (about one bar per 3 to 6 m2 )
(b) ‘driven or fired nails’, where the reinforcing nails are shorter and have
less strength in tension (50 to 150 kN), driven or shot into the soil face
with a closer spacing (one bar per 0.5 m2 ).
86 Deep excavations

Fig. 3.19. Locations of potential


failure surfaces in a nailed
wall11

The dimensions of soil-nailed walls depend on the strength of the retained soil.
An indication of this effect is shown in Figs 3.21 and 3.22. For the micropile
solution, the ratio of nail length to the height of vertical facing (L=H)
decreases from about 1.7 to 0.5, and the cumulative length of nails from 4.5
to 1.5 m/m2 when the soil angle of shearing resistance () increases from 258
to 558. For the driven or fired nails, the L=H ratio and the cumulative
length of bars are nearly constant irrespective of the value of : L=H is
approximately 0.5 and the cumulative nail length is from 11 to 12 m/m2 .
The choice between micropile and driven nails depends on cost, available
equipment and the nature of the soil (in terms of ease of drilling or driving
of the nails and pull-out resistance).
An important aspect of wall design is the development of adequate pull-out
resistance to the nail. This resistance, which is the aggregate of friction
between retained soil and the surface of the nails, varies according to installa-
tion method, grouting technique (with or without pressure) and soil type
(cohesive or cohesionless); the optimum conditions are dense granular soils
where dilation occurs during shearing. Where such dilatancy is restricted by
the dense soil mass, high restraining frictional resistance is set up along the
nail. These frictional resistances therefore vary from as high as 600 kPa for
dense granular soils to less than 100 kPa for loose sands. For cohesive soils,
the maximum soil/nail adhesion is less than that for granular soils and

Fig. 3.20. Enlarged distortion of


nail at the failure surface with
moment and shear applied.
Note the analogy of a pile with
horizontal load applied at its
head12
Open excavation: side slopes and soil retention 87

Fig. 3.21. Observed proportions of height of wall to Fig. 3.22. Nailed walls: cumulative length of nails per
length of nails as a function of retained soil square metre of wall face as a function of soil
strength11 strength12

ranges from 50 to 100 kPa, reducing as the soil becomes saturated. With the
French preference for pressuremeter testing, a large number of pull-out
tests were undertaken in the Clouterre research project and limit pressure-
meter pressure and soil–nail frictional resistance were correlated (Fig. 3.23).
In all cases where economical wall design is necessary, a knowledge of pull-
out tests for nails is essential, and before introducing soil nailed walls into
previously untried soil conditions there should be a short test programme
of nail pull-outs.
Movements which occur in other reinforced soil structures (such as
reinforced soil walls) similarly occur in nailed walls and are inherent in the
method due to the relatively flexible structure produced by reinforcing soil.
Displacements which occur during excavation and are largely complete
thereafter vary between H/1000 and H/3000, where H is the wall height.
Figure 3.24 shows some recorded values.

Fig. 3.23. Values of soil–nail surface friction as a function of soil strength: (a) expressed as angle of soil shearing resistance for
various soils; (b) expressed as limit pressure from pressuremeter tests for grouted and driven nails12
88 Deep excavations

Fig. 3.24. Observed range of


values of maximum horizontal
displacement of nailed walls12

Schlosser et al.12 referred to three displacements, H , V , 0 (Fig. 3.25),


and observed H to be approximately the same as V . The length  is defined
as that minimum distance from the back of the wall to the point where lateral
movement 0 reduces to zero. The value of  is a function of soil type (coeffi-
cient K), the inclination from the vertical of the wall , and H, according to
the empirical relationship
 ¼ Kð1  tan ÞH: ð37Þ
The values of H , V and K, from the expression for , are summarized in
Table 3.1.
Schlosser et al.12 reported that corrosion should be allowed for in structures
Fig. 3.25. Definitions of with an expected service life of more than 18 months. In France it is usual to
displacements H , V and 0 12 allow a sacrificial thickness for corrosion of the nails. The total thickness is
calculated so that, given the expected degree of corrosion, an adequate thick-
ness of nail remains at the end of the service life. The Clouterre recommenda-
tions specify degrees of corrosion which include resistivity and soil moisture
content, and these are included in Table 3.2 showing the recommended
extra thicknesses of steel.
The modes of failure of a nailed wall were referred to earlier, and basically
can be compared with a classical wedge failure but with modifications. In all
cases the overall stability of the wall should be checked, that is, over a poten-
tial failure surface which passes to the rear of, and beneath, the nailed volume
of soil to emerge at formation level in front of the wall. A conventional

Table 3.1 Displacements in


nailed walls Soil type Weathered rocks, Sandy soils Clayey soils
stiff soils

H ¼ V H/1000 2H/1000 3H/1000


K 0.8 1.25 1. 5

Table 3.2 Sacrificial nail


thickness allowed with respect Class Soil character Service life Service life Service life
to expected degrees of soil 18 months 1.5 to 30 years 30 to 100 years
corrosion
IV A little corrosive 0 2 mm 4 mm
III Fairly corrosive 0 4 mm 8 mm
II Corrosive 2 mm 8 mm Plastic barrier
I Strongly corrosive Compulsory Compulsory
plastic barrier plastic barrier
Open excavation: side slopes and soil retention 89

Bishop, 0 c0 analysis should be used to check that this overall stability is


adequate.
The design stages comprise:
(a) determination of the critical slip surface and the resisting force or
moment to maintain equilibrium
(b) determination of the tensile and shear load for an initial spacing and
inclination of nails
(c) checking at each level for:
(i) tension in the nail at the slip surface
(ii) pull-out of the nail
(iii) maximum bending and shear in the nail
(iv) bearing failure of soil against the nail.
BS 8006 recommends that design can be made on the basis of a limit equili-
brium approach based on the two-part wedge mechanism or a failure surface
using a log spiral.
Although there is some risk of oversimplification, the graphical analysis
shown in Fig. 3.26 using the two-part wedge may be used to define the max-
imum nail force N to preserve equilibrium. Trial failure surfaces for varying
values of inclination  of the lower wedge are used to calculate the maximum
tension in the nails Na . An overall factor of safety against pull-out of the soil
nails of 1.5 for permanent walls and 1.3 for temporary walls is regarded as
adequate where pull-out tests have been performed on a site of known or
similar ground conditions.
The design methods recommended by the Clouterre project, summarized by
Schlosser12 , comprise an extension of the classical limit equilibrium method
(method of slices) to reinforced soils, and take into account the bending

Fig. 3.26. Graphical analysis to


determine the factor of safety of
wall stability for nail length
90 Deep excavations

stiffness and shear resistance of the nails. It is the mode of failure of the stiff
nail system (micropiles, drilled and grouted) which justifies this more sophis-
ticated analysis. Stiffer nails can break in bending or shear at the potential
failure surface; the soil below the nail can fail in bearing or the nail can pull
out. Schlosser also recommended partial factors of safety to be used in an
overall stability analysis.
In addition to these analyses at ultimate limit state the wall should be
checked for the risk of movement from site measurements of walls in similar
soil conditions. In particular, the proximity of existing structures and services
should be noted and estimates made of the lateral distance to the point of zero
horizontal soil movement.
The design of the facing depends on the tension in the nails at the facing T0
and uniform soil pressure acting on the facing. T0 can be calculated directly
using recommended values of T0 =Tmax given in the Clouterre report, where
Tmax is the maximum tension in the nail in service. The Clouterre report
advises the following relationships:
T ðS  0:5Þ
¼ 0:5 þ for 1 m  S  3 m
Tmax 5
T0
¼ 0:6 for S  1 m ð38Þ
Tmax
T0
¼ 1:0 for S  3 m
Tmax
where S ¼ maxðSV ; SH Þ in metres and Tmax may be estimated as the ultimate
nail pull-out force used in the design. SV is nail vertical spacing, SH is nail
horizontal spacing.

Methods of installation and equipment


The stages in construction of the nailed wall are:
(a) excavation at the face of the wall in layers in 1 to 1.5 m deep increments
defined by anchor spacing; trimming to face of wall, to line and vertical-
ity (or batter)
(b) application of gunite (shotcrete) to the exposed soil face
(c) gunite to be steel mesh reinforced; installation of nails by drilling and
grouting or driving; making good gunite perforation by nails (where
conditions allow, it may be preferred to install nails prior to gunite
application).
Excavation adjacent to the wall face would typically be made by backhoe with
the assistance of a wheeled front-end loader to prepare a temporary platform
at each successive excavation stage. Where the wall face has a batter (a slight
batter of 58 to 108 is not unusual), more work is involved trimming to line and
batter with some hand work. The application of gunite (shotcrete) is by hand-
held hose from a small portable mixer/compressor unit. Water is applied at
the nozzle. Typical gunite thicknesses are 100 to 150 mm applied in successive
layers approximately 50 to 75 mm thick. Reinforcement mesh is pinned to the
first layer of the hardening concrete. There are two alternative reinforcement
methods:
(a) An unprotected metal rod or angle section driven from the soil face
(short nails up to 6 m long may be installed by pneumatic ‘shot-fired’
means). Unprotected nails would only be suitable for temporary or
short-term support unless corrosion is allowed for by a sacrificial
thickness.
Open excavation: side slopes and soil retention 91

(b) A steel rod section placed within a drilled hole and grouted at gravity,
or higher, grout pressure.
Option (b) requires the installation of steel rod nails, typically 32 to 50 mm dia.
bars within bores 100 to 150 mm in diameter. The bores would be horizontal
or with a small declination, and would be installed using rotary air or water
flush. The rig required for drilling the relatively short bores would be a tracked
hydraulic rig similar to that used for ground anchor installation such as the
Casagrande C6 rig. Depending on soil conditions, it would be necessary to
case these relatively short bores, the casing being withdrawn as grout is
introduced. Relatively low grout pressures would be used (less than 5 bar).
Although the drilled and grouted nail solution uses traditional ‘ground
anchor’ expertise, and the nails are spaced at wider centres than the driven
nail option, the cost of drilled and grouted nails is higher than the ungrouted
driven nail solution, providing permanent rather than temporary support.
The stages in construction of nailed, steepened soil slopes are similar to
those for nailed walls. For slopes of considerable height and depending
upon short-term slope stability, the method consists of excavating increments
of steepened cut from the crest of the slope, installing the nails by driving (for
temporary support) or drilling and grouting (for permanent support) and
fixing to the head of each nail a cage of galvanized mesh enclosing either
seeded topsoil or geotextile.

Soil retention: further Three further methods of soil retention using flexible wall construction
wall constructions deserve brief mention: ladder walls, gabions and crib walls.

Ladder walls
While soil nailing provides an integral soil and reinforcement mass, an earlier
installation, the ladder wall, used a facing tied back into fill with rods trans-
ferring their load in friction partly into the fill and partly to an anchorage
block forming a composite inclusion. Invented by Andre Coyne in 1926, the
method was subsequently used for retaining walls and dam construction in
France. Walls up to 20 m high were built using the method for river training
walls, to spillway channels and embankment cofferdam structures. The
method produced a flexible construction which allowed movements in the
wall foundation (for example, due to karstic conditions or due to seismicity)
without detriment to the wall itself. Figure 3.27 shows a cross-section of a dam
construction by ladder wall where the anchorage forms the protection to the
downstream slope of the dam13 .

Gabions
The use of gabions is a traditional method of retaining wall construction
which has some advantages over masonry or reinforced concrete wall
construction:
(a) gabion construction does not require skilled labour in its construction
(b) the permeability of the gabions allows both free drainage and soil
retention
(c) the gabion wall is very flexible and allows considerable distortion before
collapse
(d ) large walls can be built relatively cheaply where the source of stone is
local to the construction site.
The gabion wall can be constructed with a vertical, stepped or sloping front
face and is designed as a conventional gravity wall against overturning, sliding
92 Deep excavations

Fig. 3.27. Cross-section of a


‘ladder wall’ at Conqueyrac,
France13

and excessive base pressure. The steel gabion cage itself can be protected either
by galvanizing or plastic coating. Typical, economical gabion wall heights
range from 5 to 10 m depending on soil conditions. As an aid to stability it
is usual to tilt the face of the wall at a batter of approximately 1 in 10. Overall
economy may be improved by using a front gabion unit tied to a reinforced
soil tendon, in steel or a geogrid of high-strength polymer, as shown in
Fig. 3.28.
The Maccaferri company manufactures a patented, combined gabion box
Fig. 3.28. Tailed gabion wall and reinforced soil system using either zinc- or PVC-coated wire mesh. The
constructed with Tensar facing units, between 0.5 and 1.0 m thick, can be constructed as a single
polymer grid cages and Tensar
geogrid tails to form a
element and used in either a vertical or stepped face, as shown in Fig. 3.29(a)
Reinforced Earth structure and (b). The wire mesh units can be filled with soil inside a geotextile and
(courtesy of Netlon) hydroseeded to produce a green face (Fig. 3.29(c)). An example of a stone-
filled Maccaferri Terramesh wall to steepen an existing slope for a housing
development site in Taiwan is shown in Fig. 3.29(d).
The cost of gabion construction obviously depends on the availability and
cost of local stone, since the cost of the gabion cage is a small proportion of
total cost. As an indication, for medium height walls where stone cost is
limited to £20 per tonne delivered, the unit rate for gabion construction is
approximately £80 per m3 , excluding the cost of backfill behind the gabion.
For walls 8 m high, with stone costing £20 per tonne delivered, and an
allowance of £100 per linear metre for backfill, the unit cost of wall face
area is approximately £170 per m2 (at 2003 prices).

Crib walls
Timber and reinforced concrete crib walls have been used in the UK for soil
support for the last 20 years or more. Treated timber cribs, typically up to
10 m in height can provide a life expectancy of 125 years. Reinforced concrete
cribs can be used secured to reinforced soil reinforcement to improve the crib
wall stability. Alternatively, steel soldiers, universal beams concreted into pre-
drilled holes, can be used at intervals within the cribs to improve stability and
allow crib wall construction to greater heights, typically up to 12 m.
The crib wall is a flexible construction and can withstand some settlement
and movement without detriment. The crib wall is only applicable in dry
subsoil conditions or where the rear of the crib can be drained to prevent
Open excavation: side slopes and soil retention 93

Fig. 3.29. Combined gabion


box and reinforced soil system
known as Terramesh: (a)
vertical face; (b) stepped face;
(c) with a hydroseeded face;
(d) example of a stone-filled
Terramesh wall in Taiwan
(courtesy of Maccaferri)
94 Deep excavations

Fig. 3.30. Typical cross-section


through four-cell wide crib wall
in Tsing Yi, Hong Kong14

discharge through the face. The crib itself is backfilled with granular material
(old railway ballast is typical), but graded granular material is necessary
behind the crib to avoid loss of fines into the crib.
Daley and Thomson14 described practical design and construction relating
to crib walls in Hong Kong, and included a description of a major crib wall
constructed at Tsing Yi in the late 1960s. The wall, shown in cross-section
in Fig. 3.30, was used to stabilize an existing slope with slope angles varying
between 508 and 558. The crib wall was 75 m long and up to 11 m high. At its
highest the wall is concave in plan, but is convex at one end where the fill
height is lower. The considerable thickness of the wall is noteworthy.
Approximate current construction costs for the supply and erection near
London of a 9 m and 6 m high crib wall in reinforced concrete are £300 per
m2 and £240 per m2 of wall area, respectively.

References 1. Vaughan P.R. and Chandler R.J. Design of cutting slopes in over-consolidated
clay. Notes for informal discussion. Institution of Civil Engineers, London, 1974.
2. Bromhead E.N. The stability of slopes. Blackie, London, 1993.
3. Mitchell R.J. Earth structures engineering. Allen and Unwin, Boston, 1983.
4. Barley A.D. Slope stabilization by new ground anchorage systems in rocks and
soils. Proc. Conf. Slope Stability Engineering, Isle of Wight, UK, 1991, 313–
318. Thomas Telford Ltd., London, 1991.
Open excavation: side slopes and soil retention 95

5. Design methods for the reinforcement of highway slopes by reinforced soil and soil
nailing techniques. Design manual 4, section 1, HA 68/94. UK Department of
Transport, 1994.
6. BS 8006. Code of practice for strengthened/reinforced soils. British Standards
Institution, London, 1995.
7. pr EN 14475. Execution of special geotechnical works reinforced fill. British
Standards Institution, London, 2002.
8. Smith R.J. The development of Terratrel for the construction of steep slopes. Int.
Reinforced Soil Conf., Glasgow, 1990, 135–139. Thomas Telford Ltd., London,
1990.
9. McGown A. et al. The load-strain time behaviour of Tensar grids. Proc. Conf.
Polymer Grid Reinforcement, London, 1984, paper 1.2, 11–17. Thomas Telford
Ltd., London, 1985.
10. Leshchinsky D. and Boedeker R.H. Geosynthetic reinforced earth structures.
ASCE J. Geotech. Eng., 115, No. 1, 1459–1478.
11. Project National Clouterre: recommendations pour la conception, le calcul, l’exe´cu-
tion et le contrôle. Ponts et Chaussées Presse, Paris, 1991.
12. Schlosser F. et al. French research programme, Clouterre on soil nailing. Proc.
Conf. Grouting and Soil Improvement, New Orleans, 1992, Vol. 2, 739–750.
ASCE, New York, 1992.
13. Chabal J.P. A novel reinforced fill dam. Proc. 8th Euro. Conf. SMFE, Helsinki,
1983, Vol. 2, 477–480. Balkema, Rotterdam, 1983.
14. Daley P. and Thomson R.R. Some practical aspects associated with the design
and construction of crib walls. Proc. Seminar on Lateral Ground Support Systems,
Hong Kong, 1991, 21–37.

Bibliography Open Battered Excavations


Skempton A.W. and la Rochelle Pl. The Bradwell slip: a short-term failure in London
clay. Ge´otechnique, 1965, 15, No. 3, 221–242.
Wakeling T.R.M. Discussion: time-dependent behaviour of deep excavations in a stiff
fissured clay. Proc. 6th Euro. Conf. SMFE, Vienna, 1976, paper 2.2, 29–30.
Austrian National Committee, Vienna, 1976.

Reinforced Earth
Exxon Chemical Geopolymers Ltd. Designing for soil reinforcement, 1989.
Farrag Kh. and Juran I. Design of geosynthetic reinforced soil structures. Proc.
ASCE Conf. Grouting, Soil Improvement and Geosynthetics, New Orleans, 1992,
2, 1188–1200. ASCE, New York, 1992.
Ingold T.S. Reinforced earth. Thomas Telford Ltd., London, 1982.
Jones C.F. Earth reinforcement and soil structures. Thomas Telford Ltd., London,
1996.
Murray R.T. and Farrar D.M. Reinforced earth wall on the M25 at Waltham Cross.
Proc. Instn Civ. Engrs, 1990, Apr., Vol. 89, 261–282.
Reinforced Earth Co. Reinforced earth retaining walls.
Reinforced Earth Co. Reinforced earth: durability of reinforced earth structures.
Vidal H. La terre armée. Ann. de l’Institut Technique du Bâtiment et des Travaux
Publics, 1966, 19, No. 223–229, 888–938.

Soil nailing
Banyai M. Stabilization of earth walls by soil nailing. Proc. 6th Conf. S.M.F.E., Buda-
pest, 1984, 459–465. Akadémiai Kiado, Budapest, 1984.
Bruce D.A. and Jewell R.A. Soil nailing: application and practice, part 1. Ground
Engng, 1986, 19, Nov., 10–15; Soil nailing: application and practice, part 2.
Ground Engng, 1987, 20, Jan., 21–33.
Cartier G. and Gigan J.P. Experiments and observations on soil nailing structures.
Proc. 8th Euro. Conf. SMFE, Helsinki, 1983, 473–476. Balkema, Rotterdam, 1983.
Gässler G. and Gudehus G. Soil nailing – some aspects of a new technique. Proc. 10th
Int. Conf. S.M.F.E., Stockholm, 1981, 3, 665–670. Balkema, Rotterdam, 1981.
96 Deep excavations

Gässler G. and Gudehus G. Soil nailing – statistical design. Proc. 8th Euro. Conf.
SMFE, Helsinki, 1983, 491–494. Balkema, Rotterdam, 1983.
Goulesco N. and Medio J. Soutènement des sols en deblais à l’aide d’une paroi
Hurpinoise. Tunnels et Ouvrages Souterrains, No. 47, Sept.-Oct., 9–17.
Guilloux A. et al. Experiences on a retaining structure by nailing in moraine soils.
Proc. 8th Euro. Conf. S.M.F.E., Helsinki, 1983, 499–502. Balkema, Rotterdam,
1983.
Juran I. et al. Study of soil-base interaction in the technique of soil nailing. Proc. 8th
Euro. Conf. S.M.F.E., Helsinki, 1983, 513–516. Balkema, Rotterdam, 1983.
Juran I. et al. Kinematic limit analysis for design of soil nailed structures. ASCE J.
Geotech. Engng, 1990, 116, Jan., 54–72.
Medio J. et al. Exemples d’applications de soutènements cloués: dix ans de Hurpi-
noise. Sols et Fondations, 1983, Oct., 27–35.
pr EN 14490. Execution of special geotechnical works: soil nailing. British Standards
Institution, London, June, 2002.
Schlosser F. Analyses et différences de le comportement et le calcul des ouvrages de
soutènement en terre armée et par clouage du sol. Sols et Fondations, 1983, 184,
Oct., 8–26.
Vertical soil support: wall construction
4
Options for sheeting While the intended use of a large underground excavation may well define its
and walling size, shape and location, and therefore the overall method of construction, the
sheeting or walling techniques used at the periphery of these excavations may
be common to a number of underground excavation types. Excavations for
building basements, cofferdams for land-based industrial facilities, cofferdams
for bridge works, cut-and-cover structures and shaft construction frequently
share methods for wall construction to support the peripheral subsoil and
exclude groundwater. The sheeting or walling selected for a particular project
may provide temporary soil support prior to the permanent substructure con-
struction, or it may serve as temporary soil support before being incorporated
into the works as the permanent means of soil retention.
The type of peripheral sheeting used will therefore be influenced by the
substructure construction method (bottom-up or top-down) and will vary geo-
graphically due to soil and groundwater conditions, proximity to the source of
materials and the skills and preferences of local contractors. The sheeting or
walling should be selected after an accurate comparison in terms of construction
time and cost, although it rarely is. Available methods include the following.
. Plate and anchor wall; soil support during construction.
. King post wall, vertical soldiers and horizontal timber or precast
concrete laggings or reinforced concrete skin wall; generally only soil
support during construction.
. Steel sheet piling, pitched and driven, pressed in, or pitched into a slurry
trench excavation; generally only soil support during construction
although increased application for permanent walls in car park construc-
tion and building basements of modest depth. Sheet piling may be
strengthened with steel box pile, I-beam or steel tubes to form combi
or high-modulus walls.
. Contiguous bored piling; both temporary support during construction
and when lined also a permanent wall. Piles may be formed convention-
ally or by deep cement–soil mixing methods.
. Secant piling, hard–hard, hard–firm, hard–soft, provides both tem-
porary and permanent support, internal lining sometimes required
depending on use. Piles may be drilled and concreted conventionally
or by deep cement–soil mixing methods.
. Diaphragm walls
k reinforced concrete cast in situ; both temporary support during
construction and a permanent wall, sometimes with lining.
k reinforced concrete precast; provides both temporary and perma-
nent soil support.
k post-tensioned; provides temporary support during construction
and a permanent wall, sometimes with lining.
. Soldier pile, tremie concrete (SPTC), previously used in the USA with
occasional use elsewhere; providing only temporary soil support
during construction.
98 Deep excavations

Table 4.1 Typical applications of embedded walls31

Wall type Typical excavation Groundwater Comment


depth to formation level control Verticality

Cantilever Propped Temp Perm Typical Optimum

King post walls To 5 m 4.5 to 20 m No No 1 : 100 1 : 100 Not suitable in water-


bearing ground
Sheet piling To 5 m 4 to 20 m Yes Possible 1 : 75 1 : 100 Dependent upon soil
conditions
Combi wall To 10 m 8 to 20 m Hydraulic press
installation to limited
depth
Contiguous bored To 5 m 4 to 15 m No No 1 : 75 1 : 125
piling
Hard–soft secant To 5 m 4 to 15 m Yes No 1 : 75 1 : 125 Possible disadvantage
of inadequate
durability of soft pile
Hard–firm secant To 6 m 4 to 15 m Yes Yes 1 : 75 1 : 125 By continuous flight
auger – limits depth
Hard–hard secant To 7 m 7 to 30 m Yes Yes 1 : 75 1 : 300 Cased installation:
increases expense
Soldier pile/tremie 6m 5 to 15 m Yes Yes 1 : 75 1 : 100 Method used in USA:
concrete lesser extent now
In situ diaphragm Not used 7 to 30 m Yes Yes 1 : 100 1 : 150 By grab
wall Not used 20 to 35 m Yes Yes 1 : 200 1 : 500 By cutter/mill. High
mobilization cost.
Wall depth up to 70 m
Precast 5m 5 to 12 m Yes Yes 1 : 100 1 : 150 Depth limited by wall
diaphragm walls panel weight
Post-tensioned 5m 5 to 12 m Yes Yes 1 : 100 1 : 150 Depth limited by
diaphragm walls tendon geometry

An indication of the depth of excavation and verticality generally achieved by


these methods is given in Table 4.1. In practice, the use of sheet piles is often
precluded by the environmental constraints of noise and vibration although
the use of hydraulic press equipment allows sheet pile installation of
medium length. Piled walls find application particularly in basement construc-
tion of medium depth in favourable drilling conditions and diaphragm walls
are used for larger excavations of greater depth.

Plate and anchor wall The use of a conventional underpinning method, casting a reinforced concrete
by underpinning wall in short lengths and depths along the perimeter of a basement and secur-
ing each individual wall section by ground anchor, is feasible in certain soil
conditions. Figure 4.1 shows a site in Madrid where dry, relatively dense
sand subsoils allowed the excavation to proceed in short lengths without
soil loss on the site boundary, each short length of underpinning being
anchored before further excavation. It is not suggested that the underpinning
replaces the permanent basement retaining wall.
The use of underpinning to retain soil to a basement excavation has been
advanced by the practice of using ground anchors to stabilize a grouted
mass or wall of subsoil. Figure 4.2 shows a basement excavation for
the Commerz Bank in Frankfurt-am-Main. The adjacent building was
underpinned by cement and chemical grouting of the sandy gravel below its
Vertical soil support: wall construction 99

Fig. 4.1. Plate and anchor


method used for basement
excavation in dense dry sand in
Madrid

Fig. 4.2. Basement


construction at Frankfurt-am-
Main: anchored grouted soil
underpinning and anchored
Berlin wall used in sandy,
gravel subsoil (courtesy of
Bauer)

foundations; this block of grouted subsoil was then retained by soil anchors
and steel walings below the building. The photograph also shows the use of
the king post wall method to retain the subsoil to the site boundary to the
street side. A single layer of anchors was used to support the wall of con-
solidated soil, acting as a propped gravity structure, whereas the sheeted
wall was retained by a multi-layer anchor system.
Figure 4.3 shows the combined use of steel strutting and ground anchors to
support a grouted subsoil wall on a subway section in Nuremburg in Bavaria.
Sodium silicate grouting was used to consolidate the cohesionless quartz
sands. The grouted wall was 16 m high and four layers of ground anchors
100 Deep excavations

were used with three frames of steel struts and walings. The anchoring of jet-
grouted soil masses is an alternative application of the plate and anchor
method.

Vertical soldiers and This method owes much of its later popularity to three factors: the compara-
horizontal lagging, king tive cheapness of timber as a structural material; the economy of boring with
post method power augers in certain soils; and the unimpaired access for construction
works that anchoring allows. (See Fig. 4.4, showing a basement excavation
20 m deep in Medinah, Saudi Arabia).
The method consists of boring holes on the wall line, typically at 2 to 3 m
centres, placing vertical steel universal beam or column soldier piles within
the holes and concreting the base of each joist below final formation level.
Lean mix concrete is often used for this purpose. In suitable soils the steel
soldier pile may be driven. When the soldier pile is to be extracted the base
of each joist is supported by gravel backfilled to formation level. As bulk
excavation proceeds, horizontal lagging timbers or precast concrete units
are wedged between the soldier piles. Steel section walings are placed to
take the thrust from the soldier piles to ground anchors drilled at intervals
along the length of the waling beam. Alternatively, each soldier pile is
anchored and a waling is not needed.
In shallow excavations, the horizontal lagging timbers may be replaced by a
reinforced concrete skin wall spanning the soldier piles and cast successively in
short lifts as the excavation proceeds. The soldier piles may be cantilevered or
propped by raking shores.
In stiff clays, the horizontal skin wall can be curved in plan and cast against
a cut face to arch horizontally between the soldier piles.
Use of the king post method is limited to relatively dry ground or dewatered
soils which are capable of self-support as each lagging is secured. Where flow
of groundwater is limited in granular soils, inclined wellpoints may be tucked
between the horizontal timbering. The method is generally only used as
Fig. 4.3. Alternate layers of temporary soil support and requires construction of the permanent wall by
ground anchors and steel conventional methods. The very big advantage secured by the anchored
strutting used to support deep
excavation between chemically king post wall is the extent of the clear working space to the permanent
consolidated sand walls, works which is unimpeded by rakers, shoring, or soil berms. Nevertheless,
Nuremburg (courtesy of Bauer) the method does occupy a considerable thickness of construction which

Fig. 4.4. King post method:


basement excavation in
Medinah, Saudi Arabia
Vertical soil support: wall construction 101

Fig. 4.5. Overall thickness of


basement wall construction
including means of temporary
and permanent soil support,
showing proximity to site
boundary

remains around the site periphery without contributing to permanent soil sup-
port; more important, perhaps, the construction tolerances may be such that it
may not be possible, in deep basements, to use the horizontal laggings as a
back shutter to the permanent wall works. The method becomes more eco-
nomical in some urban areas where the soldier piles can be placed temporarily
outside the site boundary, behind the laggings, and extracted after use.
The use of the resulting king post wall method may not be economical in
terms of the construction width and the plan area occupied. The total width
of the temporary and permanent wall becomes the sum of each of the
following items:
(a) minimum width of the working space between the rear of temporary
wall and site boundary; applicable where the site boundary is, for
instance, the flank wall of an adjoining property
(b) construction tolerance of the temporary wall: typically 1:100 verticality
tolerance for augered piling
(c) width of the temporary wall
(d ) working space for the permanent wall, where the king post wall is not
used as a back shutter
(e) width of the permanent wall together with construction tolerance.
Figure 4.5 shows the total width of the construction.
The method realizes the full consequences of poor workmanship in subsi-
dence, excavation and instability risks. Peck1 showed alternative arrange-
ments for securing lagging to soldiers (Fig. 4.6). Poor contact between
laggings and the excavated soil face may induce movement, and Peck recorded
that although use of the method shown in Fig. 4.6(b) allows the soldier piles to
Fig. 4.6. Methods of be incorporated within the permanent wall reinforcement, loss of soil during
transferring earth pressure
from lagging timbers to soldier installation had caused settlements three times those at a basement where the
piles in Berlin walls: (a) lagging method shown in Fig. 4.6(a) was used in similar conditions. Figure 4.7 shows a
wedged against inside flanges good example of very poor construction standards. In the UK, the practice
of soldiers; (b) lagging set of allowing unsecured laggings to slip between the soldier pile flanges
behind flanges of soldiers; progressively as the bulk excavation continues is a particular example of
(c) grout or mortar filling
between lagging and soil; poor workmanship with an inherent risk of soil movement and subsidence.
(d) contact sheeting secured to Figure 4.8 shows an anchored king post wall with good tolerances and
face of soldiers1 carefully wedged horizontal timber laggings. This excavation, at Raschplatz
102 Deep excavations

Fig. 4.7. Example of poor king


post wall construction29

Fig. 4.8. Deep excavation through sandy Fig. 4.9. Dual method Berlin wall for
silts and silty sands supported by anchored multi-storey car park, Zurich (courtesy of
Berlin wall and dewatered by deep wells, Bauer)
Hannover subway (courtesy of Bauer)

subway station in Hannover, was 24 m deep and the subsoil was silty sand
varying to a sandy silt. Figure 4.9 also shows the use of anchored soldier
piles with horizontal laggings, but with dual methods of construction. The
upper part of the walls had steel H-beams placed into pre-drilled holes, with
concrete slabs spanning horizontally; the lower part was formed by bored
piles by Benoto rig. The maximum depth of the excavation was 28 m, retained
by eight layers of anchors. Note the economy in ground anchors achieved by
diagonal strutting between soldier piles near the basement corner.
Although the method is only feasible in relatively dry subsoils, if the soils
allow economical soil anchoring and dewatering, the king post wall finds
Vertical soil support: wall construction 103

Fig. 4.10. Berlin wall method


using steel soldier piles and
horizontal timber laggings and
wellpoint dewatering for a
sewage pumphouse at
Al Khobar, Saudi Arabia
(courtesy of Bauer)

use in apparently uneconomical conditions. Figure 4.10 shows an excavation


for a sewage water pumphouse at Al Khobar, Saudi Arabia. The groundwater
drawdown from 1 m below ground level to a depth of 11 m was by three-stage
wellpointing and deep wells in a sandy subsoil.
Particular attention is drawn to the risk of failure of anchored king post
walls by the vertical component of the anchor force transferring to the base
of the soldier pile. Vertical settlements of the wall require consideration.
The location of anchorages at re-entrant corners also requires attention
where interaction may occur due to the proximity of anchors from adjoining
walls at the re-entrant angle.
More recently, in the UK the use of vertical soldiers, raker support and
horizontal spanning reinforced concrete skin walls have found favour for
shallow basement excavations in London. Excavation by backacter for
soldiers and the use of trench boxes followed by backfilling and re-excavation
for successive skin wall pours have proved economical where groundwater
levels are low and fill or gravel conditions are reasonably dense. In parts of
Germany a particularly low price is obtained by the use of shotcrete between
and over the vertical soldiers in lieu of horizontal laggings.

Sheet piling The economical use of sheet piling in basement construction is influenced by
five factors.
(a) The ability to withdraw the piling after use. Where city sites require
maximum occupation of the site area by the basement, it is not unusual
for pile extraction to be precluded by the lack of working space for
extraction equipment, cranes and handling space. The difficulties of
extracting piles at an urban basement site can therefore minimize
their re-use despite the use of a back shutter or sheeting lining to
separate the permanent wall construction from the sheet piling.
(b) Depth of the basement. On urban sites the transporting, handling, pitch-
ing and driving of long piles can become onerous tasks. The storage of
104 Deep excavations

Fig. 4.11. Anchored sheet piles


in London clay for a telephone
switching centre, London

long piles on a typical city site may become a serious logistical problem
and costs of storage and moving on site may prove onerous. A typical
use of sheet piles for an urban basement construction is shown in
Fig. 4.11 for a basement extension to a telephone exchange in North
London. Subsoil conditions consisted of London clay from ground
level to a considerable depth below final formation level, although
the upper 15 m section was weathered. Telecommunications equipment
sensitive to vibration was nearby, so hydraulic pile installation by
Taywood equipment was specified. This equipment was ideally suited
to the subsoil conditions, allowing pile installation to proceed without
noise (apart from the power generator) or vibration. The use of ground
anchors and temporary embankments allowed construction works to
proceed largely unimpeded from strutting and raking shores. Unfortu-
nately, the site working space did not allow the extraction of any of the
temporary sheet piles to the basement periphery.
(c) Soil conditions and the ease of pile installation. The section of sheet pile
to be used will depend on the requirements of flexural strength and the
strength to resist driving stresses. Design calculations should therefore
be reviewed in the light of anticipated driving conditions. Costs
calculated by Potts and Day2 for sheet piling, diaphragm walls and
secant piles designed by modern methods showed favourably for the
sheet pile wall used in two underpasses and a basement project. Their
comparisons may prove to be inaccurate, however, due to under-
estimates of the section required for practicality of driving. Use of
high-frequency vibratory equipment to install sheet piles in cohesionless
soils and in soft to firm clays and hydraulic pile presses in clays allows
reduced noise levels during pile installation, but increased vibration or
reduced output result directly from any natural or man-made obstruc-
tions. Hard clays or layers of dense sands may require jetting to assist
sheeter installation. Where obstructions seriously impede progress,
pre-boring may be needed. When hard driving through obstructions
occurs there is risk of torn clutches (perhaps occurring below final
Vertical soil support: wall construction 105

Fig. 4.12. Giken pile press (courtesy of Giken Europe BV)

formation level) causing inadequate groundwater cut-off and severe


reduction in pile flexural strength.
(d ) Noise and vibration. These effects have seriously restricted driven sheet
pile usage in urban areas in recent years despite the use of acoustic
covers to some hammers. The use of high-frequency vibrators to install
sheeters reduces the effects of noise pollution and considerable reduc-
tion in vibration nuisance has been made by improvements to vibrators
by the introduction of a mechanical device, an adjustable eccentric
moment which avoids critical frequency vibrations during the start-
up and run-down phases. Where vibrators are ineffective, for example
in stiff clays, sheet pile installation has been made successfully by
hydraulic presses such as the Taywood Pilemaster and more recently
by equipment developed by the Japanese company, Giken.
The Giken press, shown in Fig. 4.12, consists of a saddle which
clamps on to installed piles and a mast on which hydraulic rams are
mounted to apply vertical force through a chuck to push single sheet
piles vertically into the subsoil. The service crane is only used to lift
the press on to the initial piles and after these reaction piles are
pushed in with the assistance of kentledge, successive single piles are
installed by the Giken press moving from pile to pile without craneage.
The Giken press can install sheet piles through medium dense sands and
stiff clays although either lubrication or high-pressure jetting may be
needed in the stiffer clays. Some increased difficulty may be experienced
in maintaining verticality tolerances compared with conventional
installation in panels, first pitched and then driven. Single sheet piles
tend to twist whilst being jacked in and modern wide sheet pile sections
tend to stretch in their width during installation. The method does,
106 Deep excavations

however, allow close installation of sheeters to existing services and


foundations and, with the ‘Zero-Piler’, hard up against projecting
obstructions. Minimum sheeter section for installation in stiff clays is
LX20 by Corus and using this section sheeters up to 16 m long have
been installed in stiff London clay and Gault clay.
The control of noise and vibration due to piling operations is
addressed in BS 5228 Part 4, 19923 , and a useful summary of published
work on the matter is contained in a British Steel report4 .
(e) Incorporation of the sheet pile wall into the permanent works. Where
permanent groundwater levels are low, it may be possible to use the
sheeters as part of the permanent wall, with suitable allowance for
corrosion risk. The recent use of sealants within the sheet pile interlocks
and protective metal covers to protect the sealant during sheeter instal-
lation has allowed piling producers to give warranties which guarantee
overall permeability of installed walls. Welding of clutches is sometimes
resorted to (but of course cannot be carried out below formation level).
This development has materially increased the use of sheet piles
in permanent construction for basement walls, particularly for car
parking use.
Sheet pile seepage resistance is discussed by Sellmeijer and Shouten6 and in
reports by Arbed7;8 .
Available sheet pile sections are shown in Fig. 4.13(a) and (b).

Combined walls
Where soil and groundwater conditions impose high loads on sheet piles, or
where permanent construction requires the minimum number of bracing
frames, it may be necessary to stiffen and strengthen sheet pile sections to
withstand the higher bending moments induced in the sheeting. A number
of methods are available.
(a) The use of combined walls. Steel circular tubes (‘combi’) or sheet pile
box sections are introduced at regular intervals along the length of
the wall. The combined section is assumed to span vertically and the
equivalent section modulus per linear metre of wall is calculated on
the basis that the deflections of the king piles and the intermediate
sheet piles are similar and the elements work compositely. Combina-
tions of sheet piles, both U and Z sections, with intermediate king
piles are shown in Fig. 4.13(c) and (d).
(b) The use of steel sheet pile sections reinforced with steel beams, such
as the Peine pile, by Arbed, or the high-modulus pile by Corus,
Fig. 4.13(e).
(c) The use of jagged walls: sheet piles with crimped interlocks and joined
by double interlock joint units, shown for U and Z pile sections in
Fig. 4.14. For walls with anchoring or bracing support, stiffeners are
required to the jagged wall units at support levels.
(d ) The use of steel sheet piles, either singly or in doubles, at regular inter-
vals with steel tubular piles, in a similar way to the use of box pile
sections.
(e) I-beams fitted with clutches on their flanges, such as Arbed (flared
flange tips) or Dawson crimped beam (flanges crimped to achieve a
flare).
An example of the use of a combined sheet pile and tubular steel section
wall is shown in Fig. 4.15. The cofferdam, built on the approach to the
River Medway crossing at Rochester, UK, was also used to fabricate the
Vertical soil support: wall construction 107

Fig. 4.13. (a) Properties of steel


sheet piles as provided by
Arbed (courtesy Arbed)
108 Deep excavations

Fig. 4.13. (b) Properties of steel


sheet piles produced by Corus
(courtesy Corus)
Vertical soil support: wall construction 109

Fig. 4.13. (c) Combination


piles: sheet pile properties of
combination section walls with
box piles using U-shaped
sections (courtesy Arbed)
110 Deep excavations

Fig. 4.13. (d) Combination pile


walls with box piles using
Z-shaped sections (courtesy
Arbed)

submerged tube tunnel units for the crossing. The depth of the cofferdam at
the river entrance was approximately 17 m, and it was necessary to provide
only one cofferdam frame as high as possible to accommodate the height of
the units passing under the frame at the time of flooding the cofferdam and
floating the units to their final location in the river crossing. Due to the
flexural strengths required from the sheet piles, a combined section of Larssen
sheeters and 1.2 m dia. steel tubes was used, the pile interlock being continu-
ously welded to the steel tube at its junction with the sheeter. The construction
sequence for driving the wall through alluvial deposits into chalk required
extremely onerous verticality and positional tolerances for the steel tubes.
The tubes, driven by vibrators from either river craft or on land, were installed
prior to the sheeters using either leaders or a mechanical template device. Con-
siderable care was required to ensure that the intermediate sheeters, driven
later, clutched for this whole length (22.5 m) between the 30 m long tubes.

Contiguous bored The use of low-cost augers and, more particularly, continuous flight auger
piling (CFA) rigs to drill successive but unconnected piles provides an economical
wall for both temporary and permanent use for excavations to medium
depth where soil conditions are amenable. The contiguous wall can only be
used where groundwater is not a hazard or where grouting or jet grouting
can be used to prevent leakage between the piles. The risk of ‘windows’
between adjacent piles can be assessed by applying maximum verticality
tolerances quoted for auger and CFA pile installation, typically 1 in 100
with depth, and allowing this maximum tolerance to adjacent piles in opposite
directions. Where soil and groundwater conditions are favourable, bored piles
can be installed as anchored piles with the space between them shotcreted to
provide soil support as excavation proceeds.
The nominal diameters of CFA piles are 300, 450, 600, 750, 900, 1000 and
1200 mm. Since grout or high slump concrete is introduced through the auger
Vertical soil support: wall construction 111

Fig. 4.13. (e) Peine pile walls: combination pile walls with steel joists, Z-section sheet piles and connectors (courtesy Arbed)
112 Deep excavations

Fig. 4.14. ‘Jagged’ pile


combination sections. Note the
use of an ‘Omega’ section to
join sheet piles delivered to site
in double units crimped
together (courtesy of Arbed)
Vertical soil support: wall construction 113

Fig. 4.15. Combined tubular


steel and steel sheet pile wall,
River Medway crossing, UK
(courtesy of Carillion)

stem, CFA injected piles do not require top casings during the grouting or
concreting process. The depth of wall constructed by this method is, however,
limited by the length of reinforcement cage that can be introduced through the
cement or grout. A commonly quoted maximum length of 18 to 20 m is
dependent upon ground conditions, granular soils detracting from ease of
construction of CFA piles of this length. Some risk exists of loss of cement
from CFA piles constructed in granular soils. A vibrator can be used with
an H-pile mandrel to insert reinforcement cages in piles longer than 12 m
and steel joist reinforcement can be used in lieu of reinforcement cages for
long piles, although this may prove expensive.
The width of the gap between piles is varied according to ground conditions
between 50 and 150 mm, typically 100 mm.
The use of in-cab instrumentation (of machine torque, concrete volume,
concrete pressure with depth) is general and is necessary to give confidence
in CFA pile construction. At present, concrete pressure is not measured at
the critical position of the point of discharge but in time no doubt this will
also be corrected. Recent IT developments include the transmission of in-
cab readouts to off-site locations in real time.
Inclined pile walls installed by mechanical auger should be mentioned.
Sloped sheeting raked towards the centre of the basement may reduce strut
loading. Schnabel9 reported that where sheeting had sloped at an angle of
about 108 from the vertical, the measured strut loads were consistently less
than two-thirds of the computed strut loads for vertical sheeting in the
same ground.
The principal advantages of contiguous pile walls are:
(a) low cost and speed of construction for temporary and permanent soil
support where drilling conditions are conducive
(b) cleanliness and comparative quietness of the installation process; low
level of vibration during pile installation
(c) for small excavation depths, the ability to minimize the distance
between the wall and existing walls
(d ) the range of soil conditions in which CFA piles can be used is wide:
granular soils are suitable with few exceptions; cohesive soils are
suitable except where penetrations greater than 8 m into hard clay are
required; intermediate c– soils are suitable; soft clays, where cu is
less than 10 kN/m2 , or weak organic soils, are unsuitable due to shaft
bulging; soft rocks, e.g. soft marls and chalk, are suitable (although
relatively small penetrations into rock chalk and hard mudstones can
cause problems); but hard rocks are not suitable.
114 Deep excavations

Fig. 4.16. Mini pile rig suitable for contiguous bored pile installation close to existing
structures (courtesy of Ingersoll Rand)
Vertical soil support: wall construction 115

The principal disadvantages of contiguous pile walls are:


(a) the risk of groundwater ingress between piles
(b) limitations of the depth of CFA piles: 18 m long cages are the practical
limit without special installation measures, although steel joist sections
can be used to reinforce deep piles
(c) relatively low efficiency of circular cross-section of pile in bending
(d ) where independent lining is not needed, the extent of additional works
needed to form an acceptable surface to the wall. A structural
reinforced concrete wall, tied to the contiguous piles may be required
or, elsewhere, sprayed concrete may suffice to fill the gaps between piles
(e) the minimum distance between the wall and existing structure is large
for deeper walls.
So, CFA rigs, without the need for temporary casings, generally provide a
cheap wall quickly in granular and clay soils and are not affected during instal-
lation by the presence of groundwater. But depths are limited to small to
medium depth excavations. Rotary rigs can develop torque greater than
50 T m and with this equipment walls can be built with casings to medium
to large depths. Vertical tolerances, however, may limit depth, and layers of
harder rock will slow progress and increase cost. The walls are not waterproof
since groundwater leakage occurs in the gaps between piles.
The development of mini rotary rigs has permitted the construction of small
(up to 300 mm) diameter contiguous pile walls within a minimum axial dis-
tance of 250 mm to an existing structure. A typical rig is shown in Fig. 4.16.
These rigs are also suitable for pile installation in conditions of low headroom.
Large rotary drilling rigs equipped with rock roller core barrels can be used to
excavate in hard rock with unconfined compressive strengths of 150 MPa
although the rate of drilling will be dependent on rock structure.

Secant piles The principal disadvantages of contiguous pile walls – the gaps between piles
and the resulting problems of lack of waterproofness – are effectively over-
come by interlocking or secant piles.
The method consists of boring and concreting primary, or female, piles at
centres slightly less than twice the nominal pile diameter. Secondary, or
male, piles are then bored at mid-distance between the female piles, the
boring equipment cutting a secant section from them. Male piles are bored
through female piles before the concrete has achieved its full strength;
should this operation be delayed, wear on the auger or the cutting edge to
the casing is likely to be much increased. Concrete quality control is therefore
important in this respect and deviations in maximum strength may be as
important as minimum strength deviation.
As for contiguous pile walls, the secant pile method suffers from the poor
efficiency of the circular cross-section in bending. In some instances reinforce-
ment may be bunched at opposite sides of the pile diameter to ensure
maximum effectiveness, but the risk of rotation, twist or displacement may
preclude this in longer piles. It is usual to reinforce only the male piles, parti-
cularly because of the risk of cutting reinforcement in a displaced cage when
the male pile is being bored. Some deep walls (in London and elsewhere secant
walls have been constructed to depths of 40 m) have utilized steel beam
sections placed into the male pile bore as reinforcement, and either rectangu-
lar reinforcement cages or steel beam sections may be used in female piles in
deep walls where such reinforcement is needed for flexural strength. Weaker
grades of concrete can be used in female piles when wall flexural strength is
not critical. Advantages are gained in both the higher output of male pile
116 Deep excavations

Fig. 4.17. Hard–soft secant pile


wall construction, London

installation allowing rigs of lower torque to bore the female piles. This form of
construction, known as hard–soft secant piling, may suffer from lack of
durability of the soft piles for permanent construction. Some difficulty may
be experienced in the supply of ‘soft’ concrete from off-site ready mix plants
and any slightly lower material cost of the ‘soft’ concrete is lost if the soft
mix has to be mixed on site. The early strength of the female piles requires con-
trol in hard–soft secant pile work and this is usually achieved with concrete
mixes using cement–bentonite or PFA. A typical hard–soft secant wall used
as a car parking basement in London is shown in Fig. 4.17. The wall was tem-
porarily anchored during construction and corner cross-braced. A reinforced
concrete lining wall was applied to provide permanent water resistance.
The use of hard–firm secant piles is made to secure durability and particu-
larly to give permanent resistance to water penetration through the female
piles. Concrete with a characteristic strength in the range 10 to 20 N/mm2
at 56 days with retardation is used to allow easier cutting of the secants
than for hard–hard piles. CFA rigs normally are used. Pile spacings must
allow for adequate secant width to ensure water resistance taking into account
pile installation verticality tolerances. Typical hard–firm pile spacings, centre-
to-centre of male piles are 900 mm for 600 mm dia. male and female piles and
1150 mm for 750 mm dia. male and female piles.
Hard–hard secant walls are installed using casings with high-torque rigs for
deep piles and with CFA rigs for piles of lesser depth. The high-torque rigs use
thick temporary casings, jointed with flush mechanical locks, installed and
extracted either by the rig itself or by casing oscillators attached to the base
of the rig. The maximum depth of installation, in the range 30 m to 40 m
(exceptionally) is dependent upon the torque required to rotate the casing
and the verticality tolerance needed to ensure the secant depth. Typical
spacings of piles to provide adequate secant depth are 650 mm for 750 mm
dia. piles, 760 mm for 880 mm dia. piles and 1025 mm for 1180 mm dia. piles.
Typical hard–hard reinforcement details are shown in Fig. 4.18 (after
Sherwood et al.10 ).
Junctions with floor and raft slabs can be made with secant pile walls in a
similar way to diaphragm walls using bend-out bars or bar connectors and
Vertical soil support: wall construction 117

Fig. 4.18. Hard–hard secant


pile wall reinforcement
details10

polystyrene blockwork secured to the reinforcement cage. Where pile re-


inforcement cages twist within the bore some difficulty may be experienced
in keeping the box-out in the correct position. Where I-beams are used to
reinforce female piles, a shear plate can be welded before installation and a
shear ‘shoe’ attached upon exposure.
Inclined walls in secant piling may also be installed at a small inclination to
the vertical. In Europe, sloping walls raked away from the basement excava-
tion to increase basement width have been used to dramatic effect. Figure 4.19
shows a typical installation.
The spacing, centre-to-centre, of secant piles remains a matter for judgement
by the piling contractor. Figure 4.20 shows support to an excavation for the
Piccadilly Line underground railway extension to Heathrow Airport, where
880 mm dia. piles through Thames gravel were spaced initially at 770 mm
centres, but in severe groundwater conditions the pile spacing was reduced to
750 mm centres to ensure weatherproofness. Although an internal lining was
intended to this wall, the uniformity of finish on the surface of the piles was
sufficiently good to leave them exposed without lining in the running tunnels.
The latest rigs, with usable torques in the range 11 to 60 T m, offer the
following advantages:
(a) a range of diameters of cased secant piles from 600 to 1800 mm is
available with production rates increasing with reducing diameter
(b) difficult conditions including obstructions can be overcome without
chiselling
(c) the introduction of CFA construction methods for hard–hard secant
walls in the range 410 to 1200 mm in diameter
(d ) the use of CFA equipment to bore hard–soft and hard–firm walls.
The secant method therefore provides a waterproof wall which can be built
to a considerable depth, in the range 30 to 40 m, and can cope with most types
of obstruction. The hard–soft version is limited to shallow excavation depths
because of the reduced flexural strength of the wall and some doubt concern-
ing the long-term durability and waterproofness of the soft piles.
Modifications to rotary equipment have minimized the working space
required between the rear of the secant pile wall and the site boundary. For
relatively shallow walls, up to 12 m deep, Bauer equipment, a modified
118 Deep excavations

Fig. 4.19. Inclined secant pile walls: (a) construction sequence; (b) Benoto rig (working dimensions)

Fig. 4.20. Secant piles,


Piccadilly Line underground
railway, Heathrow Airport
(courtesy of Lilley)
Vertical soil support: wall construction 119

Fig. 4.21. Typical guide wall


construction for secant pile wall

BG11 rig, permits installation of 410 mm dia. piles with an axial spacing from
an existing structure of 305 mm. In 1988 at Staines, UK, Bachy–Bauer
installed 410 mm dia. hard–hard walls using such a rig only 100 mm from
an existing wall. The pile diameter was restricted and so, therefore, was the
excavation depth. The piles were reinforced with universal beam sections.
Guide walls are needed for all forms of secant pile construction, and
mention should be made of the cost and time liability this represents. Typical
guide wall construction is shown in Fig. 4.21.
Closely spaced anchors, or struts, are used to support secant piling using
steel waling members or in situ concrete bearing pads. Secant pile walls
may also be used to carry vertical loads. Figure 4.22 shows the method of
use of secant piling anchored to deadmen below existing buildings.

Fig. 4.22. Secant pile wall


anchored to deadmen below an
existing structure (courtesy of
Bauer)
120
Deep excavations

Fig. 4.23. Secant piling, British Library, London: (a) plan; (b) construction
sequence; (c) cross-section; (d) typical cross-section of secant wall
construction (courtesy of Lilley)
Vertical soil support: wall construction 121

The hard–hard secant pile method has been used successfully in London for
the construction of a deep basement to the British Library (Fig. 4.23). The
basement extends to four floors, with the maximum depth 24 m below
ground level. Underground railway tunnels are bridged by a concrete raft at
the centre area of the site where the basement depth is limited to two floors.
The maximum depth of the secant pile is 30 m. Due to the high sulphate
content in the subsoil and the need to minimize the generated heat of
hydration caused by high cement contents, the concrete mix to the piles
incorporated 85% cement replacement with a ground granulated furnace
slag and an aggregate of maximum size 40 mm. Although a diaphragm wall
1.2 m thick was originally intended for use at the site, the existence of large
vertical service ducts at close centres adjacent to the wall caused high design
values of shear at the junction of floor slab and wall. The large steel joist
sections which reinforce the male piles of the secant wall ideally resist these
shear forces.

Diaphragm walls The first diaphragm walls were tested in 1948 and the first wall was built, using
bentonite slurry as a means of support, by Icos in Italy in 1950. That contract
was for a cut-off wall on a dam site. Structural walls followed, and in the late
1950s the construction of the Milan Metro allowed Icos to use diaphragm
walls extensively in cut-and-cover construction; a method of top-down work-
ing known as the Milan method was developed. The first structural diaphragm
wall in the UK, at Hyde Park Corner, London, in 1961, was also built by Icos
using its then patented process. Many contracts followed for Icos in Europe.
These contracts, to build basements, underpasses, marine structures and
metro works, were all excavated using rope grabs suspended from tripod
rigs and, later, from crawler cranes.
The development of diaphragm walling was encouraged by competition
among specialist international firms from the mid 1960s. These firms, includ-
ing Icos in Europe and the USA, Soletanche and Bachy in France, Trevisani
and Rodio in Italy, and Cementation in the UK, all developed excavation
systems of their own. Two preferences quickly became evident: rope grabs
suspended by rope either from tripod rigs or cranes; or grabs, generally
hydraulic powered, mounted on kelly bars mounted on cranes. These two
methods of excavation continued into the 1980s. The major problem which
these methods failed to solve was the excavation of rock. In the 1970s,
rotary percussive tools using reverse circulation were used to excavate,
albeit slowly, into hard rock.
In the early 1970s the reverse circulation system was used in Japan for a
new generation of rail-mounted rigs, manufactured by the Tone Drill Co.
Although these rigs were adequate for excavation in granular soils (but
less so in cohesive soils), they were the forerunners of modern reverse
circulation rigs and used shaker screens and cyclones to remove cuttings
from the slurry in a similar way to the latest Hydrofraise and Trenchcutter
rigs. Despite the introduction of these modern reverse circulation rigs,
excavation by grab still has its place, on smaller sites and in certain soil con-
ditions, but it is worth noting that kelly-mounted rigs are rarely used and
preferences are for heavy 8 tonne rope grabs and hydraulic grabs suspended
from cranes.
Incentives to develop excavation equipment (particularly in rock) have
come mainly from the need to improve output at the lowest cost; without
exception, panel excavation still remains the critical operation in diaphragm
wall contracts despite the complexity of all other operations. Minimization
of machine downtime is a vital factor in overall excavation production.
122 Deep excavations

In addition to excavation output, increasing wall depths and the require-


ments of wall verticality tolerance have placed demands on the development
of excavation equipment. To a large extent the requirements of depth and
tolerance, together with the particular soil or rock type, define the optimum
excavation equipment: rope grab or cutter. Modern cutter machines are
capable of cutting through rock with a compressive strength of the order of
200 MPa, and of excavation in soils to depths in excess of 100 m. Cutter
machines with small roller bits on the outer edge of the cutters were developed
for similar tasks. Verticality tolerances of the order of 0.3% can be made to
depths of 50 m. The manufacturers of these rigs currently include Soletanche,
Bauer, Casagrande, Soilmec and Tone.
Developments in mechanical excavation equipment have radically changed
the scale of diaphragm wall contracts that can be undertaken. The Seaforth
Dock contract (145 500 m2 of wall) and Redcar Ore Terminal (64 680 m2 ,
depth 45 m), both excavated by rope grabs, were among the largest projects
of their time (1968–1972). The Medinah car park project in Saudi Arabia,
completed in two years with three cutter working rigs and five rope grabs,
has a total wall area in excess of 320 000 m2 and to a maximum depth of
55 m. Wall outputs peaked at almost 30 000 m2 per month and consistently
averaged almost 20 000 m2 per month.
Accompanying these changes in excavation method has been the develop-
ment of other installation techniques on site:
(a) temporary stop end fabrications which allow water bar installation in
panel joints and do not require extraction during, or immediately
after, concreting; the CWS joint and similar stop ends
(b) fabricated permanent stop ends which allow the transfer of shear and
tensile forces through the panel joints
(c) reusable precast concrete guide walls
(d ) improved slurry cleaning equipment, including centrifuges and belt
presses
(e) polymers for artificial slurries
( f ) improved cage handling methods
(g) smaller reverse circulation rigs for urban sites (City Cutters)
(h) improved cutters such as rock roller bits on Trenchcutter rigs
(i ) steerable grabs, hydraulic rope suspended, manufactured by Bauer,
Soilmec and Bachy
( j ) improved in-cab instrumentation.
While site operations were improving, diaphragm wall designers were also
introducing innovations:
(a) structurally efficient plan shapes of diaphragm walls
(b) vertically post-tensioned panels
(c) precast concrete panels
(d ) combined slurry cut-off walls, permanent precast concrete diaphragm
walls and temporary soil retaining walls
(e) improved wall to slab connection details.
A review of developments in the construction of structural diaphragm walls
by Puller and Puller11 is referred to in greater detail in Chapter 8.
Advantages of diaphragm wall construction
The diaphragm wall is generally efficient in cost and construction
time where it is used for both permanent and temporary subsoil
retention for walls of medium, and greater, depth.
A previous review of the economics of basement construction5 by the Author
emphasized that cost comparisons between types of basement construction
Vertical soil support: wall construction 123

could only be made after a complete analysis of the unit costs of temporary
and permanent walls, the costs of shoring or anchoring and the costs of
internal lining walls and basic finishes. The diaphragm wall method compares
well for deep walls in a complete analysis of this type. Additional data on cost
comparisons are given at the end of this chapter.
In the UK in recent years the number of diaphragm wall contracts has
reduced and their size increased as preference has been shown, presumably
on the basis of cost, to concrete piled walls, particularly hard–soft secant
walls, for many two-storey basements. Diaphragm walls find favour for
deeper constructions in the range 3000 to 12 000 m2 of wall area. Some of
these contracts are part of new metro construction, particularly station box
and ventilation shaft construction. In London, in the years 1997 to 2002,
works for stations at Canary Wharf on the Jubilee Line, at Westminster
Station in Central London and at the Stratford Box and for the Graham
Road shaft on the Channel Tunnel Rail Link works have all used diaphragm
walls for wall depths in the range 30 to 55 m. The common features for these
latest contracts include relatively heavy reinforcement cages (the total cage
weights per panel at the Graham Road shaft reached 90 tonnes), CWS
joints with water bars in vertical panel joints, extensive use of couplers to
splice cages and joints with slab and roof steel, relatively long panels, up to
6.5 m, with large concrete pours despite traffic difficulties in built-up areas,
and the extensive use of cutters (the exception being Westminster Station
where low-headroom grabs were used). These contracts ranged in size from

Fig. 4.24 (a) New underground station construction, Westminster, London: plan30
124 Deep excavations

Fig. 4.24 (b) New underground station construction, Westminster, London: cross section of the works30

4000 m2 at Graham Road to 8000 m2 at Westminster Station to 65 000 m2 for


the Stratford Box.
The construction of the 38 m deep underground escalator hall at West-
minster Station was completed alongside very busy thoroughfares, in near
proximity to settlement sensitive structures and immediately adjacent to exist-
ing Jubilee Line tunnels in frequent use. The works were designed to provide a
new ticket hall and access to the new Jubilee Line extension and passenger
interchange with the existing District and Circle Line station. Crawley and
Glass12 described the construction works. The plan and section of the
works is shown in Fig. 4.24.
The ground conditions were typical of Central London, made ground at the
surface, overlying alluvium, Terrace Gravel and London clay. The con-
struction method comprised very large piles, 3 m diameter and 53 m deep to
support the existing station and then top-down construction following the
1.5 m thick reinforced concrete underpinning transfer slab within a 1 m
thick diaphragm wall installed partly by lower headroom rigs below the
existing station platform (see Fig. 4.25). The top-down construction sequence
is shown in Fig. 4.26. Temporary retaining walls were necessary for access to
support up to 7 m of ground using king post walls with in situ R.C. skin walls,
contiguous concrete piles, mini piles combined with cement–bentonite and
silicate ester grouting and sheet piles installed by Giken press. This scheme
must represent one of the most complex ground engineering schemes
completed to date in the United Kingdom.
Elsewhere, in major city centres, large schemes for office and retail use
occupy progressively larger and deeper floor space as land values escalate.
In Hong Kong in 1994 the LDC H-6 project in Queens Road Central
Vertical soil support: wall construction 125

Fig. 4.25. New underground station construction, Westminster, London: construction method12

consisted of a perimeter diaphragm wall of 18 000 m2 , 45 m deep and 1200 mm


thick, together with 46 barrettes with a total area of 8500 m2 , to the same
depth and thickness. The soil conditions generally consisted of a 15 m depth
of fill and alluvium containing cobbles and boulders underlain to a depth of
60 m by completely decomposed granite with SPT values from 50 to more
than 240. Within the upper fill and boulder layers, a hydraulic grab was
used for pre-excavation to an average depth of 12 m. Within this depth,
where no boulders greater than 700 mm were met, an average 7 to 10 m2 /h
could be achieved, but this was reduced to 30 to 50% of this output where
larger boulders were broken by chisel. The pre-excavated trench, backfilled
with lean mix concrete, was then excavated using Bauer BG 30 Trenchcutters.
The average penetration rate achieved on primary panels was:
. excavation in lean mix concrete: 3 min per metre, 65 m3 /h;
. excavation in completely decomposed granite, SPT less than 50: 5 min
per metre, 45 m3 /h;
. SPT 50 to 100: 7 min per metre, 30 m3 /h;
. SPT 100 to 300: 15 min per metre, 13 m3 /h.
These rates, achieved only during the period of machine operation, do not
include breakdown, waiting and working time. The production rate for
secondary panels was approximately half that of the primary panels.

Diaphragm walls allow effective transfer of vertical load from the


building superstructure to subsoil below basement level.
Early diaphragm walls in the UK did not generally include use of the walls to
transfer vertical load from the superstructure. The reluctance of designers to
allow such load transfer was due no doubt to lack of published test results.
Gradually this situation changed following publication by Fleming and
Sliwinski13 and Xanthakos14 of reviews of available test data together with
their conclusions that there was no evidence that load transfer characteristics
126 Deep excavations

Fig. 4.26. New underground


station construction,
Westminster, London: top-down
construction sequence12
Vertical soil support: wall construction 127

Fig. 4.27. Cleaning device for


removing excessive wall cake
thickness

of slurry panels differ materially from those in dry construction. They did
recommend, however, that design should be conservative, especially on sites
where panel load tests were not justified on cost grounds. Panel tests in the
UK have remained few. An early exception was for a load test at a basement
construction at Kensington Town Hall in London where Corbett et al.15 con-
cluded that compared with other large bored pile designs, the performance of
the test panels was not adversely affected by the bentonite slurry.
Elsewhere, concern has continued to be shown that the slurry has no adverse
effect on the ability of surface friction or adhesion to transfer vertical load to
the adjacent soil, particularly where an excessively thick wall cake has formed.
The maximum time period between completion of panel excavation and any
change of slurry and commencement of concreting may be specified with
advantage to minimize this build-up. Measures to reduce the effect of a thick
wall cake include use of tubes à manchette on the panel face and the use of
cleaning devices to remove the wall cake (a typical device is shown in Fig. 4.27).

Diaphragm wall construction causes minimum noise and


vibration disturbance
The extent of noise and vibration in diaphragm wall installations is limited to
that associated with normal civil engineering plant, cranes, generators,
pumps, compressors, power packs etc. for excavation in soils. Obstructions
128 Deep excavations

or rock strata may lead to chiselling and some vibration where conventional
grab excavation is used, although the use of modern reverse circulation cutter
rigs reduces this (although cutter rigs do not necessarily avoid the need for
chiselling boulders). Where such vibrations do occur, soil from the side of
the slurry trench may collapse, leading to additional panel overbreak. The
extents of chiselling and vibration using conventional grab excavation have
been reduced by the successive use of auger coring buckets along the length
of the panel excavation.
It should be remembered that even where soil conditions allow relatively
quiet plant operation the diaphragm wall process necessitates extraction of
stop ends from panels other than secondary panels unless expendable
permanent or CWS type stop ends are used, and this operation is only possible
some hours after the completion of concreting. Unless expendable reinforced
concrete stop ends or CWS type stop ends (which can be extracted at a
convenient time) some noise from cranage is inevitable due to out-of-hours
stop end extraction.

Quality control of diaphragm wall construction


Considerable skill and care is required on the part of the specialist contractor
installing diaphragm walls. Six items within the Author’s experience warrant
particular attention.

The need for continuous construction


The construction technique is continuous, from commencement of panel
excavation to installation of stop end formers and reinforcement cage, con-
creting and withdrawal of panel stop ends. Any major disruption, whether
caused by ineptitude or other circumstances, can cause loss of ground or
poor standards of wall construction which may be difficult and expensive to
rectify.
It is vital to maintain punctual completion of the construction cycle of
excavation, panel preparation for concreting, reinforcement installation, con-
creting and panel finishing. Factors which can disrupt this operation sequence
must be removed. Critical matters include the optimum panel size to ensure
compatibility with excavation and concrete outputs; permissible daily work-
ing hours allowed in the contract to facilitate daily completion of important
stages in the cycle, such as concreting and stop end removal; and high
standards of site supervision to ensure the availability of resources of plant,
labour and material.

Panel instability
The quality of the work (which is only disclosed when bulk excavation of the
site is undertaken) is directly dependent on the subsoil strata since the soil
forms a temporary shutter against which the in situ concrete is cast. In
terms of wall finish, the smoothness of the diaphragm wall concrete will
improve with reducing soil particle size. One would expect to obtain a
better surface finish for a diaphragm wall excavated in clay compared with
one in sandy gravel. Obstructions in the diaphragm wall excavation, whether
man-made or natural, may not only prove difficult to remove from the panel
excavation, but their removal may also cause some dislodgement of the soil on
the panel face, the resulting void subsequently being filled with concrete
overbreak. Where this void is exposed later during site bulk excavation,
some remedial work may be needed to remove the excess concrete.
Overbreak on the panel face may be caused, therefore, by removal of
obstructions, but more likely by collapses of the panel face caused by
instability. Panel stability, reviewed by Xanthakos14 and Puller16 , was
Vertical soil support: wall construction 129

originally studied by several authors including Nash and Jones17 , Muller-


Kirchenbauer18 , Piaskowski and Kowalewski19 , Fernandez20 and Huder21 .
Within the Author’s practical experience, panel stability is much dependent
on the ability of the soil around the panel effectively to arch during all
stages of the panel excavation. In loose sands, sandy gravel and open gravels,
large volumes of overbreak frequently occur below guide trench level. Typical
overbreak removal operations cause difficulties of access and construction
delay. The cost of removal is frequently three or four times that of the original
concrete as placed. A differential of at least 1.5 m height between slurry level
within the panel excavation and groundwater level within the subsoil adjacent
to the panel is necessary to avoid panel instability. The cause of serious panel
instability is frequently the neglect of maintaining such a differential height,
and in isolated cases the installation of a wellpoint dewatering system has
been necessary during the diaphragm wall works to improve the differential
height and so increase panel stability. Particular attention is needed where
groundwater flows underground or through a closing panel.
Where conventional grab excavation is used there is risk of loss of the panel
face, particularly in granular soils, from below guide trench level, by the
pumping action during raising and lowering of the excavation grab within
the bentonite slurry. Instances of major panel collapse caused by panel
instability in very weak clays were reported by Mastikian22 but fortunately
are rare. Some wall face collapses have also been observed in deep panels in
highly fissured over-consolidated clays, such as London clay, where fissuring
has much reduced the soil strength.
The risk of major panel instability does not appear to increase with greater
panel depth, probably due to the relative increase in slurry over pressure with
depth. The calculated stability of panels by traditional methods as described
in DIN standards and by Xanthakos14 and others has frequently been
shown to be unreliable in both deep and shallow panels. The increase in
time spent in the excavation and subsequent preparation of deep panels
does not in itself appear to lead to panel instability. Other factors appear to
dictate this, including the presence of very soft clays, excessive fluid loss
from the slurry, excessive vibration due to construction operations including
chiselling, inadequate differential height between slurry and groundwater
levels and unstable guide wall construction. Some instability risk particularly
applies to shaft construction. Where shaft panels are founded on sloping rock-
head it is essential to avoid excessive chiselling. It may be practicable to use
multiple reinforcement cages in each panel with a stepped panel toe to mini-
mize this and individual cages for each bite become increasingly important for
grab excavation onto sloping rockhead for shafts deeper than 30 to 40 m.
Where shaft excavation is made through permeable strata it is likely that a
pumping well within the shaft may be necessary to relieve piezometric pressure
and retain the stability of the closing panel excavation. Where such soils are
saturated only a small flow of filtrate from the slurry into the soil will cause
a rise of pore pressure in the soil. This increase will occur more readily
where the permeable stratum is capped above and below by impermeable
layers. Relief of any excess pore pressure by pumping ensures an adequate
differential head of slurry and allows the closing panel excavation in stable
conditions.
Where deep walls are built in tidal conditions it is necessary to monitor
groundwater conditions if there is risk of inadequate differential slurry head
at high tide. Saline groundwater may lead to flocculation of the slurry and
the use of additives is necessary to prevent this. It may be necessary to improve
the difference between slurry and groundwater levels by temporary ground-
water lowering, by increasing guide wall height or by use of a heavier
130 Deep excavations

slurry. In very weak soils it may be necessary to reduce panel lengths. Any
slurry loss from panels left open overnight must be made good regularly.

Inclusions
The combined use of slurry and high-grade concrete within the same wall
panel construction requires considerable skill and care to ensure homoge-
neous finished concrete without deleterious slurry inclusions in the panel.
Bad slurry inclusions, particularly at panel joints, can cause unacceptable
standards of structural competence or water resistance of the wall and must
be avoided. Cambefort23 showed examples of poor-quality diaphragm walls
caused by inadequate slurry quality amongst other factors. The cause of
such inclusions may not necessarily lie with poor-quality slurry; slow, erratic
concrete pours with non-cohesive and low-slump concrete may also
contribute. The risk of slurry inclusions within wall panels is increased
when tremie spacing is too large in long panels and at tee or corner panels.
The risk of inclusions in deep diaphragm walls is frequently related to a
limited number of causes, principally the use of large, wide box-outs, the
excessive density of reinforcement, standards of concrete placement and
slurry quality immediately before concreting.
The restriction of wall width by box-outs can seriously impede flow of
concrete upwards from tremies in the centre of the wall. Box-outs at the
rear of reinforcement should be avoided. Where slurry has an excessive
sand content there is risk of slurry inclusions above and below the box-outs
and risk of concrete contamination. The risk of inclusions increases rapidly
Fig. 4.28. Large mud inclusion as slurry quality deteriorates and especially so where concrete supply is
within a diaphragm wall panel irregular and flow within the panel is impeded. As an example, Fig 4.28
shows a panel void due to a slurry inclusion. In this case, at a panel joint,
the excessively thick slurry failed to be scoured by concrete and fresh concrete
from the centre of the panel folded over on the top surface of the heavy slurry
leaving a triangular inclusion for the full wall thickness.
For deep walls, particularly those excavated by cutter, it is often advanta-
geous to use separate slurries for excavation and concreting. The ‘excavating’
slurry is completely replaced in the panel by the ‘concreting’ slurry immedi-
ately prior to placing the cage in the panel. In this way the effect of calcium
contamination is restricted and ensures that there is ample time for adequate
preparation of slurry for each phase.
For deep walls it is particularly important to ensure that all heavy slurry is
removed prior to concreting especially when the cutter rig is used for this
purpose. An air lift, used where grab excavation is made, may be preferred
for this purpose. Care must be taken to avoid contamination of clean slurry
by cleaning cutter wheels and internal pipework if the cutter is used to do this.
To reduce the risk of inclusions the slurry properties must be a balance
between adequate filter cake building properties, sufficient shear strength
in order to support soil grains without excessive sedimentation, low filter
loss, and a viscosity which neither restricts flow of concrete nor requires
uneconomic pump energy to circulate it. The slurry will require care to
extend its life and fresh slurry may be needed to improve its properties after
repeated cleaning. Additives to the slurry include sodium bicarbonate to control
high pH and lignosulphinate thinners may be required to reduce viscoscity.
Polymers may be added to reduce free water in the slurry or in saline conditions
may be used as one measure to prevent flocculation. Recommended values of
slurry properties to minimize inclusion risk are shown in Table 4.2.
Before the introduction of cutter rigs, load bearing panels and barrettes
excavated to depths of 25 m or so were successfully excavated by grab and
concreted without a slurry change, achieving a satisfactory concrete/soil or
Vertical soil support: wall construction 131

Table 4.2 Recommended slurry properties for deep diaphragm walls

Fresh Ready for re-use During Before


excavation concreting

Density (g/m3 ) 1.05 Less than 1.15 Less than 1.25 1.10
Marsh value, seconds 30 to 45 30 to 50 30 to 60 30 to 45
Fluid loss (30 mins) Less than 30 ml Less than 50 ml Less than 60 ml Less than 60 ml
pH 7 to 9.5 7 to 11.0 7 to 11.5 7 to 11.5
Sand content No spec. No spec. No spec. Less than 3%
Notes.
This chart presumes that the slurry will be changed for the whole panel prior to the installation of the reinforcement cage and
subsequent concreting.
For load bearing walls the following values are recommended:
Before concreting: Specific gravity, less than 1.08
Sand content, less than 1%.

rock interface beneath the panel. Deeper panels excavated by grab or cutter
require more care, involving a slurry change and including cleaning the
base, probably by air lift. Small thicknesses of soft material trapped at the
base of the wall are unlikely to affect wall settlement performance at working
load, but the effects of sedimentation, loss of debris from the wall cake and
increased fluid loss from the slurry accentuate the problem at the base as
the panel depth increases. For deep panels which sustain load at their toe it
is a wise routine precaution to incorporate 50 mm grout pipes within the
panel, staggered in plan, at intervals along the toe to permit toe grouting
to reduce the effect of thick basal inclusions. In deep panels where large
wall or barrette settlement is unacceptable the following procedure is
recommended.
(a) On larger sites, construction of a preliminary test panel which can be
examined for basal inclusions and, where possible, load tested.
(b) Inclusion within random panels of sealed steel tubes to within 300 mm
of the panel toe to allow the interface below the panel concrete to be
cored. This method is preferred to penetration cone testing next to
the panel.
(c) Where the panel is founded on rock, acceptance criteria must be defined
for the founding depth. Experienced supervisory staff must be on site
to verify and agree the base conditions at the bottom of each bite of
excavation without delay.
(d ) Change of the slurry for the whole panel volume is obligatory at the end
of the panel excavation.
(e) The base of deep load bearing panels should be cleaned by grab and air
lift or, for panels excavated by cutter, by the cutter itself or by air lift.
( f ) Separate slurries should be used for panel excavation by cutter and
subsequent panel concreting.
(g) Reinforcement spacing should comply with the requirements of
DIN 4126.

Tolerances
There is considerable risk of wall construction outside specified tolerances.
Tolerances for verticality, concrete protrusions on the inner wall face, posi-
tions of box-outs for recesses and locations of reinforcement are detailed in
model specifications for diaphragm wall construction; in the UK the ICE
Specification for Piling and Embedded Retaining Walls24 lists the following
tolerances.
132 Deep excavations

(a) Guide trench. Line of finished face nearest to excavation: 15 mm in


3 m; vertical to within 1:200; minimum clear distance between the
faces of the guide walls shall be wall thickness plus 25 mm, and the
maximum distance shall be wall thickness plus 50 mm.
(b) Diaphragm wall. Exposed wall face and panel ends shall be vertical
within a 1:120 tolerance. In addition to verticality tolerance, a further
100 mm shall be allowed for protrusions resulting from irregularities in
the ground as excavated.
(c) Box-outs. Vertical and horizontal tolerances of 75 mm and a hori-
zontal tolerance, at right angles to the wall, 75 mm plus the horizontal
tolerance resulting from (b).
(d ) Reinforcement. Horizontal tolerance of cage head at top guide wall,
measured along the trench, 75 mm; vertical tolerance of cage head
relative to the guide wall shall be þ150/50 mm.
(e) Concrete tolerance. Where the final trimmed level of the diaphragm wall
is up to 1.0 m below the top of the guide wall, the casting tolerance
will be 600 mm above the trim level; for each additional 1 m depth of
final trim level, or part thereof, allow an additional 150 mm level
tolerance.
Similar tolerances are listed in the European Specification for Diaphragm
Walls25 .

Concrete quality
The use of larger pours (often in the range 200 to 400 m3 as required for deeper
walls) has placed an increased demand on concrete properties which generally
has led to more detailed and stringent concrete specification. The specifiers
now require longer periods of retardation, some improvements to 28-day
strength, a guarantee of concrete of uniform consistency free from balling
of cement and fine sands and pumpable, non-segregating, cohesive mixes.
Properties such as bleed of pours assume greater importance as panel depth
is increased. Additives such as super-plasticizers have assumed increasing
importance and the need to achieve economy in cost has led to an increasing
use of ground blast furnace slag in the UK.
The consistency of delivered concrete quality and the rate of delivery
remain key factors in the finished quality of diaphragm walls. The concrete
mix for deep walls must be cohesive with a high resistance to segregation.
The use of cleaner aggregates and the use of crushed angular stone makes
this more difficult to achieve and is not identified by traditional quality control
tests for consistency, such as the slump test. Improvements to cohesiveness
can be made by oversanding or the incorporation of small size rock particles
avoiding clay sizes.
The use of the slump test or the flow test to control mix workability on site
fails to accurately identify changes in free water content and provides no
effective control on bleed of free water from the mix. Bleed is susceptible to
variations in water–cement ratio especially when a gap exists at the lower
end of the sand grading. The use of ground blast furnace slag in the mix
may also increase bleed risk. An example of bleed in deep panels is shown
in Fig. 4.29. The depth of the defect is usually limited to the concrete cover
thickness but even so may be detrimental to wall durability. Some risk of
poor bond between concrete and reinforcement can also occur.
For a particular mix it can be shown experimentally that bleed, a sedimen-
tation effect within concrete, increases rapidly above a critical water–cement
ratio and control of free water within the mix quantities by slump testing is
often not adequate to prevent excessive bleed in deep panels. The remedy
Vertical soil support: wall construction 133

Fig. 4.29. Example of bleed


from concrete on face of
diaphragm wall panels

often lies in the more accurate determination of moisture content of aggre-


gates at the mixer.
The need to ensure competent tremie practice in placing concrete is vital as
panel depths increase and plan shapes vary. It is essential that the concrete as
discharged into the tremie hopper at the beginning of the pour does not mix
with the slurry within the tremie pipe as it travels to the bottom of the panel.
Several means can be used in an attempt to ensure this. Vermiculite granules
and plastic footballs are sometimes used, but full-scale trials have shown that
the high concrete velocity within the tremie pipe does not allow adequate
separation of concrete and slurry, and a plug of defective concrete is left at
the bottom end of the tremie pipe and remains at the base of the panel. The
proven procedure to avoid this requires the use of a ‘Chinese hat’ plug lowered
by wire slowly down the tremie pipe with the initial charge of concrete above
it. The tremie is then lifted 150 mm to allow the discharge of the uncontami-
nated concrete across the base of the panel. Figure 4.30 shows the details. The
location of tremies within the panel is also important; the ICE Specification
recommends a single tremie for panels up to 3.6 m in length, two tremies
for panels between 3.6 and 7.2 m long and three tremies for tee panels. This
advice appears adequate; it is essential at the commencement of a deep
pour that the concrete flows uniformly over the whole of the bottom of the
panel and a consistent level across the panel is maintained during the pour.
It is a wise precaution to discharge concrete from all tremies simultaneously
at the beginning of a pour to cause a rapid flush of concrete and avoid
incorporation of a large volume of slurry into the concrete at the wall toe.
During the pour accurate records of tremie length and concrete level should
be made to avoid splitting tremies either too late or too early.
In summary, the preventative measures to ensure adequate concrete quality
depend upon a balance between strength, workability and cohesiveness. It is
134 Deep excavations

Fig. 4.30. Cross-section


through Tremie tube with
disposable ‘Chinese hat’

important that personnel responsible for specifying and purchasing concrete


are also experienced in diaphragm wall construction.

Water-resisting construction
The water resistance of a diaphragm wall structure depends on standards of
design and construction. Design methods to achieve high standards include
the provision of an integral water bar system between the diaphragm wall
panel joints, in the construction joints between the wall and the basement
slab and within the basement slab itself. Water resistance of basement con-
struction is discussed more fully in Chapter 8. In general, the following factors
affecting water resistance should be noted:
(a) soil and groundwater conditions: the greatest risk occurs with high
groundwater in loose granular soils and very weak silts, silty clays
and highly fissured over-consolidated clays
(b) compliance with good standards of diaphragm wall excavation, slurry
control and concreting: avoid very long periods between excavation
and concreting
(c) wall thickness: avoid very thin walls less than 600 mm thick
(d ) plan shape of excavation: there is some evidence that leakage at panel
joints tends to appear near return angles of wall (joints of panels at
the centre of a long length of straight wall also tend to open during
bulk excavation and can also cause leakage)
(e) depth to formation level: the greater the basement depth the greater the
leakage risk, primarily due to increasing hydraulic head with depth
( f ) panel length: shorter panels mean more panel joints but may reduce the
risk of inclusions
Vertical soil support: wall construction 135

(g) means of diaphragm wall excavation: improved verticality tolerance


and reduction of vibration with cutter excavation should generate
better wall quality than grab excavation and in turn, water resistance
standards should be improved
(h) extent of reinforcement density in diaphragm wall: high reinforcement
density may reduce in situ concrete quality and, in turn, reduce water
resistance
(i ) top-down construction: this is likely to minimize wall movements and
improve water resistance, particularly at panel joints
( j ) the use of water bars in vertical panel joints using the CWS system or
similar type of joint is essential; effective design of the joint between
the wall and the bottom raft is critical
(k) the use of precautionary grouting of vertical panel joints.

Precast diaphragm walls


The disadvantages of using the subsoil face as the shutter for concrete placed
within the diaphragm wall panel are obvious. The French firms, Bachy and
Soletanche, now united as one firm Bachy–Soletanche, had developed
methods of using precast concrete wall panels to overcome these drawbacks.
The technique was first introduced in 1970 but has tended to be less popular in
recent years. There are no known jobs in the UK using the precast system.
Colas de Francs26 claimed the following distinct advantages for the precast
prefabricated diaphragm over other types:
(a) general appearance: no cutting back is required and the finished surface
is agreeably clean
(b) the shape of the diaphragm can be tailored to form an integral part of
the final structure, satisfying technical, aesthetic and economic consid-
erations
(c) improved concrete quality and accuracy in placing reinforcement gives
considerable savings on materials; prefabricated diaphragms are gener-
ally 30% thinner than conventional diaphragm walls
(d ) the prefabricated diaphragm can be built and installed in the ground to
finer tolerances, and openings can be positioned more accurately
(e) watertightness at the joints and in the wall itself is better than with con-
ventional diaphragms.
The improved surface finishes can be seen in Fig. 4.31, an example of the
Bachy system known as Prefasif walling. Whether the advantages of finish,

Fig. 4.31. Precast diaphragm


wall construction: a high
standard of wall finish is
obtained using the Prefasif
system, Schipol, the
Netherlands (courtesy of
Soletanche–Bachy)
136 Deep excavations

improved tolerances and reduction in wall thickness offset the cost penalty of
these methods can only be appraised on a job-by-job basis.
Panel widths and lengths are limited by the capacity of mobile lifting equip-
ment suitable for operation on construction sites, and in practical terms this is
of the order of 50 tonnes. The panels are usually cast on site; some additional
site area may be needed to do this although less space is needed to store
reinforcement since prefabrication of reinforcement cages is not necessary
with precasting. Figure 4.32 shows a panel being lifted.
The principal differences between the previous Panosol method by Sole-
tanche and the Prefasif method by Bachy were in the mechanical connection
between panels and the timing of the use of cementitious slurry to surround
the precast unit and support the subsoil. In the Panosol method, the panel
was dug under a cementitious slurry containing retarding and regulating addi-
tives, the slurry being removed without difficulty from the smooth face of the
wall during bulk excavation of the site. Slurry between the rear face of the
Fig. 4.32. Lifting a precast precast panel and the subsoil remained permanently as an inert filler material
diaphragm panel at a
nuclear power station at
between wall unit and subsoil, the final strength of the set slurry being
Nogent-sur-Seine, France designed to exceed neighbouring soil strength. In the Prefasif technique the
(courtesy of Soletanche–Bachy) slurry was not designed for the dual purpose of a digging slurry and a
medium- or long-term soil support; the panel was dug under a conventional
bentonite mud which was later displaced by the introduction of a cementitious
grout just before the precast unit was positioned in the panel excavation.
Bachy claimed that its process gave flexibility in site operations, allowed a
wide range of grout strengths to be used and was particularly convenient
when large vertical loads were being carried by the panels. The process also
avoided risk of contamination of the grout by soil during the excavation
process. The methods of panel connection for the previous Soletanche
system are shown in Fig. 4.33; Fig. 4.34 shows the technique used originally
by Bachy.
Precast diaphragm panels may be designed to span vertically between
ground anchor levels, taking care to avoid movement caused by the pre-
loading force to the anchors compressing the slurry between wall and subsoil.
Alternatively, soldier panels, themselves anchored, may support panels
spanning horizontally between them. In these cases the soldier wall units
may be extended in depth to improved bearing strata to support vertical
load. Precast diaphragms may also be used in a composite construction of
structural wall and self-hardening slurry cut-off, as shown in Fig. 4.35.

Post-tensioned diaphragm walls


Another innovation originally pioneered by Icos in the late 1960s replaced
vertical reinforcement within the diaphragm wall panel by a pre-stressing
tendon, the reinforcing steel in the cage being used principally to retain the
tendon position during lifting. The use of post-tensioning therefore reduces
steel reinforcement costs to a minimum, although its relatively small use in

Fig. 4.33. Horizontal cross-


section jointing detail for
Panosol diaphragm wall
panels: types (a) and (b) are
standard joints; types (c)–(e)
use temporary metal guides;
types (f) and (g) use pre-formed
water bars (courtesy of
Soletanche–Bachy)
Vertical soil support: wall construction 137

Fig. 4.34. Jointing details for


Prefasif diaphragm wall panels
– mechanical locking hook joint
at the foot of the panel and types
of joint in horizontal section:
joint type (a) uses a pre-formed
water bar, type (b) uses a
reinforcing key, and type (c)
uses grout only (courtesy of
Soletanche–Bachy)

Fig. 4.35. Composite wall


construction using precast wall
and slurry cut-off, Horizon
Tower at Puteaux, Paris
(courtesy of Soletanche–Bachy)
138 Deep excavations

Fig. 4.36. Post-tensioned


cantilever and single-propped
diaphragm walls showing cable
profile and bending moment
diagrams for the wall section27

the UK may reflect relatively high cost of tendons and stressing operations, as
well as design complexity to cover all stages of construction.
The feature of the method is that stressing of the panel is undertaken before
bulk excavation is carried out and while the wall panel is fully embedded.
Tendon forces and eccentricities are calculated for loading on the final struc-
ture with no tension across the concrete section, the panel movement during
stressing being reduced by the surrounding soil, the soil restraint varying
between full passive pressure and earth pressure at rest. The cable profiles
for a cantilever wall and a propped diaphragm wall are shown in Fig. 4.36.
Figure 4.37, taken from Gysi et al.27 , shows earth pressure, bending moments
and stresses in a typical wall unit before removal of the surrounding soil, the
moment applied to the wall section by the pre-stressing force being effectively
reduced by the soil stiffness whilst reducing tension developing across the
section due to pre-stress.

Fig. 4.37. Earth pressure, bending moments and stresses in a post-tensioned diaphragm wall before excavation in front of the
panel 27
Vertical soil support: wall construction 139

Fig. 4.38. Examples of post-tensioned walls27

Figure 4.38 shows further examples given by Gysi et al.27 The considerable
heights of cantilever wall shown, between 7 and 9.6 m, and the relative
slenderness of the wall section would appear conducive to excessive soil
deformations behind the wall but none are reported. It is possible that the
buttressing effect in plan of return walls at the corners of the excavation
may have been taken into account in the wall design. In practice, the length
of each basement wall section between return angles considerably influences
the amount of horizontal movement after bulk excavation at the critical
mid-point between corners of any basement.
It should be mentioned that the hoisting and lowering of diaphragm wall
cages may cause displacement of steel and, in the case of pre-stressed walls,
displacement of the tendon. Checking the tendon position, and correcting if
necessary, assumes considerable importance when the tendon force and
actual eccentricity are viewed together.
A typical tendon/reinforcement cage detail is shown in Fig. 4.39 after
Fuchsberger and Gysi28 . Although designs have been prepared to use large-
diameter high-tensile steel bars of the Lee–McCall type with anchorages
cast in to the base of the wall panel, this method has not been used to date,
to the Author’s knowledge. While the dependence on sound anchorage to
each bar is obvious using this method, the compressive stresses and yielding
induced in concrete within the anchorage zone of looped tendons at stressing
should not be overlooked.
The Author has found that a separately cast capping beam is preferred to
house the stressing heads and anchorages of the tendons. Unless this is
done the anchorages are founded in the top of the diaphragm wall where
concrete quality may not be at its best.
140 Deep excavations

Fig. 4.39. Typical diaphragm wall reinforcement cage with curved tendons, Irlam O’Heights near Manchester, UK28

Soldier pile tremie There have been some variations on the king post method of soil support; by
concrete (SPTC) using a mass concrete wall placed under bentonite slurry to replace horizontal
method lagging timbers, the SPTC method has found favour in the past in the USA.
The reinforced concrete walls are of limited span and are dependent on
subsoils of low overbreak risk. The method was reported by Peck1 .
Mesh-reinforced gunite has been used successfully to replace the mass
concrete walling used in the SPTC method. Figures 4.40 and 4.41 show
basement excavations in Stuttgart and Linz, respectively, where gunite over
mesh has retained subsoil between anchored vertical steel beams. The
absence of groundwater is a prerequisite for the successful use of gunite
over mesh.
Vertical soil support: wall construction 141

Fig. 4.40. Composite use of


contiguous bored pile walls and
gunited steel soldier beams
retained by ground anchors,
Stuttgart (courtesy of Bauer)

Fig. 4.41. Underpinned walls


(left and centre) and walls
retained by gunite overmesh
between anchored steel
soldiers, Linz (courtesy of
Bauer)

Construction Previously in this chapter the distinction was drawn between excavation
economics methods and the alternative of peripheral support systems. The permutations
of each are considerable. The choice of excavation method may well be con-
strained by factors other than cost, such as environmental matters like noise,
vibration and site traffic, short completion times or the proximity of adjacent
structures and services, etc. The choice of peripheral support system may, how-
ever, be made with greater freedom. Some constraints other than cost will no
doubt still apply (noise and vibration levels vary from one sheeting installation
142 Deep excavations

to another, for instance), but within the choice of peripheral support system
method lies considerable scope for economy and careful comparisons should
be made method by method on each excavation project.
Before this, however, a number of conclusions can be drawn regarding the
practical application of the peripheral wall or sheeting methods.
(a) High-torque CFA methods are limited to 35 m depth or so reinforced to
15–18 m depth in medium diameters, less in larger diameters.
(b) Low-torque CFA methods are limited to 20 m depths and pile devia-
tions may prove limiting.
(c) In ground without obstructions or rock, CFA methods generate the
least noise and vibration.
(d ) In particular cases (e.g. loose sand and gravel over a very stiff clay),
CFA methods can cause over-excavation and lead to settlement of
adjacent structures. Generally, however, CFA methods are suitable
for working close to existing structures.
(e) For temporary works, the hard–soft secant wall constitutes an efficient
wall solution. For permanent works, however, durability and strength
considerations may limit its application unless a reinforced concrete
lining wall is used, designed to withstand groundwater pressure.
( f ) Cased secant methods provide a high degree of risk-free excavation in
difficult soil or rock conditions or in granular soils near existing
foundations.
(g) The water resistance of hard–hard secant piling can reach that of a well-
built diaphragm wall. With increasing depth, the risk of gaps between
secant piles increases, with resulting lack of water resistance. Improve-
ment of water resistance occurs with closer-centred boring.
(h) For walls greater than 30 m deep, cased secants with high-torque rigs
and diaphragm walls are the only methods available. At greater
depths, 40 to 45 m, diaphragm walls are the only available method.
(i) Box-outs are more easily accommodated in diaphragm walls than in
secant piles.
( j ) Construction speed and job size affect wall method selection. Where
progress is unimpeded by rock or other obstructions daily production
rates vary in the range 50 to 60 m2 for a high-torque cased secant rig
and 70 to 80 m2 for a CFA secant rig. Diaphragm wall grab units
produce 60 to 100 m2 per day, while hydraulic cutters can produce in
excess of 200 m2 per day (all 12 h day shifts). It is possible to use several
rigs on larger sites if a slower technique is chosen.
(k) Sheet pile walls may prove economical for shallow basements in soils
free from obstructions: use of Giken press for sheeters of maximum
length 20 m in clay. Installation in granular soils by vibrator.
A comparison by Sherwood et al.10 of relative costs of various walling
techniques in different ground conditions and site circumstances is shown in
Table 4.3. Bearing in mind that market fluctuations can cause periodic
changes in secant pile and diaphragm wall prices, the conclusions that may
be drawn from this table are that diaphragms and hard–soft piles are compar-
able in cost in reasonable soils, sands and fills, where obstructions are few, but
high-torque rigs using hard–hard secant methods are necessary where there
are heavy obstructions or substantial rock thicknesses. Hydraulic cutters for
diaphragm walls are favoured cost-wise for sandy formations with substantial
rock layers on large open sites.
Reference should also be made to mobilization costs for walling equipment.
A low-torque CFA rig can be mobilized for less than £10 000, while a high-
torque cased secant machine can cost £70 000 to £100 000 to mobilize and
Vertical soil support: wall construction 143

Table 4.3 Relative costs of various walling techniques in different ground conditions and site circumstances10

Ground conditions and site Wall CFA CFA Classical Cased Grab Hydraulic
circumstances thickness hard–soft hard–hard cased hard–hard diaphragm cutter
(mm) secant secant oscillator secant wall diaphragm
low-torque high-torque hard–hard high-torque wall
rigs rigs secant rigs

Sands and fine-grained fills. <650 1. 0 1.4 – 1.6 1.05 1.1


No obstructions. Open site 650–800 1. 0 1.2 – 1.35 1.0 1.1
850–1000 – 1.25 1.4 1.15 1.0 1.1
1050–1200 – – 1.05 1.15 1.0 1.05
1200–1500 – – – – 1.0 1.1
Clays and fine-grained fills. <650 1.0 1.3 – 1.5 1.0 –
No obstructions. Open site 650–800 1. 0 1.15 – 1.3 1.0 –
850–1000 – 1.25 1.4 1.15 1.0 –
1050–1200 – – 1.1 1.15 1.0 –
1200–1500 – – – – 1.0 –
Sands and fills containing <650 – 1.0 – 1.2 1.05 1.1
some wood, bricks, etc. 650–800 – 1.0 – 1.1 1.05 1.1
and brickwork. Open site 850–1000 – – 1.15 1.0 1.0 1.1
1050–1200 – – 1.0 1.0 1.0 1.05
1200–1500 – – – – 1.0 1.1
Clays and fills containing <650 – 1.0 – 1.25 1.05 –
some wood, bricks, etc. 650–800 – 1.0 – 1.15 1.05 –
and brickwork. Open site 850–1000 – – 1.2 1.0 1.0 –
1050–1200 – – 1.0 1.0 1.0 –
1200–1500 – – – – 1.0 –
Sands, clays and fine- <650 1.0 1.4 – 1.6 1.25 –
grained fills. No 650–800 1. 0 1.2 – 1.35 1.2 –
obstructions. Small 850–1000 – 1.05 1.2 1.0 1.0 –
congested site 1050–1200 – – 1.0 1.0 1.05 –
1200–1500 – – – – 1.0 –
Sands, clays and fills <650 – 1.0 – 1.2 1.25 –
containing some wood, 650–800 – 1.0 – 1.1 1.25 –
bricks, etc. and brickwork. 850–1000 – – 1.5 1.0 1.2 –
Small or congested site 1050–1200 – – 1.0 1.0 1.2 –
1200–1500 – – – – 1.0 –
Sands, clays and fills All – – – 1.0 – –
containing heavy thicknesses
obstructions including mass
concrete, some steel etc.
Open or small congested site
All sandy formations with All – – – 1.25 – 1.0
substantial rock layers. Open thicknesses
site
All clayey formations with All – – – 1.0 – –
substantial rock layers thicknesses
All formations with All – – – 1.0 – –
substantial rock layers. Small thicknesses
or congested site

demobilize on site. Diaphragm wall mobilization costs range from £50 000 for
a single rope grab unit with equipment, to more than £100 000 for a single
Hydrofraise unit with equipment. Minimum economical job sizes are there-
fore probably of the order of 1500 to 2000 m2 for grab excavation and
144 Deep excavations

5000 m2 or more for Hydrofraise or Trenchcutter work. A further discussion


of the costs of walls for basement construction is included in Chapter 8.

References 1. Peck R.B. Deep Excavations and tunneling in soft ground. Proc. 7th Int. Conf.
S.M.F.E., Mexico City, 1969, State of the art volume, 225–290. Sociedad de
Mexicana de Suelos, Mexico City, 1969.
2. Potts D.M. and Day R.A. Use of sheet piling retaining walls for deep excavations
in stiff clay. Proc. Instn Civ.Engrs, 1990, 88, Dec., 899–927.
3. BS 5228. Noise control on construction sites. British Standards Institution,
London, 1992.
4. British Steel Corporation. Control of vibration and noise during piling. British
Steel Corporation, Scunthorpe, 1994.
5. Puller M.J. Economics of basement construction. Proc. Conf. Diaphragm Walls
and Anchorages. London, 1975, 171–180.
6. Sellmeijer J. and Schouten C. Steel sheet pile seepage resistance. Proc. 4th Int.
Landfill Symposium, Cagliari, 1993.
7. Arbed Group. The impervious steel sheet pile wall. Part 1 design. Arbed, 1998.
8. Arbed Group. The impervious steel sheet pile wall. Part 2 practical aspects. Arbed,
1998.
9. Schnabel H. Sloped sheeting. Civ. Engng, 1971, 41, No. 2, 48–50.
10. Sherwood D.E. et al. Recent developments in secant bored pile wall construction.
Proc. Conf. Piling and Deep Foundations, London, 1989, 211–219. DFI, Engle-
wood Cliffs, New Jersey, 1989.
11. Puller M.J. and Puller D.J. Developments in structural slurry walls. Proc. Conf.
Retaining Walls. Institution of Civil Engineers, London, 1992, 373–384.
12. Crawley J. and Glass P. Westminster Station, London. Proc. Deep Foundations
Conf., Vienna, 1998, 5.18.1–5.18.14. DFI, Englewood Cliffs, New Jersey, 1998.
13. Fleming W.K. and Sliwinski Z.J. Use of bentonite in bored pile construction.
CIRIA, London, 1977, Publication PG3.
14. Xanthakos P. Slurry walls. McGraw-Hill, New York, 1979.
15. Corbett, B.O. et al. A load bearing wall at Kensington Town Hall, London. Proc.
Conf. Diaphragm Walls and Anchorages, London, 1975, 57–62. Institution of
Civil Engineers, London, 1975.
16. Puller M. J. Slurry trench stability. Theoretical and practical aspects. Ground
Engng, 1974, 7, Sept., 34–36.
17. Nash J.K. and Jones, G.F. The support of trenches using fluid mud. Proc. Symp.
Grouts and Drilling Muds. Butterworth, London, 1963, 177–180.
18. Muller-Kirchenbauer H. Stability of slurry trenches. Proc. 5th Euro. Conf.
SMFE, Madrid, 1972, Vol. 3, 543–553. Sociedad Española de Mechánica del
Suelo y Cimentaciones, Madrid, 1972.
19. Piaskowski A. and Kowalewski Z. Applications of thixotropic clay suspensions.
Proc. 6th Int. Conf. S.M.F.E., Montreal, 1965, Vol. 2, 526–529. University of
Toronto Press, 1965.
20. Fernandez R. Discussion. Proc. 5th Euro. Conf. S.M.F.E. Madrid, 1972, Vol. 2,
366–373. Sociedad Española de Mechánica del Suelo y Cimentaciones, Madrid,
1972.
21. Huder J. Stability of bentonite slurry trenches with some experience in Swiss
practice. Proc. 5th Euro Conference. S.M.F.E., Madrid, 1972, Vol. 1, 517–522.
Sociedad Española de Mechánica del Suelo y Cimentaciones, Madrid, 1972.
22. Mastikian L. Nouveau Port de Bristol. Proc. 6th Euro. Conf. S.M.F.E., Vienna,
1976, Vol. 2.2, 41–44. Austrian National Committee, Vienna, 1976.
23. Cambefort H. Ge´otechnique de l’ingeniur. Editions Eyrolles, Paris, 1972.
24. Specification for piling and embedded retaining walls. Institution of Civil Engineers,
London, 1996.
25. EN 1538: 2000. Execution of special geotechnical works: diaphragm walls. British
Standards Institution, London, 2000.
26. Colas des Francs E. Prefasil prefabricated diaphragm walls. Proc. Conf.
Diaphragm Walls and Anchorages, London, 1975, 81–88. Institution of Civil
Engineers, London, 1975.
Vertical soil support: wall construction 145

27. Gysi H.J. et al. Prestressed diaphragm walls. Proc. Euro. Conf. S.M.F.E., Vienna,
1976, Vol. 1.1, 141–148. Austrian National Committee, Vienna, 1976.
28. Fuchsberger M. and Gysi H.J. Post tensioned diaphragm wall at Irlams o’ th’
Height, Manchester. Proc. 8th Conf. F.I.P., London, 1978, supplementary
paper. F.I.P., Paris, 1978.
29. Barley A.D. Ten thousand anchorages in rock, part 3. Ground Engineering, 8,
Nov. 1988, 35–39.
30. Glass P. and Stones C. Construction of Westminster Station London. Proc. Instn
Civ. Engrs, Structures and Buildings, 2001, 146, Aug., 237–252.
31. Essential guide to the ICE specification. Federation of Piling Specialists/Institu-
tion of Civil Engineers. Thos. Telford, London, 1999.

Bibliography Banks D.J. et al. Construction of Riding Mill Weir. Proc. Instn Civ. Engrs, Part 1,
1985, 77, Feb., 195–216.
Bell A.L. The lateral pressure and resistance of clay and the supporting power of clay
foundations. Proc. Instn Civ. Engrs, 1915, 199, Feb., 233–272.
BS 6349. Code of practice for maritime structures. British Standards Institution,
London, 1984.
BS 8004. Code of practice for foundations. British Standards Institution, London,
1986.
BS 8002. Code of practice for earth retaining structures. British Standards Institution,
London, 1994.
BS EN 12063. Sheet piling. British Standards Institution, London, 1999.
BS EN 1536. Bored piles. British Standards Institution, London, 1999.
Calkin D.W. and Mundy J.K. Temporary works for the pumping stations at Plover
Cove reservoir, Hong Kong. Proc. Instn Civ. Engrs, Part 1, 1987, 82, Dec., 1121–
1144.
Design manual No. 7. Soil mechanics, foundations and earth structures. US Navy,
Washington, DC, 1982.
Fenoux G.Y. Le realisation fouilles en site urbain. Travaux, Parts 437 and 438, 1971,
Aug.–Sept., 18–37.
Puller M.J. The waterproofness of structural diaphragm walls. Proc. Instn Civ. Engrs,
Geotech. Engng, 1994, 107, Jan., 47–57.
Sarsby R.W. Noise from sheet piling operations – M67 Denton relief road. Proc. Instn
Civ. Engrs, Part 1, 1982, 72, Feb., 15–26.
Design of vertical soil support
5
This chapter addresses the key design items of earth (and water) pressure on
a vertical wall and the analysis of the wall to withstand these pressures.
Cantilevered, single-propped (or anchored) and multi-propped walls will be
considered for both temporary and permanent works.
Geotechnical categories for excavation works are defined in Eurocode 7
(EC7)1 as follows.
. Category 1. Walls that are small and relatively simple structures with a
retained height not exceeding 2 m.
. Category 2. Conventional structures with no abnormal risks or
unusually difficult ground or loading conditions. These walls require
site-specific geotechnical data to be obtained and analyses to be made.
. Category 3. Walls or parts of structures which fall outside categories 1
and 2 and include very large structures, ‘one-off’ structures and those
involving high risk such as unusual ground conditions or loading (for
instance, seismicity).
The design of a typically deep wall for a basement, cofferdam or cut-and-cover
structure will require decisions on walling method, use of props or anchors,
methods of dewatering/recharge, consideration of repropping during con-
struction, acceptable safety factors and allowable settlement and deformation.
The wall design will require the consideration of both ultimate and service-
ability limit states. In practical terms the following matters will need to be
addressed.
(a) At ultimate limit state:
(i) overall stability; strut anchor or waling failure, bending stress
failure in the sheeting, passive failure of soil below successive
stages of excavation, failure by lack of vertical equilibrium of the
wall
(ii) foundation heave – in soft clays, risk of failure by unloading–
bearing capacity failure
(iii) hydraulic failure – piping in cohesionless soils with high external
groundwater table.
Walls should be designed, wherever possible, so that adequate warning
is given in advance of collapse by visible signs. The risk of progressive
failure of retaining structures needs to be particularly addressed,
especially at the temporary works stage.
(b) At serviceability limit state:
(i) deformation of sheeting – avoidance of excessive reduction in
working space or space for permanent structure at construction
stage
(ii) excessive soil movements behind the wall or sheeting; vertical
settlements not to exceed the permitted settlements of existing
structures, services or highways
Design of vertical soil support 147

(iii) unacceptable leakage through or beneath the wall


(iv) unacceptable flexural cracking of the structure
(v) unacceptable cracking in reinforced concrete walls. Crack control
requirements in the UK are specified in BS 81102 or where more
rigorous water resistance is needed in BS 80073 . For highway
structures in the UK BS 54004 and Highways Agency BD 42/005
apply.
The following practical points arise concerning wall and soil movement:
. The magnitude of movements in walls and retained soil is dependent
in large measure on workmanship standards during wall installation,
excavation and wall support works.
. Generally, flexible walls with many propping levels will give similar
displacements to stiff walls with few props.
. Wall embedment depth below formation level influences soil deforma-
tion, particularly in soft clays.
. Stiffness of propping systems influences soil deformation.
. Top-downwards construction minimizes soil deformation and settle-
ments.
. Installation of a temporary or permanent support of high stiffness at
shallow depth in an excavation is an effective control of wall deforma-
tion and soil deformation.
. Ground movement cannot be estimated with precision. As a broad
indication of movements due to wall installation the following may be
helpful.
In stiff clay: For bored piles: vertical settlement at wall, 0.04% of depth.
For diaphragm walls: vertical settlement at wall, 0.05% of
wall depth.
The distance behind the wall to negligible settlement is approximately twice
wall depth. Similarly, broad indications of patterns of wall deflection and
settlements can be obtained from empirical results published by Peck6 ,
Clough et al.7 , Clough and O’Rourke8 , St John et al.9 , Carder10 , Fernie and
Suckling11 , Carder et al.12 and Long13 .
For excavations in stiff clay, observed soil settlement at the wall due to
excavation in front of the wall is generally within the range 0.1% of
excavation depth for high-stiffness walls to 0.3% of excavation depth for
low-stiffness walls. In sands, a limited number of observations show a similar
vertical settlement at the wall of 0.1 to 0.3% of excavation depth. (The
Author’s experience of several excavations to 20 m depth in medium dense
silty sands in Saudi Arabia using anchored king post walls as support
shows consistent vertical settlement at the wall of less than 0.1% following
excavation to full depth.)

Earth pressures: Mohr’s circle of stress can be used to illustrate graphically the limiting
limiting horizontal horizontal soil pressures which can be generated in active or passive states
pressure at either side of a retaining wall. The pressures will be considered in terms
of effective stress, ignoring, for the time being, the effects of wall friction or
wall adhesion.
The problem of assessing earth pressures can be solved using the Coulomb
equation for shear strength  0 ¼ c0 þ 0 tan 0 , where c0 represents effective
soil cohesion, 0 represents effective vertical stress and 0 represents the
angle of shearing resistance in terms of effective stress. Figure 5.1 shows the
introduction of effective stress parameters, the total vertical stresses and
148 Deep excavations

Fig. 5.1. Mohr’s circle values of


limit pressures and values of Ka
and Kp

water pressures (and hence effective vertical stresses) to calculate active and
passive earth pressure according to whether horizontal stress is less than or
greater than vertical stress. Mohr’s circles are seen to touch the failure envel-
ope to give limiting values in the extremes of minimum and maximum hori-
zontal stresses.
Using active and passive earth pressures coefficients,
0H ¼ Ka 0v ð39Þ
and
0H ¼ Kp 0v ð40Þ
where
1  sin 0
Ka ¼ ð41Þ
1 þ sin 0
and
1 þ sin 0
Kp ¼ ð42Þ
1  sin 0
The values of 0 and c0 can be obtained for clay samples from drained triaxial
0
Table 5.1 Estimation of  for tests or undrained tests with pore-water pressure measurement. The effects of
cohesive soils17 stress level, rate of strain for testing, the degree of weathering and of sample
Plasticity 0 swelling before the test, will all influence the test result. When test results are
index not available for 0 , the relationship in Table 5.1 may be used for conservative
values. The design parameters 0 and c0 will cause considerable variation in
15 308 values of limiting soil pressure, and while it is not unusual to assume c0 ¼ 0,
20 288 the effect of this assumption is to reduce to low values the limiting passive
25 278 pressures immediately below dredge or formation level. In turn, this tends
30 258 to produce high wall or sheeting moments and increased strut or anchor
40 228
loads near this level.
50 208
80 158 The values of 0 in cohesionless soils should be based on in situ tests where
possible. The relationship between standard penetration resistance and 0 is
Design of vertical soil support 149

Fig. 5.2. Relationship between


standard penetration
resistance and 0 17

Table 5.2 Estimation of 0 for


weak rocks19 Stratum 0

Chalk 358
Clayey marl 288
Sandy marl 338
Weak sandstone 428
Weak siltstone 358
Weak mudstone 288
Notes. The presence of a preferred orientation of joints, bedding or cleavage
in a direction near that of a possible failure plane may require a reduction in
the above values, especially if the discontinuities are filled with weaker
materials.
Chalk is defined here as unweathered medium-to-hard, rubbly or blocky
chalk.

shown in Fig. 5.2. Table 5.2 gives approximate values for the effective angle of
shearing resistance of soft rocks considered, somewhat conservatively, as a
mass of granular fragments.

Horizontal earth pressure: at rest


In an undisturbed soil mass the horizontal earth pressure, the at rest pressure,
is K0 0v , where K0 is the coefficient of earth pressure at rest and 0v is the
effective vertical stress. For normally-consolidated soils K0 is approximately
equal to the empirical expression
0H
KONC ¼ ¼ 1  sin 0 : ð43Þ
0v
For normally-consolidated clay KONC lies in the range 0.55 to 0.65. For over-
consolidated clays the removal of considerable overburden pressures within
geological time leaves horizontal stress as the major principal stress. The
reduction in vertical stress is shown at any depth by the over-consolidation
ratio (OCR), which is the ratio of the maximum vertical effective stress ever
experienced by the soil to the current vertical effective stress at that depth.
The value of 0v can be obtained from laboratory consolidation tests and
the general relationship between K0 and the OCR is given in Fig. 5.3 after
Brooker and Ireland14 . The maximum value of K0 may lie in the range 2 to
150 Deep excavations

Fig. 5.3. Relationship between


coefficient of earth pressure
at rest K0 and
over-consolidation ratio14

3 for some clays, reducing with depth since OCR reduces with depth. K0 can
also be estimated from pressuremeter test results.

Horizontal earth pressure based on long-term, drained, effective


stress values
The limiting horizontal active and passive earth pressures acting on the wall at
any depth z are given, respectively, by
p0a ¼ Ka 0v ¼ Ka ðz þ q  uÞ ð44Þ
and
p0p ¼ Kp 0v ¼ Kp ðz þ q  uÞ ð45Þ
and for a soil with a cohesion intercept
p0a ¼ Ka ðz þ q  uÞ  2c0 ðKa Þ1=2 ð46Þ
and
p0p ¼ Kp ðz þ q  uÞ  2c0 ðKp Þ1=2 ð47Þ
where  is the bulk soil density (saturated density if below water level), q is the
surcharge on ground surface, and u is the pore-water pressure. These values,
known as Rankine values, are subject to change when the effects of wall fric-
tion and wall adhesion are considered. Walls are not made of perfectly smooth
material and wall friction  and adhesion cw vary soil stresses proportionately.
The limiting effective active and passive pressures acting horizontally at a
depth z are:
p0a ¼ Ka ðz þ q  uÞ  Kac c0 ð48Þ
and
p0p ¼ Kp ðz þ q  uÞ  Kpc c0 ð49Þ
where
  1=2
cw
Kac ¼ 2 Ka 1  0 ð50Þ
c
and
  1=2
cw
Kpc ¼ 2 Kp 1  0 ð51Þ
c
and c0 is the effective shear strength.
The pore-water pressure is added to the effective horizontal earth pressure
to give the sum of earth and water pressure: pa ¼ p0a þ u and pp ¼ p0p þ u.
Design of vertical soil support 151

Fig. 5.4. Coefficient of active earth pressure (horizontal Fig. 5.5. Coefficient of passive earth pressure (horizontal
component) for a horizontal retained surface15 component) for a horizontal retained surface15

The earth pressure coefficients Ka and Kp have been calculated for various
curved failure surfaces; those frequently used are due to Caquot and Kerisel15
and are based on a logarithmic spiral surface as shown in Figs. 5.4 and 5.5.
The coefficients derived from these figures have been corrected to give
horizontal pressures although the actual pressures are inclined at the angle
of wall friction  to the horizontal. Value of Kac and Kpc can be calculated
from the expressions above.

Horizontal earth pressure based on short-term undrained stress


values
In fine-grained soils such as fine silts and clays, the relatively low permeability
only allows slow changes in moisture content, pore pressure and overall
volume. The instant that load is applied, pore pressure increases and the
shear strength of the soil, which depends on the locked-in effective stress, is
only able to increase slowly as pore pressures reduce. This immediate shear
strength, the undrained value cu , applies before any volume change and
pore pressures alter or, expressed more exactly, the value of cu can only be
used where a new system of boundary stresses is imposed but prior to any dis-
sipation of the new excess pore-water pressures. The undrained shear strength
cu is only correctly used immediately load is applied and it is strictly illogical to
use it as soon as pore pressures change. As effective stresses change with pore
pressure, shear strength also changes. In clay soils, where the retaining wall
structure deforms and attempts to move away from the retained soil bulk,
negative pore pressures are generated in the retained soil as excavation pro-
ceeds in front of the wall. In highly fissured or laminated clays the reduction
in negative pore pressure may proceed relatively quickly and the original value
of cu quickly becomes inapplicable. The use of cu in such clays can therefore
become over-optimistic in retaining wall analysis when the effective stress (on
which soil strength depends) reduces as negative pore pressure relaxes.
In soft and very soft homogeneous clays which do not benefit from drainage
paths due to fissuring and jointing, low overall permeability of the soil fabric
may prevent more efficient change in pore pressure, and the use of cu ,
although strictly applicable at the instant of stress change prior to pore
pressure change, can be more satisfactorily relied upon. Indeed, for retaining
walls founded in soft clays it is prudent to obtain both drained and undrained
soil strengths and undertake analysis of both total and effective stress
152 Deep excavations

conditions. Temporary works design in these soils, in which the wall structure
may have a limited design life, say, of six months, may therefore use total
stress methods with some confidence, and the design parameter of undrained
shear strength can be conveniently and accurately obtained from in situ shear
tests, such as vane tests. Where such short-term conditions do not apply,
either by the greater permanence of the temporary wall or due to its role in
the permanent structure, analysis of drained effective stress becomes neces-
sary. The choice between the two methods, undrained or drained, depends
strictly on the rate of change of pore pressure which, in turn, depends on
the permeability of the structure, although it must be restated that the
undrained analysis is only strictly applicable immediately before any dissipa-
tion commences. In practice, especially for small schemes, sufficient undrained
shear measurements may be more readily available than test results from
drained samples. However, the availability of test results should not excuse
the application of an illogical method of analysis. The short-term undrained
method is therefore presented here with the recommendation that its use
should be restricted to soft clay conditions for temporary works of short
duration. Schemes of longer duration should be checked by both drained
and undrained analyses.
In terms of undrained conditions the limiting horizontal active and passive
earth pressures acting on the wall at any depth z are given by
pa ¼ Ka v  Kac cu ¼ Ka ðz þ qÞ  Kac cu ð52Þ
and
pp ¼ Kp v  Kpc cu ¼ Kp ðv þ qÞ  Kpc cu : ð53Þ
The pressure coefficients are
 
cw 1=2
Kac ¼ 2 1 þ ð54Þ
cu
and
 
cw 1=2
Kpc ¼2 1þ : ð55Þ
cu

Tension cracks The active pressure near ground level will be a minimum for the retaining wall
and, in cohesive soils, surface tensions within the clay can cause cracks which
can fill with groundwater or rain water. In this case, the water pressure within
the potential crack should be allowed as a pressure on the back of the wall.
Wall adhesion will not apply over the depth of the crack, which is ð2cu  qÞ=.

Minimum values of calculated active pressure


In the UK the prolonged use, from 1951 until 1994, of the Civil Engineering
Code of Practice for earth-retaining structures16 indicated both the sound
principles on which it was based and the difficulties which have been experi-
enced in agreeing revisions to it. Although the Code was based on calculation
of earth pressure in terms of total stress and undrained shear values, it
prudently recommended the use of minimum active pressures for design
irrespective of those calculated from test values of cu using the expression
pan ¼ Ka v  Kac cu .
The Code stated that for cohesive soils the methods of calculation were not
final. In the design of structures to retain cohesive soils, a suitable addition
should be made to the calculated active pressure of the soil in order to provide
for uncertainties. Where, as may be the case with stiffer clays, the total
Design of vertical soil support 153

pressure on the wall as calculated according to the Code was small, the value
of the total pressure to be assumed for the purposes of design should not be
less than that found by assuming the horizontal pressure at any depth to be
that due to a fluid with a density of 5 kN/m3 . This safety net for minimum
pressure at the active side of the wall was most prudent but did not find
mention in some later Codes.

Softening of clays The process of clay softening is accelerated over time by tension cracks near
the ground surface, by the joints and fissures which frequently occur in stiff,
over-consolidated clays, and by the laminations of silt or sandy silt partings
that can occur in all clays. These routes for easier passage of water not only
increase the permeability of the soil fabric but can form failure surfaces them-
selves of softened planes of clay dividing stiffer, unsoftened blocks of clay
bounded by softened, lubricated surfaces. The changes in pore-water pressure
as these pressures equalize with hydrostatic groundwater levels are therefore
also associated with a loss of clay strength due to softening along surfaces
such as joints, fissures, pre-existing shear surfaces, laminations and tensile
zones. For this reason it is often prudent to assume a complete loss of cohesive
strength in terms of effective stress as the clay softens with time along
preferred drainage surfaces.

Wall friction and wall The effect of assumptions regarding wall friction in granular soils and adhesion
adhesion in clay soils on calculated values of active and passive pressure is considerable,
and while it is self-evident that wall surfaces are not smooth, design values of
friction and adhesion should be chosen carefully. The mobilization of wall
friction, to reduce earth pressures on the active side and increase earth
pressure on the passive side, necessitates the downwards movement of a soil
wedge on the active side and the upwards movement of a similar soil wedge
on the opposite, passive side, as shown in Fig. 5.6. Important changes to
these relative soil and wall movements are caused by the downwards vertical
wall movement due to vertical loads on the wall and the vertical load compo-
nents of inclined pre-loaded ground anchors, both of which tend to mobilize
passive wall friction/adhesion but tend to reduce active friction/adhesion.
Typical values of active wall friction and adhesion as quoted in Codes and
values for both active and passive friction and adhesion in the CIRIA report
for sheet piled cofferdams17 are shown in Tables 5.3 and 5.4. These values are
all based for wall friction on peak values of  and it is understandable that
some conservatism should apply since the peak values themselves may not
apply. More logically, it may be prudent to use max ¼ 0crit if the critical
state value is known. If the wall surface texture is less than the typical particle
size of the soil ðD50 Þ, lower values of  would apply.
The relative movement between wall and soil necessary to generate wall
friction or adhesion may not always occur; where the toe of the wall is

Fig. 5.6. Wall friction acting on


active and passive soil
wedges48
154 Deep excavations

Table 5.3 Values of active wall


friction angle and adhesion Construction Wall friction Wall adhesion
angle, 

Concrete, brick 208 Where cw is less than 50 kN/m2 use


cw ¼ cu
Steel piling, tar/bitumen 308 Where walling or sheeting does not
coated penetrate any appreciable depth use cw ¼ 0
Uncoated sheet piles 158 Where cw > 50 kN/m2 use cw ¼ 50 kN/m2

Table 5.4 Active and passive


values of wall friction/adhesion Wall friction angle,  Wall adhesion

Active 0.67 0.5c0 or 0.5cu >


= 50 kN/m2
Passive 0.5 . 0 .
0 5c or 0 5cu >= 25 kN/m2

founded on hard rock, wall friction/adhesion values quoted in Table 5.4


should be reduced by 50% for dense granular materials and stiff over-
consolidated clay. For overlying loose granular soil, where the wall is founded
on rock, wall friction/adhesion should be ignored. For anchored walls where
there is a tendency for the wall to move downwards relative to the soil, zero
wall friction should be allowed.
Neither wall adhesion nor wall friction should be allowed in granular sub-
soils where the proximity of machinery, railways or vehicular traffic causes
vibrations which could be transmitted through the subsoil to the wall surface.
With cast in situ reinforced concrete diaphragm walls and reinforced con-
crete piled walls cast under bentonite slurry, values of wall friction/adhesion
for active and passive pressure can be assumed to be similar to those for
concrete walls given in Table 5.3.

Magnitude of The variation in horizontal pressure, either active or passive, generated by


movement needed to movement of the wall away from or towards the soil mass, is shown in
mobilize limit Fig. 5.7. The curve, first derived by Terzaghi and Peck18 was based on tests
pressures in dense sand. Two conclusions are evident:
. the passive coefficient is much higher than the active coefficient
. the deformation needed to fully mobilize the limit passive pressure
is much greater than that required to fully mobilize the limit active
pressure.
The relative deformation needed to fully mobilize Ka and Kp by means of the
stress path followed by a soil element behind the wall in both normally- and

Fig. 5.7. Relationship between


strain required to mobilize
active and passive pressure for
medium dense sand48
Design of vertical soil support 155

Fig. 5.8. Stress paths followed


for elements behind and in
front of a retaining wall in
normally- and
over-consolidated clays48

over-consolidated clay conditions is shown in Fig. 5.8. On the active side there
is no change in vertical stress due to excavation, while on the passive side the
effective stress path in drained loading may take varying directions depending
on the relative importance of the relief of overburden to excavation and the
horizontal pressure exerted on the wall.
The stress change needed to fail a normally-consolidated clay in active
pressure is much less than that required to fail the same clay in passive
pressure. The stress path for over-consolidated clay shows the reverse to be
true, however. Terzaghi and Peck’s18 plot of wall movement against mobilized
pressure for dense sand may not therefore, accurately show the strains
necessary to generate limit active and passive pressures in both normally-
and over-consolidated clays.
BS 800219 followed the concept of limit state design, with greater emphasis
on wall conditions at the serviceability limit state than collapse conditions at
the ultimate limit state. The basis of this method is the reduction of soil peak
shear strength values on both the active and passive sides of the wall, to values
156 Deep excavations

which would be representative of mobilized shear strengths with the permitted


wall movement. BS 800219 specified that where wall displacements are
required to be less than 0.5% of the wall height, the representative undrained
shear strength should be divided by a mobilization factor M not less than 1.5.
For designs using effective stress values, the value of M was specified to be
1.2 to reduce horizontal wall movements to 0.5% of wall height. For
effective stress design, the calculation of wall equilibrium and the structural
components of the wall was specified to be based on the lesser of:
(a) the representative peak strength of the soil divided by a factor M ¼ 1:2
representative tan 0max
design tan 0 ¼
M
representative c0
design c0 ¼
M
(b) the representative critical-state strength of the soil.
The earth pressures on the active side are thus increased and similarly those on
the passive side are reduced at serviceability state by the use of the reduced
peak shear strength. The Code concludes that structural design of the wall
components can then usually be made without the application of partial
load factors to bending moments and internal structural forces.
This principle of design using soil strengths that can be mobilized at
working conditions in the serviceability limit state is logically extended in
BS 8002 to consideration of wall friction and wall adhesion. It recommends
that the design value of friction or adhesion at the interface with the wall
should be the lesser of either the values obtained from tests or 75% of the
design shear strength to be mobilized in the soil itself, that is, using
design tan  ¼ 0:75 design tan 0
design cw ¼ 0:75 design cu
and since for the soil mass M ¼ 1:2
representative tan 0
design tan 0 ¼
1:2
this is equivalent to
design  2
0 ffi
representative  3
and similarly, in total stress analysis
design cw
¼ 0:5 after taking M ¼ 1:5:
representative cu
The overall effect of reducing peak shear strengths to shear strengths which
are mobilized in working conditions by using such M values as are specified
in the Code may not be necessary when walls are designed in stiff over-
consolidated clays, where relatively small movements are necessary to
mobilize limiting active and passive pressures. On the other hand, the recom-
mendation that bending moments and internal forces derived from earth
pressures calculated from mobilized soil strengths (peak strengths reduced
by the mobilization factor M), without the application of a further partial
safety factor on earth pressures allowed in designs based on ultimate struc-
tural strengths, may not accurately reflect the variations in load that occur
in practice along the length of a wall.
Design of vertical soil support 157

Fig. 5.9. Influence of wall


flexibility on pressure
distribution: (a) rigid wall;
(b) flexible wall48

BS 8002 also specified that in checking the wall stability and soil defor-
mation all walls should be designed for a minimum design surcharge loading
of 10 kN/m2 and a minimum depth of additional unplanned excavation in
front of the wall. This additional design depth was specified to be not less
than 0.5 m or 10% of the total height retained for cantilever walls or of the
height retained below the lowest support level for propped or anchored
walls. It should be noted that BS 8002 was originally applicable to walls up
to about 8 m high, and no differentiation is made between temporary and
permanent walls. A subsequent revision increased this wall depth to 15 m.
The imposition of such rules for all walls irrespective of design life would
appear to be over-stringent.

Wall flexibility The stiffer the wall, the greater the earth pressure it attracts and, conversely,
the greater flexure of the wall the less pressure (and moment) induced in it by
the soil it retains. This phenomenon, originally demonstrated in model tests
on an anchored wall by Rowe20 will be referred to in greater detail later in
this chapter.
Consider, however, the difference in deformed shape of a relatively stiff
anchored wall and that of an anchored flexible wall, as shown in Fig. 5.9.
The flexible wall distorts outwards by a considerable amount at mid-span
relative to the stiff wall, causing a reduction in pressure and an effective reduc-
tion in vertical span of the wall due to a rise in elevation of the centre of
passive support below excavation level. Anchor loads and pressure below
excavation level therefore both increase, but bending moment in the wall
reduces due to both earth pressure reduction at mid-span and the reduction
in the span itself. Due to the movement of the flexible wall away from the
soil at mid-span, the soil tends to arch vertically and, providing neither the
anchor nor the passive support below formation level yield once mobilized,
the abutments to the arch sustain more load and relieve soil pressure from
mid-span.

Surcharge loads The effects of additional loads, surcharging the wall, must be taken into
account at both ultimate and serviceability limit states. Loads due to adjacent
highways, railways, variations in ground level, storage areas for building
materials and plant loads need to be taken into account.

Highway traffic loading


In the UK, guidance on highway traffic loading is given in the publication
BD 37/0121 . Two categories of loading are specified, HA loading that
normally occurs on highways and HB loading, corresponding to abnormal
158 Deep excavations

vehicle loads such as industrial loads. The number of HB units is generally


specified. The loadings may be summarized as follows.
. HB loading (45 units): equivalent surcharge ¼ 20 kN/m2
. HB loading (37 units): equivalent surcharge ¼ 16 kN/m2
. HB loading (30 units): equivalent surcharge ¼ 12 kN/m2
. HA loading: equivalent surcharge ¼ 10 kN/m2

Fig. 5.10. Coefficients of earth pressure (horizontal component) for vertical walls with inclined backfill for varying wall
friction : (a) active; (b) passive15
Design of vertical soil support 159

Railway loading
Railway universal loading (RU) allows for all combinations of rail traffic
which currently runs on UK track – or is expected to run in the future. RL
is a reduced loading, for lighter traffic, such as in use on underground
railways. Both loads are applied over the track width, overall the length of
sleepers. The loadings are:
. RU loading: equivalent surcharge ¼ 50 kN/m2
. RL loading: equivalent surcharge ¼ 30 kN/m2

Surcharge loading: sloping ground


The effect of a sloping ground surface can be taken into account using curves
due to Caquot and Kerisel15 reproduced in Fig. 5.10, giving values of Ka and
Kp for vertical walls, inclined backfill and wall friction varying from  ¼ 0 to
 ¼ 0 . The effect of layered strata behind the wall should be taken into
account in calculations by using modified coefficients for each layer from
the curves. The values presented assume that the sloping ground extends a
sufficient distance from the wall to exceed the lateral dimension of the poten-
tial critical slip plane of the failure wedge (approximately at an angle of
458 þ =2 to the horizontal from the toe of the wall for the active case, and
458  =2 for the passive case).
The CIRIA report17 gives a simple adjustment to the pressure diagrams
(Fig. 5.11(a) and (b)) to show the effect of berms, on the active and passive
sides of the wall in cohesionless soil. The report notes that the risk for sliding
failure should be checked at the base of the berm, especially where the berm
soil has a higher angle of shearing resistance than the soil beneath it. Where
a gravel berm overlies a stiff clay, the report suggests it may be too optimistic
to use the undrained clay strength in checking sliding.
The NAVFAC design manual22 gives a graphical method for the calcula-
tion of passive soil resistance due to an earth berm, as shown in Fig. 5.11(c).

Surcharge loading: point loads and line loads


The CIRIA report17 reproduced a simple method due to Krey for use in
granular soils. The horizontal pressures caused by concentrated point and
line loads as estimated by Krey are shown in Fig. 5.12 and should be added
to the earth pressure diagram.
An alternative method of computing the effects of point and line loads by
Terzaghi was reported in the NAVFAC design manual22 and was commented
upon in the CIRIA report17 . According to the CIRIA report the method,
a modified Boussinesq distribution taking into account field and model
measurements, possibly underpredicts the pressure and tends to place the
centre of pressure influence too low down the wall. Some caution should be
applied, therefore, in the use of the method where the applied loads are
high in relation to the weight of soil retained. It should be noted, however,
that the method presumes an unyielding rigid wall and the lateral pressures
are already approximately double the values obtained from elastic equations.
Design charts from NAVFAC22 for line and point loads are reproduced in
Fig. 5.13.

Surcharge loading: uniform rectangular surface load


The NAVFAC design guide also provides the means to estimate the lateral
pressure on an unyielding wall due to a uniform rectangular load on the
surface at the rear of the wall, after Goldberg et al.23 (Fig. 5.14). The guide
comments that any yield of the wall during application of the load will tend
to reduce the pressures calculated from the values given by Goldberg et al.
160 Deep excavations

Fig. 5.11. (a), (b) Pressure diagrams showing the effect of sloping ground on the active and passive sides of the wall17 .
(c) Culmann method for determining passive resistance of earth berm22

Wall movement The effect of wall movement on lateral earth pressures should be restated
because of its vital effect on both active and passive pressures. Earth pressures
at rest are reduced to limiting active values by movement of the wall away
from the soil; much larger movements are needed to increase the at-rest
pressure to limiting passive values. The relationship between wall rotation
and the coefficient of earth pressure for sands is reproduced in Fig. 5.15.
The curve can be used to assess the movement needed to obtain limit pressures
in sands of varying compaction state. The NAVFAC guide22 also suggests
Design of vertical soil support 161

Fig. 5.12. Effect on surcharge loading of point and line loads17

that the effects of wall translation on limit pressures can be assessed by


assuming the effect of translation is equivalent to the movement at the top
of the wall due to rotation, as shown in Fig. 5.15.
The extent of wall movement relative to the retained soil therefore deter-
mines the proportion of the limit pressure which applies. This movement
may be derived from elastic deflection of the wall or sheeting itself, the
compression which occurs in strutting or raking shores (or the curtailment
of such movement by pre-loaded supports or anchors), the movement
needed to mobilize passive resistance of a soil at an intermediate stage of
excavation in front of the wall, or the rotation or translation of a rigid wall.
With the exception of the methods specified in BS 800219 the above means
of calculating earth pressures have led to designs based on limit pressures
where these do not necessarily apply if the extent of proposed wall movements
do not justify them. Soil pressures based on the stiffness of the wall, its support
and the soil itself are evaluated more realistically by soil/structure analyses
based either on the Winkler spring theory or by finite element methods.
These methods are referred to again later in this chapter.

Design calculation Eurocode defines design procedures under four headings:


according to Eurocode 7 . by calculation
. by prescriptive measures
. from results of load tests or tests on experimental models
. by observational techniques.
New meanings are given to words within Eurocode 7 as follows.
. actions – foundation loadings
. permanent actions – dead loads, typically weights of structures and
installations
. variable actions – imposed loads, typically wind or snow loads
. accidental actions – vehicle impacts or explosions
. transient actions – a classification related to soil response, for example,
wind loading
. persistent actions – a classification related to soil response, for example,
structural or construction loads.
Eurocode 7 makes use of partial safety factors to allow for uncertainties in the
magnitude and combination of loads (actions) and characteristic values of soil
and rock parameters obtained from field and laboratory testing.
162 Deep excavations

Fig. 5.13. Horizontal pressures on a rigid wall from line loads (left) and point loads (right)22

The design values of vertically applied actions Vd to be used in geotechnical


design are calculated from the expression Vd ¼ Fk  F , where Fk is
the characteristic load and F is the partial factor of safety for the applied
action.
Vd must be equal to or less than the foundation design bearing resistance
Rd . The value of Vd is calculated from characteristic soil strengths also with
a partial factor applied. The design value of the material properties Xd is
Design of vertical soil support 163

Fig. 5.14. Lateral pressure on


an unyielding wall due to
uniform surface load22

calculated from the expression Xd ¼ Xk =m , where Xk is the characteristic


value and m is the particular factor of safety for the material properties.
The specified values of these partial factors ðm Þ together with the partial
factors F to be applied to the actions are shown in Table 5.5.
The three cases shown in this table refer to the overall stability of the struc-
ture, typically buoyancy for an underground structure (case A), the strength
of the structural members is the substructure (case B) and the geotechnical
design of foundations and earthworks (case C). The partial factors for the
ground properties for case C are higher than in cases A and B, allowing for
variations in the ground and uncertainties in the analytical work.
Eurocode 7 (EC7) gives some guidance in the choice of characteristic soil
properties, defining them as those that give the probability of a more unfavour-
able value governing the limit state occurrence of not more than 5%, and
defines a characteristic value of soil properties as a cautious estimate of mean
values within the ground influenced by the foundation loads.

Selection of design parameters


The choice of design parameters in terms of strength and stiffness is particu-
larly important because of the considerable effect on wall design at both
ultimate and serviceability states. Three grades of assessment should be
considered:
(a) moderately conservative soil parameters (case A)
164 Deep excavations

Fig. 5.15. Effect of wall rotation on active and passive pressure in loose and dense sand22

(b) worst credible soil parameters (case B)


(c) most probable soil parameters (case C)
The first category, moderately conservative values, is a relatively cautious
estimate at the particular limit state. It is a similar value to the representative
value as used in BS 8002 and to characteristic values as defined in EC7.

Table 5.5 Partial factors –


ultimate limit states in Case F for actions M for ground properties
persistent and transient
situations1 Permanent Variable tan  c0 cu qu a

Unfavourable Favourable Unfavourable

Case A 1.00 0.95 1.50 1. 1 1.3 1. 2 1. 2


Case B 1.35 1.00 1.50 1. 0 1.0 1. 0 1. 0
Case C 1.00 1.00 1.30 1.25 1.6 1. 4 1. 4
a
Compressive strength of soil or rock.
Design of vertical soil support 165

Fig. 5.16. Assessment of moderately conservative strength vs. depth profile for site on London clay. (a) Undrained shear
strengths at site; (b) SPT test results from same site; (c) SPT results from (b) and results from adjacent sites compared with the
undrained shear strengths; (d) final assessment of all results24

Simpson and Driscoll24 (Fig. 5.16) used a comparison between undrained


cohesion values of London clay, SPT values on the same site and plots of
undrained strength vs. depth from nearby sites to assess a moderately conserva-
tive strength vs. depth profile. The construction is probably best not made on a
statistical basis but following a comparison of data from other reliable sources.
The second category, the worst credible value is the worst that can be
reasonably assessed could occur, an unlikely circumstance. Simpson and
Driscoll24 suggested that these values can be regarded as 0.1% fractile.
In the third category, the most probable values have a 50% chance of
exceedance. It should not be used as the only basis of design but can be
used with the observational approach providing the second design approach
using worst credible values are used to define contingency measures should
trigger values be exceeded.
166 Deep excavations

Certain practical matters arise in preparing a testing schedule which is


appropriate to the design on a specific site. These are as follows.
(a) An adequate number of classification, moisture content and density
tests should be made.
(b) Test correlations should be made with other tests, other sites and where
possible by back analysis in similar conditions.
(c) In stiff clays, care should be exercised in the choice of strength tests,
peak, critical state, residual and drained or undrained.
(d ) Laboratory tests for unconsolidated undrained triaxial values should be
made on 100 mm dia. undisturbed samples as single-stage tests.
(e) Laboratory tests for drained shear strength should be made on con-
solidated drained and consolidated undrained triaxial tests and shear
box tests (for cohesionless soils) and peak, critical state and residual
angles of shearing resistance should be measured. Appropriate rates
of shear should ensure complete dissipation of pore pressure.
( f ) Inspection of the soil fabric should be made visually especially for soft
estuarine clays when continuous sampling may prove necessary.
(g) Soil stiffness may be measured with the best expectancy of accuracy in
self-boring or Menard-type pressuremeter tests in stiff clays and weak
to moderately strong rocks.
(h) Measurement of groundwater levels should be applied stringently,
using standpipes and piezometers to critically assess piezometric
levels, their variation and risks associated with changes in groundwater
level.

Temporary works Important differences exist between the design approach to soil-retaining
structures for permanent and temporary works. The dissimilarities between
temporary works and permanent works design may be summarized as follows.
In temporary works . . .
(a) In clays, negative pore-water pressures are present in the soil. These
dissipate with time but while they act they cause the soil to have a
greater shear strength than it has in the long term.
(b) Support conditions are often different in the temporary works before
base slabs, floor slabs or other permanent supports are constructed.
Temporary works often rely more on the soil for their support than
do permanent works.
(c) Loading conditions are often different. Applied surcharges during
temporary works construction are often not present in the long term.
(d) Ground movements are often a matter of concern during construction,
but there is not usually time for long-term movements to develop.
(e) A greater proportion of soil strength may be mobilized during tempor-
ary works depending on constraints operating during construction.
However, the consequences of excessive ground movements or instabil-
ity can be as severe as for permanent works.
( f ) Higher stresses may be permitted in the design of the wall in temporary
works.
Temporary works may be designed either: by assuming full equilibrium (full
drained) pore-water pressures (i.e. long-term strength), the expected loads
in the temporary phase and a very low safety factor; or by taking the higher
short-term strength and a higher safety factor. (The choice is essentially one
of modifying the parameters used and the safety factor adopted in passing
from temporary to permanent works phases.)
Design of vertical soil support 167

Mixed total and Since the original ground surface and the soil below it on the active side of the
effective stress design wall are usually more susceptible to softening than the soil below final forma-
tion level, a method of calculating earth pressures using effective stress values
on the active side and undrained values on the passive side has gained some
credibility. The benefit, in terms of calculated earth pressures, is obtained
principally from passive pressures based on the limit pressure expression
pp ¼ v þ 2c, that part of the expression relying on cohesion applying at
formation level on the passive side. The risk of clay softening on the passive
side below formation level can be estimated by reducing the undrained
strength/depth profile by 20% to 30% (using 0.7 to 0.8 of the unsoftened
design line for cu ) and ignoring the clay strength in the upper 1 m thickness
of clay at the surface of the excavation. Overall, the loss of safety for the
wall as built will depend on several design matters, which include the choice
of parameters, the choice of total or effective design method and the factors
of safety that are applied to each component of the wall, the bracing or
anchor that supports it, and the extent of embedment of the wall in the
soil below formation level. If more optimistic methods are chosen for wall
analyses (by using total or mixed total and effective stress methods) these
should reflect more stringent factors of safety in the derivation of wall
member dimensions. Conversely, if effective stress methods are used through-
out, say for temporary works design of limited design life, the final factors of
safety in the dimensioning of wall and strutting may be chosen more optimis-
tically. Recommended factors of safety will be reviewed later in this chapter.

Design water pressures The design of walls to support soil will include consideration of the support of
water pressures where there is groundwater above the final excavation level in
front of the wall. Where support walls are designed for excavations to coffer-
dams and shafts below river or sea bed, the pressures from river or sea water
levels are evident to the designer and would comprise the major part of
pressures on the wall. Where earth support walls are required for excavations
on land, the designer has to evaluate the highest ground water levels during
the design life of the structure and incorporate these with earth pressures
into design pressures for the wall. In excavations on land where groundwater
is high, the water pressure will constitute a large proportion of the horizontal
force on the active side of the wall, but is a lesser proportion than the earth
pressure on the passive side. If water levels rise, the total force on the active
side of the wall increases, while if water pressure rise below formation level
on the passive side, in most cases total passive resistance from water and
soil reduces.
Where vertical soil support walls on land penetrate granular soils with a
high water table but obtain a cut-off at depth, the water pressure on each
side of the wall is calculated without complication (Fig. 5.17). The hydrostatic
pressures on each side of the wall are calculated from v ¼ w x where w is
the density of water (in fresh water it is 9.81 kN/m3 , in salt water it is
10.00 kN/m3 ), and x is the depth below the groundwater table.
Additional pressures due to heavy rain may need to be considered in defin-
ing groundwater levels, and where surfaces are unpaved, full water pressures
should be taken in tension cracks in cohesive soils. In cohesionless soils or
where drainage is introduced at the rear of the wall, an increase in soil density
to its fully saturated value should be allowed for. If the retained cohesionless
soil has low permeability but drainage holes through the wall are adequate to
discharge the water flow through the retained soil, it is recommended that the
effect of the flow of water towards the wall and the increased total active thrust
on the wall should be taken into account by calculating total active thrust on a
168 Deep excavations

Fig. 5.17. Hydrostatic water


pressures on walls in
permeable ground with
effective cut-off: (a) gross water
pressures; (b) net water
pressures17

vertical wall:
H2 : H2
Pa ¼ Ka b þ 0 5w ð56Þ
2 2
where b is the submerged density of the soil, and H is the vertical height of the
earth retained.
It is also prudent to add that where drainage through the wall is not
adequate to avoid an increase in head of groundwater in the retained soil
during storm conditions, this increase should be allowed for and, if in extreme
conditions, the head can reach the top of the wall, this should also be allowed
for.
Other conditions that may affect static groundwater levels considered in
designs include the effect of waves on retaining walls for marine structures
and river and sea cofferdam walls, and the dynamic forces of waves breaking
on beaches which can be transmitted to retaining walls for deep excavations
below beach level. Some guidance on these matters may be obtained from
the work of Sainflou25 and Minikin26 .
Where walls are situated in tidal conditions the pore pressure will vary with
the tide; the effect of rapid drawdown, of cycles of loading and any tidal lag
may require special consideration. Guidance is available in reference27 .
Where the vertical wall for soil retention penetrates cohesionless soil but
does not secure a cut-off in an impermeable stratum at the toe of the wall,

Fig. 5.18. Flow net: flow into


base of a twin-wall cofferdam in
sand17
Design of vertical soil support 169

the effect of water seepage from the retained soil on the active side of the wall
to the passive side should be taken into account. Over time a steady seepage
state will develop. In design, the effect on pressure in these conditions may be
conveniently assessed by use of flow nets, the construction of which was
explained in Chapter 2. A flow net illustrating these conditions is shown in
Fig. 5.18. The total head at any point is
d
H ¼ H1  ðH1  H2 Þ ð57Þ
nd
where H1 and H2 are the heads of groundwater on either side of the wall, d is
the number of flow net equipotential drops to the point considered, and nd is
the total number of drops. Also, at any point the total head is
u
H¼ þz ð58Þ
w
where u is the pore-water pressure, and z is the height of point considered
above datum. Since H can be calculated by flow net and w and z are
known, u can be calculated. The flow net is also used to check the factor of
safety against piping (see Chapter 2).
A simple short-cut for calculating pore-water pressure at any point can be
made by assuming that the head difference between each side of the wall,
active and passive, is dissipated uniformly along the flow path; referring
to Fig. 5.19, head difference ¼ ðh þ i þ jÞ and flow path length is
ð2d þ h  i  jÞ.
Therefore, the head at any point x below the water table on the active side is
x
Hx ¼ ðh þ i  jÞ ð59Þ
2d þ h  i  j
and since u ¼ Hw  z, pore-water pressure at the toe of the wall is
2ðd þ h  jÞðd  iÞ
utoe ¼ w : ð60Þ
2d þ h  i  j
An even simpler first approximation is that in which the water pressure
at the toe of the wall is taken as the average at that depth for active and
passive sides of the wall. All uses of steady-state flow are best checked by
flow net; the reduction of excess head linearly along the length of flow path
at the wall can lead to over-optimistic calculated pore pressures in the
narrow flow paths that occur between the retaining walls of a narrow coffer-
dam. In wider cofferdams (where width is greater than four times the
differential water head) and single retaining walls where the seepage pressure
is free to dissipate horizontally through the passive soils, the first approxima-
tion is satisfactory.
The reduction in pore pressure from one side of the wall to the other in the
steady-state condition can be assessed by such methods of calculation. More
importantly, however, is the effect of preferential drainage paths which can be
caused by sand or silt partings in laminated groundwater flow below a vertical
wall. Where the pore-water pressure does not reduce gradually because of a
horizontal drainage route below the toe of the wall, an excessive pore-water
pressure occurs on the passive side, resulting in high net pressures and
forces on the wall. Figure 5.20 illustrates this risk.
Where groundwater flow is impeded by the periphery wall to a very large
excavation the groundwater level near the wall may rise with time. Where
the risk of this increase in pressure exists piezometers should be installed
and monitored close to the wall.
170 Deep excavations

Fig. 5.19. Water pressures – simplifying assumptions for steady-state conditions: (a) uniform ground; (b) layered ground48

Fig. 5.20. Excessive pore-water


pressure on passive side due
to underflow: (a) soil
cross-section showing pervious
interlayers; (b) water pressure
diagram48

Design methodology Wall design must satisfy the ultimate limit state and the serviceability limit
state with the specified partial safety factors applied.
The partial factor on soil strength at each limit state for both total and
effective stress designs for each of the design categories, moderately conser-
vative, worst credible and most probable, is shown in Table 5.6. The design
Design of vertical soil support 171

Table 5.6. Overall stability, cantilever and single-prop walls. Partial factors for use in design calculations49

Design approach Ultimate limit state Serviceability limit state

Effective stress Total stress Effective stress Total stress

Fsc0 Fs0 Fssu Fsc0 Fs0 Fssu

A. Moderately Conservative 1.2 1. 2 1. 5 1. 0 1. 0 1. 0


B. Worst Credible 1. 0 1. 0 1. 0 Not Not Not
appropriate appropriate appropriate
C. Most Probable 1. 2 1. 2 1. 5 1. 0 1. 0 1. 0
 
tan 0
design 0d ¼ tan 1
Fs0
c0
design c0d ¼
Fsc0
s
design sud ¼ u
Fssu

parameters derived from the application of these safety factors are then used
to obtain earth pressure coefficients, from charts such as those by Caquot and
Kerisel15 .
Either the design approach using moderately conservative parameters (with
Fs ¼ 1:2 applied to effective stress values c0 and 0 or Fs ¼ 1:5 applied to
total stress su ) or the design approach based on worst credible parameters
(with Fs ¼ 1:0) is used. An early check of the two sets of calculated design
parameters shows which design approach gives the lowest design values.
These soil strengths are used with the worst combination of applied loading
in both the short term and the long term.
Groundwater conditions which include water pressure and seepage pressure
with the most unfavourable values that could occur in the extreme or acciden-
tal situation are applied at ultimate limit stage. The accidental situation
includes an allowance of 10% of the retained height or 0.5 m, whichever is
greater, for unplanned excavation. The most unfavourable values applying
in normal conditions are then used to check the serviceability state.
Choice is made then of the analytical method, either based on limiting
values of earth pressure or on soil–structure interaction analysis. For simple
structures, such as cantilever walls, the limit equilibrium can be applied
(with reasonable agreement of the designed wall with more complex soil–
structure analysis) although for propped and anchored walls the soil structure
analysis will usually show economies in terms of wall penetration, bending
moment and prop/anchor loads. These economies may be significant.
For serviceability checks soil structure interaction analysis is necessary to
check wall deformation, soil settlements and compliance with crack criteria
for R.C. walls.
As a general basis of permanent wall design the following steps are made to
establish values of bending moments and shear force, prop/anchor loads and
wall penetration.
(a) Establish ultimate limit states (overall stability and wall strength) and
serviceability limit states (limiting values of wall deflexion, settlements,
crack width criteria).
(b) Collect soil and groundwater data.
(c) Determine wall construction method and wall type.
(d) Determine soil profile and soil parameters.
172 Deep excavations

(e) Determine load case combinations and wall geometry for ultimate limit
state (ULS) calculation.
( f ) Determine the wall depth for vertical stability. (Check on vertical load
carrying capacity or hydraulic cut-off ). Check wall friction/adhesion.
(g) Make ULS calculation by limit equilibrium or soil–structure inter-
action. Check wall penetration for overall stability.
(h) Use the deeper of wall depths for vertical stability ( f ) or for lateral
stability (g) as design depth.
(i) If wall depth is determined by vertical stability ( f ) recalculate bending
moment (BM), and shear force (SF) in wall and prop loads by limit
equilibrium or soil-structure interaction method.
( j) If serviceability limit state (SLS) calculations are required, select
appropriate soil parameters, groundwater pressures, load case
combinations and design geometry.
(k) Carry out SLS calculations to determine wall deflection, BM, SF and
prop loads.
(l ) Compare calculated wall deflections with initial wall design require-
ments.
(m) Make assessment of settlement risk to adjacent structures and services.
(n) Check risk of progressive failure.
(o) Determine ULS wall moments and shear force for structural design as
the greater of:
(i) BM and SF from step (g), ULS
(ii) 1.4 times the BM and SF from step (k), SLS
(iii) BM and SF from step (n), progressive collapse check
( p) Calculate prop loads, see later this chapter.

Structural design of wall


Ultimate limit state values of bending moments and shear forces for design of
the wall section are taken as the greatest of those calculated from application
of factored soil strengths (the worst values for case A, moderately conserva-
tive or case B, worst credible) or 1.4 times the SLS calculations.

Crack control
The application of both BS and Eurocodes at serviceability state to limiting
crack widths in retaining walls due to early age shrinkage and flexure can
be most onerous. Although the Eurocodes contain a ‘deemed to satisfy’ pro-
cedure in the absence of detailed calculations this is based on providing
sufficient reinforcement to have a greater tensile capacity than the immature
concrete whilst limiting the steel stress to a value which limits the crack
width to the required width.
The draft I.Struct.E report on basement and cut-and-cover works51 recom-
mends that different criteria are set for flexural crack widths at various stages
and suggests the maximum flexural crack widths shown in Table 5.7.
Table 5.7. Maximum flexural
crack widths51 Design stage Suggested maximum
flexural crack width (mm)

During construction and recoverable 0.35


In service 0.30
Short-term event 0.35
Drawdown of groundwater to underside of base slab 0.40
Note. Cracking caused by flexure of the entire structure, such as a tunnel built over a patch of
soft (or hard) ground or subjected to localized surcharge, passes right through the section and
should be limited in width to 0.2 mm at all times.
Design of vertical soil support 173

Fig. 5.21. Pile cross-section


showing crack width calculation
principles to EC2 if tension area
in cover thickness is ignored49

Fig. 5.22. Pile cross-section


showing calculation principles
for crack width in accordance
with BS 811049

The quantity of reinforcement required for crack control may be consider-


able and strict application of code requirements may be needed to keep steel
density to practical levels.
Referring to steel required to resist a defined crack width in bending as
calculated according to EC2 (applied to a circular section cast against the
ground) the critical parameter is the area of concrete in tension A (EC2
Part 1, 1992, Clause 4.4.2.2. (3)). Figure 5.21 shows the net area in tension
which may be used for calculation purposes if the thickness of concrete
cover is assumed to be the minimum required for durability.
Referring to crack control requirements in BS 8110: Part I, 1997, the specific
items for a typical circular pile section are:
Nominal cover to earth face: 75 mm (Part I Clause 3.3.1.4).
Durability requirements: 35 mm for C35 concrete (Part 1 Table 3.3)
(concrete in moderate exposure Table 3.2, non-aggressive soil).
Crack width (Part 2, section 3.8) should be assessed in accordance with acr , the
distance from the point considered to the surface of the nearest longitudinal
bar (Part 1, Clause 3.8.2). Often assessed as the nominal cover required for
durability and not the 75 mm cover provided for constructability. Refer to
the cross-section shown in Fig. 5.22.

Cantilever walls and The design of cantilever and single-prop walls requires that the serviceability
single-prop walls limit state is considered in terms of wall deformation and soil deformation at
the rear of the wall. The ultimate limit state is as follows.
174 Deep excavations

(a) Overall instability – the provision of sufficient embedment depth to


prevent overturning of the wall. For a single-propped wall the minimum
penetration is just sufficient to avoid forward movement of the wall with-
out preventing rotation of the wall at its toe. Such a condition is known
as the free earth support condition. As wall penetration is increased
beyond this minimum, fixity of the wall is progressively increased until
full fixity is obtained and the toe is unable to rotate when earth and
water loading is applied to the wall. This state is known as the fixed
earth condition. For the cantilever wall without prop support, full
fixity must be obtained to prevent rotation and collapse of the wall.
(b) Foundation failure – in cohesive soils, the wall penetration depth must
be sufficient to prevent basal failure after excavation to formation
level in front of the wall.
(c) Hydraulic failure – in cohesionless soils, the penetration of the wall must
be sufficient to avoid piping in front of the wall after excavation to
formation level.

Overall stability
The overall stability of both cantilever and single-prop walls has previously been
assessed using limit equilibrium methods in which a factor of safety is applied
to an analysis of failure conditions to ensure their avoidance. Such factors of
safety were applied in a number of ways and have attracted comment from
Potts and Burland28 , and Symons29 . The various factor methods are as follows.
(a) Factor on embedment. A factor of safety is applied to embedment depth.
The method is described in references 30 and 31.
(b) Factor on moments of gross pressure. This method applies as a factor of
safety to moments of gross pressure on the passive side only. Water
pressure is not factored. The method is described in references 16 and 22.
(c) Factors on moments of net pressure. The net pressure diagram is calcu-
lated and the factor of safety is defined as the ratio of moments of the
net passive and active forces about the toe of the wall.
(d ) Factors on net passive resistance. Potts and Burland28 developed this
method which defines the factor of safety as the ratio of the moment
of the net available passive resistance to the moment activated by the
retained material including water and surcharge. When using effective
stress methods with c0 ¼ 0 and where there is no surcharge on the
passive side, the active pressure diagram is modified so that no active
pressure increase occurs below formation level and any such increase
as calculated is deducted from the passive pressure diagram. Where
total stress or effective stress methods are used, reference should be
made to Potts and Burland28 .
(e) Factor on shear strength on both active and passive sides. Soil shear
strengths are reduced by dividing c0 and tan 0 by the factor of safety,
and the active and passive pressure diagrams are calculated using
these reduced values. The reduced values approximate to mobilized
values. Bending moments and prop loads derived from the calculation
can be used for wall design if they are treated as ultimate limit state
values. This method is recommended in BS 800219 .
( f ) Factor on shear strength of passive side only. The passive resistance is
factored but no factor is applied to the active side.
Comparative studies show that there is no relationship between factors of
safety applied to the methods listed. There is little doubt that the preferred
method is really defined by personal use and experience, but some useful
observations were made in reference 32 as follows.
Design of vertical soil support 175

(a) The factor on embedment is empirical and should be checked by


applying a second method.
(b) The method of factoring moments on gross pressure may give excessive
penetration at low angles of shearing resistance 0 less than 208 so use
varying factors for different ranges of 0 .
(c) The factoring of net passive pressure moments tends to give high
penetration values.
(d) The factors on modified net passive resistance as recommended by
Burland and Potts28 appear to give consistent results in a reasonable
range of soils and wall dimensions.
Factors of safety recommended by Padfield and Mair48 for use in stiff clays
with methods for factoring embedment, moments of gross pressure, net
passive resistance and shear strength on both active and passive sides are
reproduced in Table 5.8. Two approaches are used: approach A is based on
moderately conservative parameters, and approach B uses worst credible
soil parameters, geometry and loading in the design. The values in Table
5.8 had been originally intended for walls embedded in stiff clays but were
also used for walls in granular or mixed soils.

Table 5.8 Factors of safety for methods of analysing embedment48

Method Design approach A: recommended Design approach B: recommended Comments


range for moderately conservative minimum values for worst credible
parameters (c0 , 0 , or cu ) parameters (c0 ¼ 0, 0 )

Temporary Permanent Temporary Permanent


works works works works

Factor on Effective stress 1.1 to 1.2 1.2 to 1.6 Not 1.2 This method is empirical.
embedment, Fd (usually 1.2) (usually 1.5) recommended It should always be
checked against one of
the other methods
Total stressa 2.0 – –
Strength factor Effective stress 1.1 to 1.2 1.2 to 1.5 1.0 1.2 The mobilized angle of
method, Fs (usually 1.2 (usually 1.5 wall friction m and wall
except for except for adhesion cwm should also
0 > 308 when 0 > 308 when be reduced
lower value lower value
may be used) may be used)
Total stressa 1.5 – – –
Factor on moments: Effective stress 1.2 to 1.5 1.5 to 2.0 1.0 1.2 to 1.5 These recommended Fp
CP2 method, Fp 0  308 1.5 2.0 1.0 1.5 values vary with 0 to be
0 ¼ 20 to 308 1.2 to 1.5 1.5 to 2.0 1.0 1.2 to 1.5 generally consistent with
  208 1.2 1.5 1.0 1.2 usual values of Fs and Fr
Total stressa 2.0 – – –
Factor on moments: Effective stress 1.3 to 1.5 1.5 to 2.0 1.0 1.5 Not yet tested for
Burland–Potts (usually 1.5) (usually 2.0) cantilevers. A relatively
method, Fr new method with which
little design experience
has been obtained
Total stressa 2.0 – – –
a
Speculative, treat with caution.
Note 1. In any situation where significant uncertainty exists, whether design approach A or B is adopted, a sensitivity study is always recommended, so that an
appreciation of the importance of various parameters can be gained.
Note 2. Only a few of the factors of safety recommended are based on extensive practice experience, and even this experience is recent. At present, there is no
well-documented evidence of the long-term performance of walls constructed in stiff clays, particularly in relation to serviceability and movements. However,
the factors recommended are based on the present framework of current knowledge and good practice.
Note 3. Of the four factors of safety recommended, only Fp depends on the value of 0 .
176 Deep excavations

Fig. 5.23. Analysis of cantilevered wall17

Calculation can be made by soil–structure interaction analysis or using limit


equilibrium methods by program or by hand.

Fixed earth support: cantilevered wall


Certain simplifying assumptions are made due to the relative complexity of
the limit equilibrium calculation. Figure 5.23 shows an idealized pressure
diagram and typical shear, bending moment and wall deflection diagrams.
The simplified procedure is as follows:
. the pressures at the toe of the wall are replaced by a resultant force F3
acting at C
. forces F1 and F2 act through the centres of pressure of their respective
areas
. depth BC is found by assuming a level for C and calculating the
moments for the forces F1 and F2 about level C; repeat until moments
balance
. increase the depth BC by 20% to give the design penetration BD
. calculate the maximum bending moment (BM) at the point of zero shear.
This method can be used for layered soils; net water pressures should be
included in the active pressure diagram where groundwater occurs. The
factor of safety can be applied by using limit pressures and increasing the
depth of embedment. The bending moment distribution at limiting equili-
brium is assumed to correspond to working conditions. A safety factor of
value 2.0 is then applied to these values for design values of BM and shear
force (SF). A partial factor of safety is applied to characteristic soil strengths
for design of the wall cross-section. Alternatively, for design according
to BS 800219 a mobilization factor is applied to peak shear strengths to
obtain design values of BM and SF. No further safety factor is applied
other than the partial factor to material characteristic strength to obtain the
design strength.
Bending moment and shear force in the cantilever wall may also be analysed
using Winkler spring, finite element or boundary element methods. These
methods produce a more realistic soil interaction solution and provide
values of lateral wall deformation.
The importance of wall flexibility related to wall deformation and depth of
embedment was originally shown by Rowe20 . The earth distribution was
shown by model tests to vary according to the flexibility of the wall. (The
wall flexibility was defined by the expression
d 4y
wall flexibility ¼ EI
dx4
Design of vertical soil support 177

where E is Young’s modulus of the wall material, I is the section modulus of


the wall transverse cross-section per unit length, and y and x are horizontal
and vertical coordinates of the vertical cross-section of the wall). The reduc-
tion in depth to full fixity caused by increasing wall flexibility produced
reduced bending moments in the wall section and, according to Rowe,
produced savings of up to 20% by weight of the wall compared with design
methods based simply on limit pressures. The method was not widely adopted
for cantilever pile design but served as an early indicator of the importance of
wall and soil interaction.

Free earth support: single-propped walls


The earth pressure, shear and bending moment distribution and wall deflec-
tion diagrams which apply to a typical single-propped (or anchored) wall
are shown in Fig. 5.24. The method of calculation uses limit pressures or
pressures calculated on critical state values of soil parameters or on peak
shear strengths reduced by a mobilization factor as defined in BS 8002, and
is as follows.
(a) Draw earth pressure diagrams using limiting or mobilized pressures for
soil pressure in cohesionless or cohesive homogeneous or layered soils.
Include net water pressure, modified to include the effect of steady flow
beneath the wall and equalization of pore pressures either side of the
wall at the toe. Use tension cracks filled with water where applicable
in clays and apply minimum fluid pressures (5 kN/m3 ) when necessary.
(b) Calculate F1 and F2 to act through the centre of their respective pressure
areas.
(c) Calculate the penetration BD by assuming a trial depth of penetration
D and taking moments of forces F1 and F2 about the tie rod level.
Repeat with trial values of penetration D until these moments balance.

Fig. 5.24. Single-propped or


anchored wall with: (a) free
earth support; (b) fixed earth
support17
178 Deep excavations

Table 5.9 Zero moment depths


to ensure fixity Angle of shearing resistance  Depth to zero moment as a proportion of depth
from ground surface to formation level H

208 0.25H
258 0.15H
308 0.08H
358 0.035H
408 0.007H

(d) The prop load T can then be found by balancing forces T ¼ F1  F2 .


(e) Calculate the bending moment in the wall section at the level of zero
shear.

Fixed earth support: single-propped walls


The analysis by hand calculation of the fixed earth, single-propped (or
anchored) wall is made less complicated by use of work due to Blum33 in
the 1930s. He approximated the bending moment in a single-anchored wall
with fixed earth support by reducing it to zero at the point of zero net earth
pressure below formation level. Blum gave approximate depths to this point
for relatively uniform soils where the wall is sufficiently deep to ensure
fixity; these depths are given in Table 5.9. This support may be solved
graphically16 or by calculation (Fig. 5.24) as follows.
(a) Draw earth pressure diagrams using limiting or mobilized pressures for
soil pressure in cohesionless, cohesive or layered soils. Include net water
pressure modified to include the effect of steady flow beneath the wall
and equalization of pore pressures on either side of the wall at the
toe. Use tension cracks filled with water where applicable in clays.
Apply minimum fluid pressures (5 kN/m3 ) where necessary.
(b) Draw the net pressure diagram from ground level to the first point of
zero pressure below formation level.
(c) Calculate active forces F1 and F2 from the gross pressure diagram.
(d) Take moments of forces F1 and F2 and tie rod force T about the first
point of zero pressure from the net pressure diagram and equate to zero.
(e) Calculate value of force T from this expression.
( f ) Assume application of passive force F3 at point C and recalculate F1
and F2 to this depth. Take moments of T, F1 and F2 about this level
and repeat with change in depth to point C until the moments balance.
As with the cantilever wall calculation, the calculated depth to point C
is increased by 20% to give design penetration depth.
(g) Draw the shear force diagram and calculate the maximum bending
moment in the wall at the point of zero shear force.
This design procedure, as for free earth support design, is usually made by
computer program and where soil deformation prediction is required can be
the subject of soil/structure interaction programs or finite element or finite
difference computations.

Anchored bulkheads: design development


The description of a propped wall includes within this generic phrase various
constructions such as sheet piling, diaphragm walls and reinforced concrete
pile walls braced by struts or raking shores, or sheeters and walls of similar
construction with tie rods or stressed ground anchors. One of these types of
constructions, sheet piling to form a wall or bulkhead tied back by steel
rods to an anchor block, known as a deadman, or to anchor piles, was the sub-
Design of vertical soil support 179

ject of a historical keynote paper by Terzaghi34 . This method of construction,


which is widespread on waterfronts where it is used to achieve a quay line with
reclamation behind it at minimum cost, does not strictly lie within the subject
of deep excavations, but no technical discussion on braced sheeting would be
complete without reference to the paper and the controversial discussion that
followed it. The principal purpose of the paper was to identify and rectify
errors in accepted bulkhead design at that time on the basis of tests and
observations by Terzaghi and model tests by Rowe20;35;36 . The paper reached
the following principal conclusions.
(a) The identification of the type of soils and fills and their in situ properties
of uniformity, relative density and strength are vital matters which are
frequently overlooked in anchored bulkhead design.
(b) The distribution of earth pressure on the bulkhead is unlikely to con-
form to the Coulomb distribution, because of the extent of deformation
of the soil structure. This deformation depends on soil and wall stiffness.
(c) If maximum bending moments are calculated on walls in sand assumed
to extend to sufficient depth to achieve full fixity irrespective of wall
flexibility and sand relative density, errors are likely and these are on
the unsafe side.
(d) For sheet piles driven into clay, the assumption of full fixity at depth
will probably not apply as time elapses and no reduction in the
calculated maximum moment in the wall should be allowed due to
wall section flexibility to compensate for this loss.
(e) Anchor tension depends on several factors other than the properties of
the backfill material and the flexibility of the wall or sheeting. There-
fore, the anchor pull should be computed on the assumption of free
earth support. Anchor pull may be greater than that calculated using
Coulomb’s theory, and may increase due to repetition of loading
and unloading from heavy surcharge. An unequal yield of adjacent
anchorage produces variations in tie rod pull. Given these risks, more
conservative stresses should be used in anchor design than are applied
in sheet pile bulkhead design.
These conclusions broadly still apply, although alternative methods of
analysis have been developed in which soil, wall and anchor stiffness can be
modelled and deformation and induced stress in all three can be calculated.
Terzaghi concluded: ‘Because of the great variety of subsoil conditions
which may be encountered, the subject (anchored bulkheads) does, and
always will, leave a wide margin for judgment – and also for misjudgment.’

Anchorage location
Anchorages, deadmen or injected tendons must be located behind potential
failure surfaces at the rear of the wall. Figure 5.25 shows the recommended
geometry for analysis of deadmen locations (from BS 634927 ).

Foundation failure
The risk of base failure to an excavation by upward heave applies particularly
in very soft and soft clays and silty clays, typically, quick estuarine deposits.
The failure is analogous to a bearing capacity failure of foundation, only in
reverse; the failure is a shear failure in the soil below formation level, but
caused by relief of load (the relief of overburden) and not by the application
of load as occurs in a conventional foundation bearing failure.
The methods of Terzaghi37 and Bjerrum and Eide38 can be applied to cal-
culate the factor of safety against base failure; these are shown in Fig. 5.26.
Terzaghi’s method is primarily applicable to shallow or wide excavations,
180 Deep excavations

Fig. 5.25. Location of deadman


anchorage in granular
retained fill27

Fig. 5.26. Calculation of factors


of safety against basal heave in
cohesive soils: (a) deep
excavations with H=B > 1;
(b) for shallow or wide
excavations with H=B < 1 32

while the method of Bjerrum and Eide is suitable for deep and narrow excava-
tions with no nearby underlying stiff clay to inhibit failure. Both methods
neglect the effect of wall penetration below formation level and therefore
results may prove to be conservative, especially where stiffer clays exist with
depth. A third method22 for predicting the basal safety factor where stiff
clays exist at depth, is shown in Fig. 5.27.
The factor of safety against basal failure is generally required to be not less
than 1.5. If uncorrected values of in situ vane tests are used, the actual factor
of safety may be close to 1.0 according to Aas39 . (A vane correction described
by Bjerrum40 is necessary to obtain more reliable values of safety factor.)
With a factor of safety, based on corrected vane results, which is less than
1.5, substantial soil deformation is likely. If such soil movement is not accep-
table, a factor of safety not less than 2.0 is recommended. Increase in move-
ment occurs as the basal factor of safety decreases, and increases rapidly as
a factor of safety of 1.0 is approached. Although basal heave is rare within
Design of vertical soil support 181

Fig. 5.27. Calculation of factors of safety against basal heave in: (a) cohesionless soil; (b) cuts in clay of considerable
depth; (c) cuts in clay limited by hard stratum22

excavations in cohesionless soils, a basal heave analysis is included in


Fig. 5.27(a) for completeness, together with basal heave analysis in clay as
described in NAVFAC22 in Fig. 5.27(b) and (c). Figure 5.28 shows the
values of bearing capacity factors for use in these analyses. Field and finite
element analysis predictions of the correlation between movement and basal
failure factor of safety are shown in Fig. 5.29.
Cantilever and single-prop walls, particularly on sloping sites in soft clays
and loose granular soils, should always be checked against risk of deep-
seated circular slip failure.

Hydraulic failure
The risk of piping failure to the base of an excavation in cohesionless soils
was described in Chapter 2. Design charts for penetration of cut-off walls
to prevent hydraulic failure in sand and stratified soil are reproduced in
Figs 5.30 and 5.31.

Wall flexibility
Rowe’s work20;35;36 in the 1950s and 1960s was initially instrumental in
showing the importance of wall stiffness in design. Following a series of
182 Deep excavations

Fig. 5.28. Plot of bearing


capacity factors against angle
of shearing resistance22

Fig. 5.29. Analytical


relationship between maximum
lateral wall movement and
factor of safety against basal
heave from field data, free end
and fixed end walls, various
sites (note is for finite element
analysis data)

model tests on sands of varying relative density in Manchester, UK, Rowe was
able to show that interaction between soil and wall was different for steel sheet
piles and reinforced concrete sheet piles because of the greater flexibility of the
steel sheet pile. This greater flexibility causes a redistribution of earth pressure
which differs considerably from the Coulomb distribution, as shown in
Fig. 5.32. The flexure of the wall causes reduction in pressure at mid-height
and causes the resultant passive force to rise with an increase in fixity for
the flexible pile. These changes reduce the design bending moment for a
flexible pile, although too often such reductions are not applied in practice
to ensure the pile does not crumple during driving.
Design of vertical soil support 183

Fig. 5.30. Penetration of cut-off


wall to prevent hydraulic failure
in homogeneous sand:
(a) in sands of infinite depth;
(b) in dense sand of limited
depth22

Materials and stresses


For cantilever and single-propped walls the component parts are designed
either from limiting equilibrium hand calculations or computer program
outputs using either limiting equilibrium or soil–structure interaction. The
input soil parameters are those based on moderately conservative parameters
with safety factors at ULS as shown in Table 5.6 (BS 8002 mobilization
factors are similar) or at serviceability limit state with a partial safety factor
of 1.4 (a check that moderately conservative parameters are more severe,
that is lower, than worst credible parameters at ULS being made at the
beginning of the design). Crack widths are calculated for serviceability limit
state in reinforced concrete walls. Characteristic strengths of steel used for
cantilever and single-prop sheet pile walls are given in Table 5.10.

Multi-prop walls The above description of design methods for cantilever and single-propped (or
anchored) walls referred to computations using limit pressures and the appli-
cation of factors of safety. The methods due to BS 800219 have introduced
design using earth pressures at the serviceability limit state. Using these
methods the bending moment in the walling can be estimated relatively
quickly by hand calculation (adopting Blum’s methods33 for cantilever and
184 Deep excavations

Fig. 5.31. Penetration of cut-off


wall to prevent hydraulic failure
in stratified soil: (a) coarse sand
underlying fine sand; (b) fine
sand underlying coarse sand;
(c) very fine layer in
homogeneous sand22

fixed earth support walls) or even more conveniently using finite element,
finite difference or Winkler spring analytical methods. Soil deformation
behind the wall may be predicted, if needed by the finite element or finite dif-
ference programs. Design requirements and analysis methods for multi-prop
walls are a different matter, however. The method of construction for these
walls is usually sequential, installing the sheeting or walling and excavation
in stages followed by installation of the prop or anchor at each installation
stage. The sheeting or walls will, in all likelihood, penetrate the ground
below the final excavation level. The extent of wall deformation in this
sequence of operations is restricted, although the passive resistance of soil
below excavation level at each stage is mobilized to support the wall prior
to installation of the bracing or the anchor at that level. Despite the frequent
Design of vertical soil support 185

Fig. 5.32. Deflection of a sheet


pile and redistribution of active
earth pressure50

Table 5.10 Characteristic


strengths for steel sheet piling Designation EN 10027 Minimum Minimum Minimum elongation
yield strength tensile strength on a gauge length
(N/mm2 ) (N/mm2 ) of L0 ¼ 5:65S0 (A%)

EN 10248 S270 GP 270 410 24


EN 10248 S355 GP 355 480 22

support for the wall, therefore, horizontal deformation of the wall occurs at
each passive soil zone prior to installation of the prop or anchor at that
level. The wall distorts inwards to mobilize this passive resistance, the wall
movement occurring below each stage of excavation. Pore pressure dissipa-
tion may occur in cohesive soils during the period needed for strut or
anchor installation at successive levels.
The extent of wall movement also depends on the stiffness of the prop or
anchor once installed at each level. Where the soil is relatively stiff, say
dense sands or gravels, the extent of forward movement of the sheeting at
each excavation stage to mobilize soil passive pressure will be relatively
small, and active earth pressures on the wall will considerably exceed
Coulomb active values above dredge level; redistribution of earth pressure
will occur between the lowest strut and formation level. In wide excavations
in soils such as soft clays or loose sands where stiffness is low, the successive
deformations below excavation level at each propping formation level are
186 Deep excavations

considerable. Load is redistributed between the struts, and the sum of the
maximum strut loads considerably exceeds the Coulomb values.
Where pre-stressed ground anchors are used at each excavation stage the
earth pressure on the wall is determined by the pre-stress levels and sub-
sequent relaxation, and by relative wall and soil stiffnesses. Design methods
that have been used for many years have been based on calculations of
anchored walls using a Coulomb distribution assuming no pre-stress applied.
This non-pre-stress value of anchorage at each excavation stage is sub-
sequently used as the actual value of pre-stress applied to the tendon. This
empirical method successfully restricts soil movement but in turn inhibits
the Coulomb active earth pressure distribution, on which the calculation is
based, from developing, the actual pressures on the retained side of the wall
being higher (and nearer K0 or Kp values) than those calculated. More
recent methods using Winkler spring and finite element programs allow an
assumed anchor pre-stress load to be introduced to the analysis, from
which the actual earth pressures are calculated on the basis of the soil
movement permitted by wall and soil stiffness and the extent of the anchor
prestress.
The ultimate limit states for multi-prop walls are similar to those for
cantilever and single-prop walls:
(a) overall stability – risk of strut failure, bending stress failure in sheeting
or passive failure of soil below stage excavation level or final formation
level
(b) foundation heave – in soft clays, risk of failure by unloading; bearing
capacity failure
(c) hydraulic failure – piping in cohesionless soils with high external
groundwater table.
The serviceability limit states are as follows.
(a) Deformation of sheeting – the acceptable limits of sheeting deformation
will depend on the purpose of the excavation and whether the works are
temporary or permanent, or a combination of both. Where walls or
sheeting are temporary the deformation must not exceed that which
would occupy space required for the permanent works nor cause
difficulties with sheeting removal if this is intended. For permanent
works, deformation of the wall or sheeting must neither impair the
durability of the substructure nor cause visual offence.
(b) Soil movement behind wall or sheeting; vertical settlements – the extent of
settlement behind the support for the excavation must not exceed the
permitted settlement of existing structures, highways or services,
unless the consequences of this can be estimated accurately and on
re-assessment are acceptable.
(c) Cracking in reinforced concrete walls – at the serviceability state, crack-
ing will occur on the tension face of reinforced concrete walls due to
application of load, in particular earth and water pressures and sur-
charge loading, and also on each face of the wall due to early thermal
cracking of the concrete. For building substructures the provisions
of BS 81102 will apply to crack control in walls or, where more rigorous
waterproofing is needed, BS 80073 may be specified. For highway struc-
tures in the UK, design flexural and tension cracks complying with the
BS 54004 are specified. Design crack widths of 0.25 mm, complying with
‘severe’ conditions are usual although the pressure of saline or sea water
may reduce this value to 0.15 mm. Additional longitudinal steel may
therefore prove necessary in diaphragm walls to control crack widths
Design of vertical soil support 187

caused by loading, although the application of rules to minimize


vertical crack widths due to thermal shrinkage of concrete in panels
of limited individual length may prove over-strenuous. Such rules for
reinforced concrete works are referred to in the UK Department of
Transport’s Standard BD 28/8741 . Eurocodes EC2 and EC7 refer to
similar crack control requirements in reinforced concrete walls.
The available methods of design for multi-propped walls are:
(a) empirical methods; based originally on strut load envelopes proposed
by Peck for three categories of soil: sands, soft to medium clays and
stiff clays. Twine and Roscoe42 more recently proposed new strut
load envelopes for soft to firm clays, stiff to very stiff clays and dry
clay or submerged granular soils. The use of empirical methods is
recommended as a check on computed strut loads
(b) limit equilibrium programs are a simple solution without addressing all
the matters of influence. Programs based on Winkler Spring theory are
perhaps nowadays the most widely used methods
(c) full soil–structure interaction analysis by finite element, boundary
element or finite difference methods: used where prediction of soil
deformation and soil settlement requires calculated estimates
(d ) pseudo finite element programs.

Empirical method based on strut load envelopes


The original empirical method, due to Terzaghi and Peck18 , was applied to
both temporary works (including piled and diaphragm walls permanently
anchored or braced by floor construction, as in top-downwards construction)
and permanent works. The strut load envelopes due to Terzaghi and Peck are
shown in Fig. 5.33. Note that these diagrams are not intended to represent
actual earth pressure or its distribution with depth but load envelopes from
which strut loads can be evaluated. Clay is assumed to be undrained and
only total stresses are considered. Sands are assumed to be drained (through
the sheeting) with zero pore pressure. Where drainage is precluded behind a
non-permeable wall, hydrostatically distributed water pressure is added to
strut loads. Sheeting or walling was then designed using the Coulomb earth
pressure distribution with hydrostatic water pressure added except where
drainage occurred through sheeting to relieve water pressure.
In 1969, reviewing his empirical method, Peck pointed out that his recom-
mended method for strut design was less satisfactory in soft to medium clays

Fig. 5.33. Apparent pressure


diagrams for computing strut
loads in braced cuts18
188 Deep excavations

than for excavation in sands. Peck did point out, however, that the behaviour
of clays and the bracing system depends on the stability number
D
Nc ¼ ð62Þ
cu
where cu represents the clay beside and beneath the cut to the depth that
would be involved if a general failure were to occur due to the excavation.
Peck added that when the depth of excavation corresponds to values of Nc
greater than 6 or 7, extensive plastic zones have developed, at least to the
bottom of the cut, and the assumption of a state of plastic equilibrium is
valid. The movements are essentially plastic and the settlements may be large.
The empirical strut load envelopes reviewed by Peck (Fig. 5.33) do not
include the effects of the toe of the sheeting or walling extending below the
final formation level, and yet from the point of view of reduction of strut
loads in the lower struts (or anchors) and the improvement to ingress of
groundwater at formation level, an extension of sheeting vertically below
formation level was desirable and often achievable. There were two design
methods which allowed this penetration of walling and sheeting to be taken
into account in strut load calculation.
The first method assumed that the passive reaction below formation level
resisted active pressure below that level together with a portion of the load
from the strut load envelope between the lowest strut level and formation
level. This method applied to strutted excavations in uniform soil conditions
of reasonable strength such as medium-dense to dense granular soils and stiff
clays. In less competent soils (loose sands and gravels and soft clays), passive
resistance against the sheeting below formation level might be less effective,
and in such cases Goldberg et al.23 advised that the sheeting should penetrate
to such depth to avoid piping failure (as discussed in Chapter 2) and the
sheeting should be designed as a cantilever from the lowest strut level.
The second empirical method, the source of which is not known, has been
used since the mid 1950s by the Author and has been adequately confirmed.
The procedure is shown in Fig. 5.34. An additional ‘strut’ is assumed to act
on the strut load envelope to represent the passive resistance acting on the
sheeting below formation level. The level of this ‘strut’ is determined from
the net pressure diagram, as shown in Fig. 5.34. The ‘strut’ load calculated
from the strut load envelope is then compared with the available passive resis-
tance using the limiting pressure from the net pressure diagram for the selected
penetration depth. The bending moment in the sheeting is calculated from the
net pressure diagram.
The strut loads calculated for each successive construction stage are
summarized and the highest value at each strut level is used for strut and
waling design purposes. Similarly, maximum moment and shear values are
calculated for the sheeting or walling for each construction stage from the
net pressure diagram and the critical values used for sheeting design.
More recently, empirical data and recommended methods of strut load
estimation have been published by Twine and Roscoe42 . Eighty one case
histories, all from the UK, were reviewed. The walls were supported by
both timber and steel temporary props, not anchors, and the study did not
differentiate between single- and multi-propped walls. Excavations in rock
were not considered. The cases were classified according to the ground
retained as follows:
. Class A: normally- and slightly over-consolidated clay soils (soft and
firm clays)
. Class B: heavily over-consolidated clay soils (stiff and very stiff clays)
Design of vertical soil support 189

Fig. 5.34. Construction of


trapezoidal strut load envelope
for braced excavation to take
into account passive resistance
below formation level

. Class C: granular, cohesionless soils


. Class D: walls retaining both cohesive and cohesionless soils.
Case histories for soft and firm clays (class A) and for stiff clays (class B) were
further sub-divided into flexible walls (timber, sheet pile and soldier pile walls)
and stiff walls (contiguous, secant pile and diaphragm walls). The character-
istic distributed prop load diagrams proposed by Twine and Roscoe are
shown in Fig. 5.35.
Twine and Roscoe referred to the range of field measurements compared to
values calculated by deformation methods. Field measurements were made on
46 props at four sites. The comparison is shown in Table 5.11. The range of
calculated to required values was generally between 1.4 and 10 times. This
very wide scatter has been referred to by others for measurements in two
deep excavations in London (Batten and Powrie43 and Richards et al.44 ).
Both these site measurements of strut loads showed considerable variation
with temperature.
Twine and Roscoe concluded that, providing the load–strain behaviour of a
prop is ductile, any temperature increase which causes the ultimate capacity of
the prop is unlikely to cause sudden prop failure because of plastic deforma-
tion of the prop and the transfer of load to nearby props. Because of this, they
recommended that the design of steel props should not include temperature
effects provided that certain SLS and ULS criteria are satisfied. This
serviceability criterion ensures that the temperature effect does not cause
190 Deep excavations

Fig. 5.35. Characteristic


distributed prop load diagrams
proposed by Twine and
Roscoe42

yielding of the prop steel and thus avoids permanent deformation and the
ULS criterion requires the ultimate capacity of the prop under the tempera-
ture effect to be greater than the total ULS design load.
Potts45 reviewed analysis methods and concluded that all simple methods of
analysis were flawed and empirical methods were only of some value if they
were verified by observation. These conclusions, in 1992, remain currently
valid although the increased use of programs to calculate wall moments,
deflections and strut loads has been rapid and empirical methods appear
Design of vertical soil support 191

Table 5.11 Comparison of


measured and calculated prop Project No. of Range of Range of ratio:
loads42 monitored measured prop Measured load
props loads (tonnes) Design load
55 Moorgate 6 62.5 to 109 0.16 to 0.28
Almack House, London 15 0 to 118 0.0 to 0.36
A406 Chingford, London
(i) Piled wall 10 45 to 160 0.28 to 0.82
(ii) Wedge excavation 4 45 to 70 0.19 to 0.39
(iii) Diaphragm wall 6 140 to 220 0.7 to 1.37
A12/M11 Hackney, London 5 623 to 1395 0.51 to 0.83

little used. This change of emphasis appears unfortunate in the Author’s view;
the equivalent strut load diagnosis of Peck and Twine and Roscoe should still
be used as a critical comparison compared with calculated results.
A summary of the advantages and drawbacks of the four types of program
is given in Table 5.12, they are based on analytical methods as follows.
(a) Limit equilibrium (e.g. STAWAL and REWARD). A stability analysis
in which calculation is made of the statical equilibrium of arbitrary slip
surfaces by resolving forces or moments and results in the calculated
soil strength mobilized to obtain equilibrium. The internal stress condi-
tions within the soil wedge are not considered.
(b) Subgrade reaction/beam on springs (e.g. WALLAP, LAWALL). Soil
behaviour is modelled by a set of unconnected vertical and horizontal
springs. Any structural support is modelled by simple springs. In
LAWALL the soil is modelled as a simple boundary element using
Mindlin’s equations for horizontal displacement due to a horizontal
force applied within a semi-infinite elastic continuum. Berms are
difficult to model; anchors and raking props are difficult to model
accurately.
(c) Pseudo finite elements (e.g. FREW). Soil is modelled as an elastic
continuum with soil stiffness matrices calculated using a finite element
program. Soil model restricted to linear increasing stiffness with depth.
Only wall deformation is calculated.
(d) Finite element and finite differences (e.g. PLAXIS, CRISP, FLAC). Full
soil–structure interaction with modelled construction sequence. Some
programs are now able to model 3D excavations. Wide choice of soil
models available; Mohr–Coulomb, and more complicated failure
criteria to represent variation of stiffness with strain. Use of low
strain values of soil stiffness are necessary.
Comparative analysis using both empirical methods and computer analyses
shows a comparatively wide scatter of results for strut loads, wall penetration
and wall bending moment. The lack of validation from field observations and
the extensive use of programs nowadays gives cause for concern.

Factors of safety: multi-propped wall design


The attention given to comparative methods of applying factors of safety
to the overall stability of cantilever and single-prop (or anchored) walls has
not extended to multi-prop walls. The design should always consider
hydraulic failure by piping from the base of the excavation and the risk of
failure of the base of the excavation in soft clays. Overall, stability considera-
tion in a multi-propped wall will rely on three matters: the risk of circular slip
instability below the whole excavation in clay conditions on sloping sites;
192 Deep excavations

Table 5.12. Summary of advantages and disadvantages of analytical methods49

Method Basis Advantages Disadvantages

Limit equilibrium Equilibrium conditions used Needs only soil strength Does not model soil–structure
(Typical programs: for the failing soil mass. parameters. Simple and interaction, wall flexibility or
STAWAL, Full mobilization of soil straightforward. Estimate of construction sequence. Multi-
REWARD) strength, factors of safety deformation possible by propped walls, non-uniform
applied to lateral stresses at relating mobilized strength, surcharges and berms require
failure soil strength and wall considerable idealization.
rotation Only give an order of wall
deformation and surface
settlement. Better for
cantilever than propped walls
Subgrade reaction/ Soil modelled by a set of Soil–structure interaction Soil behaviour idealization is
beam or springs unconnected vertical and taken into account. Wall likely to be less than perfect.
(Typical programs: horizontal springs. movements are calculated. Results are sensitive to pre-
LAWALL, Alternatively some Simple to apply. excavation stress state and
WALLAP, programs model soil as a Construction sequence spring stiffness. Any failure to
M. SOILS) simple boundary element modelled provide overall stability is
using Mindlin’s equations usually only shown by a
for horizontal displacement failure to converge
Pseudo finite element Soil modelled as an elastic Soil–structure taken into Limited to soil with linearly
(Typical program: continuum with soil stiffness account. Wall movements increasing stiffness at depth.
FREW) matrices calculated using a are calculated. Relatively Results sensitive to pre-
finite element program straightforward. excavation stress state.
Construction sequence Experience shows an
modelled overtendency to predict wall
deformation
Finite element and Full soil–structure Full complex soil models Time-consuming to set up.
finite difference. interaction with modelled can represent variation of High quality data required to
(Typical programs: sequence of construction stiffness with strain and realise full potential. Use of
OASIS SAFE (FE) anistropy. Can model Mohr–Coulomb soil model
PLAXIS (FE) complex wall and excavation may give unrealistic ground
CRISP (FE) geometry. Can model deformation
FLAC (FD) consolidation at construction
ABACUS (FE) stages as soil moves from
undrained to drained
conditions. Some programs
can model 3D conditions

the risk of inadequate prop capacity; wall penetration below final formation
level.
The wall penetration depth should always be regarded as a minimum when
calculated by methods described in this chapter. Adequate embedment depths
will depend on the risk of hydraulic failure as well as passive resistance to the
sheeting and relief of load from the lowest frame of bracing. In all multi-
propped (or anchored) walls the principal risks of failure in soils other than
very soft clays will stem from the adequacy of embedment depth of the wall
or sheeting and the sufficiency of strutting or anchoring. To counter these
risks in temporary works design requires care and experience.

Other design Plastic design


considerations Recent work by Kort46 on sheet pile design for sheet piles has highlighted
potential savings in sheet pile tonnage using plastic design methods. Kort’s
Design of vertical soil support 193

work includes some validation of the methods by full-scale tests in soft soils
and also includes allowance for the phenomenon of oblique bending that
may occur in U-shaped, Larrsen sheet piles. Kort’s work and its practical
application to sheet pile design is referred to in more detail in Chapter 7.

Construction of long walls in short lengths


It has long been realized that safer construction with less wall deformation
and reduced strut or anchor loads apply when trenchworks and deep
excavations can be taken out in short lengths and the permanent works are
built in stages progressively within these lengths. This avoidance of keeping
excavations for greater length and greater time with only temporary support
than necessary is good construction sense. This method of working in short
lengths can particularly apply at the deepest section of cofferdams with
short length working during excavation and installation of the bottom
frame and then the subsequent replacement of the temporary framing,
bottom upwards, by the permanent support. This method is very often
most desirable at this depth and in some instances, essential.
Until finite element analysis was applied in three dimensions it was not
possible to model short length working and its benefits by calculation.
Recently research work by Gourvenec et al.47 has used the CRISP FE pro-
gram to compare the deformation of a long embedded cantilevered diaphragm
wall excavation using both 3D and plane strain analyses with field measure-
ments of deformation. The wall was 800 m long with a cutting for highway
construction. The upslope side of the cutting was supported by diaphragm
wall panels 22.85 m deep, 1.5 m thick and 7.5 m long. The downslope wall
panels were 12.87 m deep, 1 m thick and 5 m long. An earth berm, 6.36 m in
height and 5.17 m wide at its base with a 60 degree face was used to provide
temporary support to the upslope wall. The wall height at the upslope side
was 7.86 m above formation level. The soil retained was a medium dense
sandy silt overlying a stiff overconsolidated fissured clay. Pore pressures
were hydrostatic below the sand/clay interface. Both diaphragm wall
installation and excavation of the berm were modelled together with realistic
construction times for each part of the staged construction.
The wall deformation measured on site compared favourably with the 3D
FE analysis deformed wall shape but the plane strain analysis predicted
total deformation at the top of the wall as 38 mm, whereas the total from
the 3D analysis was 20 mm and broadly agreed with the field measurement.
The conclusions reached were as follows.
(a) Reduced ground movements during wall installation were calculated
with the 3D analysis compared with the 2D, although the analysis
method did not have a significant effect on subsequent wall movements
during bulk excavation in front of the wall.
(b) Up to 80% of the post-installation wall movement occurred during
bulk excavation to the initial berm profile.
(c) Wall movement when the berm was removed in stages was reduced by
about 50% using the 3D analysis compared to that using the plane
strain analysis.
(d ) Wall movements at the initial base slab prop section during removal of
adjacent sections of the berm were very small. Both in the analysis and
on site the initial section of base slab prop was effective in restricting
wall movement during excavation of the nearby berm sections.
The effect of short length working compared with bulk excavation full length
is the reduced rate of dissipation of negative excess pore water pressure
induced by bulk excavation. This, in turn, makes for greater safety.
194 Deep excavations

Fig. 5.36. ‘Heidelberg’ soldier


pile and lagging wall allowing
extraction of steelwork wallings
and soldiers (patented by
Bilfinger and Berger)

King post walls


The use of soldier piles, often supported by tiers of ground anchors with hori-
zontal timbers (or laggings) or in situ concrete placed in successive lifts or
sprayed as shotcrete, provides an economical form of soil support, particu-
larly in dry subsoils. Care should be taken, however, in assessing the passive
resistance below final formation level of such a wall system. The passive

Fig. 5.37. King post wall derivatives (dimensions in metres)


Design of vertical soil support 195

Fig. 5.38. Cofferdam standard construction details: (a) typical layouts, sheet piled cofferdams; (b) connection details;
(c) temporary anchor head detail; (d) standard details with tie rods to deadmen
196 Deep excavations

Fig. 5.39. (a) Schematic arrangements for anchored and braced piled and diaphragm walls
Design of vertical soil support 197

Fig. 5.39. (b) Typical jackable end detail for a tubular (CHS) prop (after Twine and Roscoe42 )

resistance of soil at the front of the soldier piles within their embedment depth
can only be mobilized by movement of the soldier pile towards the excavation.
Soldier pile walls are relatively flexible constructions, and in relatively dense
granular soils and stiff clays this mobilization of passive resistance without
198 Deep excavations

Fig. 5.39. (c) Typical jackable end detail for a UC prop (after Twine and Roscoe42 )

undue lateral movement can reasonably be assumed. In these circumstances it


is also reasonable to allow an effective width of soil acting to provide passive
resistance in excess of the net width of the steel or reinforced concrete pile
acting as the soldier. Where soldier spacing centre-to-centre is greater than
four times the net width of the soldier below formation level it is reasonable
to assume an effective width of soil to provide passive resistance equal to
twice the soldier width below formation level. Where the soldier spacing is
less than four times the soldier width below formation level no such increase
in calculated passive resistance is permissible because of the risk of overlap of
calculated passive zones.
It is essential to check the bearing capacity at the base of such soldiers where
inclined anchors are used. The vertical components of all anchors should be
summated as they are pre-stressed together with the soldier pile self-weight
comprising the vertical load at the soldier base.
Figure 5.36 shows a soldier pile wall constructed by Bilfinger and Berger.
Part of the wall is built in a patented form known as a ‘Heidelberg’ wall
where the walings and soldiers can be extracted for reuse.
Design of vertical soil support 199

Construction details
A selection of construction details from the Author’s files is given in Figs 5.37
to 5.39. The wall types shown include anchored sheet pile walls, king post
walls using timber laggings, cofferdam details and multi-prop walls using
secant piles and diaphragm walls. Details of typical end details where jack
installation is included are also reproduced from Twine and Roscoe’s
CIRIA report42 .

References 1. DD ENV 1997–1 Eurocode 7. Geotechnical design. Part 1: general rules. British
Standards Institution, London, 1995.
2. BS 8110. Structural use of concrete. British Standards Institution, London, 1997.
3. BS 8007. Code of practice for design of concrete structures for retaining aqueous
liquids. British Standards Institution, London, 1981.
4. BS 5400. Steel, concrete and composite bridges. British Standards Institution,
London, 1990.
5. BD 42/00. Design manual for roads and bridges, Vol. 2, section 1, (DMRB 2.1.2),
HMSO, London, 2000.
6. Peck R.B. Deep Excavations and tunneling in soft ground. Proc. 7th Int. Conf.
S.M.F.E., Mexico City, 1969, State of the art volume, 225–290. Sociedad de
Mexicana de Suelos, Mexico City, 1969.
7. Clough G., Smith E. and Sweeney B. Movement control of excavation
support systems by iterative design procedure. ASCE Ground. Engng, 1989, 1,
869–884.
8. Clough G. and O’Rourke T. Construction induced movement of insitu walls.
ASCE Special Publication No. 15. Proc. Design and Performance of Earth Retain-
ing Structures, Cornell University, 1990, 439–470.
9. St John H., Potts D., Jardine R. and Higgins K. Prediction and performance of
ground response of a deep basement at 60 Victoria Embankment. Proc. Wroth
Memorial Symposium, Oxford, 1992. Thomas Telford Ltd., London, 1993.
10. Carder D. Ground movements caused by different embedded retaining wall con-
struction. Transport Research Laboratory, Crowthorne, 1995, TRL report 172.
11. Fernie R. and Suckling T. Simplified approach for estimating lateral wall
movement of embedded walls in UK ground. Proc. Int. Symp. Geo. Aspects of
Underground Construction in Soft Ground, City University, 1996, 131–136. City
University, London, 1996.
12. Carder D., Morley C. and Alderman G. Behaviour during construction of a
propped diaphragm wall in London clay at Aldershot and road underpass. Trans-
port Research Laboratory, Crowthorne, 1997, TRL Report 239.
13. Long M. Database for retaining wall and ground movements due to deep
excavations. ASCE J. Geotech. Eng., 127, March 2001, 203–224.
14. Brooker E.W. and Ireland H.O. Earth pressures at rest related to stress history.
Canadian Geotech. J., 1965, 11, Feb., 1–15.
15. Caquot A. and Kerisel J. Tables for the calculation of passive pressure, active
pressure and bearing capacity of foundations. Gauthier-Villars, Paris, 1948.
16. Civil engineering codes of practice. No. 2. Earth retaining structures. Civil
Engineering Codes of Practice Joint Committee, London, 1951.
17. CIRIA. The design and construction of sheet piled cofferdams. Thomas Telford
Ltd., London, 1993, Special publication 95.
18. Terzaghi K. and Peck R.B. Soil mechanics in engineering practice. Wiley, New
York, 2nd edition, 1967.
19. BS 8002. Earth retaining structures. British Standards Institutions, London,
1994.
20. Rowe P.W. Anchored sheet pile walls. Proc. Instn Civ. Engrs, Part 1, 1952, 1,
Jan., 27–70.
21. BD 37/01. Design manual for roads and bridges, Vol. 1, section 3, (DMRB 1.3).
Highways Agency, London, 2001.
22. Design manual 7.2. Foundation and earth structures. US Navy, Washington, DC,
1982.
200 Deep excavations

23. Goldberg D.T. et al. Lateral support systems and underpinning, Vol. 1, design and
construction, Vol. 2, design fundamentals, Vol. 3, construction methods. Federal
Highway Administration, Washington, DC, 1976, FAWA-RD-75, 128.
24. Simpson B. and Driscoll R. Eurocode 7: A commentary. Building Research
Establishment, Watford, 1998.
25. Sainflou M. Essai sur les digues maritimes verticales. Ann. des Ponts et Chausse´es,
1928, 98, 5–48, (translated by US Corps of Engineers).
26. Minikin, R. Wind, waves and maritime structures. Griffin, London, 1963.
27. BS 6349 (part 1). Maritime structures. British Standards Institution, London,
1984.
28. Potts D.M. and Burland J.B. A parametric study of the stability of embedded earth
retaining structures. Transport Research Laboratory, Crowthorne, 1983, Report
813.
29. Symons, I.F. Assessing the stability of a propped in-situ retaining wall in over-
consolidated clay. Proc. Instn Civ. Engrs, 1983, 75, Dec., 617–633.
30. Piling handbook. British Steel Corporation, Scunthorpe, 1988.
31. Sheet piling design manual. USS Steel, Pittsburgh, 1974.
32. Review of design methods for excavations in Hong Kong. Geotechnical Control
Office, Hong Kong, 1990.
33. Blum H. Einspannungsverhaltnisse bei Bohlwerken. Ernst and Son, Berlin, 1931.
34. Terzaghi K. Anchored bulkheads. Trans. ASCE, 1954, 119, paper 2720, 1243–
1324.
35. Rowe P.W. A theoretical and experimental analysis of sheet pile walls. Proc. Instn
Civ. Engrs, 1955, 4, Jan., 32–69.
36. Rowe P.W. Sheet pile walls in clay. Proc. Instn Civ. Engrs, 1957, 7, July, 629–654.
37. Terzaghi K. Theoretical soil mechanics. Wiley, New York, 1943.
38. Bjerrum L. and Eide O. Stability of strutted excavations in clay. Ge´otechnique,
1956, 6, 32–47.
39. Aas G. Stability problems in a deep excavation in clay. Norwegian Geotechnical
Institute, Oslo, 1985, Publication 156.
40. Bjerrum, L. Problems of soil mechanics and construction on soft clays. Proc. 8th
Int. Conf. S.M.F.E., Moscow, 1973, Vol. 3, 111–159. USSR National Society for
Soil Mechanics and Foundation Engineering, Moscow, 1973.
41. BD 28/87. Early thermal cracking of concrete. UK Department of Transport,
London, 1987.
42. Twine D. and Roscoe H. Prop loads: guidance on design. CIRIA, London, 1996,
Funders report.
43. Batten M. and Powrie W. Measurement of temporary prop loads at Canary
Wharf. Proc. ICE Geotech. Eng., 143, 2000, 151–163.
44. Richards D., Homes G. and Beadman D. Measurement of temporary prop walls
at Mayfair car park. Proc. ICE Geotech. Eng., 137, 1999, 165–174.
45. Potts D.M. The analysis of earth retaining structures. Proc. Conf. Retaining
Structures. Institution of Civil Engineers, London, 1992, 167–186.
46. Kort D.A. Steel sheet pile walls in soft soil. DUP Science, Delft, 2002.
47. Gourvenec S., Powrie W. and De Moor E. Three dimensional effects in the
construction of a long retaining wall. Proc. ICE Geotech. Eng., 2002, 155, July,
163–173.
48. Padfield C.J. and Mair R.J. Design of retaining walls embedded in stiff clay.
CIRIA, London, 1984, report 104.
49. CIRIA. Embedded retaining walls: guidance for economic design. Gaba A.R. et al.
CIRIA, London, 2002.
50. Bjerrum L. et al. Earth pressures on flexible structures. Proc. 5th Euro. Conf.
SMFE, Madrid, 1972, Vol. 2, 169–196.
51. Institution of Structural Engineers, Design and construction of deep basements
including cut-and-cover structures. Draft, 2003. Inst. Structural Engrs, London.

Bibliography Bjerrum L. et al. Earth pressures on flexible structures. Proc. 5th Euro. Conf.
S.M.F.E., Madrid, 1972, Vol. 2, 169–196. Sociedad Española de Mechánica del
Suelo y Cimentaciones, Madrid, 1972.
Design of vertical soil support 201

Burland J.B. et al. The overall stability of free and propped embedded cantilever
retaining walls. Ground Engng, 1981, 14, No. 5, 28–38.
Carder D.R. and Symons L.F. Long term performance of an embedded cantilever
retaining wall in stiff clay. Ge´otechnique, 1989, 39, No. 1, Mar., 55–76.
EAU 90. Recommendations of the Committee for Waterfront Structures, Harbours and
Waterways. Ernst and Son, Berlin, 6th English edition, 1993.
Egger P. Influence of wall stiffness and anchor prestressing on earth pressure distribu-
tions. Proc. 5th Euro. Conf. S.M.F.E., Madrid, 1972, Vol. 1, 259–264. Sociedad
Española de Mechánica del Suelo y Cimentaciones, Madrid, 1972.
Garrett C. and Barnes S.J. Design and performance of the Dunton Green retaining
wall. Ge´otechnique, 1984, 34, Dec., 533–548.
Head J.M. and Wynne C.P. Designing retaining walls embedded in stiff clay. Ground
Engng, 1985, 18, Apr., 30–38.
McRostie G.C. et al. Performance of tied back sheet piling in clay. Canadian Geotech.
J., 1972, 9, 206–218.
Potts D.M. and Burland J.B. A numerical investigation of the retaining walls of the Bell
Common tunnel. Transport Research Laboratory, Crowthorne, 1983, Supplemen-
tary Report 783.
Potts D.M. and Fourier A.B. The effect of wall stiffness on the behaviour of a
propped retaining wall. Ge´otechnique, 1984, 34, Sept. 383–404 (Discussion 35,
119–120).
Simpson B. et al. An approach to limit state calculations in geotechnics. Ground
Engng, 1981, 14, No. 6, 21–28.
Symons L.F. Behaviour of a temporary anchored sheet pile wall on A1(M) at Hatfield.
Transport Research Laboratory, Crowthorne, 1987, Research report 99.
Terzaghi K. Evaluation of coefficients of subgrade reaction. Ge´otechnique, 1955, 5,
No. 4, 297–326.
Winkler, E. Die lehre von elastizitat und festigkeit, Prague, 1867.
Cofferdam construction
6
Design and Packshaw1 defined a cofferdam as a temporary structure built to exclude earth
construction and water from a construction area and thus permit the work inside to be
responsibilities carried out in the dry. Nevertheless, he conceded that ‘in the dry’ should
not be taken too literally; seepage upwards through the formation and
possibly leakage through the sheeting of the cofferdam walls may bring
considerable water quantities into the working space. Packshaw said that to
fulfil the purpose of the cofferdam:
(a) the walls of the dam and any internal bracing must withstand the loads
and stresses imposed on them
(b) the amount of water entering the dam must be controlled by reasonable
pumping and must not interfere with the permanent construction inside it
(c) it must be possible to excavate down to the required level without
causing the ground to boil, heave or flow into the dam in an uncon-
trolled manner
(d ) the walls must not deflect inwards so much as to interfere with the
permanent construction inside the dam
(e) the cofferdam must have overall stability against unbalanced earth
pressure or ground movements such as circular slips.
The responsibility for the safety of cofferdam structures is frequently defined by
statutory regulations. In the UK the Construction Regulations2 and Construc-
tion (Design and Management) Regulations stipulate obligations on both
designers and contractors. Such regulations require adequate and safe support
irrespective of depth. Previous limits of minimum depths requiring timbering no
longer apply and all excavations are assessed on a site specific basis. Safety
barriers are required where any risk exists and any excavation susceptible to
risk of flooding must have adequate means of escape. All excavations require
to be regularly inspected by a competent person and adequate precautions
are required to avoid the presence of harmful, noxious gases in confined spaces.
In terms of the share of contractual responsibility for cofferdam design and
construction, other than health and safety laws, this depends on the nature of
the contracts between Employer, Engineer and Contractor. In practice, both
design and construction are largely carried out by the Contractor, with whom
these responsibilities mainly lie, although not completely so. In the UK for
instance, a well used contract form, the ICE Conditions of Contract3 requires
the Engineer’s approval to temporary works design, but such approval
does not relieve the Contractor from responsibility under contract. It
should also be remembered, however, that Engineer and Contractor have
duties of care under common law in the UK and damages may be applied
to both parties should, for instance, the Engineer be aware of deficiencies in
design or construction which result in failure and harm to life or property.
With increasing use of diaphragm wall construction in cofferdam work in
both temporary and permanent works phases, it is frequently necessary for
Contractor and Engineer to accept responsibility for design in each phase,
Cofferdam construction 203

the Contractor for the temporary works and the Engineer thereafter for the
permanent works. Where stresses induced during the temporary works
period cause later distress there is every opportunity for dispute between
the parties.
Successful cofferdam work implies minimum cost because of its temporary
nature. The design is frequently made two or three times: at feasibility stage
by the Engineer, at bid stage by the Contractor, and at construction stage by
the successful Contractor. The Contractor’s design engineer is finally set the
task of designing a stable cofferdam that complies with the permanent works
requirements at minimum cost in terms of construction, maintenance and
final removal. The cofferdam must be safe, and appear so to those who work
in it. Normal design criteria for permanent works may not prove economical
for temporary construction and factors of safety and soil design parameters
may require much care in selection. In particular, the choice of working stresses
for construction materials which have been used before and suffered
some deterioration will require judgement. Although cofferdam works are
frequently considered short-term structures, the period of their use may
range from months to years depending on the scale of the contract and any
extensions to the contract period. The requirement for structural durability
with time therefore depends on assessment of risk and the consequences of
failure to ‘life and limb’ and construction damage. Only the latter may be
influenced by insurance cover to temporary and permanent work.
The quality of design data from site investigation should also be referred to.
Packshaw1 stated the following basic rules.
. The site investigation should be deep enough to determine soil and
water pressures, especially artesian pressures. A CIRIA report4 sug-
gested an investigative depth to at least 1.5–2 times the excavation
depth, possibly more in weak soils.
. The site investigation should extend around the cofferdam site to a
distance equal to its depth to formation level.
As a prerequisite to the investigation, accurate records of existing levels, site
history, seasonal variation in groundwater levels, tides and flood frequency
are essential. Unless these data are available the cofferdam designer and
constructor cannot fulfil their responsibilities irrespective of those defined
by statute and contract.

Types of cofferdam Cofferdams may be divided into three categories:


(a) sheeted types, usually with external shores, anchors or internal bracing
(b) double-skin types, where sheeting is used to form cells, circular in plan
shape, or with parallel walls, each containing fill material; the strength
of these structures depends on the composite action of fill, sheeting and
the underlying soil support
(c) gravity and crib types, where structures made from mass concrete, often
precast, or from soil and rock fill resist, by their own weight, disturbing
forces of river and sea water flow and groundwater pressure.
Design ingenuity has developed many variations on these types as cofferdam
locations, sizes and soil and water depths have demanded. The choice of
cofferdam type is wide, and analysis of construction expense and time will
frequently be needed for a particular site before a choice can be made. The
head of water to be retained, soil properties above and below formation
level and depth to rockhead are often critical matters influencing the choice
of method. In some instances the principal choice remains between cofferdam
204 Deep excavations

and caisson construction, especially where excavation depths are considerable


in poor ground. Improvements to construction methods and mechanical
equipment have tended to increase use of cofferdam construction to greater
depths in recent years. Typically, in the 1960s and 1970s the economical
depth of cofferdams would have been in the range 15 to 20 m of water,
depending on subsoil type. In later years, however, the economical limit of
cofferdam construction frequently extends 30 to 40 m or more, and caisson
work continues to decline in popularity.

Sheeted cofferdams Where space allows, battered excavations may be economical, particularly in
clay subsoil or where dewatering is feasible. Where space is confined or
groundwater is less easily dealt with, the use of a temporary wall, cantilevered
or supported, becomes necessary at the curtilage of the permanent works.
The use of cantilevered sheeting, in steel sheet piling, bored piles or
diaphragm walling, will only serve excavations of limited depth, although
this depth may be effectively increased by either a short batter at the top of
the wall or a temporary berm of soil against the wall at formation level.
This berm is removed in short lengths as construction of the permanent
wall proceeds. Cantilevered walls are particularly economical where relatively
dry cohesionless soils overlie clays of reasonable strength at moderate depth.
Such conditions frequently occur in London, with the top surface of London
clay occurring at depths of 5 to 7 m. Although the upper horizon of the
London clay may be weakened by weathering, a rule of thumb is often used
to make the length of pile embedment equal to the depth from ground level
to formation level. Excessive horizontal movement at the head of cantilevered
walls can be reduced by use of in situ capping beams at the head of diaphragm
walls or bored piles. In other instances, the buttressing effect of return angles
in plan in the sheeting may limit excessive deformation to allowable values or
restrict the plan length, requiring temporary propping at intervals between
return angles in the wall.
Where depth to formation level within a large excavation precludes support
from cantilever sheeting because of unacceptably high sheeting moments and

Fig. 6.1. Sheet piled excavation


supported by raking shores
with temporary berm support
being removed by tracked
excavator, London (courtesy of
AMEC)
Cofferdam construction 205

deflections, the sheeting is temporarily supported by horizontal struts


spanning wall-to-wall, by raking shores, spanning where possible from
partially completed permanent work or, where space allows and easement is
gained, by ground anchors. Figure 6.1 shows a sheet piled excavation sup-
ported by raking shores, the temporary berm being removed as the rakers
are secured. The relatively high expense of the double-support operation,
which is uneconomic for small-scale excavation works, should be noted.

Circular cofferdams
The extent of internal bracing within a sheeted excavation can be reduced by
the adoption of a cofferdam that is circular in plan. The plan shape of the
permanent works needs either to fit accurately within the circular cofferdam
or be reappraised to do so.
The advantages gained by circular structures rest in the comparative
economy of structural rings built to resist hoop compression. However, the
CIRIA report4 restated that circular walings used to support sheet piles in
circular plan forms require relatively small distortion to reach critical
instability, and advised the empirical rule d  D=35, where d is the difference
between outer and inner radii of the ring beam and D is the diameter of the
cofferdam to the inner face of the piles.
Stiffness is therefore the essential criteria for circular walings. For this
reason reinforced concrete is frequently preferred for circular waling construc-
tion. Although steel walings are used more frequently for cofferdams over
water, the empirical proportions of d  D=35 must be observed and care
taken to restrain the inner flange of the waling beam to prevent buckling.
The CIRIA report gave values for safe waling loads for reinforced concrete
waling sections suitable for cofferdams between 5 and 35 m in diameter.
Although reinforced concrete waling sections cast against sheet pile, bored
pile or diaphragm wall sheeting have the practical advantage of avoiding
the use of timber packings between waling and sheeting, they also have
disadvantages in the curing time needed before they contribute to cofferdam
strength, delaying continuous excavation and possibly impeding extraction
of sheet piles for reuse.
In relatively shallow excavations it may prove economical to anchor a
square sheet piled enclosure to an outer circumscribing reinforced concrete
ring waling, as shown in Fig. 6.2. Deeper excavation will require successive
internal ring walings with sheeting of circular plan form. Figure 6.3 shows
a land cofferdam with an upper waling cast externally to the sheeting and
a lower waling cast inside.
Packshaw1 stated that the practical limit of a circular cofferdam was of the
order of 45 to 60 m in diameter using sheet piling. Developments in piling
equipment and the increasing international use of diaphragm walling, often
incorporated into the permanent structure, have extended the use of contigu-
ous and secant piles and diaphragm walling into cofferdams of circular plan
shape. An example of a large pumping station at the Isle of Grain, UK, is
shown in Fig. 6.4. The diaphragm walls, designed to span vertically between
circular walings, form a circle of 75 m diameter.
The design and construction of a circular diaphragm wall for a ventilation
building at the end of an immersed tube tunnel construction (the Ted
Williams) tunnel in Boston, MA was described by Kirmani and Highfill5 .
Soil conditions consisted of 15 m of fill overlying 7.6 m of Boston blue clay
overlying glacial till and weathered argillite. Groundwater was 1.8 m below
ground level. The cofferdam was almost 80 m in diameter and the depth to
formation level was 24 m. The circular cofferdam, consisting of a segmented
cylinder of diaphragm wall panels embedded into rock for lateral stability
206 Deep excavations

Fig. 6.2. Square sheet piled


cofferdam anchored to outer
circumscribing reinforced
concrete ring waling1

was lined with an internal concrete wall cast successively in 2.4 m deep ring
beams. The feature of both the design and construction of this cofferdam
was the requirement for two large openings at each side of the cofferdam.
The following assumptions were made for the load-carrying properties of
the cofferdam structure:
(a) hoop stresses would be resisted by the combination of diaphragm wall
and lining wall in proportion to their thicknesses
(b) horizontal bending would be resisted by the ring beams
(c) vertical bending would be resisted by the diaphragm wall
(d ) buckling of the cylindrical cofferdam would be checked with the
combination of both diaphragm and lining wall.
Some indication of the complex steps involved in cofferdam construction are
shown in Fig. 6.5. The junction of the immersed tube tunnel and the circular
cofferdam is shown in plan in Fig. 6.6. The seal, which had to be watertight,
was achieved using two hinged steel gates that spanned between the tunnel
and buttresses built in diaphragm walling at the outside of the circular coffer-
dam. The gates consisted of horizontal steel beams with a continuous vertical
steel plate. The beams were spaced at 1.2 m centres and a hinge was made by
installing a steel pin through the web of each beam and through a seat
bracketed off the tunnel. A continuous vertical rubber seal was attached to
the opposite end of the gate and positioned against a steel channel section
embedded in each buttress.
Cofferdam construction 207

Fig. 6.3. Circular cofferdam


with reinforced concrete ring
walings, outer waling
uppermost, inner waling above
formation level
208 Deep excavations

Fig. 6.4. Circular diaphragm


wall with internal reinforced
concrete waling, pumping
station, Isle of Grain, UK
(courtesy of Laing/O’Rourke)

Fig. 6.5. Circular cofferdam for


ventilation building, Boston,
MA. Stages in construction5
Cofferdam construction 209

Fig. 6.6. Plan view of the secondary cofferdam adjacent to the main cofferdam5

A circular cofferdam, also with a lining wall, was constructed as part of


recovery works for a collapsed tunnel at Heathrow Airport, London. This
cofferdam, 60 m in diameter and 30 m deep to formation level, was con-
structed in secant piling, 40 m deep; the secant piles extended to a depth of
20 m and below this depth the piles were contiguous. The upper part of the
pile over the secant length, was constructed as a 900 mm dia. contiguous
pile encased in a soft annulus of cement–bentonite grout as shown in
Fig. 6.7. The internal lining wall was constructed against the piles in successive

Fig. 6.7. Circular cofferdam:


Heathrow Airport, London,
secant pile wall construction
(courtesy Mott MacDonald)
210 Deep excavations

pours 1 m deep. The secant pile construction was noteworthy for its specified
verticality tolerance of 1:200; in the event an average tolerance of 1:400 was
achieved and the cofferdam was within 50 mm of a perfect circle at a depth of
30 m. It is likely that the adoption of a secant pile solution in preference to use
of a diaphragm wall was favoured because of the availability of piling plant on
site.
Figure 6.8 shows a vertical section through a deep circular cofferdam built
for a pumping station at Weston-Super-Mare, UK, in difficult ground condi-
tions with a high water table. Frodingham No. 4 sheet piles 23 m long were

Fig. 6.8. Vertical cross-section


of a circular cofferdam showing
reinforced concrete ring waling
construction, pumping station,
Weston-Super-Mare, UK1
Cofferdam construction 211

Fig. 6.9. Figure-of-eight plan


shape diaphragm wall water
intake, Cagliari, Italy (courtesy
of Icos)
212 Deep excavations

supported by five ring walings in a cofferdam 30 m in diameter. Circular


cofferdams on land require uniformity of soil and water pressure around
the peripheral sheeting. Ground levels should therefore be sensibly level,
soil strata relatively level and groundwater pressure constant around the
cofferdam.
In a small number of diaphragm wall examples the limited overall diameter
of the circular structure and tight control of panel verticality tolerance have
allowed the hoop compression to be transferred efficiently from one panel
to the next, avoiding ring walings completely. Due to the absence of bending
stresses in these wall panels only limited quantities of reinforcement are
required when hoop compression can be mobilized in this way.
Some variations on the circular plan shape can be economically used. Icos6
documented the successful use of elliptical plan shapes for diaphragm wall
cofferdams at Palermo, Legano and at the Quero hydroelectric plant, all in
Italy. Double circular plan shape diaphragm walls in the form of a figure-
of-eight were used by Icos6 in Cagliari and at Beckton in London. Soil
conditions and the general arrangement of the walls at Cagliari are shown
in Fig. 6.9.
Two 36 m dia. tanks were recently built at Blackpool in the United King-
dom7 in close proximity to each other. The tanks are shown in cross-section
in Fig. 6.10. A two-dimensional analysis using the FLAC finite difference
program was made to assess interaction between the tanks. The interaction
analysis was used to estimate the induced additional movements of the dia-
phragm walls at serviceability limit state, to estimate additional structural
forces and bending moments in the diaphragm walls at both serviceability
and ultimate states and check the reinforcement design due to these additional
forces and moments. The design required tight tolerances for construction of
the 40 m deep diaphragm wall. The maximum verticality tolerance allowed
was 1 in 300, that is a 133 mm maximum offset allowing guide wall tolerances
for each panel over 40 m. This maximum tolerance required the use of a cutter
machine although rope-suspended grabs were used for diaphragm wall

Fig. 6.10. Circular tank construction, Blackpool, UK7


Cofferdam construction 213

Fig. 6.11. Circular tank


construction, Blackpool, UK:
panel configuration7

excavation down to the glacial sands. A three-bite primary panel and a single-
bite secondary panel configuration was used with cut joints as shown in
Fig. 6.11.
To achieve the required verticality the Hydrofraise machine was equipped
with instrumentation that allowed 3D real-time positional monitoring of
the cutter head, the alignment being corrected continuously. The Hydrofraise
machine has two means of adjusting alignment, by varying the individual
cutter drum speeds or the angle of the cutter drums themselves, see
Fig. 6.12. The instrumentation used is shown in Fig. 6.13. The maximum
inclination of the panels as measured after excavation was 1 in 500 or
80 mm at an excavation depth of 40 m with a measured verticality of 1 in
1000 on most panels.
Excavation for the tanks was made after wall construction by two 16 t 3608
excavators of 12 t capacity which were lifted to ground level by hydraulic
Fig. 6.12. Hydrofraise
cranes of 80 t capacity. Using this method, an excavation rate of 1000 m3 /
alignment correction7 day was achieved on each tank for the full excavation depth.

Fig. 6.13. Hydrofraise cutter


real-time positional monitoring
‘Evolution 2’ machine7
214 Deep excavations

Fig. 6.14. Forth Road Bridge, south pier construction: (a) bracing frame floated into position prior to pitching sheet piles
(courtesy of AMEC); (b) construction sequence25
Cofferdam construction 215

On the Forth Road Bridge in Scotland cofferdams for the south main river
pier were built in a figure-of-eight, two circular plan forms being connected by
cross-bracing. Figure 6.14 shows the bracing frame arrangement and the
construction sequence, including the sinking of caissons from the floor of
the cofferdams.
Circular cofferdams, in pairs, were built for the east and west piers of
the first Severn Bridge, UK. The marl into which the cofferdam sheeters
were founded was too hard for pile driving. The toes of the sheet piles were
therefore bolted to a concrete anchor ring which had been cast on to the
rock surface between tides. In turn, the anchor ring was tied to the marl by
dowel rods. The rock was excavated by mechanical loader after blasting
with small charges, and a precast concrete segmental lining was grouted
into place. The excavation and lining construction were carried out tidally.
Excavation continued through the underlying mudstones and when a hard
layer was met the cofferdam area was cleaned and covered with blinding
concrete. Both ends of the cutwater pier were built to the high tide level
within the cofferdams, the sheeting and bracing removed, and the central
portion of the pier completed between tides. Figure 6.15 shows the cofferdams
and the construction sequence.

Braced cofferdams
The use of cofferdam plan shapes other than circular or elliptical requires
structural support to the sheeting by internal bracing for all but shallow
excavations. Before considering construction of braced cofferdams over
water and land, the most frequent causes of failure should be considered4 :
(a) failure to make allowance for variation of water levels due to seasonal
effect, constriction of flow by the cofferdam itself, and overtopping by
high tide or flood conditions
(b) failure to provide and use balancing and/or flooding valves
(c) variations between the ground conditions assumed for design and those
revealed by the excavation
(d ) failure to control the excavation levels, at all stage of construction, to
those indicated in the design
(e) failure to provide adequate support to frames and lateral restraint to
compression flanges of waling beams
( f ) use of frames to support plant loads, such as pumps and generators, in
excess of design allowances
(g) frame members damaged by impact from skips and grabs and
inadequately repaired
(h) unauthorized strut removal or substitution
(i) uncontrolled water ingress through separated interlocks into the
passive zone below formation level
( j) ill-fitting struts carrying unintentional eccentric loads.
Overall failure is more likely to occur as a result of inadequate strutting or
passive soil failure due to inadequate sheeting embedment rather than flexural
failure of the sheeting itself. Packshaw1 examined the reasons for cofferdam
failure by soil type and concluded that the cause of failure is almost invariably
below formation level; either the soil inside the cofferdam is unable to resist
external forces or it undergoes a change in its properties as a result of these
forces. Base failure takes several forms as follows.
(a) In permeable ground, if a ‘blow’ occurs because of inadequate cut-off or
excessive pumping from sumps it is likely that the bottom of the piling
will be forced inwards due to lack of passive resistance of the soil below
216 Deep excavations

Fig. 6.15. Severn Bridge, west


pier construction: (a) twin
circular cofferdams with
segmental ring construction
below sheet pile walls (courtesy
of AMEC); (b) construction
details26
Cofferdam construction 217

formation level. As excessive load is transferred to the lower frames,


progressive collapse follows to the whole structure.
(b) In highly permeable soil, the upward flow of water will reduce the
passive resistance of soil below formation level. This may be sufficient
to overload the lowest frame even if quick conditions do not develop
at the formation.
(c) Variations in loading, say tidal variations, not only test the stiffness and
security of the bracing but, by repeated movement, may reduce passive
resistance of the subsoil below formation. The lowest frame must be
designed to allow for any transfer of load and should be located as
near formation level as possible.
(d ) In soft clays excavated below a critical depth the piling may deflect
inwards.
(e) Uplift due to water pressure may cause base failure. Clay below forma-
tion level with artesian water in sand lenses or within a cohesionless
substrate are typical risk conditions.
The cause of failure of a major cofferdam on the River Thames did not fall
into these categories and deserves mention. The cofferdam, for a river wall
improvement, had five frames with one face towards the river and was 15 m
deep below high water, 15 m wide and approximately 80 m long. All strutting
and walings were of steel section, and standards of workmanship and
maintenance were generally above average. One particular strut on the
bottom frame was, however, badly fitted; the cut end of the strut was not
square to its axis and without shims the eccentric load placed on this strut
was sufficient to cause failure at the highest tide after full excavation. The
strut bowed badly and, without knowledgeable supervision during a weekend,
no remedial action was taken. The next high tide caused progressive failure to
the passive soil resistance below formation level and also to each of the four
frames in turn above the bottom frame. Bad workmanship, particularly on
struts, and avoidance of progressive failure should be added to the above list.

Braced cofferdams over water


In the case of braced cofferdams over water the sheeting is invariably
steel piling, and the bracing, of walings, struts, king piles and puncheons, is
nowadays usually in structural steel. Only where large quantities of fill
material are economically available to build soil islands can diaphragm wall-
ing and secant piling be considered for use as cofferdam walls.
The minimum strength of sheet piles for braced cofferdams is the greater
strength required either in flexure to resist soil and water pressure or to
resist driving stresses. Sheet piles are manufactured in various profiles: the
flexural strength of Z-type sheet piling with interlocks at the flanges can be
assumed to be the strength of the complete section; those sections clutched
on the centre-line (U profile) depend on frictional resistance to shear within
the interlocks to develop the full cross-section strength in flexure. Williams
and Little8 considered this matter in depth. BS 80049 recommends that
where piling is cantilevered substantially above the first frame and passes
through very soft clay, or pile cut-off is limited by shallow bedrock, the flex-
ural strength of the whole clutched section should be obtained by welding the
clutches. To summarize the advantages and disadvantages of each section, Z
profiles tend to have more closely fitting interlocks which are watertight for
marine use, and deflections tend to be less when used to cantilever. Z profiles,
however, reduce in modulus if they are allowed to rotate during driving. In
approximate terms, a rotation of 58 results in a 15% modulus reduction. U
profiles, on the other hand, have a greater single section modulus and are
218 Deep excavations

less prone to deviate in penetration of dense soils and can be reused more
often. The rotation permitted by the interlocks is greater for the U profile
than the Z profile (98 compared with 38), so the U section is better when driving
sheeting to a tight radius.
The sheet piles are pitched and driven in panels. Both U or Z section sheet
piles should be driven in pairs wherever possible. The toes may need to be
strengthened for very hard driving, although this may make extraction diffi-
cult at a later stage. Controlled blasting of a narrow trench in the rock may
assist where sheet piles need to be driven into strong bedrock.
After pitching in pairs the piles are usually driven in echelon order (1–3–5–2–
4–6). In difficult conditions the piles may be driven in two stages, first to part
depth and then to full depth with a larger hammer. Piles may also be driven in
panels by pitching and driving the first pair to part, but firm, penetration, then
pitching the remainder of the panel in pairs. The last pair is fully driven and
then the remaining pairs are fully driven successively back to and including
the first pair. Jetting and pre-boring may sometimes be used to overcome
difficult drilling; jetting should be replaced by conventional driving in the
final metre of penetration to avoid loose subsoil at the pile toe.
The initial panel in a cofferdam should begin with the pile pair adjacent to a
corner pile, the last panel concluding with the corner pile, all the piles being
pitched and interlocked in this last panel before any are driven. For small
cofferdams all the piles should be pitched before driving is begun. Cofferdams
in water may be built by using the top frame, supported by temporary piles as
a template. This template should be built to the design cofferdam dimensions,
whereas lower frames should be made to reduced dimensions to allow for
sheet pile verticality tolerances. The setting-out line of the top frame should
therefore allow:

Table 6.1 Types of pile driving equipment4

Type of driver Soil Noise Quiet Vibration Extract Rate of Pairs/single


compatibility output versions output penetration driving

Double-acting All soils High No Low Yes Medium/low Both


air hammer except stiff
clay
Single-acting All soils High No Medium/low No Medium Pairs
diesel
Double-acting All soils Medium/low No Medium/low No Medium/low Pairs
diesel
Cable All soils Medium/low Yes Medium/low No Medium/low Pairs
operated drop
Hydraulic All soils Medium/low Yes Medium/low No Medium Both
drop
Vibrator Granular Medium N/A Low Yes High Both
and soft
clays
High-speed Granular Medium/low N/A Low Yes High Both
vibrator and soft
clays
Hydraulic Granular Low N/A Nil Yes Slow/ Single/
thrust with jetting medium panel
and clays
Cofferdam construction 219

. total verticality and positional tolerance of sheet piles


. a width dimension to allow cofferdam drainage at formation level
. adequate working space and accommodation for shuttering, where
used.
The choice of driving equipment is large and depends on the size and weight
of the sheet piles to be driven, embedment length, subsoil conditions, and
constraints on noise and vibration. In the UK, legal power is given to local
municipalities under the Control of Pollution Act 1974 to impose limits on
site noise where this adversely affects the quality of life. BS 522810 gives
guidance on vibration and noise control due to piling. It is worth remembering
that while noise gives rise to complaint, vibration gives rise to structural
damage. A summary of pile hammer types and their use is given in Table 6.1.
Impact hammers nowadays are usually hydraulic or diesel; drop hammers
and air hammers finding less use than previously. The characteristics of
typical current diesel hammers is shown in Table 6.2 and similar characteristics
of both small and large hydraulic hammers are listed in Table 6.3. As a general
rule, it is best to be on the heavy side with hammer capacity; light hammers tend
to spread the top of the pile without achieving increased penetration. A typical
hydraulic hammer, a Junttan, is shown mounted on a modern piling rig, with
hydraulic leaders in Fig. 6.16.
High-frequency vibrators, see Fig. 6.17, are most effective in loose to
medium dense granular soils particularly for sheet pile installation, although

Table 6.2 Characteristics of


Delmag diesel hammers. The Model Mass of ram Energy range Striking rate
data listed in this table were (kg) (m kg) (blows per minute)
obtained from manufacturers’
brochures and websites and D2 220 251–120 60–70
are current in March 2003. D4 380 503–156 50–60
Reference should be made to D5 500 1268 42–60
the manufacturers for up-to- D6-32 600 1455–873 39–52
date data and availability D8-22 796 2494–345 38–52
before use. Comparative data
with other hammer types can be
D12 1250 3118 42–60
obtained from manufacturers D15 1500 3867 40–60
and website D12-32 1282 4323–2170 36–52
www.pilehammerspecs.com. D22 2204 5501 42–60
The information is provided to D16-32 1603 5570–2615 36–52
show the range of energy D19-42 1904 5931–2840 37–53
available from modern diesel D19-32 1904 5931–2840 37–53
hammers. Other manufacturers D22-23 2204 6721–3395 38–52
include APE, ICE, HMC, D30 3006 7518–3298 39–60
MITSUBISHI, IHI, FEC, BSP and D25-32 2505 8072–4136 37–52
LINKBELT
D30-23 3006 9160–4670 36–52
D30-32 3006 9686–4903 36–52
D36-23 3608 11 516–5266 37–53
D36-32 3608 11 624–5667 36–53
D44 4318 12 056–5958 37–56
D46-23 4610 14 550–6707 37–53
D46-32 4610 14 852–7242 37–53
D55 5512 16 213–8661 36–47
D62-22 6636 22 865–10 942 36–50
D80-23 8863 31 180–17 487 36–45
D100-13 10 732 41 574–21 818 36–45
D200-42 20 029 69 240–43 652 36–52
Note. The hammers listed above are single-acting. Some manufacturers have introduced
double-acting hammers incorporating a bounce chamber to increase the blow rate.
220 Deep excavations

Table 6.3 Characteristics of Junttan hydraulic hammers. The data listed in this table were obtained from manufacturers’
brochures and websites and are current in March 2003. Reference should be made to the manufacturers for up-to-date data and
availability before use. Comparative data with other vibrator types can be obtained from manufacturers and website
www.pilehammerspecs.com. The information is provided to show the range of hammer weights, energy and strike rate
available from modern hydraulic hammers. Other manufacturers include: IHC, Menck, APE, BSP, Twinwood, Bruce, HPSI,
Pilemer, Nissha, HMC, Dawson

Model Mass of ram Maximum energy Strike rate


(kg) per blow (m kg) (blows per minute)

HHK-3A 3000 – 40–100


HHK-5A 5000 6000 40–100
HHK-7A 7000 8400 40–100
HHK-9A 9000 10 800 40–100
HHK-12A 12 000 14 400 40–100
HHK-14A 14 000 16 800 40–100
HHK-18A 18 000 21 600 40–100
Note. The hammers listed above have provision of assistance to the ram during its downward travel; the rated energy includes
the effect of the motive fluid pressure in addition to the free-fall gravity mass. Other hammers are either single-acting or double-
acting. Hydraulic hammers for off-shore operation have been developed with greater energy capacity and for underwater use.
Manufacturers of such hammers include Menck.

Fig. 6.16. Modern hydraulic


piling hammer mounted on
base machine (courtesy of
Junttan Piling Equipment)
Cofferdam construction 221

Fig. 6.17. High-frequency piling


vibrator

penetration of soft to firmer clays can also be achieved for limited penetration
depths, as may apply in mixed strata. Characteristics of typical vibratory
hammers are listed in Table 6.4.
The detrimental noise and vibratory effect of sheet pile installation in urban
areas has for some years restricted their use. A report previously published by
the British Steel Corporation11 , gives a review of published recommendations
on minimum distances of pile installation to existing structures and services
and both BS 5228 (part 4)10 and the European code of practice on sheet
piles EN 1206312 give advice on noise levels.
Traditionally, acoustic covers have been used to reduce noise from drop
and hydraulic piling hammers. The use of acoustic covers has reduced as
hydraulic presses have remedied both noise and vibratory objections to
sheet piling. A typical hydraulic press, by Giken, is shown in Fig. 6.18.
The hydraulic press installs single sheeters from a clamped reaction to pre-
viously installed piles. The system which is virtually vibration-free and noise-
less, apart from crane and power-pack noise, allows sheet pile installation
immediately adjacent to existing services and allows sheet piling to be installed
without environmental problems in city centres. The presses, which can also
be used to extract sheet piles, are limited to a downwards thrust of the
order of 100 t or so and either lubrication with water introduced through a
metal tube welded to the sheeter or water jetting may be needed to assist
pile installation in stiff clays. Pile penetration is limited to the order of 15 m
or so in stiff clays. A sheet pile of minimum section LX20 is often needed to
resist the force inserted on the single pile. There is some tendency for the
sheet piles to spread in width and to twist as they are installed. Installation
tolerances are somewhat more difficult to control with single-pile installation
222 Deep excavations

Table 6.4 Characteristics of PTC vibratory hammers. The data listed in this table were obtained from manufacturers’
brochures and websites and are current in March 2003. Reference should be made to the manufacturers for up-to-date data and
availability before use. Comparative data with other vibrator types can be obtained from manufacturers and website
www.pilehammerspecs.com. The information is provided to show the range of mass, frequency and power requirements of
modern vibrators. Other manufacturers include: Tramac, HMC/Movax, MKT, Dawson, ICE, HPSI, Muller, APE, Vulcan, PVE,
Delmag/Tunkers, H and M, Foster

Model Power of vibrator Eccentric moment Max. frequency Max. amplified


(kW/hp) (m kg) (Hz/rpm) (mm)

Excavator mounted
1HF1 – 0.7 50/3000 6
3HF3 – 3 45/2700 11
7HF4 – 6.5 38/2300 20
10HF5 – 10 35/2100 15
15H2 – 13 28/1680 19
Standard
7H5 47/63 6.5 33/1980 20
15H1 108/146 15 28/1680 24
25H1A 133/181 23 29/1740 21
30H1A 193/262 30 28/1680 23
50HL 305/414 50 25/1500 31.2
High frequency
7HF3 92/125 6.5 28/2300 14.5
13HF3 124/169 13.5 38/2300 22
15HF3 167/227 15 38/2300 23
23HF3A 184/250 23 38/2300 20
30HF3A 292/397 27 38/2300 22
46HF3 447/607 46 38/2300 13
Variable frequency
10HFV 95/129 0–8 38/2300 10.6
15HFV 140/190 0–15 38/2300 15
15HFV5 200/272 0–15 38/2300 15
17HFV 240/326 0–17 38/2300 11
23HFV 222/302 0–23 38/2300 13

by hydraulic press compared with conventional panel installation by diesel or


hydraulic hammer.
When pile installation by driven hammer becomes difficult and tearing of
the piles and stripped clutches are likely to occur, a strong shoe can be
welded to the end of the sheeter to assist penetration and avoid damage (see
Fig. 6.19). Pressure water jetting can be incorporated into the shoe to further
assist penetration. Pre-boring or pre-augering may also be used along the pile
centre-line to overcome difficult driving conditions.
The depth of embedment of cofferdam sheeting is dependent on the need to
effect a cut-off in an impermeable strata, to reduce risk of boiling and the
quantity of pumping in cohesionless soils, to alleviate risk of base heave
where hard strata underlie soft clay and silts, and to generally reduce sheeting
movements by mobilizing passive resistance of subsoil below formation level.
The height of sheeting driven for tidal works will normally be based on the
highest tide levels estimated during the period of the works, with allowance
for flood risk. The freeboard will take into account wave action caused by
wind and, in navigable rivers, will allow for the wake caused by passing
vessels.
In exposed conditions on the sea shore it is at times economical to use half-
tide cofferdams, with no works being carried out at peak tides, the cofferdam
being pumped out after the tide. As with any cofferdam over water, adequate
Cofferdam construction 223

Fig. 6.18. Giken hydraulic


press for pile installation

Fig. 6.19. Piling shoe for heavy


duty installation of sheet piling
(courtesy Dawson Con Plant
Ltd)

sluice valves to allow discharge of water are essential when overtopping is


considered. In addition, internal bracing must be sufficiently rigid to allow
some reverse head and avoid loosening of strutting. Tie rods and external
walings are sometimes installed above the top frame, not to resist reverse
water heads after overtopping but to give some means of tightening the coffer-
dam after dewatering.
The struts of all cofferdams are often made deliberately short to allow
wedging or paging with timber wedges, preferably in elm, between sheet piles
and the rear face of the walings. This shortening depends on frame depth
and the driving accuracy of the sheet piles, which directly depends on quality
of workmanship and subsoil conditions. The practice of using short lengths
224 Deep excavations

of reinforced concrete to pack the gap between strut ends and the face of
walings should be used cautiously since this cube of concrete can be dislodged
by accidental blows from excavating plant and time is needed for the concrete to
cure before the strutting becomes effective. A well-built and maintained coffer-
dam in tidal conditions will therefore have each pile pan timber paged against
the back of the waling, these wedges being driven tight at each tide, the end-
plated struts remaining square to the face of the waling.
The struts, typically from tubular steel, box piles, universal column or
battened twin joist sections, are designed at spacings to comply with the
overall geometry of the cofferdam. In addition, strut centres must allow
lower frame members to be threaded through them and allow excavating
grabs and excavation plant to pass vertically without striking the frames.
The number of frames will depend on the imposed loading, the strength of
the sheeting and the frame material available. The method of excavation
within the dam will be controlled directly by the vertical height between
frames; efficient excavation equipment such as tracked loading shovels can
only be used with reasonable headroom. Frame levels will also condition
lift heights of concrete pours for permanent works within the cofferdam.
This factor applies particularly where the permanent works are built against
the sheeting and where projecting vertical reinforcement lies below walings.
Bar couplers may be necessary in these circumstances to reduce the difficulties
of excessive or inconvenient reinforcement splice lengths.
The cofferdam frames are supported by steel brackets welded to the sheet-
ing. The use of hanging rods or chains to suspend frames beneath the top
frame, which is supported on brackets, is inadvisable; all frames should be
supported on an adequate number of brackets welded to the sheeting. Long
span struts may require support from king posts which are pre-bored or
pre-driven below formation level.
In recent years, increased use has been made of steel bracing walings and
struts made to standard sizes and available from construction equipment
hire companies. The use of these members is restricted to excavations of
only modest depth, although spans of up to 20 m are available for both
walings and struts and with adaptation are available for bracing shafts of
shallow depth. The struts are usually extended mechanically, although
hydraulic struts are also available for higher loads. Typical applications are
shown in Fig. 6.20. It should be noted that considerable care is needed to
ensure competence of connection between struts and walings in such systems.
The static calculations for the cofferdam structure will take into account
soil and water pressures. However, loads due to construction plant such as
pumps being placed on the frames, the dynamic effect of waves, accidental
collision from vessels and blows caused by excavation grabs and equipment
on struts, must also be considered or avoidance measures taken during the
construction phase.
The effects of scour on cofferdams in fast flowing rivers will be referred to
later. Measures to prevent scour during the construction period include river
bed protection by rock fill, precast concrete blocks and tetrapods, rock-filled
gabions and grouted mattresses. Routine observation of the depth of scour
is essential, particularly where there is risk of under-scour to river bed
protection.
In cofferdams built in waters with a high tidal range the sequence of load
application and reduction causes movement within the sheeting which is
unavoidable and can cause heavy leakage of water into the dam through
the pile clutches. Traditional, practical methods of reducing this problem
included lead-wool caulking, asbestos-string caulking and, more crudely,
the spreading of fly-ash or sawdust on to the water outside the cofferdam as
Cofferdam construction 225

Fig. 6.20. Use of mechanical bracing strut (courtesy of Mabey Ltd)

the tide ebbs. As the dam is pumped out quickly the fine material is taken into
the clutch and helps to seal the leak. More recently, durable elastic sealants
have been applied to the interlock at site or as a complete filler to the interlock
prior to pile despatch from the works.
The construction phase where excavation has been made to a level just
below the level at which a frame is to be built was mentioned earlier. The
risk factor at the lowest frame level depends on the time taken to build the
frame in position, to prop the sheeting and make secure. All materials will
therefore be ready to drop into position and weld and page up tightly as the
last excavation is taken away. Where conditions are suspect, caution may
dictate that this should be done at low tide where possible and also, of
course, in short lengths, taking only sufficient excavation to fix one length of
waling. Where high bending moments in the sheeting are unavoidable it may
be necessary to fix frames, usually after the first frame, below water. Frames
may be slung, fixed and packed by divers prior to pumping out the dam.
226 Deep excavations

Fig. 6.20. Continued (courtesy of Mabey Ltd)

River cofferdams: construction examples


The river cofferdam examples that follow illustrate some of the constructional
principles to follow, and the dangers and faults to avoid.
(a) The Thames Barrier cofferdams were described by Grice and
Hepplewhite13 . The flood defence barrier was built by the municipal
authority, the Greater London Council, to prevent flooding of some
45 square miles of London during critical tides and winds.
Cofferdam construction 227

Fig. 6.21. Thames Barrier:


geological section across the
site13

The geological section across the Thames at the barrier site is shown
in Fig. 6.21, and a general view during construction is shown in
Fig. 6.22. Eleven deep river cofferdams were built between 1974 and
1980 to house the two abutments and nine piers. Cofferdams 4 to 8
are among the largest river cofferdams built in the UK to date. The
seven southern structures were founded on the chalk and the remaining
four northern structures were built in the Thanet sand, a very dense
grey–green silty fine sand. On the south side, the upper surface of the
chalk is about 10 m below datum and dips gently northwards. On the
north bank, the chalk is about 25 m below datum and is overlain with
the Thanet sand. The chalk, which is fissured and jointed with flints,
is classified as upper chalk and is generally unweathered where covered
with sand but is moderately to severely weathered elsewhere.
As shipping access was needed through the works, the south side of
the river was closed first for construction of piers 6 to 9 and the abut-
ment; construction work then moved to the north-side cofferdams.
Fig. 6.22. Thames Barrier: The southern cofferdams were built from Larssen No. 6 grade 50B
general view during cofferdam sheet piles 34 m long, fully driven in pairs by piling hammers ranging
construction13 from the Delmag D44 to the Delmag D62. Final driving sets varied
from 400 to 1200 blows per metre. The construction sequence was to
install the top frame at level þ2.1 OD and then excavate underwater
to 18.00 just below bottom frame level. The excavation was completed
to 23.75 OD, the formation prepared and the 5 m thick concrete plug
placed by tremie. Intermediate frames were installed in cofferdams 6 to
8 after pre-assembly above low water in a position under the top frame.
The dams were then dewatered. Access for the south-side cofferdams is
shown in Fig. 6.23. The part plan and cross-section of the pier 6 coffer-
dam is shown in Fig. 6.24. The walings were fabricated from twin
914  419 mild-steel universal beams and the struts from beams or
tubes. To minimize obstruction to permanent work construction, no
vertical bracing was used. The variable gap between the sheet piling
and the back of the walings was packed by divers using a grout bag
secured on a mild-steel mesh framework hung from the waling. The
bag was then grouted with an expanding grout from a surface mixer.
A short length of steel beam was used where the gap at the rear of
the waling was excessively large. To facilitate removal of the bag at a
later stage, a 25 mm hole was left to allow the placing of a small
explosive charge.
228 Deep excavations

Fig. 6.23. Thames Barrier: river


access to southern cofferdams
(courtesy of Carillion)

Fig. 6.24 (a) Cofferdam frame


at level 5.50, pier 6;
(b) bending moments in sheet
piles after excavations and
dewatering (courtesy of
Carillion)
Cofferdam construction 229

Fig. 6.24 (c) View of south-side


cofferdam showing heavy
leakage (courtesy of Carillion)

The construction of all southern cofferdams was slowed by excava-


tion difficulties in the chalk. The alluvium overlying the chalk was
removed by rope grabs from cranes, and then kelly-mounted augers
and hydraulic grabs were used on the chalk itself. The chalk shoulders
left beneath the wide walings were removed with some extra difficulty
by kelly-mounted grabs, by chisels and hydrojets. Blasting was also
used, with charges of Polar ammon 80% gelignite limited to 250 g of
charge per cubic metre of chalk blown. The charges were placed at
least 0.5 m from any chalk/water or chalk/pile interface, with separate
detonation by Cordtex fuse. After completion the engineers responsible
suggested that a cofferdam slightly larger in plan size may have been
more economical and would have reduced the difficulty of removing
the chalk from under the walings despite the additional excavation
and temporary works.
After excavation of the chalk, air lifts were used to sweep and clean
the chalk surface to final level, the whole finished surface being
inspected and probed by Mackintosh probes from a diving bell.
With the exception of the southern abutment, the concrete plug
weight was insufficient to resist the upward pressure and an extensive
pressure relief system was needed under each base to achieve a factor
of safety against uplift of at least 1.3. Typically the system consisted
of 20 tubes, each 865 mm in diameter, temporarily supported from
the bottom cofferdam frame. After concreting the plug, and prior to
dewatering the cofferdam, the tubes were bored out by auger to 10 m
below formation level and filled with gravel. Piezometers and stand-
pipes were installed and monitored to check actual uplift pressure.
The cofferdams were pumped in stages with progressive checks on the
effectiveness of the pressure relief wells and the integrity of each succes-
sive bracing frame. Sluices in the sheeting allowed for reflooding if this
was deemed necessary. Considerable leakage through the clutches in all
dams was only partly sealed with sawdust and fly-ash. Some split
clutches had to be plated over, and water ingress was sufficient to
warrant hanging heavy plastic sheets from the walings to reduce
nuisance to operatives in the cofferdam (Fig. 6.24(c)).
Additional works were required to the pier 7 cofferdam. One of the
fault lines in the chalk crossed pier 7 diagonally, and several split
clutches were discovered during diving inspections. After the water in
230 Deep excavations

the dam had been pumped out, high readings were obtained for uplift
pressures on the bottom of the plug and a fine crack was found
across the top of the tremied concrete. The dam was reflooded and a
reverse filter installed on the outside of the sheeting, with a grouted
cut-off on the far side of the filter to prevent further ingress of water
from outside the cofferdam. The solution was successful.
The task of placing the concrete plug by tremie to all cofferdams on
the scheme was a major operation. Early trials had established a high
slump concrete mix incorporating Pulverized Fuel Ash as half of the
cementitious content to reduce the heat of hydration. In each cofferdam
pour of 8000 m3 the site-mixed concrete was poured continuously
through tremies handled by tracked cranes. The tremies, 300 mm in
diameter, were arranged on a 7.5 m grid as shown in Fig. 6.25. The
ends of all tremies were submerged in the concrete throughout the
pour. Concrete was delivered to the tremies by 3 m3 Trucrete vehicles
to 125 mm mobile concrete pumps. The pour was brought up in
layers of 2 m maximum thickness, the concrete being built up at one
end of the pour and advanced forward on the chalk subsurface to
scour any silt remaining to the far end of the dam. The major south-
side pours, working night and day, took five days to complete.
The north bank cofferdams required extensive sheet piling through
the Thanet sand. Experience at pier 2 showed a large number of split

(a)

Fig. 6.25. Thames Barrier:


(a) longitudinal section through
cofferdam showing
arrangements for tremie
concreting; (b) view of tremie
works in progress (b)
Cofferdam construction 231

clutches in the Larssen No. 6 sheet piles. Pre-boring was used to


alleviate the hard driving for the sheet piles on cofferdam 1. For the
remaining sheeting at piers 3 to 4, Peine piles made from grade 50
steel, but with clutches of higher grade, were driven into pre-augered
secant bores that had been filled with a low-strength Pulverized Fuel
Ash–cement–bentonite grout. The clutches of each pile were stitch-
welded 10% to prevent slippage during driving the 38.4 m long piles.
Each pair weighed 21 t and the toe of each pair was sculpted with toe
plates to reduce skin friction during driving.
A part plan and cross-section through the pier 4 cofferdam shows the
Peine piles (Fig. 6.26). Their clutches proved very efficient and their
double skins gave an open box which permitted filling or grouting as
an additional sealing measure if needed. The increased strength of the

(a)

Fig. 6.26. Thames Barrier:


(a) general arrangement of
frame to pier 4 cofferdam, part
plan and cross-section;
(b) bending moments in sheet
piles after excavation and
dewatering13 (b)
232 Deep excavations

Fig. 6.27. Thames Barrier;


pier 4 cofferdam, details of
rocker bearing: (a) plan;
(b) vertical section13

Peine piles allowed the use of only one frame composed of walings and
struts from twin 914  419 universal beams in high-yield steel.
The maximum design load on the walings was 150 tonnes per metre
run, with a maximum axial compressive strut load of 1200 t from the
walls. With the risk of cofferdam collapse due to dislodgement of
members of the single frame, higher factors of safety were used in the
frame design.
Figure 6.27 shows the detail of a rocker bearing that was used to dis-
tribute the applied loading from the sheet piling equally to each beam of
the twin waling member, the rocker bearing accommodating the
deflected form of the sheet piles. A secondary waling with header
beams set into the pile heads of 305  305 column section was used
to smooth out pile deformation during re-strutting of the sheet piles
against the permanent works as the main strutting was removed.
For the 9005 section Peine piling, driven using D62 hammers on two
Menck MR 60 rigs working 16 hours per day, the period of driving was
nine weeks for 235 pairs. The average final blow counts were 600 blows
per metre.
With the exception of cofferdam 5, the cofferdams formed from Peine
piles were closed with pre-welded Larssen clutches. The seal was then
enhanced by grouting within the blister. This precaution was adopted
because of the rigidity of the Peine section and the double clutching
incorporated in it.
Cofferdam construction 233

Fig. 6.28. Thames Barrier,


north cofferdam construction:
(a) plan of cofferdam showing
relief well positions;
(b) vertical section through
relief well; (c) details of
piezometer and standpipe
installation

Excavation of the northern cofferdams was made easier by the single


frame used to support the Peine piles. Pier 3 was founded on Thanet
sand, piers 4 and 5 on chalk. In the Thanet sand, excavation was by
augering and airlifting. Once into the chalk, augers were used to
break up the chalk which was then removed by 500 to 660 mm dia.
air lifts aided by a heavy reverse circulation rig. Finally, the chalk sur-
face was cleaned by air lift sweeps. Figure 6.28 shows the plan of the
north cofferdam and details of the construction pressure relief system.
(b) River Hull Barrier. Flood barriers built across other English rivers in
the late 1970s also serve to illustrate construction techniques in river
cofferdams. The tidal surge barrier built across the River Hull and com-
pleted in 1979 required deep cofferdams for both the monoliths on
which the towers were founded and the cills. The works were described
by Fleming et al.14 ; Fig. 6.29 shows the cofferdams during construction.
The feature of the works (unlike the Thames Barrier) was that the
length of the piles was large in relation to the plan area of the cofferdam.
The foundation solution for the barrier towers was to found them on
mass concrete monoliths within dense glacial sand and gravel approxi-
mately 20 m below the river bed level. To avoid disturbance to the
subsoil below formation, the contract specified that excavation for
the monoliths should be underwater and that the water level in the
cofferdam should be maintained at least 0.5 m above the tidal level out-
side. Mass concrete to a depth of at least 14 m had to be placed by
tremie to form the monolith before piping was guaranteed not to
occur. A cofferdam sheet piled in two stages was originally considered,
but a single-stage cofferdam was constructed with Larssen No. 6
sheeters supported by five bracing frames. The 34 m long sheet piles
were the longest rolled by the British Steel Corporation at that time.
234 Deep excavations

Fig. 6.29. River Hull tidal surge


barrier: (a) pitching 34 m long
Larssen No. 6 piles to monolith
cofferdam; (b) completed cill
cofferdam ðaÞ

ðbÞ
Cofferdam construction 235

Fig. 6.29. (c) Sheeting and


bracing to monolith cofferdam
(courtesy of Dawson) (c)

The top three frames were to be fixed above water and the lower two
below water, by divers. Due to the small radius of the monolith coffer-
dam walls it was necessary to reduce the friction in the interlocks during
pitching and driving. This was done by crimping every in-pan pile so
that interlocks were in line. In addition, it was necessary to ensure
that piles were pitched and driven vertically. This was ensured by pro-
viding substantial temporary supports at close centres, by instrument
checks for verticality and by temporary welding of each in-pan pile to
the supports. Driving was carried out in small stages with the toes of
the sheeters all at approximately the same depth below ground level.
Suspended drop hammers of ram weight 5 t and 7.5 t were used to
begin driving, which was completed by a Delmag D46. Excavation
was by rope grab, and the two lower frames were suspended from
frame 3 before the dam was flooded. The packing from the two lower
frames to the sheeters was of pieces of universal column, measured indi-
vidually by the divers and fitted with a single bolted clamp.
Air lifting was used for the final excavation. During the bottoming-
up of the west monolith cofferdam a blow of approximately 80 m3 of
loose sand occurred through a split in the intermittent welding of one
of the corner piles. The split, about 50 mm wide, started about 2.5 m
above formation level. With the risk of further loss of ground if the
sand had been removed, it was left in place and later consolidated by
grouting after the tremie concrete had been placed. During the final
excavation work for the east monolith cofferdam a digging grab
impact caused the lower three frames to drop to the bottom of the
excavation. The lowest frame was refixed in an intermediate position
and there was no inward movement of the sheet piles. Due to the
head of water maintained inside the cofferdam, a differential of about
1 m at all times, the net loading in the frames was very small and the
two dislodged frames were not refixed.
The monolith was concreted underwater by tremie to a depth of
about 16 m. A total of 1550 m3 of ready mix concrete, retarded for
7 h, was placed through four tremies in 34 h continuous pours for
each monolith.
236 Deep excavations

Fig. 6.30. View of Fobbing


Horse Barrier job site (courtesy
of AMEC)

(a)

Fig. 6.31. Fobbing Horse


Barrier: (a) south-west pier
cofferdam; (b) north-east pier
cofferdam (courtesy of AMEC) (b)
Cofferdam construction 237

(c) The Fobbing Horse Barrier. Figure 6.30 shows an aerial view of works
on the Fobbing Horse Barrier in Essex, part of the Thames Flood
Defence Works completed in the early 1980s. The bracing details of
each cofferdam are shown in Fig. 6.31. The south-west pier is strutted
by diagonal steel tube bracing, while the north-east pier uses the
alternative method of cross-strutting with box pile sections. Both
steel tubes and box pile sections are efficient strut members but suffer
the disadvantage of low flexural strength compared with the universal
beam and column section.
(d ) Forth Road Bridge. Successive stages of cofferdam works to a suspen-
sion bridge tower construction are shown in Fig. 6.32. The corner
braces used to provide horizontal support to each of the three frames
are augmented by one cross-strut to each frame on the longer side of
the cofferdam. Note the increased structural stiffness provided by
relatively light diagonal steel bracing between each of the frames on

(a)

Fig. 6.32. Forth Road Bridge:


(a) south-side tower cofferdam
construction (note diagonal
bracing); (b) tower base
construction within cofferdam
(courtesy of AMEC) (b)
238 Deep excavations

Fig. 6.33. Kingsferry Bridge,


Isle of Sheppey; (a) view of site
(courtesy of AMEC)1 (a)

the longer side, and between the puncheon and upper and lower frames
across the cofferdam.
(e) Kingsferry Bridge. The layout of cross-bracing and diagonal bracing to
river cofferdams for a bridge construction at the Isle of Sheppey in Kent
is shown in Fig. 6.33. Diagonal braces are used, it will be noted, to give
maximum space for the passage of excavation grabs and equipment and
to minimize expense in cross-strutting.

Braced cofferdams on land


Cofferdams constructed on land and braced for support include excavation
for generating plant, pumping stations and underground facilities for various
industrial purposes. The problems that arise with such cofferdams and their
method of solution are sometimes common with other deep excavations on
land for building basements and cut-and-cover construction for transporta-
tion systems. Nevertheless, because building basements and tunnels built in-
trench possess some special features they are considered separately in
Chapters 8 and 9. The purpose of this section is to describe constructional
features of braced land cofferdams, typically those used in pumping stations
and similar industrial subsurface structures.
The five main differences between constructional features of land and
marine braced cofferdams are as follows.
(a) Land cofferdams require much less work to provide temporary access
for plant, labour and materials than many jobs over water.
(b) While the choice of sheeting for schemes over water is often limited to
steel sheet piling, a wider range of sheeting and walling methods are
available for work on land. The location, depth of excavation, depth
to the groundwater table, subsoil conditions and depth to bedrock, if
present, will all help to determine the most economical sheeting
system. In addition to steel sheet piles, the choice for work on land
includes diaphragm walling by in situ reinforced and post-tensioned
concrete and precast methods, contiguous piles, secant pile walling
using hard–hard, hard–firm and hard–soft methods, and temporary
Cofferdam construction 239

Fig. 6.33. (b) Cross-sections


through river pier cofferdam
and construction sequence1 (b)
240 Deep excavations

soil support from soldier walls (with soldier piles and horizontal lagging
timbers or concrete skin walls).
(c) The means of bracing land cofferdams includes ground and rock
anchors in addition to internal bracing from steel struts and walings.
(d ) The proximity of existing structures to the excavation requires greater
emphasis on problems of soil deformation around the excavation
periphery for cofferdams built on land.
(e) The loads applied to land cofferdams are not likely to include tidally
varying groundwater but allowance must be made for superimposed
loading from other structures, and from traffic and site plant around
the excavation periphery. In addition, in less temperate lands, the
effect of freezing soils on soil pressures and bracing loads must be
considered.

Choice of sheeting
Whereas methods such as the use of king posts provide temporary soil support
to allow construction of the permanent wall, secant or contiguous piling and
diaphragm walling allow the temporary walling to be incorporated into the
permanent wall structure. The saving in construction time and cost has led
to the increased use of these latter systems in land cofferdams.
In other instances, the experience of specialist contractors with diaphragm
wall construction has encouraged the introduction of composite sheeting
methods. Although only used for temporary support, sheet piles pitched
rather than driven into slurry trench excavations have proved attractive
cost-wise and avoid the environmental problems of noise and vibration to
adjoining structures. It is often necessary to place tremie concrete below
formation level to provide sufficient passive resistance to sheet piles pitched
in this way.
An example of king post walling is shown in Fig. 6.34. The method, for a
pumping station in Jubail, Saudi Arabia, found favour because of a low
groundwater table, an essential prerequisite for king post walls unless
Fig. 6.34. Berlin wall method
using anchors to scale pit dewatering is to be employed, and possibly due to lack of competitiveness
building, depth to formation by specialist diaphragm walling firms for a relatively small job which would
level 14 m, groundwater table at be distant from plant and resources.
a depth of 3 m, steel plant, Figure 6.35 shows the use of sheet piling to the substructure of an ash pit at
Jubail, Saudi Arabia (courtesy West Thurrock Generating Station, UK15 . The works, in difficult ground
of Bauer)
conditions, were sited adjacent to a reinforced concrete chimney already
partly built. The ash pit, founded on the gravel some 16 m below ground
level, was built in cofferdam in preference to sinking a caisson because of
the risk of undermining the foundations to the chimney if a soil blow were
to develop under the cutting edge. The soft marsh clay was comparatively
impermeable and had a shear strength of about 30 kN/m2 . The underlying
gravel contained an artesian head of groundwater which almost reached
ground level and rested in turn on fissured chalk which also contained ground-
water. Excavation in the dry would have been very difficult and it was decided
to excavate to third frame level only in the dry and take out the remaining
6.4 m depth under water prior to concreting a plug between levels 8.54
and 13.72. Pressure relief wells were also installed to relieve the artesian
head in the gravel. These wells consisted of 670 mm dia. bores filled with
gravel on a 4.6 m grid, and were installed before the excavation started.
After excavation down to just below second frame level, with top and
second frame fixed, sand and silt began to flow from the relief wells which
had previously discharged satisfactorily. The cofferdam was immediately
flooded and six deep wells with submersible pumps were installed outside
the cofferdam and 6.5 m below the clay. The cofferdam was pumped out
Cofferdam construction 241

(a)

Fig. 6.35. West Thurrock


Generating Station: (a) plan and
longitudinal section of ash pit
disposal cofferdam showing
deep wells, sand drains and
anti-flotation plug;
(b) transverse section of
cofferdam15 (b)

and the third frame fixed in position. The cofferdam was then reflooded to
level 1.82 and the excavation completed underwater with the artesian
head relieved by the deep wells. The cofferdam was made rigid by welding
sheet piles, walings and struts. By taking these measures, diagonal bracing
was not needed. The concrete plug was formed by grouting a gravel layer
5.18 m thick underwater using the Intrusion Prepakt method, and after
completion the cofferdam was pumped dry. A 380 mm layer of blinding
concrete and a 1.2 m reinforced concrete floor slab, both anchored into the
anti-flotation plug, completed the ash pit bottom.
242 Deep excavations

Fig. 6.36. (a) Section and (b)


plan of cofferdam and framing,
Post Office underground
station, London16

This cofferdam was built just before the introduction of diaphragm walling
and secant piling into the UK: in the soil and groundwater conditions that
existed at the West Thurrock site it is doubtful whether these alternative
methods would have shown any advantage over sheet piling had they been
available. However, it is likely that jet-grouting methods would have been
used in lieu of the Intrusion Prepakt system.
Measures to reduce settlement around land cofferdams deserve discussion.
The extent of such deformation in a stable cofferdam depends on soil bracing
and sheeting stiffness and the vertical spacing of bracing frames. Diaphragm
wall and reinforced concrete pile sections are stiffer than conventional sheet
pile sections and their use is likely to minimize soil deformation adjacent to
the excavation. Further precautions may prove necessary, however, and
these include installation of frames or other support at frequent vertical
centres, pre-loading of struts, use of pre-stressed soil and rock anchors, and
the use of jet grouting and pin piling or mix-in-place piles to stiffen artificially
the soil and increase passive resistance to the sheeting below formation level.
The use of flat jacks to stress the frames of a land cofferdam to reduce soil
movement and the resulting subsidence was reported by Collingridge and
Cofferdam construction 243

Fig. 6.36. Continued

Tuckwell16 . The excavation, in London, was for a new subsurface station for
the Post Office. It was 61 m long, 29 m wide and 22.3 m below ground level. At
the time it was one of the largest excavations made in Central London.
The contract specified earth pressure at rest in the temporary works design.
The ground conditions, typical of London, were made ground and Thames
gravel overlying London clay. The specified values of K0 were 0.75 in the
clay and 0.5 in the fill and gravel. To exclude groundwater in the gravel and
provide continuous support for all soil above the London clay, a peripheral
sheet pile wall was specified to be driven from below basement level of existing
buildings to a penetration of 2.5 m into the clay.
The method of excavation support is illustrated in Fig. 6.36. Steel H soldier
piles 25 m long were driven to the batter of the outer face of the permanent
wall at 1.6 m centres, achieving penetration 4.6 m below formation level of
the permanent structure. An intermediate, 0.23 m thick reinforced concrete
wall spanned the soldier piles. Four horizontal frames and a top raking
frame were used to support the H sections, needle joists being provided
between the soldier piles and the walings. Provision was made for jacking
each needle joist. As each frame level was reached during bulk excavation,
the steelwork was assembled and stressed to a load calculated to retain the
subsoil in a state of rest. The aim of needle joists between the main struts
244 Deep excavations

Fig. 6.37. System for pre-


loading cofferdam frames, Post
Office underground station,
London16

(Fig. 6.37) was to allow a more uniform distribution of pressure to the soil
than the application of load to the sheeting at the strut positions. Freyssinet
hydraulic flat jacks, with small travel and high capacity, were linked by
hydraulic connection and pressure was applied simultaneously in four
operations per frame. One operation used 16 jacks at one end to stress the
frame longitudinally, and three operations using 12 jacks on each side were
used to stress the frame transversely. The two jack sizes used were 27 cm
dia. rated at 64 t and 35 cm rated at 114 t. The total jack load was applied
in increments of 25%, and a complete jacking operation was completed in
half a day. Allowance for temperature stresses avoided overstressing in hot
weather, jack load being reduced by 10% for every 5.58C above 218C at the
time of stressing.
Collingridge and Tuckwell16 concluded that pre-loading of steel frames
with calculated jack loads proved a satisfactory method of preventing
ground movements, the uniform transfer of strut load to the soil being very
important.
The use of flat jacks to pre-load frames in an attempt to avoid settlement
damage has reduced to some extent since the development of ground anchors,
which also restrain soil movements. Nevertheless, easement arrangements
for anchor installation below neighbouring land and highways frequently
experience problems, and it is in these particular sites that pre-loading
methods still find application.
The use of pre-tensioned anchors through vertical sheeting or walling
to reduce movements caused by large excavations is only effective where
the anchorage zone lies outside the soil movement zone associated with
the excavation. Long-term soil movement due to consolidation within the
anchorage zone limits the application of soil anchors for permanent
anchorage to clay soils. Tomlinson17 has reported a summary of observations
of maximum horizontal movement of soil support for excavations in
normally-consolidated, over-consolidated clays and gravels. These results
Cofferdam construction 245

Table 6.5 Observed values of maximum horizontal deflection of sheeting to excavations on land17

Soil type Wall type Number Maximum horizontal Range of


in sample deflection/excavation excavation
depth (%) depth (m)

Range Average

Soft to firm normally- Anchored diaphragm wall 3 0.08–0.58 0.30 9 to 24


consolidated clays Strutted diaphragm wall/secant 4
pile wall
Strutted sheet pile, soldier pile 5
and concrete infill
Stiff to hard over- Anchored diaphragm wall 2 0.06–0.30 0.16 10 to 30
consolidated clays Strutted diaphragm wall/secant 6
pile wall
Strutted sheet pile, soldier pile 1
and concrete infill
Sands and gravels Anchored diaphragm wall 2 0.04–0.46 0.19 7 to 20
Strutted diaphragm wall, secant 5
pile wall
Strutted sheet pile, soldier pile 1
and concrete infill
Anchored sheet pile, soldier pile 4
with concrete infill or timbered

are summarized in Table 6.5. The prediction of vertical deformation adjacent


to excavations is reviewed in Chapter 11.
An example of pumping station construction using diaphragm walls of
varied plan shape is shown in Fig. 6.38. The site, at Redcar, UK, was close
to an existing quay wall built in diaphragm wall construction. The ground-
water level was high and the subsoil conditions were fine sands overlain by
filling composed of broken blast furnace slag. Groundwater was lowered by
deep wells in the fine sand located within the cofferdam, and despite large
overbreak to the diaphragm walling within the fill material, the works were
completed successfully. Figure 6.39 shows the individual panel components
of the construction; the action of the heavily reinforced capping beams in
limiting movement by any out-of-balance forces on panel units should be
noted.

Double-skin cofferdams Double-skin cofferdams built from steel sheet piling and enclosing soil or rock
fill divide into two categories.
(a) Double-wall cofferdams. These consist of two parallel lines of sheet
piling tied by steel rods at one or more levels between external walings
and a fill material, such as sand, gravel, hardcore or broken rock,
between the sheet piles. The filling requires adequate drainage which
may be maintained by internal sluices or deep wells below the fill.
Berms may be used to reduce lateral movements in the pile–fill soil
structure and increase stability. If bedrock occurs at shallow depth
and the sheeters cannot be driven into it, this type of cofferdam may
not prove adequate to resist seepage nor will it be economical because
of high moments in the sheet piles.
(b) Cellular cofferdams. These are enclosures, often circular, made from
straight web steel sheet piling to contain a filling of sand, gravel and
246 Deep excavations

Fig. 6.38. View of construction of diaphragm wall to pump house, steelworks, Redcar, UK (courtesy of Lilley)

Fig. 6.39. Redcar pumping


station: panel layout to 1 m thick
diaphragm walls spanning
vertically between base slab
and capping beam without
bracing
Cofferdam construction 247

Fig. 6.40. Cellular cofferdam


with berm to cut off canal water
from a large lock excavation,
Canal Albert, Belgium
(courtesy of Arbed)

broken rock. The straight piles, which have a high interlock strength,
resist pressure from the filling and contained water by circumferential
tension in the piling. Adequate drainage of the fill is needed to reduce
pressure on the sheeting and to avoid reduction of the shear strength
of the filling. The cellular cofferdams are economical for greater
water depths, large retained heights, long structures, and where bracing
and anchoring is not possible. The additional requirement in sheet pile
area is compensated for in the length of piling and pile section and the
absence of walings and anchors. The fill quantities may outweigh
advantages in pile tonnage, however. An example of the typical use
of a cellular cofferdam, to exclude river water from a large excavation,
is shown in Fig. 6.40. The works, for a new lock at Kwaasdmechelen on
the Canal Albert in Belgium, were completed in the early 1970s.
These two types of double-skin cofferdams are now examined in further detail.

Double-wall cofferdams
Examination of the possible modes of failure of double-wall cofferdams serves
to illustrate design and construction precautions.
(a) Sheet pile flexure; tie rod breakage; passive soil failure at foot of sheet
piles. To avoid overstress in bending or excessive flexural deformation
of the sheeters, the pressure from the contained fill and groundwater
should be as low as possible. In particular, it is essential that the fill
should be drained adequately at all times during the cofferdam life.
The drainage system may consist of weep holes or sluices in the inner
line of sheet piles, possibly connected to an internal filter layer of
gravel within the filling. In addition, deep wells through the filling
may be used; these wells, with vertical filters extending upwards into
the filling, also serve to reduce the exit gradient of water seepage
under the dam. Efficient maintenance of the drainage system is vital,
and new weep holes must be cut where migration of fine soil within
the fill causes blockage.
The extent of pile penetration into the subsoil or rock beneath the
dam and the location of the lowest tie rods determine the passive
248 Deep excavations

Fig. 6.41. Chola cofferdam,


India: (a) plan and assumed
support conditions; (b) basis of
design, stability considered by
shear induced on a vertical
plane and assuming inner and
outer walls act as a retaining
wall and anchor wall18

forces mobilized in front of the sheeting. The lowest ties should be


installed as low as possible, at low tide level in tidal rivers. Pile penetra-
tion will depend on subsoil conditions. Penetration into rock may not
be possible; if this can only be achieved by heavy driving, split clutches
may occur which in turn could lead to seepage problems and loss of
ground. To alleviate problems of piling into rock it may be possible
to excavate a trench into rockhead at low tide followed by concreting
the pile toe into the trench. In such conditions, construction costs
increase and the use of double-wall cofferdams may be uneconomical.
An example of a cofferdam founded on rock at a power station site in
India was given by Ellis18 . The cofferdam, shown in section in Fig. 6.41,
was built to allow a dry, unrestricted area for the construction of a
cooling water pumphouse at Chola in Southern India. Although a
feature of the work was the small penetration of the piles into bedrock,
it should be noted that the second row of ties was at a relatively low
elevation. Note also the plan shape of the dam, which was such as to
buttress the structure substantially at each end, reducing deformation
within the dam and stresses in sheet piles and ties.
It is sometimes worth incorporating anchor cells within long lengths
of double-wall cofferdams. The cells contribute to construction sequen-
cing but, most importantly, will confine damage to an isolated length of
dam should any failure or deformation occur, easing remedial work. A
typical arrangement is shown in Fig. 6.42. It may be noted that the cell
acts as a strongpoint in the cofferdam and reduces deformation and
stresses in piles and soils.
(b) Sliding. The total horizontal thrust from river water and ground
pressure on the outside of the dam will be resisted by passive resistance
Cofferdam construction 249

Fig. 6.42. Typical detail in plan


of strongpoint in a double-
walled cofferdam27

of soil below formation level inside the dam and shearing resistance
mobilized under the cofferdam due to the weight of fill on cohesionless
subsoil or due to the cohesive strength of clay subsoil. The condition is
unlikely to dictate cofferdam geometry if pile embedment is sufficient to
counter all other modes of failure.
(c) Shape deformation. The risk of the rectangular cross-section of the
cofferdam adopting a lozenge shape will only be reduced by the follow-
ing measures:
(i) cofferdam fill of adequate uniform quality placed to ensure resist-
ance to tilting and to shearing forces mobilized within the filling
(ii) adequate drainage within the filling, to ensure maximum shear
strength of fill at all times
(iii) adequate penetration of internal and external lines of sheet piles to
mobilize reaction to shearing resistance between fill and face of
sheet piles.
The height-to-width ratio necessary to avoid tilt may be low, of the
order of 1:1. In addition, substantial soil or rock berms may be neces-
sary to avoid unacceptable deformation at the top of the dam. Figure
6.43 shows the cross-section of the double-wall cofferdam used
during construction of the outer entrance to Gallions Lock on the
Thames in London. It is interesting that despite the extent of the
berms and the penetration of the sheeters into the chalk, horizontal
deformation of the cofferdam crest at high water reached 0.36 m.
Packshaw1 suggested that the deflection of the cofferdam may have
been aggravated by two factors: firstly, the hard-core filling, deposited
through water, may have been rather loose; secondly, the rather

Fig. 6.43. Gallions Lock,


double-wall cofferdam at outer
entrance showing dimensions
in berms, piling and filling1
250 Deep excavations

Fig. 6.44. Cross-section of the


entrance cofferdam,
Immingham dry dock19

compressible chalk on which it was built may have contributed to the


movement by consolidation of the chalk. The row of deep wells into
the chalk should be noted, reducing the piezometric head in the filling
and the exit gradient of seepage water under the dam.
Chalk was also used as filling to a double-wall cofferdam during dry
construction at Immingham, UK19 . The cross-section of the dam is
shown in Fig. 6.44. There was concern at design stage of the extent of
maximum friction which could be mobilized between the chalk fill
and the sheet pile face, so tests were made in a large shear box, with
a rusted sample of steel in the lower part of the shear box and chalk
in varying compaction states in the upper half. A design value for
angle of wall friction of 468 was used as a result of the tests.
A cofferdam built across the entrance to St Katherine’s Lock in
London during replacement of the lock cill had a width-to-height
ratio of 0.81. The cross-section and plan of the cofferdam are shown
in Fig. 6.45. The dam, founded on Thames gravel, was drained by
sluices in the inner line of piling and by two deep wells with submersible
pumps in the gravel. The use of puddle clay seals to achieve watertight-
ness against the existing jambs of the lock entrance is a rather outdated
method, but proved successful. The seals provided some support to the
dam which had very limited length, but deflection at the top of the
piling under high water reached 0.3 m.
(d ) Piping. Like any other cofferdam retaining water or groundwater,
double-walled cofferdams founded on granular subsoil are at risk
from piping failure by flow beneath the structure into the dewatered
area. To avoid piping, the emergent hydraulic gradient can be reduced
by adequate embedment of the sheet pile walls; stabilizing berms also
increase the length of the drainage path to reduce this gradient.
(e) Bearing capacity failure. In very soft clays at the founding level of the
cofferdam, the risk of bearing failure may be such to require removal
of the clay between the piling to a level where better ground occurs.
( f ) River bed scour. The risk of scour to the outer sheet piles in swiftly
flowing rivers, or where flow is constricted, must be examined in the
same way as for any other river cofferdam.
The construction of three double-wall cofferdams at Alton, IL, was described
by White and Prentis.20 At the site, 25 miles north of St Louis, the Mississippi
River is about half a mile wide and almost 10 m deep, with a current of just
Cofferdam construction 251

Fig. 6.45. Double wall entrance


cofferdam at St Katherine’s
Lock, London: (a) plan;
(b) vertical cross-section
252 Deep excavations

Fig. 6.46. Detail of streamlining


fin, dam 26, Mississippi River,
Alton, IL20

over 2 m/s. Model tests were used to design streamlining lead-in piling to the
cofferdams to reduce scour. The streamlining fin to cofferdam 1 is shown in
Fig. 6.46. The 18.3 m long steel sheeting piles were pitched and driven from
a barge to a timber framework suspended from timber piles previously
driven. Added protection was obtained by guying the streamlining fin to the
cofferdam and by dumping Riprap at its upstream end. The works were
successful in preventing scouring and erosion of the cofferdams without
causing harm; sand and silt were deposited along the entire length of the
river leg. The cross-section of the double-wall cofferdam is shown in Fig. 6.47.

Cellular cofferdams
Cellular cofferdams are constructed of flat sheet pile sections with high inter-
lock tensile strength. They offer the advantage that they can be designed as
stable gravity structures even where embedment is difficult, for example
where rock occurs at founding level or the subsoil prevents pile penetration.
In plan, cellular cofferdams are typically circular, diaphragm or clover leaf
types. Circular types may also be joined together. These four plan forms are
shown in Fig. 6.48.
Circular cells have the advantage that they can be individually built and
filled; the smaller linking areas are built later. The width of circular structures
increases in proportion to their height, and as the diameter increases the inter-
lock tension also increases. When the diameter becomes excessive and the
interlock tension exceeds permissible limits, flat cells may be used. Individual
flat cells are not stable on their own, and unless special measures are taken,
cofferdams of this type must be filled in stages. The use of intermediate strong-
points is desirable in long lengths of cellular cofferdams built with flat cells,
particularly when there is risk of rupture or storm damage. Unless isolated
strongpoints are provided, failure in an individual cell can lead to the progres-
sive collapse of several neighbouring cells.
Heights and loadings being equal, flat cell cofferdams require greater weight
of steel per linear metre than circular cofferdams. Comparisons based on total
pile tonnages can mislead, however, as the total cost of cofferdam construc-
tion greatly depends on the cost of fill as placed. While flat cell construction
may require larger pile tonnages than circular cells, flat cells are probably
Cofferdam construction 253

Fig. 6.47. Double-wall


cofferdam on the Mississippi
River, Alton, IL: plan and
cross-section20
254 Deep excavations

easier and quicker to build and use fewer expensive special connecting piles.
The flat cell type also has the advantage that the effective width may be
increased without increasing tension in the interlocks.
The earliest cellular cofferdam was built in the early 1900s at Black Rock
Harbour, Buffalo, NY. Each of the 27 cells was just over 9 m square in
plan, with a similar unsupported height. As all the walls were straight, large
deformations were expected, but in one cell an excessive bulge of more than
1 m occurred between cross-walls, although inward movement of the coffer-
dam crest did not exceed 25 mm. In 1915–1916, the Black Rock Harbour
cofferdam was repeated in a dam at Troy, NY. This time the straight outer
walls were replaced with curved walls between the cross-walls to form the
flat cell type cofferdam.
In 1910, a cellular cofferdam with 20 circular cells was used to raise the
battleship Maine from the bottom of Havana Harbour in Cuba. This coffer-
dam, the first to use circular cells, was built on soft silts and mud overlying
medium soft clay; the cell filling was clay. Inward deflections began to
occur during pumping out of the cells and it was necessary to build a berm
on the inside of the structure. During final stages of dewatering it became
Fig. 6.48. Types of cellular
cofferdam plan forms:
necessary to strut the cofferdam from the ship’s hull. The cofferdam was,
(a) circular cells connected by nevertheless, a success and led to the increased use of the cellular cofferdam
arcs; (b) semi-circular cells with and steel sheet piling.
straight diaphragm cross walls; Terzaghi21 stated that the design of cellular cofferdams based on a founda-
(c) clover leaf cells from four tion other than rock requires more judgement than the design of a double-wall
circular arcs of sheeting fixed
on two transverse walls cofferdam with a broad inner berm on a similar base. It may also be said that
perpendicular to each other and the construction requires more judgement and experience.
connected by small arcs; Cellular dams may fail in the same way as double-wall types: by sliding or
(d) circular arcs joined to each tilting, by failure of the base, by piping failure or by scour. In addition, they
other may fail by breakage of the interlocks and bursting of the cells. This latter
cause, the vulnerability of the interlocks, is the major failure risk, and many
of the failures by bursting which have occurred at the time of filling the
cells (or immediately afterwards) have been blamed on driving out of lock.
Considerable care and diligence must be applied in handling piles and in pitch-
ing and driving them to avoid ruptured interlocks.
Before any installation begins, pile dimensional tolerance, pile straightness,
quality of fabrication of special piles, cleanliness of interlocks and any possi-
ble flaws in steel quality must be thoroughly checked. Careful inspection of
welding quality in the fabrication of special piles is essential. These inspections
should be made on secondhand piles in particular, but tolerance and quality
inspection is always necessary, even on new deliveries. A surplus number of
piles must be ordered as a contingency to allow for reject piles if the progress
of the works is not to suffer.
One of the most important checks is the gauging of all interlocks. Typical
permissible interlock tolerances are summarized in Fig. 6.49. The need for
interlocks as delivered to comply strictly with tolerances is self-evident
when the large proportion of the mass of the pile section concentrated at
the interlock is examined in cross-section. Figure 6.50(a) shows sections of
flat web sheet piles manufactured by Arbed with high-strength interlocks.
The strict dimensional tolerance necessary to achieve three-point interlock
contact is obvious. Figure 6.50(b) shows alternative single- and triple-point
interlocks previously available.
The extent of pre-excavation before cofferdam construction begins depends
largely on depth of site overburden. The objective of prior excavation is two-
fold: to remove shallow obstructions and to minimize the penetration of the
driven pile. Where shallow overburden exists, a site strip 1 m deep will be
sufficient to remove tree stumps and boulders. If the overburden is deep it
Cofferdam construction 255

Fig. 6.49. Typical permitted


interlock tolerances27

may be advisable to take out 6 to 7 m to reduce driving depth. Where bedrock


outcrops it is usual to place 2 m of free-draining fill to provide a toe for the
piles during pitching and driving. If overburden consists of very soft silts or
clays or soils containing many cobbles and boulders, it must be removed.
An internal template must be used to set the sheet piles to a cellular coffer-
dam. The templates are usually made with two rings and supported by at least
four bearing piles. The template is usually made 200 mm smaller in diameter
than the net driving line to allow the piles to rotate slightly and adjust to the
correct arc during pitching. Figure 6.51 shows piles pitched to a template.
256 Deep excavations

Fig. 6.50. Flat pile sections:


(a) modern sections with
high-strength interlocks;
(b) alternative sections
previously available showing
single-point (left) and
three-point contact (right) at
(b)
interlock (courtesy of Arbed)

To pitch or set the sheet piles to the template it is usual to use the four
special junction piles as ‘key piles’. With large cells, and in strong current
conditions, additional key piles may be needed to stiffen the sheeting during
pitching and subsequent driving. These key piles consist of a straight web
pile with a steel joist section welded to the inside of the web. The sheet piles
are pitched working away from the key piles, blocking off alternate piles
from the template. Guy lines to the tops of evenly spaced piles control
verticality, especially in windy conditions and strong currents. Sheeting
cannot be accurately pitched in flows faster than 1.2 m/s unless current
deflector bulkheads are used, cantilevered from the adjacent completed cell.
Closures during pitching should be made mid-way between the key piles;
sheeting is then picked up, several piles together, and ‘shaken out’, especially
Cofferdam construction 257

near the closures. The purpose of shaking out is to ensure that the interlocks
run freely and allow some rotation of individual piles to give a smooth arc
against the template.
Sheeters should be driven in pairs, ideally with a hammer of energy 12 to
20 kNm. Larger hammers should only be used carefully to avoid split
interlocks, especially on long piles. Piles are best driven in increments, the
maximum increment of one pile compared with its neighbour not exceeding
2 m. Jetting methods can be used to good effect in sands but should be
used simultaneously both inside and outside the cell to maintain verticality.
Vibratory hammers are efficient in granular soils.
Where piles have to be spliced because of their length, this is done by driving
the bottom section to full penetration and burning a staggered splice line, the
stagger being 1.5 to 2.0 m between adjacent sheeters. Driving flat web sheet
piles longer than 15 m in one piece is difficult. Where piles are being driven
to achieve a cut-off into a sloping rockhead, final penetration should be
Fig. 6.51. Cellular cofferdam made on single piles to reduce the risk of ‘windows’ under the piles.
construction, flat sheet piles After all piles have been driven and released from the template, filling may
pitched and driven to a template
(courtesy of Arbed)
begin: cell filling should be granular, free-draining material with a reasonable
proportion of fines (say, a maximum of 15% passing a 100 sieve and a max-
imum of 5% passing the 200 size). In large cells, small boulders up to 300 mm
would be acceptable. The filling, which in circular cells should be made from
the centre of the cell to avoid deformations, may be placed hydraulically, from
grabs, by conveyor or by end dumping from trucks. Flat cells are best filled by
grab or dragline, and to avoid distortion the differential fill height between
adjacent cells should not exceed 1.5 m.
If cellular cofferdams are likely to be overtopped by high tides or flood, the
filling must be protected on top. A concrete cap 200 to 300 mm thick is often
provided, and this also protects the fill from occasional high waves.
It is essential to provide sufficient flood gates to allow drainage of the
enclosed area in the event of flooding. After filling and during dewatering,
weep holes should be burnt through the inner sheeting to allow efficient
drainage of the fill: 25 mm dia. holes at 2 m centres in every fifth web regularly
rodded will suffice. The rate of pumping out should be regulated by the rate of
draining the cell fill; in large dams a maximum rate of 1.5 m per 24 hours
should apply.
Cell deformation will have occurred during cell filling operations, the cells
barrelling at a distance of two-thirds of the cell height from the crest as the
slack in the interlocks is taken up. As the cofferdam is pumped dry there is
further settlement of the cell fill, and with further flooding and dewatering
the total settlement in high cells may reach 150 mm. Horizontal deformation
of up to 1% of the unsupported height can be expected at the cofferdam crest
after dewatering.
Where a number of circular cells on a curved plan form span-wide openings,
the end cells transfer horizontal thrust to an existing wall or natural abutment.
Figure 6.52 shows the plan and section of the circular cell cofferdam used on
the Bangor–Brewer Bridge, ME, USA. Horizontal bracing was used to strut
the tops of four key cells in the structure.

Gravity type cofferdams The earliest, and simplest form of cofferdam over water was the earth or rock
dam, built to isolate a construction area which can then be pumped dry for
works and bridge foundations. The design of such cofferdams must take
into account the slope stability of the embankment, allowing for seepage
forces, the need for scour protection to the river bed and the embankment,
and sufficient crest height against overtopping by waves or floods.
258 Deep excavations

Fig. 6.52. Circular cell


cofferdam complex with
horizontal bracing, Bangor–
Brewer Bridge, ME, USA23

The economics of embankment cofferdams obviously depend on the


availability of soil and rock for filling and slope protection. In addition,
however, pumping costs to keep the enclosure dewatered will depend on the
type of soil fill and the permeability of the soil under the embankment. If
pumping costs are excessive it may prove economical to install a cut-off
through the embankment and into the underlying soil or rock to reduce
seepage. Relatively shallow cut-offs may be constructed from sheet piling;
deeper walls may be built economically using the slurry trench technique.
During the 1930s several large sand embankment cofferdams were built
across the River Nile to allow construction of barrages for irrigation
purposes. A major scheme was described by Lee22 for remodelling works on
the Assuit Barrage. The works involved new masonry, sluice gates, lift bridges
and improved apron slabs to the original barrage at a site where the Nile is
over 800 m wide. Figure 6.53 shows typical sections through the upstream
and downstream embankments, known as sudds. The cut-offs were made
from Larssen No. 2 sheet piles driven through the pumped sand banks. The
sand, impregnated with silt on the upstream face outside the piling, was
graded to a batter of 1:3; on the downstream side, where the sand was coarser
and cleaner, the gradients were 1:7 or steeper. Dewatering was carried out
slowly to allow the sudds to drain gradually, after which there was little
seepage through the sudds. The work was carried out in four low-water
seasons in successive years, working progressively across the Nile. At the
end of each season, the sudds were slowly rewatered and the sheet piles
extracted. The greater part of the sand fill to the sudds was removed by the
scour of the high river waters.
At sites where timber is cheap, timber cribs can be used to form a gravity
cofferdam. The cribs, sometimes shaped to the river bed profile and with
pockets up to 3.5 m square, are launched and floated into position and filled
with rock and finer material to reduce permeability.
Cofferdam construction 259

Fig. 6.53. Typical sections, upstream and downstream of sudds for reconstruction of Assuit Barrage, River Nile22

Concrete blocks can also be used to form a gravity dam structure to exclude
river water from lock and dock constructions. Gabions filled with rock and
smaller material to reduce leakage can be used for the same purpose.
One of the most dramatic sequences of work occurs on river or marine
construction during closure of a cofferdam to block all flow. This situation
occurs typically where a barrage or dam is to be constructed within a
cofferdam across a flowing river. Often the dam is built in several stages
and as the river flow becomes progressively restricted, scour of the river bed
and of the exposed extremes of the cofferdam structure increases rapidly. In
the final closure work, time is absolutely critical. White23 reported some of
the measures used in such closures.
(a) On rivers with sandy and silty beds, large dredgers have been used to
replace material scoured away.
(b) On swifter flowing rivers with gravelly bottoms, rocks have been
dumped from trestle bridges to keep the trestle anchored and stop the
flow of the stream. This done, sand and clay are placed against the
upstream side of the rock fill to improve watertightness.
(c) Where the depth of the river has been so great as to prohibit access
trestles, cableways have been used to drop heavy rocks or concrete
blocks into the river.
The second-stage closure of the cofferdam for the Chief Joseph Dam on the
Columbia River was only achieved by the retention of large rocks, each
weighing from 10 to 30 t, by individual cables, held in place by a large cable
spanning upstream of the 15 m closure gap. River flow at the time of closure
exceeded 7 m/s.
Mackintosh24 referred to other closure works for dams on the Columbia
River. At the McNary Dam the original 800 m river width had been progres-
sively narrowed to a gap just over 70 m wide between two steel cells. The river
level had to be raised 5 m in order to pass through the spillway blocks, and this
caused a maximum water velocity of more than 9 m/s. The closure was made
over a period of 37 days by dropping a total of 2088 concrete tetrahedra, each
weighing 12 t, into the gap from a cableway.
Mackintosh also gave details of the closure of the Dalles Dam, where
dumped rockfill, incorporated into a permanent embankment, was used to
260 Deep excavations

Fig. 6.54. Cross-section of


Dalles Dam, Columbia River,
USA24

close a 150 m wide channel. At the deepest point the river was 55 m below
low water and in the first stage the gap was reduced to just over 16 m by dump-
ing rock from barges. To begin the final closure operation, end dumping
of rockfill was carried out on a 70 m width, working outwards from the
right-hand river bank until the channel was only 7.6 m wide. This last opera-
tion took five weeks in a river flow of 3120 m3 /s and a velocity through the
final channel of 3.7 m/s. Rock mass ranged from 100 to 259 kg. Rock filling
was followed by end dumping smaller rock and filter material and then a
sand blanket was placed, partly from barges, to form the section shown
in Fig. 6.54. The final closure operation, concentrating all resources on a
stockpile of heavy rock, was achieved on a width of 18 m in a period of only
1.5 hours.
A spectacular closure on the River Saguency in Canada should also be
mentioned. A vast, 28 m high concrete monolith was toppled by blasting
into the river. The base of this mass was pre-shaped to conform with river
bed soundings and weighed 11 000 t. The river flow was greater than 9 m/s,
but the monolith was accurately placed and successful in its purpose.

References 1. Packshaw S. Cofferdams. Proc. Instn Civ. Engrs, 1962, 21, Feb., 367–398.
2. The Construction Regulations (Health, Safety and Welfare). HMSO, London,
1996.
3. Conditions of contract for civil engineering works. Institution of Civil Engineers
Federation of Civil Engineering Contractors, London, 1955, 1979, 1991, 1999
edns IV, V, VI, VII.
4. CIRIA. The design and construction of sheet piled cofferdams for temporary works.
Thomas Telford Ltd., London, 1992.
5. Kirmani M. and Highfill S.C. Design and construction of the circular cofferdam
for ventilation building no 6 at the Ted Williams Tunnel. Civ. Engng Practice,
Spring/Summer, 1996, 31–47. Boston Society of Civil Engineers, MA, 1996.
6. The Icos Company in the underground works. Icos, Milan, 1968.
7. Wharmby N., Kieren B., Duffy L. and Puller D. Stormwater tank construction at
Blackpool. Proc. Conf. on Underground Construction. Institute of Mining and
Metallurgy, Hemming Group Ltd., London, 2001.
8. Williams S.G. and Little J.A. Structural behaviour of sheet piles interlocked at
the centre of gravity of the combined section. Proc. Instn Civ. Engrs, 1992, 94,
May, 229–238.
9. BS 8004. Code of practice for foundations. British Standards Institution, London,
1986.
10. BS 5228, part 4. Code of practice for noise control applicable to piling operations.
British Standards Institution, London, 1986.
11. British Steel Corporation. Report on noise and vibration due to sheet pile installa-
tion. British Steel Corporation, Scunthorpe, 1994.
Cofferdam construction 261

12. EN 12063. Code of practice on steel piling. British Standards Institution, London,
2001.
13. Grice J.R. and Hepplewhite E.A. Design and construction of the Thames Barrier
cofferdams. Proc. Instn Civ. Engrs, 1983, 74, May, 191–224.
14. Fleming J.H. et al. The River Hull tidal surge barrier. Proc. Instn Civ. Engrs,
1980, 68, Aug., 417–454.
15. Wakeling T.R.M. and Hamilton R.J. Discussion on cofferdams. Proc. Instn Civ.
Engrs, 1963, 24, Jan., 106–108.
16. Collingridge V.H. and Tuckwell R.E. Underground station western district post
office, construction. Proc. Instn Civ. Engrs, 1960, 15, Feb., 95–104.
17. Tomlinson M.J. Foundation design and construction. Longman, Harlow, 1995.
18. Ellis L.G. Discussion on cofferdams. Proc. Instn Civ. Engrs, 1963, 24, Jan., 110–
112.
19. Hausser P. et al. A comparison of the design and construction of dry docks at
Immingham and Jarrow. Proc. Instn Civ. Engrs, 1964, 27, Feb., 291–324.
20. White L. and Prentis E.A. Cofferdams. Columbia University Press, New York,
1950.
21. Terzaghi K. Stability and stiffness of cellular cofferdams. Trans. Am. Soc. Civ.
Engng, 1945, 110, 1083–1119 and 1120–1202.
22. Lee D. Deep foundations and sheet piling. Concrete Publications, London, 1961.
23. White R.E. Foundation engineering. Ed. G.A. Leonards, McGraw-Hill, New
York, 1962, ch. 10, 894–964.
24. Mackintosh I.B. Discussion on cofferdams. Proc. Instn Civ. Engrs, 1963, 24, Jan.,
108–109.
25. Anderson J.K. Forth Road Bridge. Proc. Instn Civ. Engrs, 1965, 32, Nov., 321–
512.
26. Gowring G.I. and Hardie A. Severn Bridge, foundations and substructure. Proc.
Instn Civ. Engrs, 1968, 41, Sept. 49–67.
27. EAU 90. Recommendations of the Committee for Waterfront Structures, Harbours
and Waterways. Ernst and Son, Berlin, 6th English edn, 1993.

Bibliography Banks D.J. et al. Construction of Riding Mill weir. Proc. Instn Civ. Engrs, Part 1,
1984, 77, 195–216.
BS 6349 part 1. Code of practice for maritime structures. British Standards Institution,
London, 2000.
BS 8081. Code of practice for ground anchorages. British Standards Institution,
London, 1989.
Calkin D.W. and Mundy J.K. Temporary works for the pumping stations at
Plover Cove reservoir, Hong Kong. Proc. Instn Civ. Engrs, 1987, 82, Dec., 1121–
1144.
Coates R.H. and Slade L.R. Construction of circulating water pump house at Cowes
generating station. Proc. Instn Civ. Engrs, 1958, 9, Mar., 217–232.
CIRIA. Control of groundwater for temporary works. CIRIA, London, 1986, Report
113.
Dondelinger M. Interlocking H-sections for cofferdams to resist high pressure. Deep
Foundations Inst. J., 1985, Spring, 17–28.
Gerrard R.T. et al. Barking Creek tidal barrier. Proc. Instn Civ. Engrs, Part 1, 1982,
72, 533–562.
Hutchinson D.G. and Smith R.A. Brecon flood alleviation scheme. Proc. Instn Civ.
Engrs, Part 1, 1986, 80, 121–143.
Jessberger H.L. et al. Ground freezing. Geotechnical Engineering Handbook. Vol. 2.
Procedures. Ernst and Son. Berlin, 2003.
Mann T. and Dunn M. Lock construction and realignment on the Sheffield and South
Yorkshire navigation. Proc. Instn Civ. Engrs, Part 1, 1986, 80, 1183–1210.
McGibbon J.I. and Booth G.W. Kessock Bridge: construction. Proc. Instn Civ.
Engrs, Part 1, 1984, 76, 51–66.
Patterson J.H. Installation techniques for cellular structures. Proc. Conf. Design and
Installation of Pile Foundation and Cellular Structures. Envo Publishing, Lehigh
Valley, USA, 1970.
262 Deep excavations

Piling handbook. British Steel Corporation, Scunthorpe, 1997.


Pokrefke T.J. et al. Unique utilization of a physical model to develop a construction
sequence of a major cofferdam. Proc. Hydraulics and Hydrology in the Small
Computer Age, American Society of Civil Engineers, New York, 1985, 1130–1135.
Porter D.L. Innovative repairs to steel sheet pile structures. Proc. Conf. Innovations in
Port Engineering in the 1990s. American Society of Civil Engineers, New York,
1986, 703–712.
Quinn W.L. Foyle Bridge, construction of foundations and viaduct. Proc. Instn Civ.
Engrs, Part 1, 1984, 76, 387–409.
Rossow M.P. Sheetpile interlock tension in cellular cofferdams. ASCE J. Geotech.
Engng, 1984, 110, Oct., 1146–1158.
Sarsby R.W. Noise from sheet piling operations – M67 Denton relief road. Proc. Instn
Civ. Engrs, 1982, 72, Feb., 15–26.
Schwab J.P. and Howe J.B. Sherman Island hydro cofferdam. Proc. Conf. Water-
power ’85. American Society of Civil Engineers, New York, 1986, 1288–1297.
Skempton A.W. and Ward W.H. Investigations concerning a deep cofferdam in the
Thames Estuary clay at Shellhaven. Ge´otechnique, 1952, 3, Sept., 119–139.
Specification for steel sheet piling. Foundation of Piling Specialists, London, 1991.
Steel sheet piling design manual. Kawasaki Steel Corporation, Tokyo, 1982.
Terzaghi K. Anchored bulkheads. Trans. Am. Soc. Civ. Engng, 1954, 119, 1243–1281.
Townsend G.H. and Greeves I.S. Design and construction of Gallions surface water
pumping station. Proc. Instn Civ. Engrs, 1979, 66, Nov., 605–624.
Vasconcelos A.A. and Eigenheer L.P. Xingo rockfill dam. Proc. Conf. Concrete Face
Rockfill Dams. American Society of Civil Engineers, New York, 1985, 559–565.
Cofferdam design
7
Braced sheeted In plan the design must allow for installation tolerances of sheeting or walling
cofferdams and sufficient space to accommodate supporting members such as walings and
anchor heads to the sheeting or walling. The design of member sizes will pri-
marily address ultimate limit state conditions, but the serviceability limit state
of deformation must be applied to ensure that deformations of the cofferdam
members, particularly the sheeting, are not excessive and thus impair con-
struction of the permanent works to the required dimensions and thicknesses.
The design life of the cofferdam will inevitably influence the type of
materials used and their durability, but temporary works of very short dura-
tion must not be built with inferior materials that are unable to support the
construction loads and the earth and water pressures imposed upon them.
The sufficiency of secondhand construction materials for cofferdam works
is particularly important in this regard.
The cost of the cofferdam must be minimized irrespective of the cost allow-
ance previously made at tender or job negotiation stage. This cost, however,
must take into account the efficiency (and cost-effectiveness) of construction
works within the cofferdam. For example, cost-effectiveness may be optimized
by spending more on an anchored cofferdam than a braced alternative,
obtaining permanent works construction unimpaired by bracing and the
need to reprop bracing. The construction cost of the cofferdam must include
the costs of its removal; extraction and reuse of the materials elsewhere will be
influenced by the design. The incorporation of the temporary works into the
permanent works should always be considered even if contractual arrange-
ments are not particularly conducive to do so. The obvious example of the
application of such a strategy is the use of structural diaphragm or piled
walls to provide both temporary soil support during construction and there-
after to form the permanent substructure walls.

Site investigation
The frequent inadequacy of site investigations for heavy civil engineering
works has been detailed elsewhere1 , but this shortfall of data for permanent
works design often follows the discovery by the temporary works designer
of an even greater shortfall in the availability of soils, groundwater and
tidal information. The site investigation, conducted to a national code such
as BS 59302 or DIN 40203 , requires an initial definition of scope to include
number, depth and type of borings and in situ tests, and a laboratory testing
programme which satisfies the needs of both temporary and permanent works
designers. A CIRIA report on the design and construction of sheet piled
cofferdams4 recommends the following checklist for information required
from the investigation:
. location of boreholes with respect to the cofferdam
. date of boring and ground level at boring based on same datum refer-
ence as construction drawings
264 Deep excavations

. diameter of boring, drill rig used and whether water was added to the
boring
. depth and thickness of all soil strata to at least twice the proposed
excavation depth, possibly more in weak soils
. bulk unit weights for each soil type
. standard penetration tests and other relevant in situ tests, such as cone
penetrometer tests
. grading curves, Atterberg limits for clay, undrained and, where
appropriate, drained shear tests
. vane test results for soft clays, with sensitivity values
. groundwater strike, rate of ingress, standing levels, any permeability test
data, casing depth at all stages of boring
. position and detail of piezometers, type and readings
. tidal variations and lag
. site geology
. previous site history
. report interpretation.
The design life of the cofferdam works will also influence the designer’s choice
of soil strength parameters, based on undrained or drained effective stress
values for clay. The prime matters affecting the period of full drainage and
effective stress analysis are the drainage path length to permeable strata and
the permeability of the consolidating stratum and the permeable drainage
strata. Where, for example, an otherwise relatively impermeable clay structure
in the active or passive zones is interlayered with permeable sandy beds or
laminations, the rate of drainage will increase and pore pressures will return
to hydrostatic pressures more rapidly, possibly within the design life of the
cofferdam works. For application of total stress, undrained analysis is there-
fore valid only for a very short period after the application of load to the
cofferdam sheeting and support, but the period for full dissipation of excess
pore pressure may vary from days to months. In these circumstances the
designer’s best option will probably lie in the use of undrained analysis in
homogeneous clays, particularly soft clays, using effective stress parameters
as a check. The use of effective stress analyses without a total stress analysis
would then be reserved for areas of good soil drainage, laminated soils or
strata of shallow depth, and analyses made at the end of a long construction
period or for a later permanent works phase.
Due to the relative cost of drained triaxial tests and their duration it is likely
that the number of available test results for a particular clay stratum will be
limited and the designer may have little confidence in their statistical average,
especially for soils where previous results are not available. This is less likely to
be the case for undrained triaxial test results. In this situation the designer may
prefer to keep to undrained total stress analyses using higher safety factors.

Design parameters for soil


Earth pressure calculations, as detailed in Chapter 5, rely on parameters for
soil density and soil strength in terms of cohesion and the angle of shearing
resistance. Preliminary designs may be made using empirical values for a
known soil, and some assistance is given in selecting these in the literature5 .
Calculation values are advised in Table 7.1. The angle of shearing resistance
in terms of effective stress can be estimated from the clay plasticity index,
which is shown in Table 5.1. Note that clays with laminations or seams of
silts or sands show lower plasticity values than clay alone, and care must be
exercised in tests in such clays. If in doubt, the higher plasticity of those
given by testing and from Table 5.1 should be used.
Cofferdam design 265

Table 7.1 Preliminary design soil parameters5

Type of soil Density Final strength Initial Modulus


strength of volume
Above Submerged, Angle of internal Cohesion cohesion of change,
water, cal  cal  0 friction, cal 0 cal c0 undrained soil, cal Es
(kN/m3 ) (kN/m3 ) (degree) (kN/m2 ) cal cu (kN/m2 ) (MN/m2 )

Non-cohesive soils
Sand, loose, round 18 10 30 – – 20–50
Sand, loose, angular 18 10 32.5 – – 40–80
Sand, medium dense, round 19 11 32.5 – – 50–100
Sand, medium dense, 19 11 35 – – 80–150
angular
Gravel without sand 16 10 37.5 – – 100–200
Coarse gravel, sharp edged 18 11 40 – – 150–300
Sand, dense, angular 19 11 37.5 – – 150–250
Cohesive soils (Empirical values for undisturbed samples from the North German Area)
Clay, semi-firm 19 9 25 25 50–100 5–10
Clay, difficult to knead, stiff 18 8 20 20 25–50 2.5–5
Clay, easy to knead, soft 17 7 17.5 10 10–25 1–2.5
Boulder clay, solid 22 12 30 25 200–700 30–100
Loam, semi-firm 21 11 27.5 10 50–100 5–20
Loam, soft 10 9 27.5 – 10–25 4–8
Silt 18 8 27.5 – 10–50 3–10
Soft, org. slightly clayey sea 17 7 20 10 10–25 2–5
silt
Soft, very org. strongly 14 4 15 15 10–20 0.5–3
clayey sea silt
Peat 11 1 15 5 – 0.4–1
Peat under moderate initial 13 3 15 10 – 0.8–2
loading
cal 0 ¼ calculation value of the angle of internal friction in cohesive and non-cohesive soils
cal c0 ¼ calculation value of the cohesion, corresponding to cal 0
cal cu ¼ calculation value of the shear strength from undrained tests in saturated cohesive soils

Appertaining angle of internal friction is to be assumed at cal 0u

For preliminary calculations, 0 for sands and gravels can be estimated


from values of standard penetration tests, as shown in Fig. 5.2. Similarly,
values of undrained cohesive strength cu can be estimated from standard
penetration tests on over-consolidated clays from the empirical expression
cu ¼ nfs where n is the test value and fs is a coefficient which varies with the
plasticity index6 . Values of cu as a function of effective overburden pressure
pn (for normally-consolidated clays) were given by Skempton7 :
cu
¼ 0:11 þ ð0:0037  plasticity indexÞ: ð63Þ
pn
Kenney8 gave preliminary values for the relationship between 0 and plasticity
index for normally-consolidated clays based on observations of more than 60
soils (Fig. 7.1).
The value of 0 for weak rocks may be estimated for preliminary design pur-
poses from a rock description given in BS 80029 (Table 5.2). These indicative
values are considered conservative, being based on granular fragments rather
than intact rock, taking into account closely jointed rock with a very low value
of rock quality designation (RQD).
266 Deep excavations

Fig. 7.1. Relationship between


plasticity index and angle of
shearing resistance41

Design of soil support and structural members to the cofferdam


The use of limit equilibrium methods to design cantilever and propped walls
for overall stability was described in Chapter 5. A factor of safety is achieved
by increasing embedment, decreasing soil strength or factoring moments of
pressure diagrams, to give walls of sufficient depth to avoid overturning.
The earth pressures on which these analyses have been based are Coulomb
values of limit pressures, that is, they are maximum passive pressures and
minimum active pressures which are generated by wall movement at failure.
These limit pressures occur, therefore, at the ultimate limit state of wall
collapse by overturning. The applied factor of safety, which increases wall
embedment, takes into account that in the serviceability condition, when
wall and soil deformation is less, the soil passive resistance will be less than
that generated in the ultimate limit state and the active pressure will be greater
than at failure. Recommended factors of safety for cantilever and propped
wall with free earth support are shown in Table 7.2.
For propped walls with fixed earth support the penetration will always be
greater than that required for free earth support, and will give an adequate
factor of safety against rotation about the prop. If a simplified method is
used for the calculation of penetration as explained in Chapter 5, 20%
extra penetration is used not as a factor of safety but the additional depth
required to correct for the simplification used in the method.
In geotechnical analysis of overall stability of cantilever and propped walls
it is customary to apply best estimated values of applied loads, unit weights
of soils and water pressure in the calculation of limit pressures. Although
contrary to the application of limit state design in its entirety, it is not recom-
mended that partial safety factors greater than 1.0 should be applied to the

Table 7.2 Recommended


factors of safety, cantilever and 0 Effective stress analysis Total stress
propped walls, free earth analysis
support 208 20–308 308

Factor on strength Fs ¼ 1:2 Fs ¼ 1:2–1.1 Fs ¼ 1:5


Factor on moment : :
Fp ¼ 1 2 Fp ¼ 1 2–1 5 . F p ¼ 1: 5 Fp ¼ 2:0
Factor on net pressure :
ðFsp Þ ¼ 2 0 Not recommended
Burland–Potts Fr ¼ 1:5 Fr ¼ 2:0
Cofferdam design 267

various loads on the retaining structure to produce factored bending moments


for the wall. As previously stated, unfactored loads with application of
increased embedment, factored soil strength or factored moments of the
earth diagram should be used according to choice.
The wall, dimensioned according to one of these methods, with an applied
factor of safety, will be checked for sliding, basal failure and hydraulic failure,
the factor of safety for each being calculated using limit pressures where
earth pressure calculation is required. Next is the calculation of the sizes of
structural members. Two methods of design are available for the design of
the sheeting and its support by walings, struts and anchors: these are the
permissible stress method and the limit state method.
. In the permissible stress method, best estimates of load are used. Limit
pressure diagrams (or strut load envelopes based on the total pressure
from limit earth pressure diagrams) are used to calculate bending
moments and shears in the wall/sheeting and in the walings, and thrusts
or tensions within the struts or anchors. Permissible stresses based on
ultimate stresses reduced by a factor of safety are then calculated
from the moments, shears, thrusts and tensions to define the size of
these structural members. Permissible stress design is therefore based
on limit pressures which do not occur in the serviceability state. Further,
the ultimate limit state failure conditions related to these limit pressures,
which are based on the monolithic failure of earth masses, may
be unrelated to the ultimate limit state conditions (and therefore
the earth pressures at these conditions) of collapse of the structural
components.
. The limit state method uses conventional methods of design for
reinforced concrete, steelwork and timber structural members in
which the application of factored loads produces factored bending
moments, shears, thrusts and tensions, which are then used with the
ultimate or characteristic strength of the material to define their size.
Although the logic of this method is flawed when using limit pressures,
it appears to be less flawed than using permissible stress design with such
pressures. For this reason, and because of the general use in structural
design of limit state design, its application is adopted here.
The design procedure is as follows:
(a) Using ULS values for the more onerous of the design approaches, that
is either moderately conservative or worst credible parameters and
applying the appropriate partial factors thereto (see Table 5.6) calculate
the sheeting moments, shears and prop loads.
(b) Calculate values of moment, shears and prop loads for serviceability
state using unfactored soil parameters (if serviceability state has been
analysed) multiplied by a factor of 1.4.
(c) Adopt greater values for (a) and (b). Consider risk of progressive
collapse.

Material stresses
A summary of the properties of materials used in cofferdam construction is
given in Table 5.10.

Plastic design of sheet pile walls


Recent developments in the design of sheet piles, particularly sheet piles in
soft soil have been described, together with full-scale validation by Kort10 .
The use of plastic design methods for sheet piles has been practised mainly
268 Deep excavations

in Denmark and North Germany since the 1950s following the publication of
a method of plastic design for sheet pile walls by Brinch Hansen in 1953.
Kort refers to significant savings that can be obtained using plastic design
compared with conventional elastic analysis. Allowing both the full plastic
modulus of the pile section but without allowance for development of a plastic
hinge may result in a saving of 15 to 20% for a typical anchored wall using a
Frodingham type section (without risk of slippage of the clutch). Allowing
rotation of the plastic hinge results in a similar saving which can reach a
total of 35% or more for plastic design. Some reduction in this saving may
be necessary to allow for local buckling of the sheet pile cross-section. The
facility to make any reduction will, of course, be dependent on the driveability
of the pile section but this would not be likely to be critical in soft soils.
Corrosion risk may also need to be assessed in reducing sheet pile thickness.
The recent work by Kort also considers the effect of clutch slippage, which
may critically affect the flexural strength of Larssen piles, clutched on the
centre line of the sheet pile wall. This aspect of design is referred to as oblique
bending and results in a reduction of the saving permitted by plastic design for
the Larssen section.
Kort used a subgrade reaction model which allowed moment redistribution
and a plastic hinge within the wall height when a certain maximum moment is
reached. The development of the plastic hinge would then occur without det-
riment to wall stability providing embedment depth was sufficient to absorb
the increased fixed end moment due to this redistribution. A large wall deflec-
tion would result of course. In the event, it does not follow that sheet pile
lengths are necessarily longer for plastic designed walls than those elastically
designed because the fixed end moment would not always be fully developed.
In multi-propped walls the prudent positioning of struts in a plastically
designed wall can be of greater effect than the embedment depth.
The effect of a reduction in flexural strength due to slippage at the clutch of
Larssen sheet pile sections was investigated by Kort and a design rule was
developed to take oblique bending into account. This was done using a sub-
grade reaction model which in turn was verified by a bending test on two
double U-piles and a 3D simulation using the finite element program DIANA.
The progressive development of sheet pile design for both temporary and
permanent works design on works other than minor schemes is encapsulated
in Kort’s work insofar as this may identify more realistic predictions of
deformed wall shape; a change of sheet pile section to facilitate economical
pile design may follow, together with greater efficiency in design in multi-
braced walls for strut levels. All this work would comply with the European
Codes, Eurocode 3 (EN 1997) and Eurocode 7 (EN 1997). The value of the
original work of Brinch Hansen may finally be realized; the reader is directed
to the work of Kort10 .

Overall stability
As has been mentioned, the overall instability of a cofferdam construction
should always be checked. Such risk particularly occurs in sloping ground
or, typically, in a riverfront cofferdam where the rear wall of the cofferdam
supports retained ground and the front face supports tidal river conditions.
Where the difference in height is small it may be possible to transfer load
from the higher to the lower side by keeping the top frame as low as possible
(Fig 7.2). Where the differential height is greater it may be expedient to rake
the top frame from one side to the other. Alternatively, anchor the sheeting at
the higher side.
Where cohesive soils extend to considerable depths on sloping sites, it is
necessary to check that a deep-seated potential slip surface is not a failure
Cofferdam design 269

Fig. 7.2. Cofferdam


construction in sloping ground4

risk. Where such a failure risk exists stability can be increased by driving the
sheeting to greater depth to intercept the potential failure surface; jet-grouted
columns may be installed below cofferdam formation level before excavation
to achieve the same objective.

Bottom failure by piping and basal heave


Repeating earlier advice, the risk of hydraulic failure by piping should be
checked for narrow cofferdams that do not achieve a cut-off in cohesionless
soils with a high external water table. Risk of basal failure in cofferdams in
soft clay should also be checked.

Aggressive site conditions: marine and river cofferdams


The extreme exposures to which cofferdams in both river and sea waters are
subjected vary according to geographical location and the size and depth of
the cofferdam. Matters to be assessed include the effect of wave forces on
the face of the structure, overtopping of the cofferdam walls, scour, protection
from vessel impact, and the impact and pressure of ice on the face of the
cofferdam.
The effects of wave action on the face of the cofferdam are discussed in BS
634911 and in EAU 905 . These loads on cofferdams are due to waves in deep
water, where waves are reflected, and those in shallower water where waves
break on the structure or some distance from it. Sainflou12 and Minikin13
determined the forces due to reflected and breaking waves, respectively,
referring, to waves on backfilled waterfront structures with standing water
within the backfill. These methods should, however, be used cautiously
when determining forces on sheeted structures. The very high loads that can
be imposed on structures due to breaking waves, with impact pressures of
10 000 kN/m2 or more, identify the critical nature of this loading, especially
since there is neither a reliable method of calculation nor an empirical solution
for determining such loads. BS 6349 refers to breaking wave forces calculated
by Minikin’s method as high as 18 times those calculated for non-breaking
waves. Estimates from computer modelling and measurements from physical
models may assist where the scale of the cofferdam works justify such a study.
Waves caused by ship movement may also require consideration. In
restricted waters the action of a headwater wave caused by water displacement
in front of a vessel may require assessment in determining the maximum head
of water on the face of a cofferdam structure; conversely, water drop occurring
to the stern of a vessel may also cause variation in loading along the length of a
long cofferdam. These matters, together with the action of bow and stern
waves, are discussed in reference 5.
Reference should be made to the effect of pressure transmitted through
coarse-graded beach deposits to the walls of cofferdams sited between high
and low water, and to the dynamic effect on the flow of water below wave-
facing cofferdam walls where a cut-off is not obtained by the toe of the
walls. In neither case is there guidance from published work and only caution
270 Deep excavations

can be advised in the selection of factors of safety on strut or anchor design


and on the risk of hydraulic failure at the base of the cofferdam excavation.
The risk of overtopping of the cofferdam sheeting depends on the freeboard
allowed to assessments of maximum tide heights, surge and wave heights from
natural and man-made causes. It is essential that where there is any risk of the
cofferdam flooding from overtopping, the cofferdam sheeting should be tied
to prevent it bursting outwards under the action of a water-filled cofferdam.
Adequate sluices with safe locations for operating controls are essential to
avoid the risk of a cofferdam remaining filled with water on a falling tide.
Scour protection will be needed where cofferdams are sited in fast flowing
rivers or tidal conditions. Large cofferdams for bridge piers in fast flowing
rivers may benefit from tests using physical models to determine the extent
of scour and any increased risk due to obstruction of the flow by adjacent
cofferdams. The upstream and downstream ends of a cofferdam may be
shaped to reduce scour by providing cutwaters. Where protection is necessary
to avoid erosion of the river or sea bed, rock or concrete blocks may be
suitable. Grout mattresses, weighted by rockfill, may be adequate for river
cofferdams. Where the scour risk is lower it may be sufficient to design the
cofferdam sheeting embedment and strutting to ensure that collapse does
not follow scour action.
In busy navigable waterways it will be necessary to protect the cofferdam by
fendering or strongpoints built into the cofferdam sheeting to avoid damage
by vessels. Barge traffic, in particular, appears to cause high collision risk.
River authorities frequently define the extent of fendering required for river
cofferdam works and specify the signage needed.

Durability of steel sheet pile walls


A report published in 2002 by Corus14 , reviewed corrosion rates in temperate
climates for unprotected sheet piles. These rates, shown for each side of both
atmospheric and marine locations in Fig 7.3 are based on investigations by
British Steel (now Corus) and others. Corrosion rates are quoted as mean
values since Corus considers these to be most relevant to design of most struc-
tures. Higher values are also quoted based on 95% probability for use where
the designer deems necessary.
In their report, Corus refer to the risk of localized corrosion considerably in
excess of those shown as mean or 95% probability values (Table 7.3). This
local corrosion, known as ‘accelerated low water corrosion’ or ALWC in
the UK and as ‘concentrated corrosion’ in Japan is not restricted to a particu-
lar section or steel manufacturer. The corrosion, based on bacterial growth in
colonies, is generally confined to outpans of sheet piles in a zone at the mean
water level, or just below. On U-shaped piles, the attack is most severe at the
corner of the outpan whilst for Z-shaped piles the attack is concentrated on
the corners or legs of the outpans. Corrosion rates within the range of 0.3
to 0.8 mm per year have been observed and in exceptional cases attack has
caused local holes and slits in the sheet piles. It must be emphasized that
the frequency of occurrence is small, and associated with marine structures;
ALWC has not been observed to attack permanent sheet piling used for base-
ment construction in the UK. Papers by Johnson et al.15 , Tsuchida et al.16 and
Fukute et al.17 should be consulted regarding the risk associated with ALWC.
Methods of increasing the effective life of steel sheet piles are also reviewed
in the British Steel Piling Handbook18 , and are summarized as follows:
. Use of a heavier section with allowance for a sacrificial thickness.
. Use of high yield steel as an alternative to mild steel, allowing increased
bending stresses.
Cofferdam design 271

Pile face Environments

Fig. 7.3(a) Fig. 7.3(a) Fig. 7.3(c) Fig. 7.3(b) Fig. 7.3(c) Fig. 7.3(c) Fig. 7.3(b) Fig. 7.3(b) Fig. 7.3(a)
A B HJ DF I K E G C
Side 1 Atmospheric Atmospheric Splash or low water Splash or low water Tidal Immersion Tidal Immersion Soil
Side 2 Atmospheric Soil Splash or low water Soil Tidal Immersion Soil Soil Soil
Corrosion ratesa 0.07 Mean 0.05 Mean 0.15 Mean 0.09 Mean 0.07 Mean 0.07 Mean 0.05 Mean 0.05 Mean 0.03 Mean
(mm/year)
Corrosion ratesa — — (0.33) (0.18) (0.19) (0.25) (0.11) (0.14) —
(mm/year)
95% Probability value
a
Total loss of section (side 1 þ side 2)
Notes
1. The corrosion rates quoted are based upon investigations carried out by Corus and others on steel exposed in temperate climates. For most environments mean values are quoted
since they are considered to be most relevant to the design and performance of most sheet piling structures. However, in some circumstances the designer may wish to take account of
higher values. It is suggested that in these circumstances a reasonable practical upper corrosion rate limit would be that corresponding to a 95% probability value. (Values given in
parentheses).
2. All corrosion rates quoted in the table are measured as total loss of section thickness from both sides, taking into account both environments.
3. For combinations of environments where low water corrosion is involved, in a small number of locations, higher rates than those quoted have been observed at or just below the low
water level mark and Corus recommend that periodic inspection is undertaken. In the case of uncertainty, please contact Corus for advice.
4. A maximum value is quoted for soil corrosion and this applies to natural undisturbed soil or well compacted and weathered fill ground where corrosion rates are very low. Recent fill
ground or waste tips will require special consideration.
5. Corrosion losses due to fresh water immersion are generally lower than for sea water, however, fresh waters are very variable and no general advice can be given to quantify the
increase in life.

Fig. 7.3. Corrosion rates for unprotected sheet piles in temperate climates18

Table 7.3. Factors affecting localized corrosion14

Mechanism Cause

1. Macro-cell effect Potential difference: anodic at low water zone. Macro-cathode within the tidal zone
allows available oxygen for the cathodic reduction reaction to cause a corrosive
environment. Dependent upon local conditions
2. Removal of the corrosion Continual removal of the abrasion product by the action of fendering systems,
product by erosion or propellor wash, bow thrusters,waterborne sands and gravels or repeated stresses
abrasion can lead to localized corrosion. The area where the rust layer is removed becomes
anodic to the unaffected areas, particularly in the low water zone where macro-cell
effects are strongest
3. Micro-biological activity Corrosion products from affected structures can contain compounds such as
sulphides which stimulate local corrosion
4. Bi-metallic corrosion Can occur where steel is connected to other steels, metals or alloys. Corrosion is
concentrated in the less noble steel often at the junction of the dissimilar materials
5. Fouling by animals/plants Can cause acceleration of corrosion in localized areas, resulting from the formation
of differential aeration cells or possibly by biological processes
6. Stray currents From improperly grounded DC power sources. Localized damage where current
leaves the structure
272 Deep excavations

. Coatings: British Steel (now Corus) recommended tar vinyl and high-
build epoxy pitch coatings. Coatings for potable water in the UK
must meet byelaw approval and are approved by the Water Research
Centre. Most comprise bitumen or two-pack epoxy resin systems.
. Concrete encasements: for marine use.
. Cathodic protection: usually the work of specialist firms.
A review of coating systems for use on sheet piles is made in a publication by a
specialist firm, SIGMA Coatings19 .

General layout of the cofferdam


The plan shape of the cofferdam will usually conform approximately to the
plan shape of the permanent works to be built within it. In general terms,
the cost of cofferdam construction of constant depth is directly proportional
to its plan area. Any wastage of space in the choice of cofferdam plan shape
therefore will affect cost. The design of circular cofferdams is reviewed later,
but in terms of relative economy only square-shaped plan structures or
those with circular plan shape storage wells, such as pumping stations, can
be accommodated in circular cofferdams cost-effectively.
For rectilinear cofferdam plan shapes the most economical arrangement
uses maximum straight runs of piling with the minimum of return angles.
The length of strutting across the cofferdam will define the need to support
struts by king posts (traditionally known as puncheons) and prop them
laterally to reduce the effective lengths of strutting and the risk of buckling.
Maximum strut length in steelwork is likely to be of the order of 40 m, and
distances in excess of this may need to rely on raking shores from a completed
central raft section or support from ground or rock anchors; 458 diagonal
corner braces should always be used in frames unless this hinders the
permanent construction. Some typical arrangements of framing in plan are
shown in Fig. 7.4. The vertical location of frames must be such that the
maximum bending stress induced in the sheeting between frames is approxi-
mately constant. Overstressing of the sheeting must be avoided, particularly
below the lowest frame, which should be located as low as possible to avoid
risk of passive failure in the soil. The location of the penultimate frame is
also important: the vertical height between final formation level and the
penultimate frame should not be excessive in order that sheeting stresses
and passive soil stresses are not exceeded immediately prior to placing the
bottom frame.
Where soil strengths are low, near or below formation level, and earth and
water pressure loading on the sheeting is relatively high, it may be necessary
(particularly in river and marine cofferdams) to avoid excessive sheeting
stresses and risk of passive failure of the soil below formation level by placing
the lowest frame underwater by divers having flooded the cofferdam. The
frame levels should generally be chosen to allow concrete lifts to be poured
economically. Care may be needed to avoid starter or splice steel reinforce-
ment from one pour being obstructed by the waling of a cofferdam frame.
This applies particularly to the lowest frame impeding vertical starter steel
in the kicker from the base slab. It may be possible to locate this front and
rear starter steel to the wall construction on either side of the temporary
waling, or couplers may be needed to extend the steel using the shortened
height of the starter steel.
It is essential that load is transferred efficiently from soil to sheeting to
waling to strutting, without any doubt as to the direction of the transfer.
The principal reasons for failure of sheeted cofferdams are poor workmanship
in connections causing insecure transfer of load, inadequate strut sections,
Cofferdam design 273

Fig. 7.4. Typical cofferdam


framing in plan and in vertical
cross-section4

inadequate embedment of the sheeting and overload due to inadequate


allowance for surcharge loading. These modes of failure should never be
forgotten.

Use and design of ground anchors


The economical use of anchors in land cofferdams depends on the strength of
the subsoils in which the anchor is to be founded. Dense sands and gravels
would be preferred to cohesive subsoils at a similar depth.
The use and design of ground anchors is described in two codes of practice,
BS 808120 and the European Code BS EN 1537: 200021 .
The opportunity to use anchors will depend on the ownership of land at the
periphery of the cofferdams and on permission to found anchors in neigh-
bouring land. The presence of existing substructures or basements may
obstruct anchor installation.
274 Deep excavations

The use of removable anchors may be necessary to avoid obstruction to


future use of adjacent sites. Systems which allow removal of the unbonded
steel tendon in the face length have been available for some time but only
the use of a loop of totally unbonded strand allows the complete removal
of the tendon. The method, developed by Keller, requires the pre-bending
of plastic-coated strand at its midpoint to an internal diameter as small as
60 mm. The base of the loop complete with a thimble inside the loop is
placed at the bottom of the bore. Successive loops at shallower depths may
be placed as part of the single-bore multiple-anchor system. The anchor is
then grouted. After use the whole looped tendon is destressed and each
unbonded strand jacked or winched out of the anchor one by one.
Where internal angles occur in the line of sheeting, anchors from adjacent
walls have to be located carefully to avoid anchors obstructing each other
(Fig. 7.5).

Fig. 7.5. Arrangement of


ground anchors at re-entrant
angles to retaining walls in
plan20
Cofferdam design 275

Earth pressures acting on an anchor installation in a temporary cofferdam


depend not only on soil strengths but also on wall and soil stiffnesses, anchor
spacing, anchor yield, the pre-stress locked into the anchors as installed and
loss of pre-stress with time. The anchored wall may be designed with active
and passive earth pressure diagrams using limiting pressures, as described in
Chapter 5 for multi-propped walls. The loads in the anchors can be obtained
from the empirical trapezoidal strut load envelope diagrams. Alternatively,
methods based on Winkler spring analysis may be used with computer
programs in which earth pressures are computed from trial anchor loads
input to produce deformed wall profiles, the anchor loads being varied in
successive runs to produce acceptable maximum wall deflections.
Temporary anchors used in cofferdam construction have stressing tendons
to transfer load from the fixed anchorage on the face of the cofferdam wall to a
fixed anchorage within the subsoil outside the potential failure wedge at the
rear of the wall. Figure 7.6 shows typical details and geometry. In cohesionless
soils the fixed anchorage is formed by pressure-grouted techniques, while in
cohesive soils, tremie methods of pouring grout may also be used. BS 808120
provides a comprehensive review of the design and construction of both
temporary and permanent anchors in soil and rock (it also has an extensive
list of references). It depicts five anchor types.
(a) Type A: anchorages consisting of tremie, packer or cartridge-grouted
straight shaft boreholes temporarily lined or unlined depending on
hole stability.
(b) Type B: anchorages consisting of low pressure (typically injection
pressures less than 1000 kN/m2 ) grouted boreholes with an increased
diameter within the fixed length.
(c) Type C: anchorages consisting of boreholes grouted to high pressure
(typically injection pressures more than 2000 kN/m2 ). The length of
these anchors is enlarged by hydrofracturing of the ground to give a
grout fissure system beyond the nominal borehole diameter. Post-
grouting using tubes à manchette is frequently used with only a small
quantity of secondary grout being required.
(d ) Type D: tremie-grouted boreholes consisting of a series of bells
or under-reams mechanically formed within the fixed length of the
anchor.
(e) Type E: other types of anchor formed by jet grouting or similar
techniques.
BS 8081 refers to four items that have to be addressed in the design of these
anchors: overall stability, depth of embedment, fixed anchor dimensions,
and group effect. To check overall stability in cohesionless soils, several
analysis methods are detailed in BS 8081. The simplest method is based on
a planar failure surface. The failure wedge may be assumed as based on
Coulomb wedge with an angle to the back of the wall of  ¼ 458 þ =2
(Fig. 7.6). A further method, known as the sliding block method, is shown
in Fig. 7.7 with a subsequent variation shown as Fig. 7.8. The fixed anchor
dimensions based on safety factor methods as detailed below comply with
the recommendations of BS 8081. Safety factors for temporary and perma-
nent anchors are reproduced from BS 8081 in Table 7.4. (The relatively
high values for temporary anchors compared with those used in Continental
Europe may have led to their reduced use in the UK.)
From BS 8081, for fixed anchor design in rock, for type A anchorages, the
ultimate load carrying capacity Tf (kN) is estimated from
Tf ¼ DLult ð64Þ
276 Deep excavations

Fig. 7.6. Ground anchors:


wedge method of analysis20
Cofferdam design 277

Fig. 7.7. Ground anchors:


sliding block method of overall
stability analysis20 (BS 8081):
(a) as modified by Locher and
Littlejohn; (b) as modified by
Ostermayer

where ult is the ultimate bond or skin friction at the rock/grout interface
(in kN/m2 ), D is the diameter of the fixed anchor (in m), and L is the length
of the fixed anchor (in m). Bond values recommended for design in rocks
and soft rocks are shown in Table 7.5.
For fixed anchor design in cohesionless soils for type B anchorages
Tf ¼ Ln tan 0 ð65Þ
0
where  is the effective angle of shearing resistance (degrees) and n is a factor
to allow for drilling technique, depth of overburden, fixed anchor diameter
and grout pressure. Littlejohn and Bruce22 suggested that n ranges from
400 to 600 kN/m for coarse sands and gravels, and from 130 to 165 kN/m
in fine to medium sands. These values were measured in borehole anchor
diameters of approximately 0.1 m. Where the design diameter is larger, n
should be increased in linear proportion. Alternatively,

Tf ¼ A0v DL tan 0 þ Bh ðD2  d 2 Þ (side shear and end shear) ð66Þ
4
278 Deep excavations

Fig. 7.8. Ground anchors:


revised sliding block method of
overall stability analysis20

where A is the ratio of contact pressure at the fixed anchor/soil interface to the
average effective overburden pressure (it has a value between 1 and 2, depend-
ing on construction technique),  is the unit weight of soil overburden (use
submerged weight below water table) (in kN/m3 ), h is the depth of overburden
to the top of the fixed anchor (in m), 0v is the average effective overburden
pressure adjacent to the fixed anchor (in kN/m2 ), B is the bearing capacity
factor and is equivalent to Nq =1:4 (refer to Berezantzev’s curve, Fig. 7.9), d
is the diameter of the grout shaft above the fixed anchor (in m) and D is the
diameter of the fixed anchor (in m).
For type C anchorages, in cohesionless soils, design methods have relied
heavily on field tests in a range of soils rather than load capacity expressions.
Figure 7.10 shows load-carrying capacity in sandy gravels and gravelly sands
for increasing anchor length. Sherwood and Harris23 proposed an empirical
design method for regroutable anchors which used their experience over 15
years to establish limit shaft friction values in clays and silts and in sands
and gravels over the groutable length of the proprietary TMD anchor from
Bachy. They argued that while BS 8081 recognizes that pressure-grouted
Cofferdam design 279

Table 7.4 Minimum safety factors recommended for design of individual anchorages20

Anchorage category Minimum safety factor Proof


load
Tendon Ground/grout Grout/tendon or factor
interface grout/encapsulation
interface

Temporary anchorages where a service life is less 1.40 2.0 2.0 1.10
than six months and failure would have no serious
consequences and would not endanger public safety,
e.g. short-term pile test loading using anchorages as a
reaction system
Temporary anchorages with a service life of, say, up 1.60 2.5a 2.5a 1.25
to two years where, although the consequences of
failure are quite serious, there is no danger to public
safety without adequate warning, e.g. retaining wall
tieback
Permanent anchorages and temporary anchorages 2.00 3. 0 b 3.0a 1.50
where corrosion risk is high and/or the consequences
of failure are serious, e.g. main cables of a suspension
bridge or as a reaction for lifting heavy structural
members
a
Minimum value of 2.0 may be used if full-scale field tests are available.
b
May need to be raised to 4.0 to limit ground creep.
Note 1. In current practice the safety factor of an anchorage is the ratio of the ultimate load to design load. The table defines
minimum safety factors at all the major component interfaces of an anchorage system.
Note 2. Minimum safety factors for the ground/grout interface generally lie between 2.5 and 4.0. However, it is permissible to vary
these, should full-scale field tests (trial anchorage tests) provide sufficient additional information to permit a reduction.
Note 3. The safety factors applied to the ground/grout interface are invariably higher compared with the tendon values, the
additional magnitude representing a margin of uncertainty.

anchorages are superior to other types in terms of load capacity, BS 8081 does
not differentiate between anchors installed with uniform injection pressure
and those with repetitive and selective high-pressure injection. Sherwood
and Harris concluded that factors of safety recommended by BS 8081 may
be appropriate to less sophisticated ground anchors but are much too
conservative for anchor systems which produce more uniform anchorage
performance as a result of their installation procedure. They reasoned that
preliminary anchor tests should be used as a means of refining design and
reducing the factor of safety.
For type D anchorages, BS 8081 does not suggest a design method for
cohesionless soils due to lack of published data but warns that if shaft
enlargements are required to take all the anchor load, the shear across the
interface between the nominal diameter shaft and the enlargements will
require examination.
For fixed anchor design in cohesive soils, for type A anchorages, load-
carrying capacity may be estimated from
Tf ¼ DLcu ð67Þ
2
where cu is the undrained shear strength (in kN/m ) and  is the adhesion
factor (in the range 0.28 to 0.36 for stiff clays and 0.48 to 0.6 for stiff to
very stiff marls).
For type C anchorages in cohesive soils, BS 8081 reproduces the work by
Ostermayer24 as a design guide for borehole diameters between 0.08 and
0.16 m. The effect of post-grouting pressure on skin friction is shown in
Fig. 7.11.
280 Deep excavations

Table 7.5 Rock/grout bond values which have been employed in practice22

Rock type Working Test Ultimate Factor of safety Source


bond bond bond
(N/mm2 ) (N/mm2 ) (N/mm2 ) Measured Design

Igneous
Basalt 1.93 6.37 3.3 Britain—Parker (1958)
Basalt 1.10 3.60 USA—Eberhardt and Veltrop (1965)
Tuff 0.80 France—Cambefort (1966)
Basalt 0.63 0.72 Britain—Cementation (1962)
Granite 1.56 1.72 Britain—Cementation (1962)
Dolerite 1.56 1.72 Britain—Cementation (1962)
Very fissured felsite 1.56 1.72 Britain—Cementation (1962)
Very hard dolerite 1.56 1.72 Britain—Cementation (1962)
Hard granite 1.56 1.72 Britain—Cementation (1962)
Basalt and tuff 1.56 1.72 Britain—Cementation (1962)
Granodiorite 1.09 Britain—Cementation (1962)
Shattered basalt 1.01 USA—Saliman and Schaefer (1968)
Decomposed granite 1.24 USA—Saliman and Schaefer (1968)
Flow breccia 0.93 USA—Saliman and Schaefer (1968)
Mylontized prophyrite 0.32–0.57 Switzerland—Descoeudres (1969)
Fractured diorite 0.95 Switzerland—Descoeudres (1969)
Granite 0.63 0.81 Canada—Barron et al. (1971)

Metamorphic
Schist 0.31 Switzerland—Birkenmaier (1953)
All types 1.20 Finland—Maijala (1966)
Weathered fractured 1.56 1.72 1.1 Britain—Cementation (1962)
quartzite
Blue schist 1.52 1.67 1.1 Britain—Cementation (1962)
Weak meta sediments 1.10 1.23 1.1 Britain—Cementation (1962)
Slate 0.43 Britain—Cementation (1962)
Slate/meta greywacke 1.57 1.73 1.1 Britain—Cementation (1962)
Granite gneiss 0.36–0.69 Sweden—Broms (1968)
Folded quartzite 0.51 Australia—Rawlings (1968)
Weathered meta tuff 0.29 USA—Saliman and Schaefer (1968)
Greywacke 0.34 Germany—Heitfeld and Schaurte (1969)
Quartzite 0 93–1.20
. 1.02–1.32 1.1 Britain—Gosschalk and Taylor (1970)
Microgneiss 0.95 Italy—Mantovani (1970)
Seridite schist 0.05 Italy—Berardi (1972)
Quartzite/schist 0.10 Italy—Berardi (1972)
Argillaceous and calcareous 0.63 Italy—Berardi (1972)
schist
Slate 0.95 1.24 1.3 Switzerland—Moschler and Matt (1972)
Highly metasediments 0.83 1.08 1.3 USA—Buro (1972)
Slate and greywacke 1.08 1.40 1.3 Germany—Anon (1972)
Various metasediments 1.57 Germany—Abraham and Prozig (1973)
Micaschist/biotite gneiss 0.53 0.80 1.5 USA—Nicholson Anchorage Co. Ltd
(1973)
Slate 0.60 0.90 1.80 1.5 3.0 Britain—Littlejohn and Truman-Davies
(1974)
Sound micaschist 1.74 2.16 USA—Feld and White (1974)
Micaschist 0.52–0.74 1.24 USA—Feld and White (1974)
Very poor gneiss and mud 0.07 USA—Feld and White (1974)
band

Carbonate sediments
Loamy limestone 0.63 Italy—Berardi (1960)
Fissured limestone and 1.08 1.19 1.1 Britain—Cementation Co. Ltd (1962)
intercalations
Limestone 0.65 Switzerland—Muller (1966)
Poor limestone 0.32 France—Hennequin and Cambefort (1966)
Massive limestone 0.39–0.78 France—Hennequin and Cambefort (1966)
Karstic limestone 0.54 France—Hennequin and Cambefort (1966)
Tertiary limestone 1.00 2.83 2.8 Switzerland—Losinger and Co. Ltd (1966)
Cofferdam design 281

Table 7.5 continued

Rock type Working Test Ultimate Factor of safety Source


bond bond bond
(N/mm2 ) (N/mm2 ) (N/mm2 ) Measured Design

Limestone 4.55–4.80 Switzerland—Ruttner (1966)


Marly limestone 0.03–0.07 Italy—Berardi (1967)
(average)
0.21–0.36
(measured)
Limestone 0.27 USA—Saliman and Schaefer (1968)
Limestone 0.28 Italy—Berardi (1969)
Dolomitic limestone 1.80 Canada—Brown (1970)
Marly limestone 0.39–0.94 Italy—Berardi (1972)
Limestone 0.26 Italy—Berardi (1972)
Limestone/puddingstone 0.44 France—Soletanche Co Ltd (1968)
Limestone 1.18 1.42 1.2 USA—Buro (1972)
Chalk 0.70 Britain—Associated Tunnelling Co. Ltd
(1973)
Dolomite 1.66 Canada—Golder Brawner (1973)
Dolomitic siltstone 0.43 USA—White (1973)
Limestone and marly bands 0.37 0.55 1.5 Italy—Mongilardi (1972)

Arenaceous sediments
Sandstone 1.44 1.58 1.1 Britain—Morris and Garrett (1956)
Hard sandstone 1.42 1.56 1.1 Britain—Cementation Co. Ltd (1962)
Bunter sandstone 0.95 0.98 1.03 Britain—Cementation Co. Ltd (1962)
Sandstone 0.76 0.84 1.1 Britain—Cementation Co. Ltd (1962)
Sandstone 0.74 Czechoslovakia—Hobst (1968)
Sandstone 0.31 0.40 1.73 1.29 5.6 USA—Drossel (1970)
Sandstone 0.80 USA—Thompson (1970)
Poor sandstone 0.40 Germany—Brunner (1970)
Good sandstone 1.14 Germany—Brunner (1970)
Sandstone and breccia 0.38 France—Soletanche (1968)
Sandstone 0.95 Australia—Williams et al. (1972)
Bunter sandstone 0.60 1.20 2.0 Britain—Littlejohn (1973)
Sandstone 1.17 Australia—McLeod and Hoadley (1974)

Argillaceous sediments
Shale 0.62 Canada—Juergens (1965)
Marl 0.10 0.28 2.8 Italy—Berardi (1987)
Shale 0.30 0.63 2.1 Canada—Hanna and Seaton (1967)
Very weathered shale 0.39 USA—Saliman and Schaefer (1968)
Shale 0.13–0.24 USA—Koziakin (1970)
Grey siltstone 0. 62 Britain—Universal Anchorage Co. Ltd
(1972)
Clay marl 0.14–0.24 0.21–0.36 1.5 Germany—Schwarz (1972)
Shale 0.62 Canada—McRosite et al. (1972)
Argillite 0.82 Canada—Golder Brawner (1973)
Mudstone 0.63 0.88 1.4 Australia—McLeod and Hoadley (1974)

Miscellaneous
Bedded sandstone and shale 0.20–0.50 Italy—Beomonte (1961)
Porous, sound goassamer 1.57 1.72 1.1 Britain—Cementation Co. Ltd (1962)
Shale and sandstone 0.07 0.10 1.5 USA—Reti (1964)
Soft rocks 0.75 Sweden—Nordin (1966)
Sandstone and shale 1.82 Poland—Bujak et al. (1967)
Siltstone and mudstone 1.65 Australia—Maddox et al. (1967)
Fractured rock (75% shale)
Poor 0.24 USA—Saliman and Schaefer (1968)
Average 0.35 USA—Saliman and Schaefer (1968)
Good 0.75 USA—Saliman and Schaefer (1968)
Limestone and clay breccia 0.20–0.23 Italy—Berardi (1972)
282 Deep excavations

Fig. 7.9. Berezantzev’s curve of


bearing capacity factor Nq
against angle of shearing
resistance

Fig. 7.10. Ultimate load capacity as a function of fixed length in sandy gravels and gravelly sand42

For type D anchorages in cohesive soils, the ultimate load capacity of multi-
under-reamed anchorages is given by

Tf ¼ DLcu þ ðD2  d 2 ÞNc cub þ dlca
4
(side shear þ end bearing þ shaft resistance) ð68Þ
where D is the diameter of under-ream (Fig. 7.12) and Nc is the bearing capa-
city factor (a value of 9.0 is often used), cub is the undrained shear strength in
Cofferdam design 283

Fig. 7.11. Influence of


post-grouting pressure on skin
friction in a cohesive soil24

Fig. 7.12. Diagram of


multi-under-ream anchorage at
ultimate capacity20

the clay at the top end of the fixed length, and ca is the shaft adhesion (in the
range 0.3 to 0.35cu ) (in kN/m2 ). BS 8081 comments that under-reaming is
best suited to clays with an undrained cohesion cu greater than 90 kN/m2 .
Poor under-reams are likely if this value reduces to 60 to 70 kN/m2 , and
under-reaming becomes almost impossible if cu is less than 50 kN/m2 .
Under-reaming is difficult in low plasticity soils where the plasticity index is
less than 20.
BS 8081 suggests that to limit interaction between fixed anchors, the spacing
between anchors, centre-to-centre, should not be less than four times the
enlarged fixed anchor diameter; in practice a minimum spacing of 1.5 to
2.0 m is usual. The tolerances of drilling and borehole enlargement should
be considered when anchor spacing is decided, particularly with long
anchors.
284 Deep excavations

Table 7.6 Typical sizes and


characteristic strengths for Type of steel Nominal Specified Nominal
pre-stress tendon design20 diameter characteristic steel area
(mm) strength (kN) (mm2 )

Non-alloy steel
Wire 7. 0 60.4 38.5
7-wire strand 12.9 186 100
15.2 232 139
15.7 265 150
7-wire drawn strand 12.7 209 112
15.2 300 165
18.0 380 223
Low alloy steel bar
Grade 1030/835 26.5 568 552
32 830 804
36 1048 1018
40 1300 1257
Grade 1230/1080a 25 600 491
32 990 804
36 1252 1018
Stainless steel
Wire 7 44.3 38.5
Bar 25 491 491
32 804 804
40 1257 1257
a
This grade is not covered in BS 4486.

High grout pressures should be avoided for anchors founded at shallow


depth where subsoil movement due to these pressures would be detrimental
to existing structures or services.
The maximum values of grout/tendon adhesion recommended by BS 8081
for cement grout with a minimum compressive strength of 30 N/mm2 before
stressing are:
. 1.0 N/mm2 for clean plain wire or bar
. 1.5 N/mm2 for clean crimped wire
. 2.0 N/mm2 for clean strand or deformed bar
. 3.0 N/mm2 for locally noded strands.
Minimum recommended tendon bond lengths for cement or resin grouted
anchorages are 3.0 m where the tendon is homed and bonded in situ, and
2.0 m for tendons bonded under factory-controlled conditions.
For pre-stressed anchors the tendons consist of steel bar, strand or wire,
either singly or in groups. Pre-stressing steel is covered in sections 2 (for
non-alloy steel wire) and 3 (non-alloy 7-wire strand) of BS 589625 and in
BS 448626 (low alloy steel bar). Typical sizes and specified characteristic
strengths from BS 808120 are reproduced in Table 7.6. It is usual to proof-
load temporary works anchorages to a load equal to 1.25 times the required
unfactored working load and then lock-off the anchor at 1.1 times the
unfactored working load.
Design of the bearing plate beneath the stressing head requires care. A typi-
cal arrangement is shown in Fig. 7.13. The bearing is a thick steel plate, often
stiffened to span between wedge-shaped plates which transfer the load from
the stressing head through the bearing plate to the steel waling or soldier mem-
bers. Where the bearing head is bedded on to the concrete wall surface (as for
piled or diaphragm walls), BS 8081 recommends that the bedding mortar
Cofferdam design 285

Fig. 7.13. Anchor bearing plate,


detail from Berlin wall
installation

thickness should be limited to 10% of the plate width, with a maximum of


100 mm. The design bearing stress should not exceed 30% fcu (where fcu is
the characteristic strength, the 28-day compressive strength) of the grout.
Where concrete pads are used to bed the bearing plate on to the wall
surface and the depth exceeds 10% of the bearing plate width or 100 mm,
the pad should be designed to resist the stresses induced by the bearing
plate in the same way as a post-tensioned concrete block is designed as part
of a wall.
Methods of improving the load capacity of anchorages in both rock and
soil were discussed by Barley27 . Other developments in anchor construction
include the use of removable tendons. This is necessary where further develop-
ment at the curtilage of site is likely to involve underground excavations
that would be impaired by tendons used for temporary support. Specialist
anchor firms, including Keller in the UK, have developed systems whereby
tendons may be extracted after use with relatively low forces. Keller also
originally developed a single-bore multiple-anchorage system, and a typical
application in chalk was discussed by Barley et al.28 The system comprises
the installation of multiple anchorage units within a single borehole 100 mm
to 200 mm in diameter. The anchorages are located at staggered depths
in the borehole and each unit anchorage transfers load in a controlled
manner to a specific length of borehole. A comparison of the load transfer
mechanism of a normal and single-bore multiple anchor is shown in
Fig. 7.14. This method, which is particularly suited to anchorage in weak
rocks, clays, and mixed cohesive and granular materials rather than non-
cohesive soils alone, allows the utilization of a considerable length of borehole
as a fixed anchor. As shown, it can materially add to the load capacity of an
individual anchor; a limit of approximately 10 m is usual for fixed-length
anchors.
The design and installation of 10 000 anchorages in rock was discussed by
Barley29 who gave detailed information regarding anchor dimensions and
286 Deep excavations

Fig. 7.14. Comparison of the load transfer mechanism: normal and single-bore multiple anchors

working and test bond values from a large number of sites in chalk, mudstone,
siltstone and marls, sandstone and other rock including strong rocks such as
granites. The paper also referred to the hazards of downward movement
of sheeting caused by the vertical component of anchor loads and the
inadequate transfer of load by poor waling systems. Barley mentioned that
inclined twin-channel walings have proved to be the most trouble-free
system. Below is the summary of Barley’s paper.
Since Littlejohn’s initial correlation of standard penetration test (SPT)
values with a limited number of bond stress values in chalk some 20
years ago, a steady drive to substantiate and extend this approach and
to develop similar correlations in other rocks has continued. However,
to allow continual assessment of bond stress/SPT and bond stress/rock
strength relationships in order to estimate the safe capacity of rock
anchorages, the basic requirement for thorough site investigation opera-
tions prior to anchorage design must be satisfied. This requires SPT
values, core recovery percentages, RQD values, core logs, unconfined
Cofferdam design 287

compressive strength (UCS) and point load index values, all to a depth of
almost 10 m into the bedrock. In particularly weak and varied rock (that
is coal measures) where the presence of good anchorage rock (sand-
stones) may exist, the full thickness of these bands should be investigated
to allow economic design of efficient anchorages.
Chalk, initially considered to be a problematic anchoring rock, has
proved to be a very satisfactory founding medium, generally without
the requirements of under-reaming. Test anchorages using the described
installation techniques have confirmed a reliable correlation between
SPT values and ultimate bond stress with F (F ¼ bond stress/SPT
value) chalk factors at failure of 20 to 30.
Installation of over 1500 working anchorages, all without failure, has
confirmed a reliable guideline for a safe working bond stress of 5  SPT
value to 7  SPT value. Working loads in the range 400 to 600 kN in
medium grade chalks and of the order of 1000 kN in upper grades are
generally realistic.
Weak mudstones, siltstones, shales, marls and other weak fine-grained
rock continue to be rather unreliable founding strata for straight shaft
anchorages, unless particularly low bond stresses are considered.
Reasonable correlation exists between SPT and bond stress, albeit at
much lower values and over a much broader envelope than those exhib-
ited by chalk anchorages. F mudstone factors at failure generally range
from 2.4 to 6.0 with ultimate bond stresses in the 200 to 400 kN/m2
range. The use of an F mudstone value of unity for permanent straight
shaft anchorages, and 1.5 for temporary straight shaft anchorages, is
recommended for the initial estimation of working bond stresses. Work-
ing loads of shaft anchorages lie in the 150 to 350 kN range. The use of
under-reaming techniques to form the positive lock in weak mudstone
will allow mobilization of much greater loads and reduce failure rates.
Where enlargements are designed and spaced correctly they will reduce
the load loss due to creep. In stony mudstones some satisfactory shaft
anchorages have been constructed with temporary loads in the 600 to
800 kN range. Sandstones, when their properties and conditions are
adequately investigated and reported contribute highly to the success
of rock anchorages. Working bond stresses range from 200 to 1000 kN/
m2 , depending on particle grading and rock strengths. Although there
is very limited data to relate SPT values in very weak sandstones to
bond stresses, it is likely that the F sandstone factors will equate to, or
be greater than, F chalk values of 20 to 30. The few failures encountered
in sandstones have generally been attributed to artesian conditions
unknown at the time of anchor installation or to unexpected depths of
weathering. Generally, under-reaming is unnecessary in sandstone
anchorages and working loads well in excess of 1000 kN have been
frequently employed.
BS EN 1537: 2000 was prepared to address the installation, testing and
monitoring of permanent and temporary ground anchors where the load
capacity is tested. An annex to the standard describes alternative limit state
and serviceability limit state design principles in compliance with Eurocode
EC7. The Code describes the corrosion protection required for both tempor-
ary and permanent anchors. For temporary anchors, the Code requires a pro-
tection to the steel components which exhibit corrosion for a minimum design
life of two years. In summary this is fulfilled by:
. tendon bond length: a minimum 10 mm cement grout cover to the
boreholewall.
288 Deep excavations

. tendon free length: low-friction system to allow movement of the tendon


within the borehole by sealed or compound filled plastic or steel sheaths
. transition between anchor head and free length: free length sheath
sealed to the bearing plate/anchor head or a metal sleeve or plastic
duct may be sealed to the bearing plate
. anchor head where anchor head is accessible, a coating or tape protec-
tion; where inaccessible, a metal or plastic cap is required filled with
compound for extended use.
A longer period of protection or aggressive conditions require additional
protection in order to comply with the Code.

Tremied plugs and grouted bases


Where outer sheeting or walling to a cofferdam fails to achieve a cut-off due to
the depth of permeable strata, it may prove economical to seal the cofferdam
with a temporary concrete plug installed below the base slab of the permanent
works. This concrete plug would be installed below water by tremie methods
of placing following the bulk excavation of the cofferdam, with the cofferdam
flooded to existing ground level or, for river cofferdams, to the level of the
water outside the sheeting. This concrete plug would be designed to withstand
the head of water acting as an upward pressure on its underside. The weight of
the plug needs to balance or exceed this pressure in order that water within the
cofferdam above the plug can be safely pumped out. In wide cofferdams the
span of the plug and its effectiveness in withstanding the upward pressure
may be improved by the installation of tension piles or anchors which hold
the plug down. Figure 7.15 shows typical construction details.
The concrete plug in fact performs a dual role; in addition to preventing the
ingress of groundwater into the cofferdam, it also provides an effective prop
between the sheeting below formation level prior to the removal of water
from the flooded cofferdam, and therefore reduces design bending moments
in the outer sheeting or walls and reduces the loads in the lowest and the
penultimate frames.

Fig. 7.15. Cross-section of


cofferdam showing anchored
plug
Cofferdam design 289

Some difficulty may be experienced in obtained a completely dry seal to the


cofferdam with a tremied plug, and care is needed to avoid a ‘blow’ through
any tremied base concrete of low quality. The concrete plug should wherever
possible be placed in one continuous operation without disruption. Any small
leakage between the plug and the outer sheeting once detected is remedied by
local grouting.
The alternative to placing a concrete plug by tremie is the installation of a
horizontal curtain by injection or jet-grouting methods. These curtains can be
anchored vertically by tension piles installed prior to grouting, the piles
extending to formation level for later incorporation into the basement raft
of the permanent works. In a similar way to the concrete plug, the horizontal
curtain excludes (or largely excludes) groundwater from the excavation and
provides improved passive resistance to the outer sheeting when the excava-
tion within the cofferdam is made to the full depth, reducing the lower strut
loads and possibly even reducing the number of cofferdam frames. Typical
details are shown in Fig. 7.16.

Fig. 7.16. Grouted soil plug to


metro station excavation, Cairo
Metro, 1994 (from Tunnels and
Tunnelling
290 Deep excavations

Fig. 7.17. Installation of a


horizontal grout curtain,
Leipzig

Drilling in Leipzig for the installation of a horizontal silicate grout curtain


is shown in Fig. 7.17. The subsoils consist of river sands and gravels and two
forms of construction have evolved locally, a deep grouted curtain as shown
or a shallow jet grouted curtain restrained by tension piles.

Sheet piling: selection of section


Steel sheet piling is supplied to comply with EN 10248: 1995 in two grades as
shown in Table 7.7. Where steel sheet piling is used for the outer walls of a
land or marine cofferdam, the choice of section depends both on the flexural
strength needed to resist earth and water pressures and the strength needed to
resist driving stresses. Although these latter stresses may be reduced dramati-
cally by jetting through granular soils, this provision may not necessarily be
available. The CIRIA report4 provides some assistance to the designer in
assessing the pile section needed for driveability in both cohesionless and
cohesive soils. Table 7.8 from the report is based on experience of sheet
piles, approximately 500 mm wide, driven in panels. The two criteria for
sufficiency of the pile section during driving or vibration are that the head
of the pile should not be damaged unreasonably and the toe of the pile
should not be damaged and become declutched. Resistance to pile penetration
in granular soils is primarily at the toe and the effect of skin friction on a

Table 7.7 Properties of steel used in sheet piling18

Designation EN 10027 Classification Minimum yield Minimum tensile Minimum elongation


EN 10020 strength, ReH strength, Rm on a gauge length of
Steel name Steel number (N/mm2 )b (N/mm2 )b L0 ¼ 5:65S0 (A%)

EN 10248 S270GP 1.0023 BSa 270 410 24


EN 10248 S355GP 1.0083 BS a
355 480 22
a
BS: Base Steel.
b
The values in the table apply to longitudinal test pieces for the tensile test.
Cofferdam design 291

Table 7.8 Guide for the selection of pile size to suit driving conditions in granular soils4

Dominant SPT Minimum wall modulus Remarks


(N value) (cm3 /m)

BSEN 10 025 BSEN 10 025


Grade 430A, Grade 510A,
BS 4360 BS 4360
Grade 43A Grade 50A

0–10 450 Grade FE 510A for lengths greater than 10 m


11–20 450
21–25 850
26–30 850 Lengths greater than 15 m not advisable
31–35 1300 Penetration of such a stratum greater than 5 m not advisablea
36–40 1300 Penetration of such a stratum greater than 8 m not advisable
41–45 2300
46–50 2300
51–60 3000
61–70 3000 Some declutching may occur
71–80 4200 Some declutching may occur with pile lengths greater than 15 m
81–140 4200 Increased risk of declutching. Some piles may refuse
a
If the stratum is of greater thickness use a larger section of pile.

moving pile is not severe. Pile penetration is therefore a function of the relative
density of the soil at the pile toe. In contrast, for cohesive soils, resistance to
pile installation is primarily due to soil adhesion to the pile face and little resis-
tance occurs at the pile toe. The resistance to penetration in clay is therefore a
function of clay strength cu and the length of pile within the clay. Damage to
the pile toe is unlikely but resistance to pile buckling is necessary when the
driving resistance, in terms of clay adhesion, is overcome by a hammer of
sufficient capacity. Table 7.9 provides guidance for pile driving in cohesive
soils. A review of methods of sheet pile installation has been carried out by
the Technical European Sheet Piling Association (TESPA)30 .
The previous use of sheet pile installation equipment such as Taylor
Woodrow’s ‘Pilemaster’ machine was particularly suited to pile installations
in stiff clays. A more versatile type of machine, developed by Giken presses
a single pile from a reaction gained from previously installed piles. This

Table 7.9 Guide to selection of


pile size to suit driving Clay description Minimum wall modulus (cm3 /m) Maximum
conditions in cohesive soils4 length (m)
BSEN 10 025 BSEN 10 025
Grade 430A, Grade 510A,
BS 4360 BS 4360
Grade 43A Grade 50A

Soft to firm 450 400 6


Firm 600–700 450–600 9
Firm to stiff 700–1500 600–1300 14
Stiff 1600–2500 1300–2000 16
Very stiff 2500–3000 2000–2500 18
Hard (cu > 200) Not recommended 4200–5000 20
Note. The ability of piles to penetrate any type of ground is also a function of attention to good
pile driving practice and this table assumes that this will be the case.
292 Deep excavations

press is capable of installing sheeters into some cohesionless soils. Installation


in stiff clays is assisted where necessary by lubrication with water or by water
jetting. Some risk of loss of soil may occur in the use of jetting in loose soils
such as fine sands and silts. During jetting in stiff clays there may be some
risk of clay softening, although Giken states that there is no evidence of
this in the long term. (Traditional use of jetting in stiff clays to assist sheeter
penetration has sometimes been precluded in the last one metre of pile pene-
tration for this reason.)
It should be noted that Giken presses may not be able to install standard
corner sheet piles and special corners and junctions need to be fabricated.
Corners are installed by the Giken press using two dummy piles, usually
shorter piles, which are subsequently extracted.

Design of the bracing


Walings
Where walls or sheeters span vertically, walings are needed to transfer loads
from the sheeting to the struts, which provide the bracing, or to the anchors,
which retain the sheeting. The walings need not be continuous as, for example,
in hammer head struts used against diaphragm wall panels where separate
waling reinforcement may be included within the panel reinforcement of the
wall. Alternatively, with anchored diaphragm walls it is common to incorpo-
rate waling rebar steel to the full panel width without external walings. Where
secant pile walls are used in cofferdams, and where every pile or alternate piles
are anchored walings may not be necessary. Common waling arrangements
are shown in Fig. 7.18.
Where steel walings are used and subjected to heavy loads it may be
convenient to use steel beams in pairs in order to provide adequate width
on which to seat the bracing struts. It is often convenient to weld end plates
to each length of waling to connect them together. It is vital that where
rakers or sloping struts are used the tendency for the waling to turn on its
support must be resisted. Figures 7.19 and 7.20 show a typical detail.
Where anchors are used with walings, the spacing between steel beams must
be sufficient to accommodate the inclined tendon between them, or the pair
of beams must be inclined, with an adequate gap.
Where steel sheet piling is used it is usual to make the walings continuous
over two supports. Unless the piling can be driven to good tolerances in
vertical and horizontal alignment it is prudent to allow walings to cantilever
mid-way between struts without connecting one to the other. Where toler-
ances are likely to be well maintained, it is advantageous to connect the
ends of walings behind the incoming strut (Fig. 7.21). Where walings are
continuous over two spans and joined behind struts the design moment is
WL=10, but where they cantilever to half span the design moment becomes
WL=8.
The values of waling load for limit state design as calculated using mobili-
zation factors from BS 8002 or the ULS values using the partial factors
applied to soil parameters for moderately conservative parameters, load
case A (or the more onerous load case B for the worst credible parameters)
as detailed in Chapter 5, are used as design values without further factoring.
Comparison should always be made with waling loads calculated by empirical
methods.
Where steel sheet piles are braced by steel walings any irregular alignment
of the steel piles is rectified by steel packers or hardwood wedges. Where the
alignment is particularly poor, concrete infilling can be used between the
waling and the sheeters. If diagonal struts transfer longitudinal thrust into
the waling, the waling must be designed to take both this thrust and bending
Cofferdam design 293

Fig. 7.18. Typical waling


details, steelwork and
reinforced concrete

stresses due to the span between struts. It may be necessary to weld steel angles
to the back of the walings prior to erection in order that horizontal acting
shear keys can be formed by concreting the leg of the angle into the pan of
the sheet pile. This will be required if the available length of waling is short
and therefore the frictional resistance between waling and sheet pile is
insufficient to transfer the thrust (Fig. 7.22). Alternatively steel shear plates
occupying the whole pile pan can be welded to the waling and the face of
the sheet pile.
Where heavily loaded struts or highly loaded anchors bear on steel walings
it will be necessary to use web stiffeners to avoid web buckling of the waling. A
typical detail is shown in Fig. 7.23.
294 Deep excavations

Fig. 7.19. Typical steelwork


detail at junction of rakes and
waling with bracing to prevent
rotation of waling

Fig. 7.20. Typical connection


detail to avoid rotation of waling
at junction with strut

Where reinforced concrete diaphragm walls use internal walings within


the reinforcement cages, such cages must be reinforced for shear where
the anchor or strut bears on the waling but must avoid impeding the
passage of tremie tubes through the waling beam reinforcement. Typical
arrangements of both steel and reinforced concrete walings are shown in
Fig. 7.24.
Walings for use with inclined anchors may themselves be inclined, with the
anchor plate bearing directly on the face of the waling. Gusset plates welded
to each pile face incline the waling. Alternatively, the waling may bear directly
on to the sheeting with the bearing plate inclined. Fig. 7.25 illustrates the
arrangements.
Cofferdam design 295

Fig. 7.21. Detail of typical


waling connection behind strut
in light cofferdam steelwork

Fig. 7.22. Sheet piled


cofferdam construction: shear
keys at rear of waling transfer
thrust from diagonal strut into
short waling length and through
sheeters into soil at rear of piles
296 Deep excavations

Fig. 7.23. Typical detail of web


stiffener in waling to avoid web
buckling

Passive anchors
The design of anchored walls is described in Chapter 5. Tie rods from the
sheeting wall are anchored to deadmen, an anchor wall or an A-frame of
driven piles. Anchor walls are designed on the basis that net available passive
resistance is equal to passive pressure less active pressure. No allowance
should be made for surcharge being available in front of an anchor wall in
this calculation, and wall friction should be ignored because of the risk of
vertical movement of the wall to the detriment of this friction. Tie rods,
based on a factor of safety of 1.5 to 2.0, are designed using the following
working stresses:
. Mild steel (BS 4360 grade 43A or BS EN 10025 grade Fe 430): 110 N/mm2
. High-yield steel (BS 4360 grade 50B or 50C or BS EN 10025 grade
Fe 510): 140 N/mm2 .
Tie rods can be housed in the bottom of pipework to avoid the effects of fill
settlement and can be wrapped in Denso tape to reduce corrosion.

Struts
The most likely collapse mechanism of a braced cofferdam is the buckling of
its strutting, beginning with the lowest frame and continuing progressively to
the highest frame. The collapse of the lowest frame may be associated with
inadequate penetration of the sheeting and passive failure below formation
level; it may occur during extreme loading, such as high water for a river
cofferdam or high waves in storm conditions for a cofferdam in open water.
It may also be associated with poor workmanship in bracing or piling, or
both. Collapse due to failure of walings or flexural failure of the sheeting
Cofferdam design 297

Fig. 7.24. Waling systems used


with diaphragm walls

itself is much less likely. The extra care that is necessary in the design, detailing
fabrication and fixing of cofferdam struts is self-evident. Any economy in the
design of struts, especially long struts, can be false. It is essential that a design
check is made against progressive collapse. This is best done by removal of
any one strut and an analysis of the safety factors against collapse of the
remaining structure.
Struts are generally laterally spaced so that excavation grabs may pass
safely between them, so that construction materials can be lowered between
them, and so that lower frames, when struck, can be threaded out through
the frames above. Struts are therefore usually between 4 and 6 m apart. The
most versatile strut is the steel tube because of its efficiency in buckling
(although less so in bending) and its smooth outline which precludes snagging
from muck grabs and construction materials as they are lowered. Other
sections used include steel universal column sections, battened pairs of steel
beams and box piles. Long tubular struts used in the cofferdam works at
298 Deep excavations

Fig. 7.25. Waling details with


inclined anchors, shown with
typical Berlin wall

the Dartford Creek Barrier are shown in Fig. 7.26. Battened steel beams are
shown in use on an earlier Thames cofferdam in Fig. 7.27.
Economy in the use of struts in a rectilinear cofferdam can often be achieved
by using diagonal compression members in the corners, but care is needed
to avoid obstruction to construction plant, such as piling plant, and installing
permanent works in the corners.
It is essential to support struts adequately at the walings at each end of the
strut. Steel location angles may be usefully welded to the strut end plates prior
to bolting or welding into position. It is vital that the struts are square to the
walings in plan and the end plates bear uniformly on the walings to avoid
eccentric loading. (It is, nevertheless, worth checking the effect of eccentricity
of the thrust in the strut by, say, 10% of the strut width or depth in each
direction.) The effect of materials and plant loads placed on the strut
should be added to the self-weight of the strut in considering the combined
effect of compression load and bending.
Design values for struts using limit state design are those similar to walings
and are calculated in a similar way. If the mobilization factors are applied as
specified in BS 8002 no further factors are required on these values for limit
Cofferdam design 299

Fig. 7.26. Tubular steel struts


used to brace cofferdam at
Dartford Creek, UK (courtesy of
AMEC)

Fig. 7.27. Battened steel beams


used as struts and braces on
River Thames cofferdam,
London

state design. If the design process as detailed in Chapter 5 is used, the


calculated loads, based on ULS case A moderately conservative values are
used without further factoring (if ULS case B, worst credible parameters
appropriately factored are more severe than case A these should be used).
Comparison should always be made with strut loads calculated by empirical
300 Deep excavations

Fig. 7.28. Plan of cofferdam


construction showing effective
lengths of struts used in design

methods. The effective length of the strut is assumed to be equal to its actual
length unless braced laterally or vertically at midspan (Fig 7.28). Temperature
effects should be considered; the effect of sunlight on long steel struts can be
reduced by painting them white.
It is essential that cofferdams are stiffened along their width and length by
diagonal bracing to avoid risk of collapse into a lozenge shape. Typical steel-
work bracing examples are shown in Fig. 7.29.
The vertical spacing of struts depends on both the strut capacity and the
flexural strength of the sheeters or walling. In practical terms, the minimum
frame spacing needs to be sufficient to allow mechanical excavation plant to
pass under the frame prior to placing the next frame below and to allow
sufficient excavation of the cofferdam. The placing of the lowest frame in a
multi-frame cofferdam is the period of greatest risk to the bracing. At this
stage the excavation is usually close to final formation level. It is likely that
the next highest frame will be highly loaded and the factor of safety against
passive failure will be at its lowest for all stages of the excavation. The bending
stress and deformation of the sheeting will be at their highest values, and
although two-dimensional analysis will be unable to show any benefit, it
will be a practical advantage to carry out excavations for the lowest frame
in short lengths where stresses in the soil, bracing and sheeters are very high
at this stage. It is essential that all bracing components are fabricated ready
for installation in this bottom frame, to avoid the cofferdam remaining
unpropped at the lowest level for any lengthy period, especially at high
tides, where these apply. It should be noted that the use of anchors on the
bottom frame does not allow the sheeters to be secured speedily at this critical
lowest level since the anchor grout requires time to achieve sufficient strength
to pre-stress and thus secure the tendons.
Cofferdam design 301

(a) (b)

(c)

Fig. 7.29. Diagonal bracing to cofferdam steelwork: (a) tubular steel bracing between walings, cofferdam to north pier, Forth
Road Bridge; (b) light diagonal bracing between walings of circular cofferdam, Severn Bridge west pier; (c) diagonal bracing,
Dartford Creek Barrier cofferdam (courtesy of AMEC)

The use of thickened, and where necessary reinforced, blinding slabs to


act as props in the period before the base slab or raft is cast, may prove
advantageous especially where the span of such props is small.
Where reinforced concrete diaphragm walls are used to retain soil at the
curtilage of a cofferdam it is possible to use the strength of the wall panel,
designed as a plate, with support from passive soil resistance at the bottom
302 Deep excavations

Fig. 7.30. Typical reinforcement detail for diaphragm wall panel (courtesy of Bauer)
Cofferdam design 303

of the panel and point supports from hammerhead struts at the vertical joints
between panels. This arrangement obviates the need for either external or
internal walings, but high shear stresses may occur within the panel near
the stub end of the strut and require shear reinforcement. The use of
reinforced concrete struts for this arrangement suffers from the comparatively
high self-weight of the units, the difficulty of altering the reinforced concrete
section for use elsewhere, and the high cost of disposal.

Walling details
A typical reinforcement detail for an anchored diaphragm wall panel is shown
in Fig. 7.30. Items that need to be addressed in detailing the wall reinforce-
ment include:
. lifting and bracing steel
. method of joining sections of cages
. method of lifting
. lateral spacing between reinforcement cages of adjacent panels
. type of box-out for junction with floor slabs
. details of starter steel for floor slabs
. access for concreting tremie pipes
. use of couplers
. detail of access through panel for ground anchors
. inclusion of reinforcement for walings
. minimum spacing of reinforcement steel and provision of cover
. consideration of water bars in vertical panel joints.

Sheeting details
Although Larssen U-type sheet piles possess certain advantages over Froding-
ham Z-types (in stacking, driving, etc.) care is needed with the Larssen section
which has clutches along the centre-line, or neutral axis, of the section. In
certain conditions, described in BS 8004, insufficient friction may develop
within the clutch to develop the full flexural strength of the section. These
conditions are:
(a) where the piling passes through very soft clay or other weak material
(b) where the piling is prevented by rock from penetrating to normal cut-off
depth
(c) where the piling is used as a cantilever, or if it has a substantial
cantilever height above the upper waling
(d ) if backfill is placed on one side of the piling after it is driven.
In these circumstances it may be necessary to intermittently weld pile pairs at
their interlocks to develop full strength31 .

Circular cofferdams
Circular plan shape cofferdams benefit from the hoop compressive stresses
that are induced in waling and walling sections and the ease of providing
steel or concrete to resist these stresses. The lack of internal strutting provides
unrestricted working space and the circular section improves efficiency in
terms of hydraulic performance and scour in river and marine works. It is
essential that the induced compressive stresses and earth and groundwater
pressures are uniformly applied to the circumference of the cofferdam.
These conditions may not apply where ground conditions vary locally, in
sloping ground, where local surcharge loading could apply, or where ground-
water is flowing swiftly round one side of the cofferdam to the other. As
referred to before, the economy of the cofferdam relies on the extent to
304 Deep excavations

which the permanent works occupy the circular space provided. Bridge piers
and pumping stations can use the circular space particularly well.
Several types of construction can be applied to circular cofferdams: steel
sheet piling spanning vertically between circular walings in reinforced con-
crete or steel, and diaphragm wall units or bored piles also spanning vertically
between walings or diaphragm walls spanning vertically or circumferentially.
Although considered as a true circle in theory, some deviation from this
shape may occur during construction and subsequent loading of circular
walings. Care should be used in the application of the theoretical formula
for ring beams due to Timoshenko and Goodier32
KEI
Wu ¼ 3 ðkN=mÞ ð69Þ
R  105
where Wu is the ultimate radial waling load, K is a coefficient depending on the
stiffness of the retained medium, R is the radius of the cofferdam (in m), E is
the Young’s modulus for the waling material (in N/mm2 ), and I is the moment
of inertia about the vertical axis (in cm4 ). A value of K ¼ 3 is used where the
retained medium is water, as in a marine cofferdam, and increasing values
apply with increasing stiffness, which becomes soil stiffness for land coffer-
dams. It is common, however, to use a value of K ¼ 3 for both land and
marine conditions, and applying a factor of safety of 2.0 the safe radial
waling load W (in kN/m) becomes
1:5  EI
W¼ 3 : ð70Þ
R  105
Again, the CIRIA report4 suggests that ring waling stiffness is highly impor-
tant and recommends the use of the empirical rule for reinforced concrete
walings d ¼ D=35, where d is the depth of the ring beam (in m) and D is
the diameter of the inner face of the cofferdam sheeters (in m). Table 7.10
gives safe loads for reinforced concrete walings of specific size and reinforce-
ment for cofferdams of varying diameter18 . Care is needed to avoid uneven
loading of the waling either at the top or lowest level of the waling due to
non-verticality of sheeters. Any uneven distribution of applied load over the
depth of the waling could induce torsional stress within it.

Table 7.10 Safe load (in kN/m)


in reinforced concrete waling to Diameter of Size of waling d  b (mm):
circular cofferdams18 cofferdam,
D (m) 450  300, 600  400, 750  500, 900  600, 1050  700,
six 20 mm ten 20 mm ten 25 mm fourteen 25 mm twelve 32 mm
dia. bars dia. bars dia. bars dia. bars dia. bars

5 280 500
10 140 250 390
15 90 165 260 375
20 125 195 280 380
25 155 225 305
30 185 255
35 215
Based on: (i) permissible compressive stress in concrete not to exceed 5.2 N/mm2
(ii) waling load (in kN/m) ¼ 1.5EI/105 R3 ; E ¼ Young’s modulus, for concrete,
E ¼ 13 800 N/mm2 , I ¼ moment of inertia about xy axis (cm4 ), R ¼ cofferdam
radius (m)
(iii) depth of waling d to be not less that D=35
(iv) need to check tension in waling beam if sheet piles distort under load and con-
centrate load on top and bottom of waling beam.
Cofferdam design 305

The design of circular cofferdams with walls that span vertically between
walings can therefore be made by considering the walling to span between
the circular walings and the usual multi-braced wall analytical methods,
spring programs can be used, the circular walings being designed according
to the methods of Timoshenko and Goodier or the CIRIA report. For circular
cofferdams with walls that are designed to withstand hoop compression, and
especially those with large openings in the walls or uneven loading, a more
refined approach is necessary. The risk of buckling of the wall panels must
be checked. A three-dimensional finite element approach to the design of a
large circular cofferdam built with diaphragm wall panels with a reinforced
concrete lining wall was described by Kirmani and Highfill33 . The lining wall
was cast as a series of ring beams successively with depth as the excavation
proceeded. In the analysis, the load-carrying mechanism of the diaphragm
wall-lining wall structure was somewhat arbitrarily divided as follows.
. Hoop compressive stresses would be resisted by the combination of
diaphragm wall and lining wall in proportion to their thicknesses.
. Horizontal bending would be resisted by the lining wall.
. Vertical bending would be resisted by the diaphragm wall.
. Buckling of the cylindrical cofferdam would be checked with the
combination of the diaphragm wall and the lining wall.
The peripheral walls of circular land cofferdams may be built in contiguous or
secant piling or diaphragm walling. Such walls are economical where they are
used as temporary support during construction and as permanent walls to the
final structure. Secant or contiguous piles, spanning between ring walings in
reinforced concrete, are limited to 30 m or so in depth, depending on
ground conditions. Maintaining the secant connection between adjacent
piles may prove difficult in some ground conditions and require the use of
large piling plant and casing oscillators.
Diaphragm walling has advantages and disadvantages over bored pile
walls. In depth, for instance, with modern reverse circulation trenchcutters,
diaphragm walling can be installed in a range of soils and soft rocks to
depths in excess of 50 m. In shafts of small diameter, the length-to-breadth
ratio of each wall segmental panel may be such that the wall is kept in
hoop compression by earth and groundwater pressures and little or no
bending occurs within the wall panel. In these circumstances, the wall requires
minimal reinforcement; in fact the concrete strength may be adequate to
withstand the hoop compression. In practice, it is best to consider the risk
of at least one panel joint failing to transfer all its load to its neighbour
because of poor panel verticality. Sufficient vertical steel and space should
be allowed for the introduction of an emergency waling.
Where the circular cofferdam diameter is larger, however, the deviation
angle between the centre-line of each diaphragm wall panel is reduced and
horizontal bending occurs within the panel in addition to hoop compression.
In this case it is generally more economical to span the wall vertically and
introduce temporary walings. In top-downwards construction, permanent
floor slabs provide both temporary and permanent support to allow the
wall to span vertically between them. Figure 7.31 shows a simplified method
of stress analysis for diaphragm walls without walings designed to span
circumferentially.
Care is needed when using temporary circular stop ends to ensure circular
walls of limited diameter remain properly connected at the joints. Flat-section
temporary stop ends cannot be used. Cut joints are needed below temporary
stop ends for deep walls. Full-depth cut joints are suitable but only where the
deviation angle between panels is less than 12 to 158.
306 Deep excavations

Fig. 7.31. Circular diaphragm wall design: calculation of hoop compression and bending stress from a three-pinned arch
Cofferdam design 307

Fig. 7.31. Continued


308 Deep excavations

Practical difficulties may also be caused by instability of the last panel


where the construction is in water-bearing ground. Groundwater tends to
rise within an enclosure and temporary wellpointing may be necessary to
avoid a differential head of groundwater between the inside and outside of
the closure panel.

Sheet pile walls across dock entrances


There are several means of building a cofferdam across the entrance to lock or
dock construction to exclude external river or sea water in order to excavate
the floor of the dock or replace existing cills and gates. For new construction,
cellular cofferdams or twin, parallel-walled cofferdams may be used. To
exclude water from an existing dock, raking struts to the dock floor or walings
from the sides of the dock, with diagonal struts to divide the waling span, are
useful options. For an existing dock entrance it is often feasible to span
walings across the whole entrance width by driving sheet piles to an arc
circular in plan, braced by arch walings of steel or reinforced concrete,
using the existing walls at the entrance as ‘abutments’ to the arch walings.

Estimating costs of temporary cofferdams


Many design curves have been published for cantilever and braced cofferdam
construction. Those due to Packshaw (Fig. 7.32) show the section modulus of
the piling and the number of bracing frames of normal construction in average
soil conditions; bending moments induced in cantilever cofferdams for various
heights and conditions; and the section moduli, waling loads and maximum
penetration depths for one- and two-frame cofferdams in cohesionless soils.
The greatest contribution of these graphs may be to prevent serious errors
in the estimation of sheeting and bracing requirements during cost estimating
by the engineer or contractor.

Fig. 7.32. (a) Curves for


estimating purposes: number of
frames and sheeting section
modulus for average
cofferdams as depth
increases49
Cofferdam design 309

Fig. 7.32. (b) Section modulus of cantilever sheeting as depth increases for land and water cofferdams in cohesionless soil
310 Deep excavations

Fig. 7.32. (c) Section modulus


of cantilever sheeting as depth
increases, cantilever piles in
cohesive soil

Double-wall Double-wall cofferdams are gravity structures consisting of twin parallel lines
cofferdams of sheeters driven below dredge level, tied together at one or more levels by
steel ties and filled with selected material, preferably cohesionless soil. The
width-to-height ratio of the structure is at least 0.8, and it is usual to place
a berm of granular soil on the inside face to extend the drainage path of
water passing beneath the cofferdam and avoid piping near the inside line
of sheeters. The stability of a double-wall cofferdam depends on the strength
of the sheeters and the ties, on the shear strength of the fill material and the
Cofferdam design 311

Fig. 7.32. (d) Section modulus,


waling load, cofferdams with
one bracing frame,
cohesionless soil

soil at foundation level. This type of cofferdam is not suitable where strong
bedrock occurs at shallow depth below formation level as it is necessary for
the sheeters to penetrate sufficiently to avoid passive failure in front of
them. It may be necessary to reduce the level of water in the fill material
between the piles to increase the effective shear strength of the fill and
reduce the pressure on the sheeters. Submersible pumps may be needed for
this drawdown. In any case, it is essential that sluices are provided on the
inner line of piles to reduce the level of the phreatic surface as much as
312 Deep excavations

Fig. 7.32. (e) Section modulus,


waling load, cofferdams with
two bracing frames,
cohesionless soil
Cofferdam design 313

possible. Where the length of the cofferdam exceeds its width by four or five
times, it is usual to drive a cross-wall of sheeters to connect the inside and out-
side line of sheeters. This assists craneage access for construction and reduces
the consequences of failure by forming compartments of restricted size.
Cofferdams of this type find application at the entrances to locks and docks,
both for new construction and the maintenance of existing structures, such as
remedial works on dock gates and cills. The major load on the cofferdam is the
head of river water on the outside sheeters. As discussed earlier, wave heights,
the effect of water levels caused by passing vessels and the risk of collision by
river traffic, must all be considered when assessing this horizontal loading. A
concrete slab cover should be provided on the top of the cofferdam where
waves can overtop the outside sheeting and increase the water level in the
filling.
The possible modes of failure were reviewed in Chapter 6. Each of these
failure modes is addressed below for the design of double-wall cofferdams.
(a) Tie rod design and water pressure. Where the sheeters do not achieve
penetration below the minimum required for free earth support, due
to earth and water pressures within the retained fill, the structure will
act as a gravity structure when loaded with the outside head of water.
In this situation the sheet piles should be designed using at-rest pressures
for the fill (using the coefficient for each pressure at rest K0 for the filling)
because deformation of the sheeters is restricted by the ties and it is
intended to restrict movement at the head of the cofferdam. Walls
should be designed for the most severe assumptions of internal water
pressure. Where, for instance, hydraulic filling is used to place sand
backfill between the sheeters, the design water level within the cofferdam
should be considered vis-à-vis the rate of drainage possible from sluices
of flap valves on the inside face. In the worst situation, the water level
within the filling may reach the level of the top of the sheeters if the
rate of pumping the fill is high and the rate of drainage is low.
(b) Sliding. Resistance to sliding is provided by the passive resistance of the
soil on the inside face of the cofferdam, the shear strength of the sheeters
and the frictional resistance beneath the material filling the cofferdam.
Referring to a typical cross-section (Fig. 7.33):
resistance to sliding ¼ Pp þ S1 þ S2 þ Ssoil
Pp þ S1 þ S2 þ Ssoil
and factor of safety against sliding failure ¼ :
Pa
ð71Þ
(c) Shear failure of the cofferdam filling. To establish the efficiency of the fill
material between the sheeters in resisting any tendency for the top of the

Fig. 7.33. Sliding and resisting


forces in double-wall
cofferdam, vertical cross-
section
314 Deep excavations

Fig. 7.34. Stability analysis of double-wall cofferdam, Gallions Lock entrance, London34
Cofferdam design 315

cofferdam to move horizontally, the stability analysis should use a force


polygon as suggested by Bishop34 in the discussion to Packshaw’s
paper. This analysis was first used for a double-wall cofferdam at the
entrance to Gallions Lock in London. The analysis expresses the
factor of safety against shear failure between blocks of filling (as
shown in Fig. 7.34) as the ratio between the tangents of the angle of
shearing resistance for the fill and the mobilized angle of shear on the
vertical face of the blocks. Friction between the filling and the internal
face of the sheeters is also taken into account, and the factor of safety
against slippage between sheeters and fill is expressed as the ratio of the
tangents of the angle of wall friction and the mobilized wall friction as
shown on the force polygon. Bishop pointed out that the tie-rod pull did
not appear as a term in the analysis because its value is small and it was
safe to omit it in examining the factor of safety against shear between
the vertical elements. There is no other reason, however, for excluding
this term, the value from the analysis of the sheet pile walls being
included in the force polygon. Factors of safety of the order of 1.5
would be regarded as satisfactory for this stability analysis.
(d ) Overall stability of the cofferdam structure. A method of checking the
overall stability of the double-wall cofferdam is described in reference
5. Figure 7.35 shows the cross-section of a cofferdam founded on soil

Fig. 7.35. Calculation of double-wall cofferdam stability: (a) cofferdam embedded in load-bearing soil, stability analysis;
(b) investigation of the anchorage of the inner line of sheet piles5
316 Deep excavations

Fig. 7.36. Flow net for


double-wall cofferdam,
calculation of exit gradient35

with free earth support to the outside sheeters. The failure plane at the
base of the wall is approximated by a logarithmic spiral centred at point
0. The factor of safety against overturning is then the ratio of overturn-
ing and resisting moments about the centre of the most unfavourable
failure surface. The minimum factor of safety must be at least 1.5. If
this is not obtained in the trial, the required stability may be obtained
by increasing the width, improving the quality of the fill or deeper
driving of the sheeters, or a combination of all three.
(e) Piping at the inner face of the cofferdam. Where the cofferdam is
founded on relatively permeable material the flow of water from the
riverside to the landward side beneath the cofferdam may cause
piping failure near the inside line of piling. This condition is examined
by use of a flow net, as illustrated in Fig. 7.3635 . A berm of granular
material, including where necessary a blanket of filter material, is
effective in increasing the drainage path from the river side and thereby
preventing piping.
( f ) Bearing capacity failure at founding level. The bearing capacity of the
soil below the fill placed between the sheeters must be sufficient to
support the weight of the filling and resist the lateral force due to the
Cofferdam design 317

riverside water head and impact forces due to waves and accidental
collisions. Where soft clays or silts exist at this level and a stronger
soil underlies the soft material, excavation between the sheeters
should improve the bearing prior to placing the fill.
(g) Scour on the outside of the cofferdam. Where river currents cause risk of
scour, the outer face of the cofferdam must be protected to avoid
instability of the outer line of piles. This protection may consist of
rock or precast concrete blocks placed against the outer face, or grouted
mattresses laid on the river bed.
Lateral deformities at the head of a double-wall cofferdam are frequently
greater than similar gravity structures when loaded with river water at a
high elevation on the outside face. These deformations are associated with
mobilization of the shear strength of the cofferdam filling and the friction
between the filling and the inside surface of the sheeters. The deflection of
the top of the riverside sheeting at the Gallions Lock cofferdam was
350 mm after it had been subjected to the full water load; the height between
high water and the inside cill was 12.8 m. A similar displacement was
measured at the head of the St Katherine’s Lock cofferdam described in
Chapter 6. The use of strongpoints, as was shown in Fig. 6.42, minimizes
such horizontal movements.

Cellular cofferdams The construction of cellular cofferdams was described in Chapter 6. Their use
extends to piers, dolphins and breakwater structures, but cellular cofferdams
also provide economical soil retention temporarily and permanently in deep,
wide excavations and exclusion of river and sea water from deep excavations
for lock, gravity dock and similar massive, large plan area structures. These
gravity structures, depending on the weight of the retained fill and the tensile
strength of the sheeters that retain it, are often used where pile driving condi-
tions preclude deeper sheet piling for braced cofferdam construction and
where internal bracing or sheeting is unacceptable or impractical. The most
popular plan shapes are circular cells with one or two connecting arcs, and
diaphragm cells with outer arcs and straight cross-walls. Circular cells consist
of independent self-supporting structures, the diameters of which are a func-
tion of the interlock strength of the sheet pile secton; the greater the supported
height of soil or water, the larger the cell diameter and, in turn, the greater the
pile interlock tension. These factors encouraged the development of a straight
web pile with three-point contact at the interlock to give increased interlock
tensile strength compared with single-point contact interlocks.
Before design can begin the necessary data must be collected. For a
cofferdam within a river such data would comprise:
. tidal data and rate of flow, and prediction of both during the design life
of the cofferdam
. scour behaviour
. river bed profiles
. soil profile and test data to define strength, permeability and consolida-
tion properties
. borrow areas for suitable filling, and spoil disposal areas
. water quality for cofferdams with a long design life
. previous site use and obstructions
. collision risk from river traffic
. river regulations; permission needed for cofferdam construction.
For maritime works, wind and wave data would be needed, including wave
height and period, tide dates and storm risk; also ship collision assessment.
318 Deep excavations

Fig. 7.37. Modes of failure of cellular cofferdams11

The design of cellular cofferdams was reviewed by Dismuke36 and summar-


ized in BS 634911 . Design of a cellular cofferdam should address the following
modes of failure (shown in Fig. 7.37):
(a) excessive tilting or rotational failure on a curved rupture surface at or
near the base of the cell: internal stability
(b) interlock and connection failure
(c) instability of base and sliding
(d ) loss of cell fill due to piling rise
(e) overturning.
Each of these failure modes is now addressed in more detail.

Internal stability
The following methods of analysis of cell stability were compared by
Dismuke.
Cofferdam design 319

Fig. 7.38. Equivalent rectilinear


width and design geometry for
cellular cofferdams of circular,
diaphragm and cloverleaf
cells36

(a) The vertical shear method developed in the late 1930s and 1940s,
proposed by Terzaghi37 and later developed by US Tennessee Valley
Authority engineers38 .
(b) The horizontal shear method developed by Cummins39 . This method
was introduced because of inconsistencies in the vertical shear
method. The method assumes that horizontal shear planes develop
within the cell fill and implies that fill on the unloaded side of the cell
could be reduced without affecting stability. This conclusion is not
practically sound and should not be used in design to reduce fill levels
within cells.
(c) Methods due to Brinch-Hansen40 and described in detail by Ovesen35 .
Two variations, known as the equilibrium and extreme methods,
resulted from observations that a circular rupture surface occurred at
the base of a model double-wall cofferdam which was loaded to failure:
the extreme method is recommended in reference 5 and is the basis of
design for cell diameter or width described in the following part of
this section.
Before describing the extreme method due to Brinch-Hansen it is necessary to
refer to the method of changing a cellular cofferdam plan shape to a rectilinear
shape to reduce computation. Figure 7.38 shows how this can be done.
The extreme method assumes that the cell is filled with granular material
and is founded on a rock or granular soil base. To simplify the rather compli-
cated calculations of the internal forces on the rupture line at the base of a
cofferdam founded on rock, the kinematically true circular rupture line is
substituted by a logarithmic spiral satisfying the polar equation r ¼ r0 e tan  .
Such a spiral has a characteristic that its radius vector at any point makes
an angle  with the corresponding normal. In cohesionless soil, with an
angle of shearing resistance , the resultant of all internal forces within the
spiral will thus be directed towards the rotation point of the spiral. This
calculation for cofferdams on rock involves the following steps (Fig. 7.39(a)):
(a) generate log spiral locus line
r ¼ r0 e tan  ð72Þ
for angle of shearing resistance  with spiral through the feet of the
walls.
(b) compute the external and gravity forces ðP; bH; Wp ; Wf Þ and reactions
ðSh ; Sv Þ
320 Deep excavations

Fig. 7.39. Stability analyses for


cellular cofferdams by the
extreme method: (a) founded on
rock (Wp ¼ weight of piling.
Wf ¼ weight of fill below
rupture surface); (b) founded on
rock overlain by soil; (c) founded
on soil, additional check for
deep embedment with concave
failure surface

(c) take moments about the rotation point of the spiral and find the factor
of safety
M
f ¼ stabilize ð73Þ
Mdisturb
(d ) Find the critical moment, that is the moment for which f is a minimum,
by changing the position of the spiral. If minimum f is greater than
unity the cofferdam is stable, although a minimum value of 1.5 is
required for design acceptance.
If the cofferdam is founded on rock which is overlain by soil (Fig. 7.39(b)) or if
the cofferdam is founded in soil, the disturbing forces are increased by the
active earth pressure on the outside of the cofferdam and reduced by the
passive earth pressure on the inside. Since deformation will be small it is
usual to limit passive pressure to the at-rest value K0 ¼ 1 for sheeting with
shallow embedment, and to calculate Kp setting wall friction equal to zero
for sheeting with deep embedment.
Where sheeting is driven deep to provide stability, a check for concave
failure planes is necessary (Fig. 7.39(c)). The spiral is then located so that
its centre of rotation does not lie beyond the line of action of passive force
Pp with the angle of wall friction equal to zero.

Interlock and connection failure


The cell hoop force outside the connecting arcs and the hoop stress may be
calculated from
ta ¼ p  ra ð74Þ
Cofferdam design 321

Fig. 7.40. Calculation of hoop


forces in circular cofferdam
cells: (a) circular cell
cofferdam; (b) circular cell
bulkhead36

and
tc1 ¼ p  rc ð75Þ
where ta is the hoop or interlock force for connecting arcs, tc1 is the hoop or
interlock force for cells outside arcs, and p is the lateral unit pressure (taken as
earth pressures at rest, K0 ¼ 1  sin , at the base of the excavation).
Dismuke36 pointed out that the greatest interlock force, located just inside
the arc connection, is frequently overlooked. The cell hoop force at the arc
connection is
tc2 ¼ pL sec  ð76Þ
where tc2 is the circular cell hoop or interlock force between arcs, L is half
the centre-to-centre distance between cells, and  is the angle between the
centre-lines of the cells and a line from centre of a cell to the point on cell
periphery where the arc connects.
The relative hoop forces, at any level in the cell, are shown for circular and
diaphragm cells in Figs. 7.40 and 7.41, respectively.

Fig. 7.41. Calculation of hoop


forces in diaphragm cells36
322 Deep excavations

Mass Minimum ultimate strength of interlock (tonnes)

Section B (mm) t (mm) Per linear metre Per square metre Grade 43A steel ASTM-A328 steel Grade 50A steel Weights of standard
(nominal) of pile (kg) of wall (kg) T-junctions (kg/m)

SW-1 413 9.5 55.3 134.0 285 299 384 83.6


SW-1A 413 12.7 63.8 154.5 285 299 384 96.3

Section Nominal Web Deviation Perimeter Steel section Mass per m Mass per Section modulus Moment of inertia Guaranteed interlock
width, thickness, angle, 8 of a single of a single of a single m2 of wall of a single pile of a single pile strength (t/m)
L (mm) e (mm) pile (cm) pile (mm) pile (kg/m) (kg/m2 ) (cm3 ) (cm4 ) Steel grade

PAE PAE PAE


270 360 390

A500-12 500 12.0 6 138 92.1 72.3 145 47 180


A500-12.5 500 12.5 6 138 94.8 74.4 149 47 180 300 400 500
A500-12.7 500 12.7 6 138 95.8 75.2 150 47 180

Note: All the sections interlock together

Fig. 7.42. Straight web pile sections and junctions produced in Europe (courtesy of Corus (Frodingham straight web sections,
upper figure) and Arbed (Arbed straight web sections, lower figure))

Arc connections between circular cells and diaphragm cells are the most
highly stressed part of the sheeting and are the principal point of failure
risk. The connections are usually made through T and Y junction piles. Dis-
muke indicated the theoretical direction of loads acting on T and Y junctions
and showed that the formula for tc2 gives consistently good results when used
to evaluate interlock forces between arcs.
A summary of typical straight web pile sections and junctions produced in
Europe with guaranteed interlock strengths is shown in Fig. 7.42.

Stability of base and sliding


For cells not founded on rock it is essential to check the risk of foundation
failure; for cofferdams on clay, a bearing capacity failure could be caused by
lateral loading on the cofferdam, or pile settlement on the inside line of piles
could cause excessive under seepage of water. Where the weight of the cell
filling causes excessive settlement slippage of the sheet piles, loss of interlock
friction normally maintained by tension at the interlock could occur.
Figure 7.43 gives expressions for assessing the bearing capacity of the soil
below the cofferdam, treating the cell as a rigid body. The figure also gives
Cofferdam design 323

Fig. 7.43. Stability of base soils below a cellular cofferdam36

the method for calculating the factor of safety of internal instability due to
settlement where the base soil is a compressible clay. The factor of safety
against sliding is
W tan 0 þ Pp
f ¼ ð77Þ
P
where W is the weight of fill and piling, 0 is the angle of shearing resistance of
cell fillings, Pp is the passive resistance at the inner line of the piling, and P is
the lateral force from soil and water on the outer line of the piling.
A value of 1.25 to 1.3 would be considered satisfactory for temporary
works. It is not usual for sliding to be a critical mode of failure for cells,
except for those founded at shallow depth on rock.

Piling pull-out
There is risk to sheet piles on the outside of cells of pull-out as a result of over-
turning moment due to lateral loading by soil and water. If this occurs, cell fill
may be lost and the cell would fail if the quantity were large. For cells founded
324 Deep excavations

on rock, the factor of safety against piling rise is


 
Hb H Pa Hs
f ¼ bðPw þ Pa Þfgs þ Pp Pw þ ð78Þ
3 3 3
where b is equivalent cell width, fgs is the coefficient of friction between the
cell filling and sheet piling, Hb is the berm height, Pw is the lateral force due
to external water pressure, Pa is the lateral force due to external active soil
pressure, Pp is the lateral force due to passive resistance due to the berm, H
is the height of the water head above formation level on the outside face,
and Hs is the overburden height on the outside face. Values between 1.25
and 1.3 would be acceptable for temporary works.
For cells founded on sands and clay the factor of safety against pile pull-out
is
resistance to pull-out per unit length of cofferdam Cp
f ¼ ¼ : ð79Þ
pull-out force per unit length of cofferdam Fp
The pull-out force Fp is
Pw H þ Pa d  Pp Hb
Fp ¼ ð80Þ
3b½1 þ ðb=4LÞ
where d is the pile embedment below formation level, and L is the cell module
(see Fig. 7.44).
The resistance to pull-out Cp for cells on sand bases is
Cp ¼ 12 ka d 2 tan 2rc : ð81Þ

Fig. 7.44. Piling rise and pile


pull-out calculations for circular
cofferdams: (a) founded on
rock; (b) founded in sand and
clay
Cofferdam design 325

On clay bases
Cp ¼ ca 2rd ð82Þ
where ca is the adhesion of clay to sheet piles. The factor of safety should not
be less than 1.25 for temporary works.

Overturning
If the cell is assumed to rotate about its toe, in a similar way to a gravity wall,
the resultant moment due to cell weight and lateral forces should be restricted
to the middle one-third of the cell width. The factor of safety against over-
turning is
Mrestoring ðWb=2Þ
f ¼ ¼ : ð83Þ
Mdisturbing ðPH=3Þ

The relative complexity of the statical calculations for cellular cofferdams is an


opportunity for a finite element approach. This itself would involve similar
simplifying assumptions, would have to be made in three dimensions and
would only be considered reliable if the results compared with those of the
semi-empirical approach of the statical method described above.

Gravity cofferdams Precast concrete blockwork can provide weight stability for cofferdam
structures and is economical in circumstances where the temporary works
has a long design life and where alternative means of construction may be
unavailable. Blockwork construction is traditional for quay walls and both
horizontal and inclined bedding units have been used extensively. Similar con-
struction methods can be used to exclude water from a deep excavation in a
river bed. Blockwork cofferdams find particular application at the entrances
to docks and locks where excavation and maintenance work is required to
the floor of the existing facility. Their use is, of course, limited to sites
where adequate load bearing soil is present at reasonably shallow depth, or
where the existing soils can be improved in situ at low cost by such methods
as vibro-compaction or dynamic consolidation.
The dimensions and weight of the blocks will depend on the head of river
water to be excluded from the excavation and the plant, labour and material
resources available at the site. Access to the site of the cofferdam and working
areas may favour or preclude the use and transport of heavy blockwork. The
blocks themselves should be as durable and watertight as possible with a
strong dense concrete mix. They may be cuboid or wedge shaped but in all
cases should include a key to ensure interlock between adjacent blocks.
The prepared soil bed on which the initial course of blocks is laid should be
at least 0.5 m thick, and consist of crushed graded rock accurately levelled
with the assistance of divers. In fine-grained subsoil it will be necessary to
provide a soil or geotextile filter between the subsoil and the bedding to the
units. To improve the stability of the blockwork against overturning and
sliding it will frequently be necessary to construct a berm on the inward
side of a blockwork wall built to exclude river water. At the same time this
reduces the risk of hydraulic failure at the inside of the cofferdam by length-
ening the drainage path of water flowing under the blockwork. A filter blanket
should be incorporated at the base of the berm.
The design of a blockwork cofferdam, like other traditional gravity
structures, consists of checks on sliding, overturning and bearing values,
and restricting the resultant moment due to lateral forces and the weight of
the blockwork to within the middle-third of the base width to avoid tension
326 Deep excavations

on the lowest course of blocks. These checks are made, course-by-course, from
the top of the cofferdam to the underside of the initial course of blocks.
The head of water to be supported by the blockwork will be equal to the
high tide level at a spring tide plus an allowance for wave height. An allowance
for reflected waves due to wind and river craft is required: for reflected waves,
this is half the height of the highest wave with reference to the most unfavour-
able standing water level. Where there is the risk of breaking waves on the
cofferdam face (where water depth at the wall is greater than 1.5 times wave
height it can be assumed that only reflected waves apply) it is unlikely that
a blockwork solution will be adequate to resist the dynamic pressures. It
should be noted that wave pressures act not only on the face of the wall but
are also transmitted to the joints between blocks. Although this increase in
pressure is limited in duration, additional measures to restrict blockwork
movement where this force momentarily exceeds the effective block weight
should be carefully considered.

References 1. Site Investigation, Steering Group. Without site investigation ground is a hazard.
Thomas Telford Ltd., London, 1993.
2. BS 5930. Code of practice for site investigations. British Standards Institution,
London, 1981.
3. DIN Standard 4020. Geotechnical investigations for construction purposes (draft).
Deutsches Institut fur Normung, Berlin, 2002.
4. CIRIA. Design and construction of sheet piled cofferdams. Thomas Telford Ltd.,
London, 1993, Special publication 95.
5. EAU 90. Recommendations of the Committee for Waterfront Structures, Harbours
and Waterways. Ernst and Son, Berlin, 6th English edn., 1993.
6. Stroud M.A. and Butler F.G. The standard penetration test and the engineering
properties of glacial materials. Proc. Symp. Engng Behaviour of Glacial Materials,
Birmingham, 1975, 124–135. Midland Soil Mechanics and Foundation Engineer-
ing Society, Birmingham, 1975.
7. Skempton A.W. The consolidation of clays by gravitational compaction. Quart.
J. Geol. Soc. Lond., 1970, 125, 373–411.
8. Kenney T.C. Geotechnical properties of glacial lake clays. ASCE J. S.M.F.E.,
1959, 85, June.
9. BS 8002. Earth retaining structures. British Standards Institution, London, 1994,
(amended 2001).
10. Kort D.A. Steel sheet pile walls in soft soil. DUP Science, Delft, 2002.
11. BS 6349. Maritime structures. British Standards Institution, London, 2000.
12. Sainflou M. Essai sur les digues maritimes verticales. Ann. des Ponts et Chausse´e,
1928, 98, 5–48 (translated by US Corps of Engineers).
13. Minikin R. Wind, waves and maritime structures. Charles Griffin, London, 1963.
14. Durability and protection of steel piling in temperate climates. Corus, Scunthorpe,
2002.
15. Johnson K. et al. Low water corrosion of steel piles in marine waters. EUR
17868en, 1997. European Committee for Standardisation, Brussels, 1997.
16. Tsuchida E. et al. Studies of the corrosion of steel materials in a marine environ-
ment. Permanent International Association of Navigation Congresses, 1985.
17. Fukute T. et al. Steel structures in port and harbour facilities. Permanent Inter-
national Association of Navigation Congresses, 1990.
18. Piling handbook. British Steel Corporation, Scunthorpe, 1997.
19. Protection of steel sheet piling. SIGMA coatings, Uithoorn, Holland, 1999.
20. BS 8081. Ground anchorages. British Standards Institution, London, 1989.
21. BS EN 1537: 2000. Execution of special geotechnical work ground anchors. British
Standards Institution, London, 2000.
22. Littlejohn G.S. and Bruce D.A. Rock anchors: state of the art. Foundation
Publications, Brentwood, 1977.
23. Sherwood D.E. and Harris R.R. Regroutable ground anchors. Conf. Retaining
Structures, Institution of Civil Engineers, London, 1992, 448–456.
Cofferdam design 327

24. Ostermayer H. Practice in the detail design applications of anchorages, Institution


of Civil Engineers, London, 1976, 55–61 (discussion 62–78).
25. BS 5896. Specification for high tensile steel wire strand for the prestressing of
concrete. British Standards Institution, London, 1980.
26. BS 4486. Specification for hot rolled steel bars for the prestressing of concrete.
British Standards Institution, London, 1980.
27. Barley A. Drilling and grouting methods and modes of enhancement of
anchorage capacity. Proc. Ground Modification Seminars, Sydney, Melbourne,
Brisbane, 1992.
28. Barley A. et al. Design and construction of temporary ground anchorages at
Castle Mall development, Norwich. Conf. Retaining Structures. Institution of
Civil Engineers, London, 1992, 429–439.
29. Barley A. Ten thousand anchorages in rock. Ground Engng, 1988, 21, No. 6,
Sept., 20–20; No. 7, Oct., 24–35; No. 8, Nov., 35–39.
30. Methods of sheet pile installation. Technical European Sheet Piling Association
(TESPA). Luxembourg, 1995.
31. Williams S.G. and Little J.A. Structural behaviour of sheet piles interlocked at
the centre of gravity of the combined section. Proc. Instn Civ. Engrs, 1992, 94,
229–238.
32. Timoshenko S.P. and Goodier J.N. Theory of elastic stability. McGraw-Hill,
New York, 1970.
33. Kirmani M. and Highfill S.C. Design and construction of the circular cofferdam
for ventilation building No. 6 at the Test Williams Tunnel. Civ. Eng. Practice,
Spring/Summer, 1996, 31–47. Boston Society of Civil Engineers, 1996.
34. Bishop A.W. Discussion on cofferdams. Proc. Instn Civ. Engrs, 1963, 24, 112–
116.
35. Ovesen N.K. Cellular cofferdams: calculation methods and model tests. Danish
Geotechnical Institute, Copenhagen, 1962, Bulletin 14.
36. Dismuke T.D. Cellular structures and braced excavations. Foundation engineer-
ing handbook. Eds. H.F. Winterkorn and H.Y. Fang, Van Nostrand-Reinhold,
Princeton, NJ, 1975, part 14, 445–480.
37. Terzaghi K. Stability and stiffness of cellular cofferdams. Trans. ASCE, 1945,
110, 1083–1119 (discussion 1120–1202).
38. Steel sheet piling cellular cofferdams on rock. US Tennessee Valley Authority,
1957, technical monograph No. 75, vol. 1.
39. Cummins E.M. Cellular cofferdams and docks. Trans. ASCE, 1957, 125, 13–34
(discussion 34–45).
40. Brinch-Hansen J. Earth pressure calculations. Institution of Danish Civil
Engineers, Copenhagen, 1953.
41. Das Braja M. Principles of Foundation Engineering, 4th ed, Brooks/Cole, Califor-
nia, 1998.
42. Ostermayer H. and Scheele F. Research and ground anchors in non-cohesive
soils. Rev. Française de Ge´otechnique, 1978, No. 3, 92–97.
43. Locher H.G. Anchored retaining walls and cut-off walls. Losinger Ltd., Berne,
Switzerland, 1969, 1–23.
44. Littlejohn G.S. Soil anchors. Proc. Conf. on Ground Engineering, ICE, London,
1970, 33–44.
45. Littlejohn G.S. Ground anchors. Proc. Review of Diaphragm Walls, ICE,
London, 1977, 93–97.
46. Kranz F. Uber die Verankesung von Spundwanden, Ernst and Son, Berlin, 1953,
55–61.
47. Ranke A. and Ostermayer H. Beitrag sur Stabilitatsuntersuchung mehrfach
verankerter Bangrubenumschliessugen. Die Bautechnik, 45, 10, 1968, 341–350.
48. Cheney R.S. Permanent ground anchors. US Dept of Trans. Federal Highways
Admin. Report FHWA/DP/68, 1984.
49. Packshaw S. Cofferdams. Proc. Instn. Civ. Engrs, 1962, 21, Feb., 367–398.

Bibliography Ayers J.R. and Stokes R.C. Design of flexible bulkheads. Trans. ASCE, 1953, paper
2676, 373–383 (discussion 384–402).
328 Deep excavations

Belz C.A. Cellular structure design methods. Proc. Conf. Design and Installation of
Pile Foundations and Cellular Structures, Envo Publishing, Lehigh Valley, USA,
1970, 319–337.
Boyer W.C. and Lummis H.M. Design curves for anchored steel sheet piling. Trans.
ASCE, 1953, paper 2689, 639–657.
BS EN 1537. Ground anchors. British Standards Institution, London, 1999.
BS EN 12063. Sheet piling. British Standards Institution, London, 1999.
Carle R.J. High strength interlock sheet piling in cellular structures. Proc. Conf.
Design and Installation of Pile Foundations and Cellular Structures, Envo Publish-
ing, Lehigh Valley, USA, 1970, 367–379.
Cedergen H.R. Seepage, drainage and flow nets. Wiley, New York, 2nd edition, 1977.
Clayton C.R. and Milititski J. Earth pressures and earth retaining structures. Blackie,
London, 1993.
Cornfield G.M. Direct reading nomograms for design of anchored sheet pile retaining
walls. Civ. Engng Public Works Rev., 1969, 64, Aug., 753–756.
Dismuke T.D. Stress analysis of sheet piling in cellular structures. Proc. Conf. Design
and Installation of Pile Foundations and Cellular Structures, Envo Publishing,
Lehigh Valley, USA, 1970, 339–365.
Gray H. and Nair K. A note on the stability of soil subject to seepage forces adjacent
to a sheet pile. Ge´otechnique, 1967, 17, Mar., 136–144.
Kaiser P.K. and Hewitt K.J. The effect of groundwater flow on the stability and
design of retained excavations. Canadian Geotech. J., 1982, 19, May, 139–153.
King G.J.W. Design charts for long cofferdams. Ge´otechnique, 1990, 4, Dec., 647–
650.
Manual for the design of reinforced concrete building structures. Institution of
Structural Engineers, London, Second edition, 2002.
Manual for the design of steelwork building structures. Institution of Structural
Engineers, London, Second edition, 2002.
Patterson J.H. Installation techniques for cellular structures. Proc. Conf. Design and
Installation of Pile Foundations and Cellular Structures, Envo Publishing, Lehigh
Valley, USA, 1970, 393–411.
Potts D.M. and Day R.A. Use of sheet pile retaining walls for deep excavations in stiff
clay. Proc. Instn Civ. Engrs, 1990, 88, Dec., 899–927.
Rossow M.P. Sheet pile interlock tension in cellular cofferdams. ASCE J. Geotech.
Engng, 1984, 110, No. 10, Oct., 1446–1458.
Schnabel J.R. Sloped sheeting. ASCE J. Civ. Engng, 1971, 41, Feb., 48–50.
Seed H.B. and Whitman R.V. Design of earth retaining structures for dynamic loads.
Proc. Conf. Lateral Stresses in the Ground and Design of Earth Retaining Structures.
American Society of Civil Engineers, New York, 1970, 103–947.
Swatek E.P. Summary – cellular structure design and installation. Proc. Conf. Design
and Installation of Pile Foundations and Cellular Structures, Envo Publishing,
Lehigh Valley, USA, 1970, 413–423.
Terzaghi K. Anchored bulkheads. Trans. ASCE, 1954, 119, paper 1720, 1243–1280
(discussion 1281–1324).
Tomlinson M.J. Foundations design and construction. Prentice Hall, Harlow, 7th
edition, 2001.
Basement construction and design
8
Engineering an A number of factors control the relative difficulty of basement construction.
excavation Very often these factors cannot be changed by the designer, and include the
location of the building, the proximity of existing buildings and services,
previous site use, and the proposed use of the basement together with soil
and groundwater conditions. The basement structure will be designed to over-
come these constraints to transfer the loads from the superstructure to the
subsoil. The method of basement construction and the type of peripheral
basement wall will be selected to support soils and groundwater at the
curtilage of the basement as economically as possible. The permitted soil
deformation around the basement construction has to be assessed and
complied with. The process was itemized by Lambe1 , as shown in Table 8.1.
Increasingly, clients and architects are demanding larger and deeper base-
ments. This chapter reviews the development of construction methods and
the range of basement walling methods available, and describes the design
problems that arise.

Construction methods Seldom does the location of a basement allow open battered excavations.
for soil support Particularly on urban sites, space is limited and insufficient to accommodate
the cut slopes of battered excavations. Land is expensive and basement con-
structions inevitably occupy as much of the site as possible. The use of open
excavations was reviewed in Chapter 3, although mention will be made here
of the need to review soil strength parameters critically for temporary cut
slopes.
In certain soils, over-consolidated clays such as London clay for example,
the soil strength characteristics are time-dependent, so the period for which
the excavation is to be kept open must be carefully assessed. Where space
allows the use of battered slopes, the cost penalty of a slope failure should
be weighed against the cost of a full soil retention system using temporary
walling. It is possible that a compromise solution, using soil nailing or similar
ground improvement methods incorporating reinforced soil, may be econom-
ically attractive where some horizontal working space is available at the rear
of the basement construction but is not sufficient to accommodate a full
battered slope. Where some space exists behind the permanent basement
wall the choice of method will be determined in permeable soils or granular
soils by the extent of groundwater flow and the feasibility and cost of control-
ling groundwater during basement construction.
An example of a battered basement excavation with a slurry trench cut-off
to control groundwater inflow and cut slopes designed on a cost against risk
basis was given by Wakeling2 . The excavation was 130 m  80 m in plan, to a
depth of 5.8 m in soft clay and gravel, extending into stiff fissured silty London
clay to a maximum depth of 14.5 m. A bentonite slurry cut-off wall into the
London clay contained groundwater in the upper gravels. Groundwater
flow from the gravel and the underlying silty sands was controlled by
330 Deep excavations

Table 8.1 Engineering an


excavation: a checklist1 Step Activity Considerations

1 Explore and test subsoil


2 Select dimensions of excavation Structure size and grade requirements,
depth to good soil, depth to floor
requirements; stability requirements
3 Survey adjacent structures and Size, type, age, location, condition
utilities
4 Establish permissible movements
5 Select bracing, if needed, and Local experience, cost, time available,
construction scheme depth of wall, type of wall, type and
spacing of braces, dewater excavation
sequence, pre-stress
6 Predict movements caused by
excavation and dewatering
7 Compare predicted with
permissive movements
8 Alter bracing and construction
scheme if needed
9 Instrument – monitor construction
and alter bracing and construction
as needed

gravel-filled counterfort drains dug down the slope during bulk excavation.
The excavation was battered with side slopes of 1:1 with an intermediate
berm at the top of the London clay. A plan and cross-section of the excavation
are shown in Fig. 8.1. The method was successful and demonstrates the use of
soil parameters based on partially drained soil conditions. Wakeling reported
that three slips occurred, all shallow-seated, in the batters within the London
clay, reaching their greatest depth between two and five months after excava-
tion. From back analysis on these slopes and using shallow-seated slides
reported by Skempton and La Rochelle3 , Wakeling reached a tentative
conclusion: for excavations in stiff fissured clays, short-term shallow-seated
slips are likely to occur when the computed failure strength exceeds the
measured undrained shear strength in the clay by approximately 20%. The
point of interest in this example is the cost-effectiveness of a relatively steep
slope batter of 1:1 with some risk of minor failure accepted during a relatively
short construction time. The use of fully drained parameters in the slope
analysis would have led to flatter slopes, albeit with less risk of slippage.
Where compromise solutions to peripheral soil support are required, that is
where some space exists at the rear of the permanent basement structure but is
insufficient to accommodate a battered slope, crib walls and anchored crib
walls can be used. A further solution is the use of soil nailing. These methods
were discussed in general terms in Chapter 4, but the relevance of soil nailing
to basement construction deserves discussion here.
The soil nailing method developed from the use of fully bonded rock bolts
for tunnel support in the 1950s and 1960s. Using the same principles of
ground support its use progressed from weaker rocks, such as marls and
weak sandstones, into cemented sands, strong clays and, later, to a wider
range of granular soils and middle-strength cohesive soils. The range of
soils in which nailing can be used is relatively wide (weak clays and loose
silts are probably precluded, and the presence of groundwater limits its
application in any soil). Although finding widespread application for general
soil support in basement schemes in France, Austria, Germany and North
America, its application in the UK has been slow.
Basement construction and design 331

Fig. 8.1. Plan and cross-section of a battered deep excavation into London clay2

Soil nailing is in fact a soil-reinforcing technique and uses short tendons,


driven or inserted into short bores in the excavated soil face, to improve the
shear strength of subsoils. The exposed face is retained and protected by a
gunite layer reinforced with a wire mesh. The technique is described by
Gässler and Gudehus4 and Banyai5 , but a complete description of its develop-
ment – the soil–tendon interaction, design, construction and specification – is
given in the report of the French Clouterre project6 . The UK code of practice
BS 8006 Strengthened/Reinforced Soils applies to soil nailing and the draft
European code on soil nailing is in course of preparation (Pr EN 14490 at
enquiry stage, July 2003).
Typical cross-sections of soil-nailed excavations were reported by Barley7
and are shown in Fig. 8.2. A soil-nailed excavation support in Pocking,
Bavaria, is shown in Fig. 8.3.
332 Deep excavations

Fig. 8.2. Cross-sections of typical nailed slopes in the UK: (a) temporary soil-nailed slope; (b) soil-nailed slope; (c) rock-bolted

The application of soil nailing to retain excavated slopes at the periphery of


basements is referred to here as temporary support, but soil-nailed slopes can
also be used for permanent works. Tendon durability and protection was
discussed by Barley7 and in the Clouterre report6 . The use of alternative
reinforcing materials such as stainless steel, carbon fibre and glass-reinforced
plastics may lead to an increase in permanently retained soil-nailed excava-
tions. The cost effectiveness of any scheme, the material used and the installa-
tion method is much dependent on job size.
The extent of working space at the rear of the permanent retaining wall
is likely to reduce the nearer the basement site is to a city centre. In the
remainder of this chapter, it is assumed that such space is limited. Soil support
systems which incorporate both temporary and permanent support are likely
to prove most efficient in minimizing the total width of soil support wall.
Alternatively, the construction of sloping sheeting can lead to economies in
construction cost where limited working space is available. Schnabel8
reported that where sheeting sloped at an angle of about 10% from the

Fig. 8.3. Soil-nailed slope at Pocking, Bavaria Fig. 8.4. Plate method being used
(courtesy of Bauer) for a basement extension to the
Technical University, Zurich
(courtesy of Bauer)
Basement construction and design 333

vertical, the measured anchor strut loads were consistently less than two-
thirds of the computed anchor loads for vertical sheeting in the same soil. A
further study9 of sloped sheeting supported by ground anchors presented
model tests results in sand, confirming Schnabel’s recommendations for soil
pressures on an inclined anchored wall. It was noted from these model tests
that inclined walls require a considerable base width if they are not to suffer
a bearing capacity failure.
Underpinning in short lengths may prove necessary to avoid settlement of
adjacent structures during basement construction. In dry soil conditions,
where the water table lies beneath basement formation level, it may be
sufficient, and expedient, to rely on the underpinning to provide horizontal
soil support during basement construction in addition to its main purpose
of vertical load transfer to depths below the new basement construction.
Unless the underpinning is braced or propped from the excavation side its
depth will be restricted in either concrete or grouted soil because of horizontal
soil pressure at the rear of the underpinning.
Where ground conditions allow successive excavation in the dry, an
anchored reinforced concrete plate can be used to provide a continuous
reinforcement wall at the periphery of the new basement. These ground con-
ditions may be obtained by grouting in certain soil conditions, given legal
consents. Figure 8.4 shows the plate method being used in a base extension
in Zurich. Each cast in situ element was retained by ground anchors,
excavated alternately at each level. The subsoil was a cohesive silty sandy
gravel.
The provision of lateral support to a deep excavation thus turns on six
factors: neighbours’ rights (reference to the Party Wall Act 1996 in the UK
is made in Chapter 1), neighbouring construction, subsoil and groundwater
conditions, neighbouring services, and the proposed construction depth and
optimization of site area to give the best financial return.
In complying with these factors the majority of urban basements sites will
not allow battered open excavations due to space limitations. Vertical periph-
eral soil support is therefore required, temporarily during construction and as
a permanent retaining wall. During construction the simplest form for either
sheeting or walling is to cantilever without propping. In typical basement
excavations in London the maximum height of cantilever is generally of the
order of 5.5 m from formation level. The extent of soil movement during
and after bulk excavation, and the presence of delicate services or important
highways at the rear of the wall, mitigate against the use of high cantilevers.
Temporary berms at the front of the cantilevered wall reduce soil movements
effectively but are often uneconomical because of the need to remove the berm
successively in short lengths and small volumes. Although propped cantilevers
provide more security against excessive wall movement, the cost penalties of
providing this support and the impedance to bulk excavation may prove
unacceptable. The economical use of peripheral steel sheet pile propped by
steel raker tubes from a completed central raft construction with a temporary
edge berm was shown in Fig. 6.1. A cost comparison on that particular site
showed little difference between propped steel sheet piling with in situ perma-
nent retaining walls and propped diaphragm wall construction.
Figure 8.5 shows why the use of temporary soil berms to reduce cantilever
wall movement is unpopular with contractors. The basement, in West
London, was large in plan area but limited in depth to 5.625 m from existing
ground level. Maximum horizontal wall deformation was specified as 25 mm
and ground conditions were medium dense sands and gravels overlying
London clay. Two schemes were prepared by a specialist diaphragm contrac-
tor, the first with a temporary soil berm and minimum wall depth, the second
334 Deep excavations

Fig. 8.5. Cantilever wall, West


London: (a) wall cross-section;
(b) soil profile; (c) calculated
maximum deformation with and
without soil berm

with a free-standing wall of greater depth. Although the analysis showed a


significant beneficial effect on wall deformation of providing a relatively
small berm, the main contractor preferred the deeper wall without the
berm, shown in Fig. 8.5(c). The cost of the additional walling was significant
but avoided later excavation of small soil volumes and impedance to the base
raft construction programme. During construction the maximum deforma-
tion of the cantilever wall without the berm was 19 mm.
Using finite element methods and assuming linear elastic perfectly plastic
soil material (with the effects of enhanced stiffness at small strains), Potts
et al.10 reached a number of conclusions on the effectiveness of berms: for
berms between 2.5 and 5.0 m high it is the volume of the berm, not its specific
geometry, that dictates soil movements adjacent to the excavation wall defor-
mation and bending movements; as the height of the berm reduces below
2.5 m, berms of equal volume, but varying geometry result in different wall
deflections and moments – deflections increase and the berm becomes less
efficient.
To avoid the obstruction of temporary berms and rakers during construc-
tion, soil anchors may be used for soil support to basements of moderate
Basement construction and design 335

depth. This solution depends on the suitability of soil conditions for anchor-
ing and the legal and practical implications of founding anchors outside the
curtilage of the site. The popularity of anchoring in the UK appears to be
less than in France and this may depend on cost which in turn may depend
on design safety factors. In France a value of 1.5 is usual, whereas in the
UK a minimum value of 2.0 has been specified by BS 8001 for temporary
works.

Water-resisting A complete review of the methods of safeguarding basements against water


basement construction and dampness was given in the CIRIA report 13911 in 1995. Later comment
on the durability and water resistance of basements is made in the Institution
of Structural Engineers report on basements and cut-and-cover construction12
in draft form in 2003. The British Standard BS 810213 , published in 1990,
serves as a code of practice on the subject.
The whole matter of waterproofness standards is a matter of potential
controversy and originates from a false belief that a basement construction
in water-bearing ground can be made watertight and completely dry. The
requirements of water resistance, a more realistic term, are described in both
the CIRIA report 139 and BS 8102 in similar classification. A summary
chart of the basic grades of water resistance as defined in both documents is
given in Table 8.2. Four performance grades are specified as follows: grade 1
basic utility, grade 2 better utility, grade 3 habitable and grade 4 special.
The water-resisting methods to be adopted to address these performance
standards are given in CIRIA report 139 as being one (A, B or C) or a

Table 8.2 Basic grades of water resistance11

From Table 1 of BS 8102: 1990 Abbreviated commentary given


by CIRIA Report 39
Grade Basement usage Performance Form of protection
levela

Grade 1 Car parking plant Some seepage Type B with RC design to Visible water and BS 8110 crack
Basic rooms (excluding and damp BS 8110 width may not be acceptable.
utility electrical equipment); patches tolerable May not meet Building
workshops Regulations for workshops.
Beware chemicals in
groundwater
Grade 2 Workshops and plant No water Type A or Type B with Membranes in multiple layers
Better rooms requiring drier penetration but RC design to BS 8007 with well lapped joints. Requires
utility environment; retail moisture vapour no serious defects and higher
storage tolerable grade of supervision. Beware
chemicals in groundwater
Grade 3 Ventilated residential Dry environment Type A or Type B with As Grade 2. In highly permeable
Habitable and working, RC design to BS 8007, plus ground, multi-element systems
including offices, Type C with wall and floor (possibly including active
restaurants, leisure cavities and DPM precautions, and/or permanent
centres and maintainable under-
drainage) probably necessary
Grade 4 Archives and stores Totally dry Type A or Type B with As Grade 3
Special requiring controlled environment RC design to BS 8007 and a
environment vapour-proof membrane,
plus Type C with ventilated
wall cavity and vapour
barrier to inner skin and
floor cavity with DPM
a
See CIRIA Report 13911 for limits on environmental parameters.
336 Deep excavations

combination of two (C þ A or C þ B); the types being as follows: A, structure


requiring the protection of an impervious membrane (i.e. tanked); B, structure
without a membrane but with structurally integral protection; C, drained
cavity (for use with type A or type B or alone).
Overall, water-resisting reinforced concrete should be used for all the grades
together with appropriate design, detailing and construction. Somerville14 ,
reviewing the durability of R.C. structures referred to the importance of the
four ‘Cs’ in terms of durability; constituents of the mix, cover, compaction
and curing. In terms of water resistance it should be recognized that reinforced
concrete inevitably cracks as the cement hydrates, through thermal movement
and later drying shrinkage. Perhaps, in terms of reinforced concrete base-
ments, the risk of leakage and the management of that leakage to ensure
compliance with the required performance standard should be based on
four ‘Ss’: simplicity in structural detailing to take into account the intended
construction methods and buildability; location of construction joints to
receive special attention; services: make design provision at an early stage;
spacing of wall reinforcement and supervision of concreting to ensure well
compacted dense concrete.
It is inevitable that reinforced concrete walls and rafts will crack, initially
due to thermal shrinkage and thence with time due to drying shrinkage.
Early age strains due to thermal action are likely to be far greater and more
important than strain due to drying shrinkage. The hydration temperatures
generated in thick reinforced concrete sections are high and any change to
thin sections is likely to generate differential movement and resulting cracking.
Similarly, restraint provided by cross-walls or similar structural restraint is
likely to cause cracks. These early cracks are important in terms of water resis-
tance because they are likely to pass right through the section. The maximum
crack width for early age thermal effects should be limited to 0.2 mm. Cracks
due to flexure under load do not generally pass through the section, compres-
sive stresses progressively reducing strains in this zone of the structural
section. The maximum flexural crack widths at the surface of the concrete
section may be greater than the maximum of 0.2 mm for thermal cracks and
typically would be of the order of 0.3 mm. The avoidance of steel congestion
in basement R.C. sections to obtain stringent crack control and only secure
poorly compacted concrete with risks of bad durability and water leakage
cannot be over-emphasized.
A recent paper by Boikan15 summarized the pitfalls in waterproofing of
basements and their prevention. As an introduction, Boikan quoted a legal
case in England, Outwing Construction vs. Thomas Weatherald, when the
High Court held that the designer takes responsibility not only for the speci-
fication of a waterproofing product, but must also assess the compatibility of
that product in conjunction with other parts of the basement design, and the
impact of inadequate workmanship and site conditions on the integrity of the
design overall. Boikan recommended that particular matters such as the use of
incompatible discontinuous products, use of products such as low tolerance to
poor workmanship, use of waterproofing products from various sources all
required special attention. He further commented that physical barriers, pre-
viously asphalt tanking but nowadays polyethylene cast in membranes, or
bentonite clay bound into a geotextile required particular attention. These
membranes should be assessed in terms of their performance in hard and
soft water and contaminants, their gas resistance, their means of bonding to
the concrete and the details of bonding one sheet to its neighbour. Boikan
included a review of the performance of hydrophilic waterstop systems
which are now much preferred to PVC waterstops in in situ R.C. basement
construction.
Basement construction and design 337

Progressive As basement excavations increase in depth, excavation methods have become


development of more complex, leading to top-downwards techniques which allow simulta-
construction methods neous basement construction and superstructure erection. These techniques
for deep basements became popular in major international city centres in the 1980s and 1990s.
This review of the development of these methods in London is largely based
on a paper by Zinn16 and shows the increasing dependence of basement
construction methods on progressively larger and more powerful piling and
diaphragm wall equipment.

Trench construction
At the beginning of the last century, basement construction in London was
restricted to major buildings. Basement walls were built as gravity structures
in deep, heavily-timbered trenches at the basement periphery. In later years
the walls were built as cantilever reinforced concrete walls in trench. Wall
dimensions that resulted from this construction method can be judged from
Fig. 8.6, a cross-section of the basement wall to the Shell Centre17 . The
wall, cantilevering for a depth of four basement storeys, was built in pre-
stressed concrete within a sheet piled trench with reinforced concrete wallings
and struts. Two separate excavation operations were required, the first by

Fig. 8.6. Cross-section of


peripheral cantilever
prestressed concrete retaining
wall in trench, Shell Centre
basement, London17
338 Deep excavations

grab within the peripheral trench, the second by excavation of the central
dumpling after wall completion. The wall construction operations were
obstructed by the temporary trench bracing but before it could be removed
the inner and outer sheeting to the trench had to be re-strutted against the
completed wall; all frame levels, reinforcement splicing levels, concrete lift
heights and re-strutting levels were interdependent.

Peripheral walls propped by floor and raft


In the late 1950s the trench method began to be replaced by more efficient
basement excavation methods. The changes exploited large-diameter bored
piles which were introduced into the UK at that time together with the
simple innovation of using the horizontal strength of floors and raft sections
as deep beams to span the length of breadth of the excavation. The Fu Centre,
Hong Kong (Fig. 8.7) was built with a pile wall, but the base of the cantilever
reinforced concrete wall was designed to span horizontally and resist all
horizontal earth pressure, allowing the removal of temporary support without
inducing purely cantilever moments into the wall.
In the tower block basement of the Hilton Hotel, London (Fig. 8.8) two
waling beams, each forming part of a structural floor, were used to tempora-
rily support the outer contiguous bored pile wall. The walings were designed
as frames and the upper waling was supported at the bored pile wall and by
bored piles inside the wall.
In the third phase of the development of this method, the Royal Garden
Hotel site (Fig. 8.9) used diagonal struts, also supported on piles, to reduce
the 72 m span of the longest side of the basement. Later, peripheral diaphragm
walls propped by successive floors, designed to span as a horizontal frame,
were used at Gardiner’s Corner, London (Fig. 8.10).

Fig. 8.7. Stages of basement


construction, Fu Centre, Hong
Kong16
Basement construction and design 339

Fig. 8.8. Stages of basement


construction, Hilton Hotel,
London16

Fig. 8.9. Stages of basement


construction, Royal Garden
Hotel, London16
340 Deep excavations

Fig. 8.10. Top-downwards


basement construction with
floors used as horizontal frame,
Gardiners Corner, London
(courtesy of Cementation)

Fig. 8.11. Stages of car park


basement construction,
Leicester Square, London16
Basement construction and design 341

Top-downwards construction
The fourth and fifth phases of the development of the basement excavation
method are illustrated by a car park at Leicester Square and the Winter
Gardens Theatre, both in London.
In the car park works (Fig. 8.11) a 458 berm was used to support the outer
bored pile wall while the central area substructure was completed. Floors
acting as walings as the berm was removed were supported by reinforced
concrete columns, cast in advance of the berm excavation, within pre-bored
shafts. Early construction of the central part of the substructure allowed
superstructure works to begin before berm excavation and completion of
the peripheral wall and slabs.
At the theatre site (Fig. 8.12) the disadvantage of removing the berm subsoil
from within the existing work was overcome with temporary steel lattice
columns to support the floor sections used as walings at each level.
The top-downwards method was, therefore, at that time almost in place; to
achieve maximum economy three criteria had to be achieved by the excava-
tion method:
(a) the retaining wall which supports the width of the excavation should
obtain support at each floor level

Fig. 8.12. Stages of basement


construction, Winter Gardens
Theatre, London16
342 Deep excavations

Fig. 8.13. Stages of basement construction, House of Commons underground car park, London: (a) site plan; (b) vertical
cross-section showing soil profile; (c) vertical section and plan of car park19
Basement construction and design 343

(b) the ground should act as a temporary soffit ‘shutter’ to floor construction
(c) removal of the excavation must be rapid and continuous.
With the development of diaphragm walling during the 1970s it was logical
that bored pile peripheral walls would be replaced with diaphragm walls to
act as both temporary and permanent soil support. Fenoux18 described the
construction of a nine-storey basement for car parking in which the super-
structure was opened for use before the completion of the substructure. In
1972 the House of Commons underground car park in London was built
with peripheral diaphragm walls with temporary support from floor slabs
cast successively with continuing excavation. The basement, shown in section
in Fig. 8.13, reached a maximum depth of 18.5 m with diaphragm wall 30 m
deep. The prime concerns at design stage were to minimize soil movements
due to the bulk excavation and limit the effect on nearby historic buildings.
The risk of soil heave was more acute since there was no superstructure
above the basement. The solution, to build a relatively stiff wall (1 m thick
diaphragm) propped at relatively small centres (storey heights) with relatively
stiff propping from in situ reinforced concrete floors was, therefore, designed
to reduce the risk of settlement of existing structures rather than to reduce
construction time. A detailed description was given by Burland and
Hancock19 ; details of soil movements, which caused only minor cracking
and movement to the adjacent buildings, were given by Burland et al.20
A more recent deep basement construction in London was described by
Marchand21;22 . This basement, constructed by top-downwards techniques to
a depth of 23.9 m from ground level to the lowest basement formation level,
was built for car parking below an eight-storey office block superstructure.
The basement is one of the deepest in London. Two details are worthy of
note: precast concrete stop ends were used in the 1 m thick diaphragm wall
construction and, although generally successful, Marchand commented:
‘Some of the joints between the precast stop ends and in situ concrete
leaked and this was dealt with by grout injection. The sealant used has been
specially developed for the mining industry and is pumped in as a fluid
which changes to a flexible mass of matted rubber particles. This material
can then flex without cracking. In a few places at low level clay had adhered
to the stop end, leaving a strip of clay up to 70 mm wide between adjacent
panels. This was raked out to a depth of 150 mm and made good in order
to provide a waterproof joint.’
Waterproofness of diaphragm wall basements is discussed later in this
chapter. Where basement walls built by pile or diaphragm wall techniques
remain unlined, the longevity of remedial measures to ensure acceptable
waterproofness remains a matter of concern.
The second noteworthy innovation was the use of five rows of pin piles
installed in front of the basement wall to stiffen the London clay and prevent
softening with time. With increasing groundwater levels the construction
would otherwise have required a substantial ground slab to prop the wall
and prevent passive failure of the wall. The stiffened soil approach allowed
the wall toe to be raised 5 m from the original design, a significant reduction
in a very deep wall, originally almost 38 m deep from ground level.
In the 1960s top-downwards construction techniques required the setting of
steel columns as part of the superstructure support within pile heads or pile
caps at basement formation level. This operation required personnel to trim
pile and set precast caps or make bases in situ for the column installation at
final basement level. On this contract, a scheme for setting the steel columns
directly into the wet pile concrete was investigated but rejected because of the
risk of inaccurate placing of the columns. Liners, 21 m long, were used to gain
344 Deep excavations

access to each pile head. The operation was costly and time-consuming and
was frequently underestimated in terms of construction time.
More recently, specialist firms have developed jigs to enable the steel plunge
column to be placed very accurately both in verticality and position within an
unlined box supported by bentonite slurry. The development of this jig there-
fore permits the use of slurry support and allows the steel column to be placed
in the wet concrete of the recently concreted pile. The general accuracy of
placing plunge columns is of the same order as bored pile construction of
the order of 1:75 but with the use of specialist jigs and good site control
placing accuracy in the range 1:200 to 1:400 can be maintained. The rolling
tolerance of the steel column itself may be critical and should be taken into
account in determining likely in situ verticality.
The top-downwards method has obvious advantages in terms of soil
movement and completion time, but important disadvantages include the
additional cost of excavation and removal of soil from beneath floor slabs
in cramped conditions compared with conventional open excavation
methods. Also, there is the congestion caused on site by superstructure and
substructure contractors working within the same programme period.
A successful application of top-downwards construction was recently made
in London on a congested site for the redevelopment of a site in Knightsbridge
for use by Harrods department store. The new building consisted of seven
storeys above ground and a seven-storey 25 m deep basement. The works
are described by Slade and Darling et al.23 , and the control of ground
movements and the application of compensation grouting by Fernie24 and
Kenwright et al.25
The new structure, built at the rear of the existing Harrods building and
connected to it by a new access tunnel, incorporated an existing façade on
one elevation. A site plan is shown in Fig. 8.14 and the construction sequence
is shown in Fig. 8.15.
The ground conditions on the site are made ground of 3.5 m thickness
overlying Terrace Gravel 6.0 m thick which in turn overlie London clay of
proven thickness greater than 50 m. The groundwater conditions comprise

Fig. 8.14. Harrods,


Knightsbridge, London.
Site plan23
Basement construction and design 345

Fig. 8.15. Harrods,


Knightsbridge, London.
Construction sequence23

an upper perched water table in the Terrace gravels and a lower aquifer in the
Thanet sands and Chalk which underlie the London clay. A hydrostatic pore
water pressure distribution was shown by piezometers within the London
clay.
The basement construction for parking, plant rooms and workshop use was
required to comply with grade 1 of BS 8102 allowing some seepage and damp
patches. Due to the demands for basement space the use of a lining wall and
drained cavity was not viable and the 800 mm thick diaphragm wall remained
unlined. Top-down construction with the ground floor initially built and then
excavation in two-storey level increments was possible and shown to be so by
finite element analysis. This procedure saved construction time and was
monitored during construction by comprehensive instrumentation, collecting
data from precise level points, electrolevel beams, in-place inclinometers,
water levels, survey targets and base traverse stations. Readings were taken
on a 24-hour basis and any change above program supervision trigger levels
was sent to the Engineers’ terminals. In the event, the trigger levels were not
exceeded.
The initial finite element analysis predicted both the deflected shape of
the basement walls and the resulting settlement profile outside the basement
346 Deep excavations

perimeter. In turn, using these deformation contours, the settlement behaviour


of nearby existing buildings was assessed and the classification due to
Boscardin and Cording26 used to make a risk prediction. Maximum settle-
ments due to basement construction were limited to 10 mm and angular
distortions to 1 in 750.
The overall stability of the basement in the long term was checked to ensure
a minimum factor of safety of 1.4 for the whole basement, taking into account
the uplift force due to soil heave and the groundwater pressure resisted by the
combination of building weight, the tensile resistance of the piles and the skin
friction on the faces of the diaphragm walls. (Skin friction to the lowest base-
ment level was used.) The total assessed long-term uplift pressure over the
whole basement area was 340 kN/m2 .
In the USA, the top-downwards method has been adopted more recently.
American practice is described by Fletcher et al.27 The excavation of a large
four-storey deep basement for the Milwaukee Centre, close to historic
structures and within 3.5 m of the Milwaukee River, demanded a cut-off and
control of groundwater, minimum soil movement and early completion. A
major bracing or raker system was judged to be too cumbersome and
costly, and a temporary freeze wall system was dismissed because a permanent
ground water cut-off would have been necessary. The top-downwards method
with a deep diaphragm wall as a cut-off was adopted and proved successful.
Scope for innovation remains. In 1993, a contract for an opera house in
Paris used the technique for a 28 m deep basement with anchored support
in lieu of lateral support from basement floors as excavation proceeded.
Large barrette sections were constructed for superstructure support. A typical
cross-section of the basement is shown in Fig. 8.16. Due to the planned con-
struction, after completion of the basement, of a Metro running tunnel on one
side of the basement and in close proximity to it, anchor tendons constructed
from glass fibre were adopted to avoid obstruction to the Metro tunnelling
machine. The brittle, low shear strength of the glass tendons ensured that
they would be easily removed by the tunnelling machine.
In Hong Kong, land values have increased demand for larger, deeper base-
ments frequently in unfavourable soil conditions with stringent settlement
criteria applied to the basement peripheral subsoils. This market demand
has brought about almost an exclusive use of top-downwards construction
generally using diaphragm walls. A recent example of the construction of
the deep basement for the Dragon Centre in the heart of the Western Kow-
loon Peninsula was given by Lui and Yau28 . The site was located on old
reclaimed land and the development, for a new retail building, comprised a
nine-storey reinforced concrete structure over a five-storey basement for
parking use. Historical records showed that the foreshore originally fronted
the site. The existing MTR tunnels are within 100 m of the site; the site
reclamation was made in 1924. Adjacent existing buildings are either sup-
ported on pad footings or driven precast piles. The ground conditions are
loose granular fill underlain by marine deposits, generally loose silty clayey
fine sand and highly developed granite (dense silty fine to coarse sand) and
granite bedrock. Groundwater level is 1.5 m below ground level. A geological
section is shown in Fig. 8.17.
The basement structure, 107 m  67 m in plan was formed from a dia-
phragm wall box, 1200 mm thick up to 40 m deep installed by cutter with
CWS joints to 30 m depth and a cut joint below 30 m. The top-downwards
method was used with support of basement floors by steel box stanchions
filled with sand–cement grout. Each internal column was supported on a
single large-diameter bored pile founded on the bedrock with an allowable
bearing pressure of 5 MPa at depths between 45 and 65 m below ground level.
Basement construction and design 347

Fig. 8.16. Paris basement construction sequence, Provence Opera, Paris: (a) barrette sections installed; (b) ground floor
construction; (c) superstructure construction and excavation below ground floor slab; (d) superstructure and substructure

The basement construction proceeded as the basement floors were cast on


grade successively downwards from the ground floor slab. The prime design
requirement was to minimize ground movements outside the site. In order
to do this, the cut-off effectiveness of the diaphragm wall was improved by
grout injection a further 10 m below the wall. A full-scale pumping test was
348 Deep excavations

Fig. 8.16. (e) Load transfer of temporary support loads to floors; (f) geological conditions; (g) site plan; (h) anchor head detail
for composite soil anchor (courtesy of Soletanche–Bachy)

made and then back analysed. The results were used as design parameters for
prediction of ground movements due to full-scale excavation. A further
multiple-well pumping test was made after the completion of the diaphragm
wall box to simulate the construction dewatering. An array of pumping
wells at about 25 m centres was used to lower the water table over the site
Basement construction and design 349

Fig. 8.17. Dragon Centre, Hong Kong. Geological section.

area to the lowest basement level which is about 27 m below ground level. The
test was maintained for eleven days until steady-state flow was obtained and
then the pumps were switched off and the groundwater allowed to recover to
its initial level. The results showed the following.
(a) The total steady rate of flow from the wells was 25 m3 per hour with
measured hydrostatic water pressure on both sides of the wall at
steady state.
(b) Readings from piezometers within 5 m distance from the wells showed
that the groundwater was lowered to 23.5–29.5 m depth.
(c) With the exception of the eastern corner of the site, the drawdown
outside the site was an average 0.5 m at standpipes and 1.5 m at
piezometers.
(d ) Ground surface settlements were recorded in the range 4 to 16 mm with
negligible settlement at adjacent buildings.
(e) The maximum lateral deflections recorded by three inclinometers were
in the range 50 to 80 mm at the top of the diaphragm walls, with general
wall rotation from the toe.
Following the cessation of pumping, there was negligible recovery to the
ground surface settlements and the lateral wall deflections recovered to
about half of the maximum wall deflections during the pumping test.
Back analysis using the finite element computer program SEEP was made
to vary the rock mass permeability so that calculated pore-water pressures
matched those at steady state in the pumping test. The matched calculated
value was 1  108 m/s.
Further back analyses were made with the computer program FREW,
which assumes a linear elastic continuum between active and passive limits
on both sides of the wall. With some difficulty the observed wall displacements
350 Deep excavations

Fig. 8.18. Brittanic House, London. Construction sequence30


Basement construction and design 351

were used to estimate a value of K0 (0.3 to 0.4) and a relationship between soil
modulus EI and SPT value N.
Using these back analysed values, the wall deflections due to the subsequent
excavation were analysed and the ground settlements estimated.
Overall, the pumping test and the seepage analysis made thereafter high-
lighted the importance of achieving a good seal by grouting between the toe
of the diaphragm wall and the intact rock. Comparative analyses could be
made between a grouted and ungrouted box. The predicted wall deflections
reasonably agreed with observed ground movements after basement excava-
tion. In general the maximum ground surface settlement as predicted was
approximately 30 to 50% of the maximum wall deflection caused by the
pumping test and basement excavation respectively. The vital importance of
carrying out pumping tests for deep basements in water-bearing soils was
established on this job and continues as standard practice in Hong Kong.
Recharge wells were installed at the Dragon Centre but were not used
extensively following observations of settlements outside the basement as
excavation progressed.

Semi top-down construction


The use of very large openings in floor slabs that are designed as a frame to
provide lateral support to the external walls together with the use of excava-
tors with long dipper arms enables excavation to proceed below the top slab
support for considerable depths without intermediate floor support. This
method, with only a skeletal structure for the working platform and one inter-
mediate floor was used for a 25 m deep excavation for station construction on
the Singapore MRT29 . Following the construction of the roof, the skeletal
plan shape of the roof was used to excavate down to concourse level using
backhoes and long arm excavation. The concourse slab was then built with
similar large openings to allow further deeper excavation whilst propping
the external walls.

Combination of bottom-up and top-down methods


Basements of large plan area can benefit from a combination of both bottom-
up and top-down methods. This was used at a very early stage of top-
downwards construction in the UK in 1962–1963 at the site of an 18 m deep
excavation at Brittanic House in central London. The work, reported by
Cole and Burland30 , involved an external diaphragm wall box supported
temporarily by an earth berm during construction of the central raft followed
by bottom-upwards construction of the central core and tower columns. The
outer floors were then constructed top-downwards below an upper strutting
floor. The sequence is shown in Fig. 8.18 taken from Cole and Burland’s
paper. A similar combination of bottom-upwards and top-downwards
construction was used for the basement of the Main Tower in Frankfurt31 .
A piled raft was used to support the 198 m high tower on Frankfurt clay,
an over-consolidated but rather weak soil. The core of this tower was built
bottom-upwards in a conventional four-frame supported excavation in
advance of construction, top-downwards of the floors around the core
within a peripheral secant pile wall. This combination, shown in sequence
in Fig. 8.19, altered superstructure core construction to advance in parallel
with basement construction, to the advantage of the overall programme.

Peripheral sheeting or The system adopted to sheet or wall the periphery of the excavation will be
walling influenced by the choice of basement construction method, the suitability of
ground and groundwater conditions, the need to build close to site boundaries
352 Deep excavations

Fig. 8.19. Main Tower, Frankfurt. Construction sequence31

and minimize wall thicknesses and, not least, by the local availability of
materials and specialist plant and equipment. Peripheral sheeting methods
are generally the following:
. anchored underpinning: reinforced concrete plates and grouted soil
. king post wall: vertical soldiers and horizontal laggings or reinforced
concrete skin wall
. sheet piling
. contiguous bored piling
. secant piling
. soldier pile tremie concrete method (SPTC) – as used previously in the
USA.
. diaphragm walls
k reinforced concrete cast in situ
k precast reinforced concrete
k post-tensioned.
The general features of each method were covered in Chapter 4, but their
particular application to basement works is reviewed below.

Anchored underpinning
Where the total excavation depth of basement work is typically in the range
8 to 12 m and ground conditions are dry and capable of supporting a face
1.5 to 2 m deep and of similar length, the anchored plate method provides
an economical temporary wall support if permission to install anchors outside
the curtilage of the site is forthcoming. In conditions where soils lack the
strength to stand unsupported to these modest depths, pre-grouting may
Basement construction and design 353

prove worthwhile in granular soils. Where foundation loads from adjacent


structures are such to necessitate transfer of load below the proposed excava-
tion depth, pre-injection of the subsoil below the existing foundations, with
anchorage to avoid lateral movement of the grouted soil mass, may prove
an economical alternative to conventional mass concrete underpinning.

King post wall


The king post or soldier pile and horizontal timbered wall, previously widely
used in North America, has become increasingly popular for basement con-
struction in Europe in recent years. For use in shallow excavations, the king
posts may be cantilevered or propped by raking shores or anchored in succes-
sive layers as bulk excavation proceeds in deeper basement works. The king
posts may be double joist or channel units battened together to allow the
anchor to conveniently pass between. Figure 8.20 shows soldier pile walls
supported by anchors constructed in basement works in Saudi Arabia to
depths exceeding 20 m. Subsoil conditions were layered washdown silts and
silty sands, with a groundwater table at formation level or a small height above.
The method requires moderately dry ground conditions with soil of suffi-
cient strength to maintain a vertical face prior to support from the horizontal
lagging being placed. King post centres vary from 1.5 to 3.5 m, depending on
soil strength, depth of excavation and surcharge loads. A popular innovation
is the use of in situ reinforced concrete skin walls cast against the exposed
soil face, with thicknesses between 150 and 200 mm. The walls, which span
horizontally between king posts, are cast in lifts between 1 and 1.5 m high,
depending on the ability of the soil to stand without support.
The king post excavation may be bored by auger rig or, where headroom is
limited, low-headroom rigs may be necessary. The toe of the king post is
usually concreted to basement formation level, although deeper king posts
surrounded by sandy gravel washed in may be preferred if the king posts
are to be subsequently extracted. It is economical to use the lined face of
the timber laggings or the face of the skin wall as a back shutter to the perma-
nent basement wall, but allowance must be made for tolerances in the king
post wall construction.
Due to the width of the king post wall and the permanent wall construction
it may be necessary to drill the king post bores close to the site boundary.
Where an existing structure is close to this boundary the minimum distance
between king post bore and site boundary will be determined by the minimum

Fig. 8.20. Anchored soldier pile walls used in dry layered sand and clayey silt soil for deep basement construction.
Note the unimpaired access for plant and site operations, Medinah, Saudi Arabia (courtesy of NCF)
354 Deep excavations

Table 8.3 Minimum distances between soil support system and site boundary for various types of installation plant

Support system Installation plant Distancea

Underpinning Conventional bulk excavation plant, e.g. Nil


hydraulic excavator with hydraulic grab
Steel sheet piling Crane and piling hammer/tracked hydraulic 500 mm, rear of sheeters
piling machine to face of boundary wall
Contiguous bored pile wall Bored pile:
tripod equipment typical 600 diameter pile 150 mm
Large-diameter rig:
Hughes CEZ 300 typical 740 pile 385 mm
Hughes CEZ 450 typical 750 pile 385 mm
Hughes KCA 100/130 typical 900 pile 450 mm
Contiguous bored pile wall CFA rig:
and hard–soft secant wall Soilmec CM 45 typical 750 pile 350 mm
Soilmec CM 48E typical 750 pile 350 mm
Rotary rig CFA:
Bauer BG 11 typical 500 pile 400 mm
Bauer BG 14 typical 600 pile 300 mm
Bauer BG 26 typical 600 pile 150 mm
Bauer BG 30 typical 750 pile 100 mm
Hard–hard secant wall Bauer BG 7 FOW method Nil
254, 273, 305, 343, and 406 mm diameter
Diaphragm wall Rope suspended grab 200 mm
City Cutter 150 mm
Hydrofraise 150 mm
Berlin walls: soldier piles Rotary piling equipment 600 diameter bore 400 mm
and horizontal lagging
Manual excavation: 200 mm up to 6 m depth
hydraulic excavator and trench box
a
Minimum distance between outer face of support system and site boundary. Distances quoted are those at ground level;
consideration must be given to verticality tolerance of support system.

Fig. 8.21. Vertical cross-section


through part of a substructure
basement, Marylebone,
London32
Basement construction and design 355

overhang of the auger rig from the rear face of the pile bore. Table 8.3 shows
the minimum dimension between an existing structure and the outer face of
the new wall for various wall types.
It is usual for king post wall construction to be used only as temporary soil
support. An exception was described by Mair32 . In Marylebone, London, a
king post wall was used as the permanent peripheral retaining wall in a top-
downwards type two-storey basement construction. A cross-section of the
construction is shown in Fig. 8.21. King post wall construction was feasible
because the whole depth of the basement, to 8 m, was accommodated
within dry sands and gravels 11.5 m deep, below which was London clay.
The water table in the sands and gravels was below the basement formation
level, at a depth of 9.5 m. The king post centres were 1.5 to 1.8 m, and mass
concrete infill was placed in 1 m lifts as excavation proceeded. As usual in
top-downwards construction, the king post wall was successively propped
by ground, lower ground and basement floors.

Sheet piling
The use of sheet piling for temporary soil support to basement construction in
urban areas has declined as environmental controls on noise and vibration
progressively strengthen. Only where sheet piles can be installed by hydraulic
means, particularly in cohesive soils, can the effects of these controls on noise
and vibration be avoided. The use of Giken hydraulic press equipment has
steadily increased in the UK in a wider range of soils. Typically used in city
centres, penetration through stiff clays such as London clay and Gault clay
is restricted to a pile length of 16 to 20 m. Water jetting or lubrication of
the sheeter surface with water may be necessary to achieve penetration.
Where noise and vibration are critical, as in most city centres, sheet piles
for deep basements can be installed by the combined use of slurry trench
and sheet piling methods. The sheet piles are pitched into slurry trenches
filled with cementitious self-hardening slurry and the toes of the piles con-
creted in by tremie pipe up to basement formation level. The technique,
although uneconomic at first sight, is environmentally friendly. The sheet
pile section can be selected on the basis of flexural stress without consideration
of driving stresses, and considerable accuracy can be achieved in pitching the
sheeters into the slurry trench. The sheeters can obtain support from ground
anchors with conventional steel walings or from bracing or raking shores.
Interlocks in sheet piles cannot be assumed to be completely watertight in
water-bearing ground unless provision is made for sealing them. Sheet piles
sealed by welding at the exposed surface after excavation or prior to installa-
tion by a steel contaminant section to allow a bituminous or polymer material
seal of the interlock can remedy lack of watertightness. The use of welding
may be necessary to augment the sealant compound. This system, now
often the subject of contractual guarantees of maximum wall permeability,
is becoming more popular in the UK and is further promoting the use of
sheet piles for permanent works. Figure 8.22 shows typical joint details
between permanent sheet piles and the basement slab.

Contiguous bored piling


The cheapest type of concrete piled wall is the contiguous piled wall. The use
of modern continuous flight auger (CFA) rigs allows high output in a wide
variety of soils. The depth of pile is limited to the order of 18 to 20 m by
the difficulty of inserting reinforcement cages to greater depths through wet
concrete and the lack of water resistance in water bearing ground due to
the gaps between piles. A structural facing may be applied as sprayed concrete
or an in situ reinforced concrete lining.
356 Deep excavations

Fig. 8.22. Typical joint details


between permanent sheet piles
and the basement slab

Prior to the advent of CFA rigs in these jobs only a short length of top
casing was necessary, separating the piles by approximately 50 mm. The
piled wall depth was limited to some extent by the verticality tolerance that
could be obtained by the augers, typically 1% with depth. In the UK many
basement walls were constructed in this way in the 1960s to the 1980s. The
walls were anchored temporarily using steel walings or were braced with
strutting or rakers. Grouting was used in permeable soils where groundwater
entered in the gap between piles. In some instances the intended use of the
basement allowed the bored pile wall constructed in this way to remain
unlined, while for high-grade basements the piled walls were lined with
reinforced concrete or an independent, non-load-bearing blockwork wall.
The advent of CFA piling rigs in the early 1980s, with their ability to
operate without casings (even without a top casing), their high output and,
for smaller low-torque machines, their ease of transport and erection on
site, produced economies which allowed them to replace conventional
augers in most soil conditions. CFA piles for wall construction are typically
300, 450, 600, 900, 1000 and 1200 mm in diameter.
Hydraulic auger cleaners, introduced to prevent soil falling on personnel
also avoid contamination of new concrete with soil. Other innovations include
a projecting tremie pipe from the base of the auger to pump concrete to a
lower level than the core of soil progressively lifted by the auger. Standards
of quality control for CFA piles are now much improved by in-cab monitor-
ing. Data referring to auger depth, torque applied during pile excavation, rate
Basement construction and design 357

of withdrawal and concrete pressure are relayed to the rig operator and a more
recent IT development enables this data to be transmitted to an off-site
terminal such as the contractor’s or engineer’s office. Further progress is
needed to measure concrete pressure near the point of discharge.
CFA rigs operate in a wide range of soil and soft rock conditions, but hard
rocks, rock chalk and strong mudstones cause obstruction and make the rigs
uneconomic. Minimum distances for rig operation from existing wall bound-
aries are shown for a range of CFA and rotary rigs in Table 8.3. Some stated
dimensions may be reduced by modifying the standard equipment. Bauer,
in particular, has introduced a purpose-made rig to operate with reduced
minimum distance.

Secant piling
Improvements in rotary rig and equipment design have, as with contiguous
pile walls, changed construction methods for secant piles in recent years.
Until the 1980s the Benoto rig was the primary method of installing secant
pile walls, cutting the concrete of female piles to interlock male piles between
them. A heavy-duty hammer grab was used with twin-wall lockable tempor-
ary casing equipped with a cutting edge for excavation, the casing being

Fig. 8.23. Configuration of


reinforcement in contiguous and
secant pile walls62
358 Deep excavations

oscillated under crowd to achieve penetration. The introduction of powerful


rotary machines equipped with CFAs or casing oscillators has enabled
much higher output rates than the traditional Benoto rig, which had been
used for almost 50 years in Europe and the UK.
Benoto rigs used bores of 880, 1080 and 1180 mm diameter. At present,
secant piles installed by CFAs have diameters in the range 450 to 750 mm
diameter, whereas cased secant piles typically range from 750 to 1180 mm
diameter. Reinforcement, usually limited to male piles, may be from cages
or, where required for shear or flexural strength, joist sections may be used.
A comparison between contiguous pile and secant pile configuration and re-
inforcement is shown in Fig. 8.23. The method provides a near-waterproof
wall for both temporary and permanent soil support and, with CFA, rotary
or hammer grab excavation, secant walls can penetrate most soils and rock
obstructions to maximum depths of 30 to 40 m. The secant pile basement
wall is capable of supporting both lateral load from soil and groundwater
and vertical load from the curtilage of the superstructure. Vertical loads
may be transmitted to the piles by a reinforced concrete capping beam or,
for lesser loads, by shear on the vertical contact face between adjacent piles.
Since 1985, the use in the UK of hard–soft and hard–firm secant piles, for
depths to 20 m by CFA rigs and for greater depths with high-torque rigs with
casing oscillators, has improved production, reduced cost and to some extent
reduced the use of diaphragm walling. The unreinforced female piles, cast in
the bentonite–cement or bentonite–cement–Pulverized Fuel Ash mix, are cut
by the male piles which are reinforced and concreted in the normal way. This
type of hard–soft secant wall may not be satisfactory for walls requiring high
standards of waterproofness or long-term durability, although internal
reinforced concrete lining walls may remedy this situation. Pile rigs used to
install secant piles before and after the 1980s are shown in Fig. 8.24.
Other innovations in recent years include mix-in-place piles. This process
utilizes cement slurry, pumped through the hollow stem of the auger, mixed
with sandy soils during boring and extraction. The unreinforced female
piles constructed in this way vary with the sand–cement ratio (with compres-
sive strength of the order of 15 N/mm2 ) and can be alternatively spaced with
male piles constructed with concrete and reinforced in the normal way.

Fig. 8.24. Piling rigs used for


secant pile installation: (a) BG
26 rig (courtesy of Bauer); (b)
Benoto rig with hammer grab
(courtesy of Lilley)
Basement construction and design 359

Fig. 8.25. CDSM piles:


suggested pile spacing35

The use of cement deep soil mixing (CDSM) to produce blocks of over-
lapping piles for nailed or self-supporting gravity-retaining structures has
recently progressed rapidly in the USA after early development in Japan.
Yang33 summarized the use of CDSM for cut-off walls and excavation
support in addition to its use in ground stabilization. A later review of deep
mixing technology was given by Porbaha et al. in 2001.34 The method, devel-
oped in the 1970s from soil–lime mixing methods uses a triple auger machine
to produce pile sections in the range 550 to 990 mm in stiffer soils. Steel H
sections are installed as flexural reinforcement in retaining walls prior to the
hardening of the soil–cement mixture. The soil–cement is designed to arch
between adjacent steel H sections. Taki and Yang35 suggested a spacing of
H sections based on empiricism, as shown in Fig. 8.25. Cement deep soil
mixed piles reinforced with steel joists for soil support and groundwater
control are shown in Fig. 8.26. A soil–cement wall for both groundwater
control and for soil retention in highly permeable coralline conditions was
used for a two-storey basement at the Marin Tower project in Hawaii. The
excavation was only 30 m from the harbour and a high groundwater level.
The partial cut-off scheme was achieved by a soil–cement wall of average
depth 14 m, using 55 cm soil mix piles installed by triple auger. The coral
limestone was ground down to gravel size by the augers without pre-drilling
for thorough mixing with the cement grout. Mix designs with cement dosages
of 300 to 500 kg/m3 of in situ soil were used.
The relatively slow take up of soil mixing processes for excavation support
in Continental Europe and the UK may be associated with the lack of
adequate QA methods. Bruce et al.36 reported current methods in 2000.

Support for piled basement walls by ground and rock anchors


Secant pile walls may be supported by strutting and walings, rakers with
walings or ground anchors. Where used, anchors may be taken axially
through the piles or through the contact face between the piles. It may be
sufficient to use anchors at alternate male piles or every fourth pile, depending
on the extent of lateral load, anchor capacity, and the available shear resis-
tance mobilized on the contact face between adjacent piles.
The decision to use ground anchors as a temporary wall support for any
wall system will be based on practicality, cost and installation time, which
are all influenced by:
. depth of the basement
. groundwater conditions during anchor installation and, thereafter,
during basement construction
. subsoil conditions and their suitability for accommodating anchors of
adequate capacity economically
. maximum permissible soil and basement wall movements and the plan
shape of the basement; the susceptibility of adjacent existing structures
to soil movement caused by the basement excavation
360 Deep excavations

Fig. 8.26. Soil–cement wall for


excavation support and
groundwater control, Tokyo,
Japan (courtesy of Raito Inc.)

. the basement construction programme


. the aggressiveness, if any, of groundwater
. the location of existing services
. the location of neighbouring substructures and/or basements
. legal permissions to accommodate anchorages outside the curtilage of
the construction site
. the risk of obstruction of future works within the construction site by
the presence of anchors.
This list, in no order of priority, may not be exhaustive on any particular site,
but indicates those items requiring earliest consideration. Some items are self-
explanatory. Subsoil conditions will indicate likely anchor capacities, compact
granular soils generally being preferred to cohesive soils in the fixed zone of
each tier of anchors. Subsoil and groundwater conditions will dictate drilling
costs for anchor installation, and the aggressiveness of groundwater and the
period of use of the anchors will dictate the need for corrosion protection.
Basement construction and design 361

The location of adjacent substructures and services will determine the


practicality of installing anchors at the required elevations, and the plan
shape of the basement may determine any difficulties caused by obstructing
the drilling of anchors from an adjacent re-entrant basement wall. Above
all, legal permissions and licences must be available from owners of adjacent
land or highway authorities to allow anchor installation outside the site area.
Where anchors are likely to obstruct future construction, a removable-type
anchor may be necessary.
Wall movement is likely to be reduced by the use of anchors that are
stressed after installation, particularly when fixed-length anchors are founded
in competent medium-dense or dense granular soil. When the anchors are
founded in stiff cohesive soils only short-term benefit may be gained.
The programme implications of anchor installation also require exam-
ination and depend on the timing of bulk excavation following anchor
installation. The sequence of drilling, tendon installation, grouting, grout
strengthening and stressing for each bank of anchors has to be phased
within the overall excavation programme. Comparison with an overall con-
struction programme using alternative forms of wall support may be neces-
sary, taking into account the improved construction outputs obtained by
unobtained work areas achieved by anchoring.
The design and construction of ground anchors, described in Chapter 7, is
explained in detail in BS 808137 , which contains an extensive bibliography on
ground and rock anchors. Littlejohn and Bruce38 reviewed the state-of-the-art in
rock anchoring and Barley39 updated this, in particular giving observed bond
stress values of both straight shafted and under-reamed anchors in chalk,
mudstones, siltstones, shales, marls and sandstones.

Soldier pile tremie concrete method


This North American practice, popular in the 1960s, of modifying the king
post wall method by excavating a panel by grab under bentonite slurry
between the king posts and filling the panel excavation with unreinforced
concrete by tremie, has rarely been used in Europe. In Germany, however,
a modification of this method, using mesh-reinforced gunite sprayed between
and over the king posts successively as built excavation proceeds, has gained
acceptance. Although the construction provides only temporary soil support
it is particularly economical in dry granular soils which can be excavated to a
vertical face for 2 to 3 m without collapse in shallow to medium-depth
basements. The overall thicknesses of temporary and permanent walls and
working space at the site boundary are usually not excessive.

Structural diaphragm walls


It may be argued that the most significant advance in recent years in basement
construction has been the introduction of the structural diaphragm wall. The
principal advantages of this form of construction, introduced into Europe
by Icos in the 1950s and 1960s, are: the dual use of the wall to provide
both temporary and permanent soil support; the efficiency in bending of the
rectangular wall section compared with the circular pile cross-section
used previously; the reduction in noise and vibration during installation
compared with percussive drilling of sheet piling; the ease of installation of
propping, strutting and anchoring against the wall face; the ease of applying
finishes to the flat wall face; the ability of the walls to transfer vertical loads;
and its use to depths generally in excess of other forms of wall construction.
The dual support provided by the diaphragm wall at construction stage and
then during the basement life was often sufficiently economical to justify it on
financial savings alone compared with other walling methods. Advantages
362 Deep excavations

such as the minimum thickness of construction required by the diaphragm


wall for both temporary and permanent soil support were a bonus.
The principal disadvantages of diaphragm walling are the risk of loss or
spillage of bentonite slurry, the relatively high cost of cleaning and the dis-
posal of the slurry, the site space needed for large reinforcement cages and
the large cranes needed to handle them. Above all, the need for continuity
in the construction process from excavation through concreting to removal
of temporary stop end formers, is a disadvantage of the method. Structural
diaphragm walls still remain the preferred method of walling for deep base-
ments and concrete piled walls, particularly secant walls, have only tended
to replace diaphragm walling in basement works of medium depth.
Icos40 gave details of a wide range of basement constructions. These base-
ments were generally of moderate depth, perhaps two or three basement
storeys. The diaphragm walls were all excavated by cable grab mounted on
tripods on rails. The wall depth attainable by this equipment was consider-
able, however – up to 28 m in one example in Paris. The panels used were
straight, L- and T-shaped and corrugated in plan. Figure 8.27 shows the
diaphragm wall options for basement excavations recommended by Icos.
As patent protection on the Icos wall waned, specialist firms in Europe
introduced alternative excavation equipment, although in later years Icos
persevered with rope grabs mounted on heavy tracked cranes. In Europe,
kelly bar mounted hydraulic grabs became popular in the 1970s and the
early 1980s. These grabs, which were capable of excavating wall widths
between 500 and 1500 mm, were mounted on single or telescopic kelly bars
to maximum depths of approximately 25 m. Panels were dug in a series of
grab ‘bites’ which were typically each 2.8 m long.
Excavation in medium-strong to strong rocks for diaphragm wall works
was difficult for all specialist firms at this time. The usual method from the
1950s to the early 1980s was to use a drop chisel progressively along the

Fig. 8.27. Alternative methods


of wall support and plan forms
of diaphragm walls for
basement construction
(courtesy of Icos)
Basement construction and design 363

(a)

Fig. 8.28. Development of diaphragm wall excavation rigs 1950s–1960s: (a) (i) Icos tripod rig; (ii) action of Else bucket scraper;
(iii) excavation with bucket scraper; (iv) hydraulic grab, Kelly mounted; (v) rock chisel
364 Deep excavations

panel under the slurry and alternatively grab the arisings and chisel again.
Progress was slow and vibration often quite severe in strong rock. In excep-
tional cases rotary core barrettes were drilled successively into the rock
along each panel under the slurry, and both Icos and Soletanche developed
rigs incorporating percussive chiselling with direct or reverse slurry circulation
to remove cuttings.
The introduction of a rail-mounted reverse circulation rig for excavation in
soil and soft rock by the Tone Boring Company of Japan in the late 1970s was
followed by the development of reverse circulation equipment known as the
Hydrofraise by Soletanche. In the mid 1980s Bauer and Casagrande
produced similar equipment. The Bauer Trenchcutter has been developed
into smaller, more manoeuvrable rigs known as City Cutters and more
recently, Mini Cutters. In turn, the Hydrofraise has been made more compact
for city sites in joint development between Soletanche and Rodio to produce
the HL 4000 track-mounted rig. Diaphragm wall rigs developed between the
1950s and 1990s are illustrated in Fig. 8.28. The improved manoeuvrability of
the Bauer City Cutter rig is shown in Fig. 8.29.

(b)

Fig. 8.28. (b) (i) and (ii) Crane-mounted grabs; mechanism of submersible motor drill; (iii), (iv) and (v) Tone Long Wall Drill;
(vi) Tone Long Wall Drill: vertical section through cutter
Basement construction and design 365

Fig. 8.28. (c) (i) Tone


Electro-Mill Drill; (ii) Bauer BC
30 Trench Cutter rig;
(iii) Bauer MBC 30 Trench
Cutter rig
366 Deep excavations

Fig. 8.29. Bauer City Cutter rig:


use of alignment device on
small base machine to improve
manoeuvrability in limited
space

Excavation for diaphragm walls is currently made by a variety of rigs


depending on job size, wall depth and soil and rock properties. The reverse
circulation machines have the added qualities of silent and vibration-free
excavation through a wide range of soils and rocks but, even so, conventional
crane-mounted rope grabs are frequently used for walls of moderate depth on
small to medium sized jobs. Heavy mechanical grabs up to 9 tonnes in weight
are favoured for smaller, shallower basements particularly in cohesive soils.
The recent re-introduction of hydraulic grabs by Bauer and others, rope
mounted, has accompanied the use of electronic monitoring to improve
installation tolerances. The usefulness of rams to allow steerage of these
rope suspended hydraulic grabs is less certain.
Cutter reverse circulation rigs saw two significant developments in the early
1990s: the use of rock roller bits on the vertical cutter wheels to allow more
efficient excavation in moderately strong and strong rocks, and the construc-
tion of a compact, low-headroom rig known as the MBC 30 Trench Cutter.
This rig (Fig. 8.30) has its own carrier system mounted on a railway bogie
or on crawlers, so a conventional crane is not needed to carry the cutter.
The overall dimensions of the complete unit are reduced to 4.7 m long,
4.1 m wide and 5.0 m high (6.0 m high when crawler mounted). Trench
widths vary from 640 to 1500 mm; the standard trench length is 2790 mm;
and the maximum cutting depth is 55 m.
Excavation in rock for diaphragm walls still remains arduous and expen-
sive. Mention should be made of a practice, typically in Hong Kong, of
Basement construction and design 367

Fig. 8.30. Bauer MBC 30 Trench


Cutter rig

underpinning deep diaphragm walls where formation level is below rock head,
using shear piles drilled through the diaphragm wall, often bundles of T50
rebar, to support the wall in the temporary condition. Figure 8.31 shows a
cross section.
The cutters of a modern Soletanche rig, in this instance for a 1500 mm wide
wall, are shown in Fig. 8.32.

Fig. 8.31. Underpinning of


diaphragm wall for excavation
of formation level below
rockhead; typical Hong Kong
practice
368 Deep excavations

Fig. 8.32. Cutters for a modern


Soletanche cutter rig (courtesy
Soletanche–Bachy)

In a recent paper, Guillaud and Hamelin41 reviewed the development of


diaphragm wall excavation plant during the 1990s and beyond, to 2002.
They describe the early sites as cluttered areas, with noisy machines, a liberal
covering of slurry over ground surfaces and nearby streets, and long road
closures. Such sites should no longer remain; modern, compact low-noise
machines and similar slurry treatment equipment have allowed environ-
mentally friendly diaphragm wall excavation to greater depths with greater
accuracy. Guillaud and Hamelin review excavation under two headings,
cutters for continuous excavation and hydraulic or mechanical grabs, rope
or Kelly suspended (or a combination of both). For cutters they refer to
recent Soletanche/Bachy cutter developments, referring to the HC03 machine,
particularly for confined, city jobs requiring less than 5 m headroom. Despite
its 90 tonne weight it can excavate to 50 m depth (see Fig. 8.33(a)). The leading
details of this machine are as follows:
. diesel engine delivery 370 kW at 2400 rpm
. max hydraulic pressure 32 MPa
. max depth 50 m
. max torque 80 kNm at 32 MPa
. suction pump: 450 m3 /hr
. max pressure on tool 25 tonnes
. total weight 93 tonnes
. cutter drums 650, 800, 1000 and 1200 mm
. excavated length 2400 mm or 2800 mm.
Regarding noise emissions, the noise from the HC03 is no more than 72 dB in
the cab and 80 dB within a hemisphere of 16 m radius.
Basement construction and design 369

(a)

Fig. 8.33. (a) Modern


Hydrofraise rig, type HC03
(courtesy Soletanche–Bachy);
(b) modern Evolution
Hydrofraise rig (courtesy
Soletanche–Bachy) (b)

For deep diaphragm walls, the Evolution Hydrofraise rig (see Fig. 8.33(b))
digs to 70 m depth. Other rigs, capable of excavation to more than 100 m
display the same accuracy, reliability and ease of erection as the smaller
machines. (The Hydrofraise used in 1997 by Obayasti Corp. at Nagoya to
build a buried LNG tank inside a circular diaphragm wall, 1.8 m thick
weighed 245 tonnes.)
370 Deep excavations

Fig. 8.34. Swivelling hydraulic


grab, type KS 3000 (courtesy
Soletanche–Bachy)

Developments to grabs are shown in the Soletanche–Bachy KS 3000 (see


Fig. 8.34), a swivelling hydraulic grab mounted on a hydraulic crane which
also powers the grab. The rig is very compact and can operate very close to
existing walls. Steerable grabs are used on the latest variants.
Both grab and Hydrofraise machines benefit from an automatic control and
reporting system called SAKSO. This system has three stages of automation:
manual (operator in full control); teaching (operator tells the system what
movements to make); and automatic (system repeats movements learnt).
The movements are fourfold: jib orientation, jib angle, grab orientation,
hose winder retract. The real-time monitoring systems of cutters and grabs
now include features to include not only deflection on the XX and YY axes,
but also rotation about the ZZ axis (corkscrewing) and deviation from the
vertical (drift).

Practical design and construction of diaphragm walls


A number of items require consideration in the pre-planning and design of
diaphragm wall works.
(a) Panel size. The panel length will vary from a minimum of one grab bite
to a multiple of grab bites typically 6 to 7 m preferably made up of a
number of whole bites with smaller widths between them. Grab bites
vary between 2.3 and 2.8 m. The panel length will include two stop
ends for the initial (primary) panels, or one stop end for mixed panels
dug next to a completed panel. Secondary panels are those dug between
two previously concreted panels. Cut joints can be used with cutter rigs,
and are particularly advantageous for deep walls. Views of the surfaces
of a cut joint from a test panel are shown in Fig. 8.35. The length of the
panels must first be assessed on the basis of panel stability (DIN
Standard 412642 gives methods of assessing panel stability for varying
subsoils and surcharge loadings). It is necessary to limit panel length,
and hence panel volume, to ensure that concrete outputs are sufficient
to fill the panel within a reasonable period taking maximum daily
working hours into account. Panels of modest depth can often be
dimensioned to ensure excavation of one panel each day (say a 20 m
deep panel 4.5 m long dug in stiff clay with no obstructions at an
average of 5.5 m2 per hour, rope grab excavation for 10 hr, one daily
Basement construction and design 371

Fig. 8.35. Surfaces of cut joint


from test panel

shift). Such an arrangement gives output a rhythm. Panel size is there-


fore a decision for the specialist contractor.
(b) Excavation sequence. The sequence of excavation to the basement walls
is planned to minimize rig movement and avoid moving pipework from
panel to panel. Where stage completions of the peripheral walls are
agreed, or where top-downwards construction encourages simultaneous
use of the site by both superstructure and substructure contractors,
detailed programming must allow access to panels for rigs, muck-away
trucks or slurry removal vehicles, service cranes and concrete trucks,
and allow curing of concrete in complete panels prior to adjacent
mixed or secondary panels.
(c) Panel stability. Working platform levels must be selected with an aware-
ness of the minimum differential between the head of bentonite slurry
in the panel excavation and groundwater level next to the panel. The
minimum value to ensure stability is 1.0 m, and many specialists
would specify a preference for 1.5 m, especially where groundwater
flows in permeable strata.
(d ) Guide trench construction. The standard of diaphragm wall construction
is itself influenced by standards of temporary guide wall construction.
The guide walls must be sufficiently robust to avoid movement due to
loads from excavation rigs, service cranes or reinforcement cages and
reaction from stop end jacking systems. Reusable precast concrete
guide wall sections have been successfully used on T-shaped panels
and cellular walls but each precast unit must be interlocked by a
bolted mechanical joint to ensure the same standard of rigidity as in
situ concrete walls. It is essential to maintain continuity between in
situ guide wall pours by reinforcement passing through the construction
joints.
(e) Wall–slab construction joints. Joints between basement rafts and
diaphragm wall, and between intermediate basement floors and wall,
can transmit vertical shear or, where necessary, bending moment. Alter-
native forms of joint construction are shown in Fig. 8.36. Threaded-end
couplers (such as Lenton couplers) can be used to develop the full
strength of reinforcement bars from within a recess in the face of the
372 Deep excavations

Fig. 8.36. Vertical cross-


sections of typical horizontal
joints between diaphragm wall
and basement floor slab:
(a) to achieve resistance to
groundwater ingress and
transmission of vertical shear;
(b) moment connection by
couplers

diaphragm wall at the junction with the slab. Bend out bars can be used
instead of couplers, although the closeness of the bar spacing and the
diameter of these bars may be limited by the ability to house them in
the face of the joint.
( f ) Box-outs. The depth of box-out will normally be limited to the concrete
cover on the main reinforcement. Although polystyrene is used as a
Basement construction and design 373

Fig. 8.36. (c) Variation for a structural hinge; (d) Variation for structural continuity

box-out material there is a risk that the box-out will be displaced unless
a timber frame with a thin plywood cover secured to the front to
ensure its rigidity is used to contain the polystyrene. It is usually a
disadvantage to extend the depth of the box-out behind the vertical
reinforcement due in part to the difficulty of adequate preparation of
this face of the joint but particularly because of the risk of entrapment
of heavy slurry below the box out during concreting.
(g) Reinforcement cage and density. It is unwise to allow the requirements
of calculated shear, moment or crack width to make the spacing of
horizontal binders and vertical main steel too small. DIN Standard
412642 advises that to ensure no slurry inclusions remain in the con-
crete, the differences between the flow resistances in adjoining zones
in plan in the panel should be kept as small as possible. Minimum
spacing of bars for both single- and double-layer vertical steel and
horizontal binders is reproduced in Fig. 8.37.
The increasing depth of walls and the higher calculated flexural
strengths required from them has led to cages of considerable tonnage.
These cages are usually joined in sections over the panel using couplers
to join bars in each section. A cage awaiting the arrival of the next
section is shown in Fig. 8.38. The vertical bars are staggered to avoid
couplers forming a block to concrete flow at one level. Cages are some-
times fabricated off-site in sections and transported to site overnight (see
Fig. 8.39). The maximum panel length may be restricted because of
transport although multiple cages may be used to allow an economical
panel length. Total cage weights of jointed sections for deep walls may
demand very large craneage for the final lift into the panel. A 90 tonne
weight cage is shown during final insertion into the panel in Fig. 8.40.
(h) Tremie operation. Reinforcement cage detailing must allow sufficient
vertical access for tremie tubes. Very small single-grab bite panels can
only accommodate a single tremie tube, and even with almost continu-
ous concrete supply from truck mixers it is difficult to obtain an average
pour rate greater than 40 m3 per hour. For larger panels a second tremie
pipe can be used and with continuous concrete delivers a concreting rate
between 60 and 80 m3 per hour. In deep, large, T-shaped panels it may be
necessary to use three tremies to maintain a uniform upper surface to the
concrete as the pour continues, but generally two tremies are sufficient.
374 Deep excavations

Fig. 8.37. Minimum spacing and concrete cover to reinforcement in diaphragm walls as recommended by DIN standard 412642
Basement construction and design 375

Fig. 8.38. Reinforcement cage


in panel excavation awaiting
jointing with a further cage
length

Fig. 8.39. Cage transportation

(i) Concreting rate. In the Author’s view there is considerable risk of poor
panel concrete when concreting rates drop below 15 to 20 m3 per hour
per tremie. Even when concrete mix quality varies from the optimum, a
high rate of concreting may be sufficient to displace the slurry, scour the
surface of the reinforcement bars and flow between the reinforcement
and around the box-outs. Where the density of reinforcement bar is
high, and especially with multi-layers of bars, a high concreting rate
and a very workable cohesive concrete are essential.
(j) Slurry reuse. The earliest diaphragm wall jobs by Icos did not have the
equipment to clean bentonite slurry after excavation and concreting.
The slurry was used once and, after storage for much of the pour
volume, was carted from site in road tankers or, on the earliest jobs,
pumped into public sewers. With the advent of reverse circulation
376 Deep excavations

Fig. 8.40. Heavy cage during


final lowering into panel. In this
panel, two cages have been
used in separate vertical
sections

rigs which used large quantities of slurry to excavate each panel, clean-
ing technology from the oil industry was used to design and build
shaker screens, hydrocyclones and at a later date, centrifuges and
presses, to clean the slurry. Compact, transportable units were made
which could be brought to site and quickly mobilized for use. Nowa-
days, slurry is cleaned and reused as a matter of routine on virtually
all diaphragm wall works to clean bentonite slurry where either grab
or reverse circulation rigs are used.
(k) Stop ends and extraction. Initially, tubular steel stop ends were used to
form semi-circular joints between diaphragm wall panels. This practice,
popular with Icos, the originator, and other European firms, gradually
changed as rectangular formers gained acceptance. Both types of stop
end were extracted as the concrete at the bottom of the panel started
to set and gain strength. In very large pours, it was necessary to use
high dosages of retarder additive within the concrete to ensure that
the set was delayed. Even so, extraction of the stop ends often began
before the concrete pour had been completed. Types of vertical panel
joints are shown in Fig. 8.41. With the advent of reverse circulation
rigs with vertically mounted cutter wheels, it became possible to cut
the vertical surface at the end of a concreted panel during excavation
of the adjacent panel. This cutting back of the concrete surface avoided
the use of temporary formers, and the new concrete in the second panel
could be poured against a true vertical surface on the first panel. This
procedure has since become less favoured because of the risk of
heavy, calcium-contaminated slurry remaining near the cut surface,
Basement construction and design 377

Fig. 8.41. Types of vertical


panel joint used in diaphragm
waling, shown in horizontal
section

which could contaminate the concrete within the second panel, near
the vertical joint. This slurry-contaminated concrete often proved to
be porous and led to leakage of groundwater into the basement after
excavation. The CWS-type stop end former developed by Bachy
incorporates single, twin or triple water bars cast into the vertical joint
and is released from the vertical surface of the concrete pour after the
concrete has hardened during excavation of the second panel. This
type of joint (and its derivatives by competing specialist firms) has now
gained widespread useage where resistance to groundwater penetration
is needed. The CWS joint details are shown in Fig. 8.42. Examples of
CWS joint construction for a 1.5 m wide wall, 55 m deep with a double
water bar are shown in Fig. 8.43. A temporary plywood former was
used on the CWS stop end to facilitate its removal.

Recent developments in diaphragm wall works


Since the conception by Veder of reinforced concrete walls cast into trenches
dug under bentonite slurry, and development of the process by Icos and
others, there have been many improvements in both application and
378 Deep excavations

Fig. 8.42. CWS joint detail: (a) pulling out stop end sideways after excavation of the adjacent panel; (b) stop end blades
installed in the CWS joint; (c) CWS joint before concreting; (d) CWS joint with water stop installed (courtesy of
Soletanche–Bachy)

mechanical plant. Developments reviewed by Puller and Puller43 in 1992 and


since in terms of plant innovation by Guillaud and Hamelin41 are as follows:
(a) the use of polymeric slurries for excavation support, avoiding effects of
some contaminants, reducing pumping energy, avoiding wall cake
thickness, reducing slurry disposal costs
(b) the development of structurally efficient diaphragm wall plan shapes
based on improved joint efficiency between panels
(c) the use of post-tensioned diaphragm walls, either precast or cast in situ
(d ) the use of precast concrete diaphragm walls
(e) the use of reverse circulation excavation equipment such as the
Hydrofraise and Trenchcutter rigs
( f ) the development of excavation equipment for work on congested sites
(g) improved design of mechanical grabs
(h) the development and use of electronic monitoring in grabs and cutters
to improve panel excavation tolerances and overall quality control
standards
(i) the development of improved stearage of cutters and hydraulic grabs
Basement construction and design 379

Fig. 8.43. CWS joint fabrication for a 55 m deep panel joint

(j) the development of improved stop end design; use of CWS joints
(k) improved standards of slurry cleaning and quality control of slurry
during use.
Items (a) and (b) are reviewed in more detail below; the remainder are referred
to elsewhere in this chapter and in Chapters 4 and 9.
While the original use of bentonite slurry to support diaphragm wall
excavations has generally persisted for excavations made by grab, the larger
quantities of slurry required for circulation purposes with Hydrofraise and
Trenchcutter equipment have brought innovation to slurry design. Using
experience from the oil drilling industry, polymeric slurries and mixes of poly-
meric and bentonite slurries have been used successfully on larger diaphragm
wall jobs where the economies of scale have proved beneficial.
Polymeric slurry behaves as a pseudo-plastic fluid and, unlike bentonite
slurry, acts in trench support without forming a filter cake. Within the
polymeric slurry a molecular lattice structure, which is different from the
thixotropic gel structure of a natural clay slurry, is built up after mixing.
380 Deep excavations

The polymeric structure is more efficient in the suspension and transportation


of soil particles and leads to reduced energy costs in pumping from excavation
to slurry station on the job site. As fluid loss from polymeric slurries is less
than that from bentonite slurries, the polymeric slurry can be used with advan-
tage in weak soils where an increase in soil moisture content would cause risk
of panel instability. Polymeric slurries cost considerably more than bentonite
slurry, but since, with care, the slurry is reusable many times, disposal costs
partly compensate for the high initial cost. The rheology of bentonite slurries
was discussed by Rogers44 , and the merits of polymeric slurries were reviewed
by Beresford et al.45
Icos began innovations in joint design to enable wall panels to be incorpo-
rated into rectilinear and arch plan shapes in the late 1960s. At Redcar, UK,
diaphragm wall panels were joined by tension connections to form cellular
structures to a new harbour wall. In plan, each of the cells measured 30 m
by 15 m. In plan, an inverted arch of diaphragm wall panels was restrained
by the tensile resistance of cross-walls, also in panels which, in turn, were
anchored to a rear arch of panels. The weight of the soil enclosed within
the cells was not used to restrain the wall in overturning because the panel
joints could not transmit vertical shear, only tension46 .
Developments by Bachy have led to patented joint forms that develop either
full flexural continuity from panel to panel (known as the Teba system) or
shear between adjacent panels. The first option, to develop a continuous
wall, is based on the use of hydraulic jacks cast within the panel concrete

Fig. 8.44. Development of continuity between diaphragm wall panels to achieve improved, structurally efficient wall plan
forms43
Basement construction and design 381

and near the panel joint. After concreting, these jacks are actuated and rods
are thrust horizontally from the cast concrete to provide continuity with the
reinforcement of the adjacent panel subsequent to excavation and concreting.
The method is rarely used and only finds application on special projects
because of the cost implications of the jack system.
The use of joints to transmit shear, however, has wider application and
more scope for development with methods of interactive soil-structure analy-
sis. The construction of walls from large T-shaped and H-shaped diaphragm
wall panels joined by shear connection enables the full flexural strength of
the multi-panel plan shape to be realized and, in addition, can utilize the
stabilizing effect of the weight of soil encapsulated between the legs of the
T-units or within the enclosed cellular areas formed by a series of H-panels.
The use of such plan shapes to form semi-gravity structures, shown in
Fig. 8.44, may only be justified in terms of economy where bracing or
anchoring of a basement wall is not possible and single-wall construction
has insufficient flexural strength to cope with high imposed cantilever
moments and shears. The wall area in T-shaped or cellular panel construction
is not cost-effective in terms of the linear wall unless special support con-
straints apply. Figure 8.45(a) shows development in joint construction to
transmit full continuity, shears and tension. The use of shear joints, however,
to form a diaphragm wall of varied castellated plan shape can be very cost-
effective as a deep unpropped cantilever for a very large basement providing
the plan area occupied by this wall can be accommodated in the site space. The
flexural efficiency of the castellated shape is shown in Fig. 8.45(b).
The use of semi-gravity structures may be justified where soil and wall
movement is critical. Such movements can be effectively minimized by the
use of a stiff wall structure and the gravity effect of the weight of soil retained
within the cells. One of the largest structural diaphragm wall contracts
completed to date utilized semi-gravity cellular plan form jointed diaphragm
wall panels for a large underground car park in Medinah, Saudi Arabia, in the

Fig. 8.45. Diaphragm wall built


to castellated plan shape: (a)
typical layout; (b) plot of section
modulus of castellated section
against overall depth of section
showing improvements to
flexural strength compared with
a straight wall
382 Deep excavations

Fig. 8.46. Medinah car park: views of unobstructed excavation supported by cellular diaphragm wall

early 1990s. The total area of structural wall exceeded 320 000 m2 to provide
the exterior walls to an excavation approximately 18 m deep from the ground
surface, 100 m wide and more than 1.5 km in total length. Views of the com-
pleted diaphragm walls are shown in Fig. 8.46. The site for the car park lies in
the centre of a bowl of washdown soils, mainly silty sands and clayey silts
from surrounding mountains. The depth to basaltic bedrock over the site
area varied from 23 to 55 m. The groundwater level had been monitored
over a period of two years prior to construction and showed some variation
with time. The average level during construction was 2 m above final excava-
tion level.
The client’s brief was to provide a basement construction with a design life
of 120 years, which could be excavated without the impediment of cross-
bracing, raking shores or temporary berms, and would give the minimum
of soil movement behind the walls and, in the long term, below the lowest
basement level in the car park. Precious historic structures were sited 20 m
from the excavated face of the basement wall and neither noise nor vibration
could be tolerated. A cellular diaphragm wall construction was adopted,
propped by two basement floors in the permanent condition but acting as a
cantilever in the two-year construction period. The cellular wall shown in
plan in Fig. 8.47 was adopted using vertical rock anchors into the basalt at
the rear of the wall to achieve stability during construction where bedrock
was shallow, but where bedrock exceeded 35 m in depth the cellular wall
was allowed to cantilever during the construction period.
The use of the cellular wall was justified on the basis of its stiffness to reduce
wall and soil movement and the opportunity it gave to undertake unimpeded
bulk excavation to the basement with staged handovers of large working
areas. Alternative designs using ground anchors, bored piles and T-shaped
diaphragm wall panels were considered but were precluded by the predicted
soil movement and the relative inefficiency of ground anchoring in the
Medinah silts and clays.
To construct the H-shaped cellular diaphragm wall units, three separate
panels were excavated and cast using precast concrete permanent stop ends
in the construction of the central web panel. Concrete was poured into the
central web excavation enclosed within a bag of plastic sheeting to avoid
leakage around the stop end which would reduce the effectiveness of the
shear and tension bond to the outer panels. The use of plastic sheeting as a
temporary measure proved effective in retaining all the concrete within the
web panel. The use of projecting reinforcement from one panel to the next
had been developed in France and Japan, but its application had been limited;
Basement construction and design 383

Fig. 8.47. Medinah car park,


cellular wall construction:
(a) plan of cellular wall
construction; (b) construction
procedure for single wall unit;
(c) wall unit construction
adjacent to completed unit

at Medinah, however, the precast joints were used successfully more than 1000
times with only five minor collapses. Construction of the central web panel
cage is shown in Figs. 8.48 and 8.49.
The rate of diaphragm wall production on this contract was impressive. The
soil conditions, which varied from medium strength silts and clays to stiff and
hard clayey silts, were conducive to grab excavation but were less economic-
ally excavated by cutter rigs until excavation depths increased below 20 m or
so. Three cutter rigs and up to five rope grabs were used on a 24-hour, 6-day
week basis. Concreted panel production averaged more than 4000 m2 per
week over much of the two-year wall construction period and reached
7000 m2 per week over several weeks at peak production.
The Medinah cellular wall was designed using a two-dimensional analysis
of the soil–wall structure taking into account non-linear elastic–plastic soil
conditions. Soil movement and stress levels in the surrounding soils were
predicted for the modelled excavation stages by finite element analysis,
taking into account the dissipation of negative pore-water pressure with
time for varying depths of rockhead and groundwater. The analysis methods
were those described by Jardine et al.47 The structure deformation results
from the two-dimensional analysis were then used with a three-dimensional
structural program to predict stress levels and design reinforcement within
the cellular wall. This work was incomplete because the analysis did not
384 Deep excavations

Fig. 8.48. Medinah car park:


cellular wall detail showing
reinforcement and permanent
stop ends to central web panel

include the deformations within the soil and the resulting changes in soil stress
levels caused by excavation of the diaphragm wall panel itself. It had been
realized prior to the Medinah design that accurate prediction of soil move-
ment and stress adjacent to a completed diaphragm wall basement excavation
had to include installation effects of the diaphragm wall panel during panel
excavation and concreting. The effects of concreting on in situ soil stress
had been shown48 in measurements which concluded that induced stress
caused by the pressure of wet concrete within the diaphragm panel, and the
resulting soil deformation, discouraged the use of low earth pressure coefficients
in wall design. More recently Lings et al.49 referred to placing temperatures
within the concrete as an important influence on wet concrete pressure.
A number of published results of soil movement caused by panel excavation
tend to show small horizontal soil movements near the panel, rapidly
reducing at short distances equal to the panel length, say, from the panel.
Measurements made at Medinah confirm this. (The exception to these
generally small movements were those observed during diaphragm wall con-
struction for the Hong Kong Metro. The decomposed granite residual soil
conditions in Hong Kong are quite different from those where the other
measurements of soil deformation adjacent to panel excavation were made.
It was concluded that the cause of the large soil movements in Hong Kong
Basement construction and design 385

Fig. 8.49. Medinah car park:


(a) fabrication of web panel with
permanent precast concrete
stop-ends to cellular wall;

was the presence of a soil with high swell potential, high permeability and a
high groundwater table.) Finite element modelling of installation effects by
Gunn et al.50 showed promise in predicting soil movement and stress during
panel excavation.
The Medinah soil–structure analysis to predict deformed wall shape and
soil movement was only partly successful because of the difficulty of modelling
the shear stiffness of a three-dimensional structure in plane strain and the
oversimplification of ignoring panel installation movement and stress in the
386 Deep excavations

Fig. 8.49. (b) Key plan of one


element; (c) isometric of web
panel

subsoil. The maximum predicted horizontal soil movement of 27.5 mm


after bulk excavation for the deepest cantilever walls was not reached, the
total observed maximum horizontal deformation for both panel and bulk
excavation for the cellular wall being less than 20 mm, allowing for dissipation
of pore pressure with time.

Composite walls and grouting techniques


Mention should be made of the use of diaphragm walls, either precast or in
situ concrete construction, as part of a soil retention and groundwater control
protection system incorporating non-structural slurry wall cut-offs and hori-
zontal grout plugs over the plan area of the basement. Figure 8.50 shows
examples of composite precast diaphragm walls incorporating temporary
king post walls cast in the upper section of one wall. The use of jet grouting
and intrusion grouting to form a horizontal grout plug to control the inflow
of groundwater to excavations was discussed in Chapter 2.

Water resistance of structural diaphragm walls


The difference between expectation and the actual performance of diaphragm
walls regarding water resistance has caused disappointment and dispute since
the earliest structural walls in the 1950s and 1960s. Then, structural walls
generally required the minimum of surface treatment to produce a dry face.
Many of the earliest diaphragm wall basements in Paris and London were
used for car parking and were either left without finishes or, at most, with
an applied sand–cement render. Where leaks occurred these were sealed by
application of surface chemical renderings such as Vandex or Xypex to
make a crystalline waterproof coating to the wall. These walls, generally
600 or 800 mm thick, were excavated by rope grab or hydraulic grab (rope
or Kelly mounted), and temporary tubular steel stop ends were used through-
out. Only in basement construction where storage of perishable goods was
Basement construction and design 387

Fig. 8.50. Composite


diaphragm wall
construction using precast
concrete wall sections:
(a) anchored Panosol wall
with slurry cut-off wall into
an impermeable stratum;
(b) anchored Panosol wall
with temporary Berlin wall
above and slurry cut-off
wall below;
(c) load-bearing anchored
Panosol wall with concrete
wall below, penetrating a
bearing stratum at depth;
(d) table of flexural strength
of typical Panosol panel
sections (courtesy of
Soletanche–Bachy)

planned, in shopping areas or office facilities, was a separate lining wall


constructed.
In London, the earliest diaphragm walls were built by Icos from 1961
onwards. By 1974 a sufficient number of contracts had been completed to
hold a keynote conference on diaphragm walling. At the conference, Sliwinski
and Fleming51 addressed the problem of water resistance:
It is therefore evident that the concrete used for diaphragm walls can for
practical purposes be considered impermeable. However, in practice the
permeability of a panel must also depend on the formation of cracks and
on any local defects in the concrete such as may result from segregation
but with normal concrete control and care, such occurrences should be
limited to isolated cases. The simple butt joint between panels cannot
be claimed to be proof against water entries but significant leakages are
rare due to the presence of soil impregnated with bentonite behind the
joint, and to the presence of some thin layer of contaminated bentonite
at the edges of the joint. Where leaks occur they can usually be ascribed
to differential deflexions between wall panels, and these differentials are
worse near corners. The whole matter of deflexion differentials (and
thereby risk of leakage) between wall panels depends on panel shape in
plan, wall height, the use of anchors, excavation procedures and other
factors. The present practice for dealing with damp joints is to allow
the leak to appear, for the differential wall deflexion, for the most part,
to take place, and then to inject cement or chemical grout into the soil
at the back of the joint, either vertically or horizontally through drillings
depending upon access. Alternatively, steel or other suitable plate,
bedded on epoxy-resin mortar, can be bolted to the concrete over the
internal face of the joint.
Sliwinski and Fleming continued to refer to panel deviations caused by
obstructions due to concrete passing beyond the tubular stop ends, but did
388 Deep excavations

not refer to this as a risk of bad water resistance of the wall below basement
excavation level. No reference was made to the occurrence of leaks at the junc-
tion of diaphragm wall and basement slab in basement or other construction.
The situation described by Sliwinski and Fleming provides a reasonable
summary of the opinion of UK specialist firms in the early 1970s. BS 200452
stated:
These joints are usually watertight but minor seepage through leaking
joints can be dealt with by grouting or may even be tolerable in certain
classes of structure.
Generally the attitude was optimistic, and risk of leakage was only con-
sidered after it had happened. The earliest model specification in the UK53
did not refer to water resistance and typical paragraphs in tender letters by
specialist contractors stated:
The diaphragm walls will be constructed so as to be substantially water-
tight on initial exposure (free from running leaks but not damp proof )
and we only accept responsibility for repairing leaks, within the exposed
height, caused by faulty workmanship and/or materials. It should be
noted that possible ingress of water into the excavation from below
formation level is not prevented by the diaphragm walls.
Overall, specialist firms were relatively optimistic about the likely occurrence
of leaks on jobs throughout the UK in a variety of soil conditions and for a
range of basement uses. The risk of overbreak, panel collapse, loss of bento-
nite, displaced box-outs, inadequate excavation rates, etc. were all critical
tender risk assessments for the specialist contractor, and wall waterproofness
and the cost of associated remedial works were not considered as important as
they are today. In the UK in the 1960s and 1970s, specialist firms were usually
awarded diaphragm wall contracts on the basis of design-and-construct after
technical discussion with a consulting engineer or architect. The contractual
risk for waterproofness (apart from damp patches) generally remained
firmly with the specialist contractor. By the end of the 1970s most major
consultancy firms were designing and specifying diaphragm wall schemes
themselves and the contracts were let on a construct-only basis. At this
stage, the overall use of the underground structure was clear to the designer,
who could incorporate measures such as non-load-bearing blockwork walls
and drainage channels to hide any persistent ingress of groundwater.
The use of basement lining walls, cut-and-cover works and underpasses has
continued in the UK, France and Germany. In some instances (Lyon Metro
1981, Eastbourne Pumping Station 1993) an in situ reinforced concrete lining
has been specified by the engineer or owner to be capable of withstanding the
full groundwater pressure acting on the wall.
As referred to previously, the guidance for water resisting design of
basements is BS 810254 and CIRIA report 13911 . The use of diaphragm
walls in basements can comply with the water-resisting methods defined in
documents in terms of method B structural integral protection and method
C drained cavity construction. To address method A (an external waterproof
membrane) specialist firms have sought over a period of some years to develop
a diaphragm wall system to include an outer plastic liner to completely
encapsulate the diaphragm wall. The method has had very little application
on site and doubts remain regarding its efficiency.
The requirements of BS 8102 to comply with type C (drained protection)
have become standard design principles for basement diaphragm wall work
in the UK. Most basements are designed with a drained cavity construction
and only those used for car parking have the option of exposed unclad
Basement construction and design 389

diaphragm wall surfaces. Building owners are not prepared to allow unlined
diaphragm walls (say for a basement for storage use) where there is reasonable
chance of a change of use during the life of the basement. Current practice is
therefore to make a drained cavity with a permanent pump provided at a
sump at the lowest level. To make this provision, the design volume of the
basement is reduced by the volume of the drainage cavity, the volume of
the blockwork lining and the volume occupied by the verticality tolerance
of the diaphragm walls (say for grabs 1:80 or 1:100), irrespective of whether
this tolerance is used by the wall or not. It is widely acknowledged that this
solution is uneconomic and often leads to a reduction in car parking spaces
in basements where lining is used.
The use of lining walls does not automatically produce protected construc-
tion because faulty drainage cavities and drainage between basement floors
often lead, over time, to damp blemishes on the exposed face of the lining
because of leaks in the hidden diaphragm wall.
The current situation in the UK is this: the specialist contractor generally
contracts to leave the exposed diaphragm wall free from running leaks (but
not damp patches) and probably has half of the full retention money held
against this for, say, twelve months from the end of the main contractor’s
contract. When leaks arise during this maintenance period the specialist con-
tractor seals them by grouting and trusts that a final inspection at the end of
the maintenance period will be the end of leakage responsibility.
The use of non-load-bearing lining walls in basements is similar in both
Germany and the USA, although a slightly more optimistic view regarding
the water-resisting efficiency of diaphragm walls may remain in France. The
water resistance of structural diaphragm walls has been reviewed in some
detail55 .
In the Author’s experience, basement wall leakages occur at any of five
locations:
. in the panel itself
. at vertical panel joints
. at horizontal bottom slab/wall joints
. at the top horizontal joint between panel and capping beam
. below formation level.
Each is now discussed in more detail.
(a) The panel. Leaks and damp areas in the panel are caused by soil or
slurry inclusions, random cracking perhaps due to shrinkage, or poor
quality concrete. Xanthakos56 concluded:
It is not difficult to produce walls with permeability of the order of
1010 cm/s. Suppose a wall is 60 cm thick, retains a hydraulic head
of 10 m and has a porosity of 15%; if we take into account a
suction pressure of 1 atmosphere to assist water flow, the quantity
of water percolating through the wall is close to 0.3 litre for
100 m2 of wall surface over a period of 24 hours.
So if this estimate is correct and the concrete is sound and homo-
geneous, only very limited dampness should occur on the exposed
surface due to concrete permeability.
If soil or slurry inclusions penetrate the full thickness of the wall in
water-bearing ground their removal can be difficult and expensive.
This matter is a particular risk in water-bearing silts of low strength.
The density of reinforcement, the depth of wall, the size of box-outs
and the thickness of the wall, concreting continuity and concrete
fluidity will all influence the risk of slurry inclusions. The risk is
390 Deep excavations

increased by any obstruction to flow or tremied concrete and any dis-


ruption, time-wise, to the flow of fluid concrete. These risks are evident
in thin, highly reinforced walls with box-outs. Entrapment of heavy
slurry below congested, large-diameter reinforcement at slab/wall
junctions are particular risk areas for slurry-contaminated concrete
and leakage. Random cracking within panels seldom appears, and
cracks are not necessarily continuous nor to full wall depth, but where
they occur the cracks can cause running leaks. There is a possibility that
the size of vertical reinforcing bars may influence shrinkage cracking in
relatively thin walls.
There is risk of soil inclusion in weak silts and silty clays and in highly
fissured over-consolidated clays, and panel length should be minimized
in such soil conditions. Overall, poor water resistance resulting from soil
or slurry inclusion may prove to be a high-cost risk because removal of
the inclusion throughout the whole wall thickness may be essential for
strength, durability and waterproofness requirements. This operation
may prove to be costly both in remedial expense and contract delays.

Fig. 8.51. Leakage path of


groundwater between water
bars at basement slab level,
diaphragm wall construction
Basement construction and design 391

(b) Vertical panel joints. This is the area at greatest risk of leakage. The
most likely causes of leakage are panel movement caused by bulk exca-
vation or application of vertical load, or pre-stress of ground anchors,
honeycombed concrete and poor concrete near the stop end due to
inclusions. Leakages can also occur through slurry contaminated with
calcium from concrete cut from the end of the adjacent panels, and
through inclusions in joints where precast concrete stop ends are
inadequately cleaned before concreting. The use of CWS joints to
reduce the risk of leakage at vertical joints by the inclusion of one or
more water bars has been shown to be effective. Nevertheless some
leakage may occur due to the passage of water between the water
bars as shown in Fig. 8.51.
(c) Horizontal bottom slab/wall joints. The next highest risk of leakage is the
horizontal joint at the basement slab. Although such leaks are possibly
a split responsibility between wall and slab constructors, precautions to
avoid leakage are necessary in the wall design. The length of the leakage
path from the underside to the top of the basement slab will be
increased if the underside of the slab is haunched to an increased thick-
ness adjacent to the junction with the wall. Special provision can be
made by securing a horizontal L-section flexible water bar to the vertical
wall surface within the slab box-out recess and connecting this to the
slab water bar system. This precaution is not successful, however,
unless the L-section water bar is connected to a water bar within the
vertical panel joint. A continuous water bar system between wall and
slab, in all horizontal and vertical joints, is expensive and success is
not guaranteed. The provision of a continuous water bar system is
more easily achieved in precast diaphragm wall construction than
conventional in situ walls. The risk of heavy slurry entrapment referred
to previously should be noted.
Where a groundwater head exists below the basement floor/dia-
phragm wall joint, it is essential that design provision is made to
resist penetration. The practice of pouring slab concrete into a recess
in the wall to receive the floor slab is simply not sufficient to ensure
water resistance even if the vertical concrete floor slab surface within
the recess is correctly prepared.
(d ) Top horizontal joint between panel and capping beam. It is not unusual to
find leakages in the horizontal joint at the top of a diaphragm wall with
a capping beam or other in situ reinforced concrete where a high
groundwater occurs or where rain water is allowed to pond in porous
backfill to the capping beam/guide wall excavations. Such leakages
can occur even when the top of the diaphragm wall is adequately cut
down to remove porous concrete and the surface is correctly prepared.
Although such work is essential, the provision of a Hydrotite water bar
strip in the horizontal joint on the earth face would overcome the risk of
water leakage.
(e) Below formation level. This risk is often ignored, although the financial
consequences of wall leakage or even loss of ground where ‘blow’
symptoms occur are likely to be very significant. The standard tender
clause used by some diaphragm wall specialists in the past has stated:
We only accept responsibility for repairing leaks, within the
exposed height, caused by faulty workmanship and/or materials.
The implication must be that the specialist contractor would not be
responsible for faulty workmanship and/or materials below formation level
in the area of highest risk. The clause was rarely queried by main contractors
392 Deep excavations

or their clients and presumably relieved the specialist contractor of consider-


able risk.
The obvious preventative measures for leakages all refer to standards of
workmanship and quality control during diaphragm wall construction. The
importance of design decisions, however, regarding wall thickness, permanent
stop end construction and provision of water bars in panel joints, should not
be ignored when risk of leakage below formation level is assessed.
Up to the 1990s the general level of acceptance of wall water resistance
in the UK was based on the criterion that damp patches on the exposed
wall surface would be accepted but running leaks would not, and grout or
surface treatment would be accepted as a remedy. Documents such as the
DIN standard 412642 , the British Department of Transport’s Specification
for Highway Works57 , and the European Code Execution of Special
Geotechnical Works: Diaphragm Walls58 , failed to make any reference to
acceptable standards of water resistance.
The ICE Specification for Piling and Embedded Walls59 , a widely used
standard specification together with a Particular Specification written for
the individual contract, makes reference to water resistance, defined as
‘water retention’ but fails to define unacceptable leakage volumes or areas
of dampness. The wording is as follows:
The Contractor shall be responsible for the repair of any joint, defect or
panel where on exposure of the wall visible running water leaks are found
which would result in leakage per individual square metre in excess of
that stated in the Particular Specifications. Any leak which results in a
flow emanating from the surface of the retaining wall shall be sealed.
In France DTU 14.160 sets more definitive limits:
The wall shall comply with the watertightness criteria of a lining wall to a
relatively waterproof structure.
The watertightness of the wall will be such that the flow of seepage and
leaks is limited to
For the whole of the outside walls in their entirety:
0.5 litre/m2 /day on yearly average
1.0 litre/m2 /day on weekly average
For all areas of 10 m2 of wall:
2.0 litre/m2 /day on weekly average.
These flows take account of the seepage flows at the joint of the raft to the
wall. The limited exactness of the ICE Standard Specification is replaced by
precise legal definitions of leakage standards in the French specification,
although some difficulty may be experienced in measuring such quantities
and applying them accurately.
A useful clause in use in the Middle East appears to strike a realistic
specification:
The diaphragm wall shall be watertight and substantially dry. Remedial
measures shall be carried out as directed by the Engineer in areas that do
not comply with this requirement. A panel will be considered acceptable
if within any area 1 m square the total damp surface does not exceed 10%
of that area. The Diaphragm Wall Contractor shall be responsible for the
repair of any joint, where, on full exposure of the wall, leaks are found.
Running leaks will not be accepted at any location.
It is the Author’s opinion that specification clauses for water resistance of
diaphragm wall works in general use are inadequate because they do not
Basement construction and design 393

specify acceptance to adequate standards, they do not specify the method of


repair and they do not approach the subject of waterproofness with the
considerable importance and detail that it deserves. Several matters are
ignored in the latest specifications. As an example, the building owner or
tenant’s interests are not served by the lack of consideration for leakages
occurring after the contractual maintenance period by, say, rising ground-
water, the failure of repairs to previous leaks, or movements between panels
or basement floor joints due to long-term soil movement such as heave.
Current specifications do not relieve the specialist contractor for leakage
due to panel movements caused by application of superstructure loading or
anchor stressing, matters that are frequently outside his control. The water
bar system in the diaphragm wall should be specified by the designer. The
wall system should be connected efficiently to the water bar system in the
base slab, and should also be specified by the basement designer.
Overall, early optimism among specialist contractors in Italy, France and
the UK has now been replaced by a realism that acknowledges that concrete
tremied into a slurry in a series of panels will not automatically produce a dry
basement. In Europe, internal lining walls are frequently used to avoid the effect
of the groundwater ingress. Unfortunately, these internal walls cover up a
continuing risk to the building owner and mask the occurrence of further
leaks or deterioration in the repairs to the original leaks.

Overall stability: Where hydrostatic groundwater pressures, during construction and within the
design for uplift design life of the structure, are at a higher elevation than the underside of the
lowest basement floor level, it is necessary to examine the overall stability of
the basement. Failure due to the lack of buoyant self-weight of the basement,
and insufficiency of frictional forces to avoid upward displacement of the
basement substructure, are fortunately infrequent, but not unknown. Vertical
anchoring of the basement raft to rock strata or strong soils below the base-
ment becomes necessary where hydrostatic uplift is severe and such strata
exist at economical depth. Drainage of a granular blanket or porous no-
fines concrete with permanent pumping may prove necessary where anchoring
is not feasible. The rise of groundwater within the design life of the structure
should be carefully assessed after consideration of the dead weight of the
structure at successive stages of superstructure construction, including final
completion; a factor of safety of the order of 1.4 should be obtained for the
sum of downward dead weight, total vertical downward anchoring force
and frictional resistance to the basement walls compared with upward
hydrostatic force on the basement underside. Where most of the downward
force consists of dead weight a modified criterion may be used:
Dead weight Friction
Upward hydrostatic force  þ
1:1 3:0
Where diaphragm walls are used for basement construction, the frictional
resistance in cohesionless material or the wall adhesion due to clay strata,
should be calculated in the same manner as frictional resistance to bored
piles in similar soils, restricting the diaphragm wall surface used in the
calculation to the inner and outer surface of the wall below formation level.
Further discussion on the overall stability of underground structure is
included in Chapter 9.

Construction The relative costs of secant pile and diaphragm walls given by Sherwood
economics et al.,61 and quoted in Chapter 4, serve only as a comparison between wall
394 Deep excavations

costs and do not consider the relative costs of alternative propping systems or
the costs of linings to the inside face of the wall. To make a logical choice of
wall, propping and lining system on cost grounds, a detailed comparison of
the cost of systems should be made for each job site.
A broad cost comparison is reproduced in Fig. 8.52. This comparison
includes completed wall construction for two- and three-storey basements

Fig. 8.52. Cost comparison


of basement walls for
temporary and permanent
works, two- and three-storey
deep cross-sections: (a), (b)
for soil profiles shown at
boreholes A and B for the
constructions described in
(c)–(f). A base index of 100
has been used for the lowest
cost construction shown in
(c).
Basement construction and design 395

in two particular soil and groundwater conditions. It would be unwise to draw


conclusions from this small sample but broadly it showed that, in the UK,
contiguous bored pile walls with an internal lining are comparatively inexpen-
sive in good piling ground; secant piling is an economic choice in less condu-
cive soil conditions and for deeper basements; where sheet piling is left in
place, an expensive wall results; anchored diaphragm walling is competitive
in deeper basements in good soil conditions.

References 1. Lambe T.W. Proc. Conf. Lateral Stresses and Earth Retaining Structures, ASCE,
New York, 1970, 149–218.
2. Wakeling T.R.M. Discussion, deep excavations. Proc. 6th Euro. Conf. S.M.F.E.,
Vienna, 1976, Vol. 2.2, 29–30. Austrian National Committee, Vienna, 1976.
3. Skempton A.W. and La Rochelle P. The Bradwell slip, a short term failure in
London clay. Ge´otechnique, 1965, 15, Sept., 221–242.
4. Gässler G. and Gudehus G. Soil nailing, some aspects of a new technique. Proc.
10th Int. Conf. S.M.F.E., Stockholm, 1981, Vol. 3, 665–67. Balkema, Rotterdam,
1981.
5. Banyai M. Stabilization of earth walls by soil nailing. Proc. 6th Conf. SMFE,
Budapest, 1984, 459–466. Akadémiai Kiado, Budapest, 1984.
6. Project National Clouterre, recommendations 1991. Presses de l’Ecole Nationale
des Ports et Chaussées, Paris, 1992.
7. Barley A. Soil nailing case histories and developments. Proc. Conf. Retaining
Structures. Institution of Civil Engineers, London, 1992.
8. Schnabel H. Sloped sheeting. ASCE J. Civ. Engng, 1971, 41, No. 2, 48–50.
9. Hanna T.H. Anchored inclined walls – a study of behaviour. Ground Engng,
1973, 6, 24–33.
10. Potts D.M. et al. Use of soil berms for temporary support of retaining walls.
Proc. Conf. on Retaining Structures. Institution of Civil Engineers, London,
1992, 440–447.
11. CIRIA. Water resisting basement construction. CIRIA, London, 1995, Report 139.
12. Report on basement and cut and cover construction. Draft. Institute of Structural
Engineers, London, 2002.
13. BS 8102. Protection of structures against water from the ground. British Standards
Institute, London, 1990.
14. Somerville G. The design life of concrete structures. Struct. Eng., 64A(2), 1986,
60–71.
15. Boikan A. Avoiding pitfalls and risk factors in below ground waterproofing.
Struct. Engr., 80, 2002, 5 June, No. 11, 16–18.
16. Zinn W.V. Economical design of deep basements. Civ. Engng Public Works Rev.,
1968, 63, Mar., 275–280.
17. Measor E. and Williams G. Features in the design and construction of the Shell
Centre. Proc. Instn Civ. Engrs, 1962, 21, Mar., 475–502.
18. Fenoux G.Y. Le réalisation fouilles en site urbain. Travaux, Parts 437 and 438,
1971, Aug.-Sept., 18–37.
19. Burland, J.B. and Hancock R.J. Underground car park at the House of
Commons. Struct. Engr, 1977, 55, Feb. 87–100.
20. Burland J. et al. Movements around excavations in London clay. Proc. 7th Euro.
Conf. S.M.F.E., Brighton, UK, 1979, Vol. 1, 13–29. British Geotechnical Society,
London, 1979.
21. Marchand S.P. A deep basement in Aldersgate Street, London, part 1: contrac-
tor’s design and planning. Proc. Instn Civ. Engrs, 1993, 93, Feb., 19–26.
22. Marchand S.P. A deep basement in Aldersgate Street, London, part 2: construc-
tion. Proc. Instn Civ. Engrs, 1993, 97, May, 67–76.
23. Slade R.E., Darling A. and Sharratt M. Redevelopment of Knightsbridge Crown
Court for Harrods. Struct. Engr, 80, 2002, 5 June, 21–27.
24. Fernie R. Movement and deep basement provision at Knightsbridge Crown
Court, Harrods, London. Conf. Response of Buildings to Excavation-induced
Ground Movements, July 2001. CIRIA, 2002.
396 Deep excavations

25. Kenwright J., Dickson R.A. and Fernie R. Structural movement and ground set-
tlement control for a deep excavation within a historic building. Deep Foundations
Institute Conference, New York, 2000. DFI, Englewood Cliffs, New Jersey, 2000.
26. Boscardin M.D. and Cording E.G. Building response to excavation induced
settlement. ASCE J. Geotech. Eng., 1986, 115, No. 1, 1–21.
27. Fletcher M.S. et al. The ‘down’ of top down. Civ. Engng, 1988, 58, Mar., 58–61.
28. Lui J.Y.H. and Yau P.K.F. The performance of the deep basement for the
Dragon Centre. Proc. of Seminar on Instrumentation in Geotechnical Engineering,
Hong Kong Institution of Engineers, 183–201. Hong Kong, 1995.
29. Mitchell A., Izumi C., Bell B. and Brunton S. Semi top-down construction
method for Singapore MRT, NEL. Proc. Int. Conf. on Tunnels and Underground
Structures, Singapore, 2000. Balkema, Rotterdam, 2000.
30. Cole K.N. and Burland J.B. Observations of retaining wall movements asso-
ciated with a large excavation. Proc. of 5th Euro. Conf. S.M.F.E., Madrid, 1972.
31. Katzenbach R. and Quick H. A new concept for the excavation of deep building
pits in inner urban areas combining top/down method and piled raft foundation.
Proc. 7th Int. Conf. and Exhibition on Piling and Deep Foundations, Vienna, 1998,
15.17.1–15.17.3. DFI, Englewood Cliffs, New Jersey, 1998.
32. Mair R.J. Developments in geotechnical engineering research: application to
tunnels and foundations. Proc. Instn Civ. Engrs, 1993, 93, Feb., 27–41.
33. Yang D.S. Deep mixing. Proceedings of the Geo-Institute Conference, ASCE,
Logan, Utah, 1997, 130–150. ASCE, New York, 1997.
34. Porbaha A. et al. State of the art in construction aspects of deep mixing technol-
ogy. Ground Improvement 5, No. 3, 2001, 123–140.
35. Taki O. and Yang D.S. Soil-cement mixed technique, Geotechnical Engineering
Congress, ASCE New York. Geotechnical Special Publication 27, Vol. 1, 298–
309. ASCE, New York.
36. Bruce D.A. et al. Deep mixing: QA/QC and verification methods. Grouting–Soil
Improvement Geosystems including Reinforcement. Editor, Hans Rathmeyer.
11–22, Finnish Geotechnical Society, Helsinki, 2000.
37. BS 8081. Code of practice for ground anchorages. British Standards Institution,
London, 1989.
38. Littlejohn G.S. and Bruce D.A. Rock anchors: state-of-the-art. Foundation
Publications, Brentwood, 1977.
39. Barley A. Ten thousand anchorages in rock. Ground Engng, 1988, 21: No. 6,
Sept., 20–29; No. 7, Oct., 24–35; No. 8, Nov. 35–39.
40. Diaphragm walls. Icos, Milan, 1968.
41. Guillaud M. and Hamelin J.P. Innovations in diaphragm wall construction plant.
Proc. D.F.I. Conf., Nice, 2002, 3–8. Presses de l’Ecole Nationale des Ponts et
Chaussées, Paris, 2002.
42. DIN Standard 4126. Cast in-situ concrete diaphragm walls. Deutsches Institut für
Normung, Berlin, 2002.
43. Puller M.J. and Puller D.J. Developments in structural slurry walls. Proc. Conf.
Retaining Structures, Institution of Civil Engineers, London, 1992, 373–384.
44. Rogers W.F. Composition and properties of oil wall drilling fluids. Gulf Publishing,
Houston, 1967.
45. Beresford J.J. et al. Merits of polymeric fluids as support slurries. Proc. Conf.
Piling and Deep Foundations, London, 1989. DFI, Englewood Cliffs, New
Jersey, 1989.
46. Fisher F.A. Diaphragm wall projects. Proc. Conf. Diaphragm Walls and
Anchorages, Institution of Civil Engineers, London, 1971, 11–18.
47. Jardine R.J. et al. Studies of the influence of non-linear stress-strain characteris-
tics in soil-structure interaction. Ge´otechnique, 1986, 36, Sept. 377–396.
48. Reynaud P. and Riviere P. Mesure des pressions developpees dans une paroi
moulee en cours de betonnage. Bull. Liaison Lab., Ponts et Chausée, Paris,
1981, No. 113, 135–138.
49. Lings M.L. et al. The lateral pressure of wet diaphragm wall panels cast over
bentonite. Proc. Instn Civ. Engrs, 1994, 107, 163–172.
50. Gunn, M.J. et al. Finite element modelling of installation effects. Proc. Conf.
Retaining Structures, Institution of Civil Engineers, London, 1992, 46–55.
Basement construction and design 397

51. Sliwinski Z. and Fleming W.G. Practical consideration affecting the construction
of diaphragm walls. Proc. Conf. Diaphragm Walls and Anchorages, Institution of
Civil Engineers, London, 1975, 1–10.
52. BS 2004. Code of practice for foundations. British Standards Institution, London,
1972.
53. Specification for cast in place concrete diaphragm walling. Federation of Piling
Specialists, London, 1985.
54. BS 8102. Code of practice for the protection of structures against water from the
ground. British Standards Institution, London, 1990.
55. Puller M.J. Waterproofness of structural diaphragm walls. Proc. Instn Civ. Engrs,
Geotech. Engng, 1994, 107, Jan., 47–57.
56. Xanthakos P. Slurry walls. McGraw-Hill, New York, 1979.
57. Specification for highway works. Department of Transport, London, 1991.
58. EN 1538. Execution of special geotechnical works: diaphragm walls. British
Standards Institution, London, 1996.
59. Specification for piling and embedded retaining walls. Institution of Civil
Engineers, London, 1996.
60. Centre Scientifique et Technique de Batiment. DTU No. 14.1. Travaux de
cuvelage. CSTB, Paris, 1987.
61. Sherwood D.E. et al. Recent developments in secant bored pile wall construction.
Proc. Piling and Deep Foundations Conf., London, 1989, 211–219. DFI,
Englewood Cliffs, New Jersey, 1989.
62. CIRIA. Embedded retaining walls: guidance for economic design. Gaba A.R. et al.
CIRIA, London, 2002.

Bibliography BS EN 12715: Grouting. British Standards Institution, London, 2000.


BS EN 12716: Jet grouting. British Standards Institution, London, 2001.
BS EN 12063: Sheet piling. British Standards Institution, London, 1999.
Pr EN 14475: Reinforcement of fills. CEN, 2002.
BS EN 1538: Execution of special geotechnical work: Diaphragm walls. British
Standards Institution, London, 2000.
BS EN 1537: Execution of special geotechnical work: Ground anchors. British
Standards Institution, London, 1999.
BS EN 1536: Execution of special geotechnical work: Bored piles. British Standards
Institution, London, 1999.
Draft EN 14679: Execution of special geotechnical work: Deep mixing. British
Standards Institution, London, 2003.
Davies R. and Henkel D. Geotechnical problems associated with construction of
Chater Station. Proc. Conf. Mass Transportation in Asia, Hong Kong, 1980, 1–31.
Ikuta Y. et al. Application of the observational method to a deep basement excavated
using the top-down method. Ge´otechnique, 1994, 44, Dec., 655–664.
Ramaswarmy S.D. Soil anchored tieback system for supporting deep excavations.
J. Instn Engrs Singapore, 1975, 14, Dec., 10–33.
Ramaswarmy S.D. and Aziz M.A. Some experiences with ground anchors for
substructure construction in Singapore. Proc. Conf. Geotech., Singapore, 1980,
170–180.
Ramaswarmy S.D. and Yong K.Y. Some case studies on problems of substructures of
high rise building. Proc. Conf. Construction Practices in Geotech. Engng, Surat,
India, 1982, 255–260. Oxford and IBH Publishing Co., New Delhi, 1982.
Smoltczyk U. Editor. Geotechnical Engineering Handbook Vol. 2. Procedures. 2.5.
Ground anchors. Ostermayer H. and Barley T. Ernst and Son, Berlin, 2003.
Washbourne J. The three dimensional stability analysis of diaphragm wall excava-
tions. Ground Engng, 1984, 17, May, 24–39.
Yandzio E. and Biddle A. Steel intensive basements. Steel Construction Institute.
Ascot 2001.
Cut-and-cover construction
9
Introduction As the name suggests, cut-and-cover construction consists of tunnel construc-
tion by deep excavation in trench, construction of the permanent tunnel struc-
ture, and subsequent backfill and reinstatement of the ground surface. The
method is economical in comparatively shallow tunnel works and is typically
applied in urban highway schemes and for urban metro stations and running
tunnel construction. This chapter therefore describes highway and metro
schemes and includes the construction of station boxes: sometimes more
exactly these stations could be classified as basements, but are included in
this chapter with other metro illustrations for completeness.
Historically the method was used as an alternative to bored tunnel con-
struction for underground railway and river-crossing highway schemes in
European cities in the second part of the nineteenth century, particularly in
London and Paris. Early photographs in 1903 of cut-and-cover works for
the Saint-Lazare station on line 3 of the metro in Paris are shown in Figs.
9.1 and 9.2. The photos, taken two weeks apart, show the rapid progress
which the method allowed. The station was excavated from below the roof
vaults of the station following roof construction. Prior to the Second World
War, metro construction in European cities such as Berlin, Paris and
London exploited cut-and-cover construction and furthered construction
techniques such as the king post method of soil support. Its use provided
an alternative to boring for underground tunnels within a range of depths,
typically 8 to 10 m. Excavation plant and craneage was largely steam driven,
and structural materials were usually timber or steel sheet piling.
The reconstruction of European cities in the 1950s, and the improvements
to public transport facilities with progressive urbanization in the 1960s and
1970s, allowed the introduction of improved methods of tunnelling, including
cut-and-cover techniques. In particular, improvements to excavation and
drilling equipment, the availability of high-quality steel sections and reinforce-
ment and the introduction of ready mixed concrete transformed construction
methods. A range of walling methods became available and alternative
methods of installation were developed. Reinforced concrete piles were now
installed by powerful rotary auger, steel sheet piling was driven by diesel
hammer, vibrator or by hydraulic equipment, and new methods of walling
such as diaphragm walling and methods of support such as ground and
rock anchoring were introduced by innovative contractors and specialists.
In the years following the 1970s many cities invested in new metro systems,
further exploiting cut-and-cover methods. In particular, the construction of
the Island Line and more recently West Rail in Hong Kong and the Singapore
Metro North East line have further used bottom-up, top-down and variant
systems and all the walling methods.
While the choice between tunnel or surface construction may be clearly
determined by the availability and value of land and the depth of the proposed
permanent construction, the choice between bored tunnel and cut-and-cover
construction methods may sometimes be less clearly defined. In other
Cut-and-cover construction

Fig. 9.1. Cut-and-cover works: St Lazare Metro Station, Paris, 190330


399
400
Deep excavations

Fig. 9.2. As Fig. 9.1. photographed two weeks later30


Cut-and-cover construction 401

instances, however, the prevalent groundwater conditions, availability of con-


struction site areas or the proximity of existing structures and their founda-
tions, may combine with the available horizontal and vertical alignments to
pre-determine the use of either bored tunnels or cut-and-cover construction.
Although the construction methods of cut-and-cover work may appear to
be more direct and free from the risks of bored tunnel construction, greater
risk of subsidence due to shallow works and the disruption of traffic and
services due to large-scale trench works may make cut-and-cover work less
attractive. In the early 1980s Megaw and Bartlett1 listed the disadvantages
of cut-and-cover in busy urban areas:
(a) Lengthy occupation of street sites with noise disturbances and disrup-
tion to access. This can be mitigated by mining excavation methods
below a roof slab constructed at an early stage on the permanent
tunnel walls. Roof slab construction allows speedy reinstatement of
highways and surface works. In special circumstances tunnelled head-
ings may be used to build the permanent walls with the minimum of
surface activity.
(b) In soft clays and silts, excavation in trench may be limited to maintain
stability and reduce heave. Short-length working will be necessary and
will increase construction and occupation time.
(c) Constraints on alignments by following existing streets may be undesir-
able, especially where small-radius curves are introduced into metro
construction. In some city areas, basements which encroach beyond
building lines may worsen the situation.
(d ) Progress and cost of cut-and-cover schemes can be badly affected by
diversion works to existing services, especially those inaccurately
recorded or uncharted and disclosed during trenchworks. These
works often require a break in the sequence of trench wall construction
to divert the service and then construct the trench wall across the
previous alignment of the service.
(e) Ground movement and subsidence of existing structures and services
have to be minimized. Methods to reduce subsoil heave, loss of
ground, and changes of groundwater level and flow entail cost and
construction time penalties. The use of pre-stressed ground anchors,
pre-jacking of struts, grouting works and groundwater recharge may
all be necessary, particularly where sensitive or old buildings are
nearby, and all have cost and time implications for cut-and-cover work.
The construction costs of cut-and-cover works increase significantly with
depth, but the effect of construction depth on the cost of bored tunnel
works is often much less. The choice of horizontal and vertical alignment
for large-scale works such as metro construction additionally involves com-
paring the capital cost of alternative alignments using varying lengths of
bored tunnel and cut-and-cover with the projected energy running costs of
trains on those alternative alignments.
Four methods are available for cut-and-cover wall construction:
(a) temporary support from braced or anchored steel sheet piling followed
by permanent reinforced concrete wall construction
(b) the soldier wall method of temporary support using soldier piles and
horizontal laggings, or sprayed concrete skin walls with bracing or
anchoring followed by permanent reinforced concrete wall construction
(c) temporary concrete walls in contiguous, secant reinforced concrete piles
or in situ diaphragm wall construction followed by permanent
reinforced concrete construction
402 Deep excavations

(d ) combined temporary and permanent wall construction from walls in


reinforced concrete secant piles, or cast in situ or precast diaphragm
wall construction.

Bottom-up, top-down or semi-top-down construction


Whilst bottom-up construction was used with timbered or sheet piled earth
support for cut-and-cover works until the 1960s and 1970s the use of top-
down for building basement construction in those times led to its use in
European cities for metro construction, particularly by Soletanche in Paris
and Icos in Milan. Every variant of walling method and geotechnical process
has been used since then, especially in the Far East, to maximize site
development potential, and reduce cost, construction time and disturbance
to traffic and urban life.
An example of recent station box construction in Hong Kong has been
described by Cook and Paterson2 for Nam Cheong station on the West
Rail work for the Kowloon Canton Railway Corporation. The original
contract design used diaphragm walls for the new station, 350 m  80 m in
plan, with an average depth of concourse construction of 15 m below
ground level. The new station straddles the existing elevated West Kowloon
expressway and the airport expressline railway, as shown in cross-section in
Fig. 9.3. The ground conditions comprised approximately 25 m of hydraulic
sand overlying a variable thickness of alluvial silt and clay underlain by a
weathered granite system. Groundwater varied between 2 m and 5 m below
ground level.
The original scheme comprised a diaphragm wall box using the top-down
method. In the event, the alternative scheme as built used less materials
at less cost to build an in situ structure within a sheet piled cofferdam, the
external walls being supported by plunge columns. Detailed vibration studies
were undertaken to assess vibration risk in driving the 28 m long sheet piles
and only in restricted lengths was this found to be excessive. At these lengths,
the sheet piles were pitched into slurry trenches. It is intended to recover all the
sheet piles after backfilling behind the permanent walls.
In order to dewater the site, the groundwater was lowered using deep wells.
A settlement prediction made prior to the works gave an estimated total
settlement below the future high-rise structure over the east box of 35 mm,
most of which would occur during construction.

Fig. 9.3. Nam Cheong Station, Hong Kong, cross-section31


Cut-and-cover construction 403

Fig. 9.4. Station construction


sequence, N.E. line,
Singapore25

The alternative design as built, also built top-down in order to achieve an


early handover of the track slab at ground level, minimized time and cost in
the use of barrettes with post-grouted shafts to improve load capacity in
lieu of bored piles and staged excavation and dewatering to control settle-
ments during construction.
A derivative of top-down construction named semi-top-down has been
applied in construction of the stations on the new N.E. line in Singapore.
Particularly because of the design requirement for civil defence purposes of
a 2 m thick station roof located 3 m below ground level, the roof itself was
used as a working platform. The access holes within this roof were large
and savings were made by using the smallest skeleton possible of permanent
works as temporary works and so the least weight of bracing and propping
to be supported by plunge columns, whilst maintaining watertightness and
avoiding settlement risk to nearby existing services and buildings. The
sequence of top-down and final bottom-up construction is shown in Fig. 9.4.

Choice of wall system The choice of walling method depends on geology, depth of excavation and
the presence of buildings or roads near the excavation. A review by Hulme
et al.3 of cut-and-cover walling methods for a large new transportation
404 Deep excavations

Table 9.1 Singapore Metro: construction methods for cut-and-cover stations3

Station Maximum Typical soil Retaining system used Special measures


depth of sequencea
excavation (m)

Braddell 14.9 1F, G4 0.6 and 0.8 m diaphragm walls


Toa Payoh 13.5 4F, 4K, G Sheet piles
Novena 14.7 1 .
2F, 14 2K, G Sheet piles
Newton 14.3 3F, 13K, G 0.8 m diaphragm walls Jet grouting
1
Orchard 21 2F, G Nailed slopes
Somerset 16.2 2F, 8K, G 0.6 m diaphragm walls or sheet piles
Dhoby Ghaut 16.1 1F, 10K, S Sheet piles
City Hall 22.3 3F, 2K, S3 King piles and shotcrete lagging
Tanjong Pagar 17.9 1
2F, S Slopes, anchors
Outram 13.9 2F, 3K, S 8 m deep sheet piles over king piles
and timber laggings
Tiong Bahru 14.1 1F, S King piles and shotcrete lagging
Bugis 18.3 1F, 34K, O 1.2/1.0 m diaphragm walls Lime piles
Lavender 16.5 3F, 20K, O 1.0 m diaphragm walls
Marina Bay 16.4 12F, 24K, O Composite H pile/sheet pile Underwater excavation
a
Key: F ¼ fill 
G ¼ granite
including weathered rock
S3 ¼ Jurong
K ¼ Kallang
S ¼ boulder bed
O ¼ old alluvium
Example: 3F, 13K, G ¼ 3 m of fill overlying 13 m of Kallang deposit overlying granite (in this case completely weathered granite)

system showed the choices for each station or section of running tunnel on the
Singapore MRT. Table 9.1 summarizes the walling methods used for the
underground stations on the system. A similar comparison of cut-and-cover
station walls on the initial Hong Kong MTR system was presented in the
1980s by McIntosh et al.4 (Table 9.2).

Sheet pile walls


The traditional use of sheet piles in temporary soil support for cut-and-cover
construction has been reduced by environmental pressures to avoid noise and
vibration due to pile driving in favour of the use of top-downwards techniques
which generally favour walling methods that use combined temporary and
permanent soil retention. Nevertheless, the use of hydraulic press equipment
and jetting to install sheet piles and the monitoring of noise and vibration in
less sensitive areas does allow increased sheet pile use. This has been shown
particularly in lengths of both the Singapore and Hong Kong metros where
excavation depths are limited to the order of 15 to 16 m and where soil con-
ditions allow economical pile driving. Appropriate applications include
river crossings, areas where groundwater is high and sites that are some
distance from existing structures and services. The lack of flexural stiffness
of sheet pile sections, which would normally require frequent propping or
the risk of large settlements can be corrected either by stiffening the sheeter
sections with soldier piles or tubes, as with the Combi wall, or the stiffening
of the retained soil by jet grouting or mix-in-place piles. These latter piles
can be drilled in a cellular plan form or as a series of buttresses behind the
sheet piles. The adequacy of space to withdraw sheeters from behind the
Table 9.2 Hong Kong Metro stations — adopted construction methods for cut-and-cover works4
Station Depth of Cover to Depth to Engineer’s assumed Temporary works Proximity to buildings Walls Constructor Special measures
excavation roof slab rock from method sequence
(m) (m) surface (m)
a
Choi Hung 20 0–3 Temporary Berlin wall Permanent walls One end only Hand-dug Top down Skeletal roof of cross-
with preboring or interlocking caissons beams with precast T
diaphragm walls beam infills
b
Diamond 22 3 Steel I sections king Permanent walls No Hand-dug caissons Top down Walls are to be
Hill piles and intermediate for steel piles and removable for future
sheet piles concrete jack arches widening
a
Wong Tai 24 max. 3.5–6.5 Diaphragm walls Permanent walls Medium height housing Diaphragm walls Top down Roof was clear spanning
Sin blocks during excavation with
concourse suspended
from it
Lok Fu 27 2 0–30 Bored tunnel Berlin wall of steel piles High-rise housing block In situ Bottom up Dewatering by ground
and concrete lagging — treatment and wells
ground anchors
b
Kowtown 18 2 Diaphragm walls Part permanent walls No Part diaphragm wall, Bottom up —
Tong — part Berlin type- part in situ
ground anchors
Shek Kip 18–24 1:5–6:5 0–30 Open cut in rock, sheet Berlin wall, part High-rise housing In situ Bottom up Short length of station
Mei piling with grouted strutted, part blocks and schools platforms in bored rock
anchors in soil ground anchors tunnel
Prince 28 2 16–30 Diaphragm walls Permanent walls High-rise commercial Benoto type secant Top down Extensive grouting was
Edward and residential piles and hand-dug used, plus dewatering
caissons and limited recharging
b
Argyle 25 3–5 Diaphragm walls Permanent walls High-rise commercial Benoto type secant Top down Columns extended to
and residential piles underlying rock and
vertically anchored;
some areas of slab also
anchored; grouting to
walls; use of recharge
wells
Waterloo 28 2 0–27 Part open cut, part Permanent walls High-rise commercial Benoto type secant Top down Underpinning to walls
diaphragm walls and residential piles to rock, then in
situ
Jordan 18–23 0–4:5 4–20 Diaphragm walls and PIP pile walls and 7 High-rise commercial In situ Bottom up Half of station anchored
rock anchors, in situ levels of steel strutting and residential to underlying rock
underpinning
Tsim Sha 17–21 3.5–7.5 9:5–13:5 Diaphragm walls and PIP pile walls and steel High-rise commercial In situ Bottom up Part of station anchored
Tsui rock anchors, in situ strutting and residential to underlying rock
underpinning
Admiralty 25 0–3 20 Diaphragm walls on Combination of open No Part diaphragm wall, Bottom up Part of station anchored
rock; rock anchors and cut, anchored sheet part in situ vertically to underlying
in situ underpinning piling and permanent rock: underpinning to
walls, also slurry diaphragm wall
trenches
Chater/ 28 3 33c Diaphragm walls Permanent walls with High-rise commercial Diaphragm walls Top down Special measures to
Pedder struts and hotels, low rise construct walls and
Cut-and-cover construction

historic buildings groundwater recharging


a b c
Not known. Not known, large boulders. But rock level not proven in some sections.
405
406 Deep excavations

constructed permanent structure may prove vital in the economic use of


sheeters even where good driving conditions exist.
To summarize, the disadvantages of using sheet piles in the continuous
walls of cut-and-cover works are as follows.
(a) Noise and vibration during installation: may be overcome by use of
Giken type presses for piles of medium depth.
(b) Support is provided only during construction and permanent works are
required for tunnel construction.
(c) Obstructions reduce driving efficiency and increase risk of damage due
to vibration.
(d ) Adequate allowance must be made for installation tolerances. The
initial piling line must make allowance for verticality tolerance to
ensure adequate width between sheet pile walls to accommodate the
permanent works.
(e) Ingress of groundwater through pile clutches and split clutches may
cause delay or even local failure. Use of sealed or welded clutches
may be feasible.
( f ) Future use of sheet piles may determine the cost-effectiveness of the
method. Extraction of sheet piles after completion of the permanent
works may prove difficult due to soil conditions and lack of working
space.
(g) Temporary bracing between sheet pile walls must be replaced as the
permanent structure is built with new struts between sheeters and per-
manent structure. The incomplete and complete permanent structure
must be designed to transfer soil and groundwater pressures in this way.
(h) Extension of sheeters for increased cut-off requires welding operations
with disruption to overall production.
Where subsoil conditions, environmental restrictions, excavation depths and
working space constraints are not severe, sheet piling is still economical for
cut-and-cover work. In the 1960s and 1970s, the method was the forerunner
of deep secant pile walls and diaphragm walls which now provide alternatives.
Reference to these earlier jobs shows some of the difficulties which were
experienced.

Historic use of sheet piles in cut-and-cover construction


The second Blackwall Tunnel, a crossing of the River Thames in London,
used sheet pile cofferdams for both north and south cut-and-cover
approaches5 . During construction of the north approach, artesian pressures
below the cofferdam in a fine dense sand were not sufficiently relieved by
pumping from deep wells and pore-water pressures caused spongy patches
to develop in the overlying London clay exposed at formation level. The
reduction in effective passive pressures supporting the sheet piles below
formation level appeared critical as overloading of the cofferdam bracing
increased. The formation was hastily reloaded, excavation works were tem-
porarily stopped and the pumps given time to relieve the artesian head.
Another feature on the north approach cut-and-cover was a short, 20 m
long in situ diaphragm walling built at the junction of the cut-and-cover
approach and deep tunnel section to avoid the risk of piling vibration loosen-
ing tunnel segments immediately adjacent to the junction.
Two tunnel crossings of the Thames at Dartford used alternative walling
systems for each of the cut-and-cover approaches. The first crossing in the
early 1960s used sheet piles driven into soft alluvial clays overlying sands
and gravels containing an artesian groundwater pressure. The strutting
system used to support the sheet piles, shown in Fig. 9.5, allowed secure
Cut-and-cover construction 407

(a)

Fig. 9.5. First Dartford Tunnel,


cut-and-cover construction:
(a) details of bracing frames;
(b) cross-section32 (b)
408 Deep excavations

frame fixing promptly after excavation to successive depths, preventing yield


of the piling and consequent increases in strut loading. The permanent cut-
and-cover was built within the trenchworks in tunnel rings, and the sheet
piles were left in.
The second Dartford Tunnel was built in 1972 and the walls were built to
resist both temporary and permanent soil and groundwater pressures using
in situ diaphragm walls.
Neither cut-and-cover sections to the Dartford crossings experienced
construction problems, unlike the north approach at Blackwall. There were
difficulties, however, with the stability of the formation at the portal of the
Clyde Tunnel, completed in the 1960s and built in a sheet piled cofferdam6;7 .
Portal construction within caissons was considered but because of the
proximity of nearby buildings and the existence of a boulder clay stratum
below formation level which could have affected caisson sinking, a sheet
piled cofferdam was preferred. In the event, the boulder clay stratum was
found at much greater depth than anticipated. Air bubbles rising within the
partly excavated cofferdam showed leakage of compressed air from the
tunnel workings and a lack of seal in the sheet piling to the cofferdam. As
excavation proceeded through the silt, cofferdam struts were overloaded

Fig. 9.6. Clyde Tunnel: plan and


vertical cross-section of
completed partial cofferdam7
Cut-and-cover construction 409

and the silt below formation level became unstable as pore water pressures
increased. A borehole put down within the cofferdam gushed water confirm-
ing that groundwater had access to the underside of the silt stratum. A well
sunk 3.6 m into the bedrock made some improvement but not enough to
allow excavation to continue over the whole base area of the cofferdam.
The cofferdam was partly flooded and an auxiliary second frame inserted
and pre-loaded. Erection of the third frame was completed in the dry, and
the fourth and final frame was built in short trenches where boulder clay
did not exist above bedrock, the silt being excavated and replaced by mass
concrete within sheet piled cells driven between the upper cofferdam frames.
Excavation was completed to formation level over the remainder of the coffer-
dam (Fig. 9.6).

Anchored or braced king post walls


Although vertical soldier piles or king posts with horizontal poling boards
spanning between them had been used most effectively in the sandy subsoils
of Berlin in the 1930s, it was the development of powerful mechanical
augers, anchoring methods and methods of spraying concrete which pro-
moted its post-war use. Unrestricted, wide, anchored excavations were now
possible, and metro schemes, particularly in Germany, adopted the method
for temporary soil support. The method is most economical where ground-
water is absent or can be reduced by dewatering. Figure 9.7 shows a typical
excavation below bracing with reduction of groundwater by pumping.

Fig. 9.7. Cut-and-cover construction using soldier pile walls: sequence of construction using cut-offs and dewatering system33
410 Deep excavations

Fig. 9.8. Composite Berlin wall


with pre-stressed precast
diaphragm wall, Fukuoka
Metro, Japan (courtesy of
Soletanche–Bachy)

The king post wall method is particularly useful as part of a composite


wall system, acting as a temporary soil support for shallow depths above
a temporary/permanent pile or diaphragm wall at greater depth. A typical
composite wall is shown in Fig. 9.8 and is referred to later in this chapter.

Contiguous bored pile walls


Contiguous reinforced concrete piles, installed by either CFA rig or power
auger with casing oscillator, are ideally used for cut-and-cover works of
moderate excavation depth, say 15 m or so, in cohesive soils with minimal
groundwater. Where ingress of heavy groundwater does occur locally through
sandy or gravel seams, jet grouting may be used at the rear of the piles to
reduce leakage.
An ideal application of the method was the cut-and-cover approach
section of the Mersey Kingsway Tunnel in Liverpool in the early 1970s8 .
The cut-and-cover structure was founded on sandstone while the 17.7 m
high walls supported boulder clay. To avoid the considerable thickness
of cantilevered walls or large overhead propping beam to span the
approach width of 26 to 31 m, a continuous arch roof structure was used
with backfill over it. The rise of this arch was designed in proportion to
the arch span and the depth of backfill so that a balance was obtained
between lateral ground pressure from the walls and the outward horizontal
arch thrust.
Excavation was initially made to arch springing level between temporary
anchored king post walls (Fig. 9.9). From this level, 2.5 m dia. contiguous
piles were augered into bedrock. A dumpling between the contiguous bored
piles could not be removed until the arch thrust had been developed from
the backfill load over the arch. Successive stages of dumpling excavation
and filling were carefully sequenced. Two hundred piles were installed at a
peak rate of four piles per day.
Cut-and-cover construction 411

Fig. 9.9. Vertical cross-section


of cut-and-cover structure,
Liverpool approach to Mersey
Kingsway Tunnel; contiguous
bored pile walls propped by
concrete arch8

Secant pile walls


Secant pile construction, alternate male and female piles interlocked to form
a hard–hard secant wall, provides an efficient and economical walling system
to moderate and greater depths in a wide range of soil and groundwater
conditions. A permanent wall is constructed to allow soil support during con-
struction. The principal advantages of the system are as follows.
(a) Permanent structural walls are constructed in one operation ahead of
excavation.
(b) The hard–hard secant pile walls are substantially watertight.
(c) Excavation methods using down-the-hole hammers, reverse circulation
drills, heavy-duty rotary augers/buckets, or temporary casings with
casing oscillators and hammer grabs are highly efficient in hard soil
and rock conditions. Excess heads of water or slurry within the tem-
porary pile casing can be used to overcome onerous groundwater
conditions. For cut-and-cover structures of modest depth higher pro-
duction rates can be obtained in less demanding soil conditions with
CFA rigs particularly for hard–firm and hard–soft secant pile walls.
(d ) Good verticality tolerances can be achieved with twin-walled temporary
casing, and casing oscillators. Tolerances of the order of 1 in 200 to 1 in
300 may be expected, depending on soil conditions. Little overbreak
may be expected and the pile finish is uniform.
(e) Pile installation is comparatively noise- and vibration-free although
some vibration is inevitable when penetrating through dense granular
or rock strata.
( f ) Loss of ground during excavation is generally small. In soft silts and
clays or where sand with a high piezometric head is penetrated, the tem-
porary casing affords continuous lateral support and the stability of the
base of the bore within the casing may be continuously retained by a
head of water or slurry within the casing to ground level.
(g) Vertical loading of secant walls is viable because of the reliability of
good soil density below the concreted base of the pile.
(h) Temporary gaps may conveniently be left in the secant pile wall to allow
service access. Piles are temporarily filled with sand after boring at these
locations and are concreted later.
The design and construction of six recent deep stations on the Copenhagen
Metro was described by Beadman and Bailey9 . The design, on the basis
412 Deep excavations

Fig. 9.10. Copenhagen Metro


Station construction:
cross-section9

of 100-year design life, was made in accordance with the requirements of


recently introduced Eurocodes. The six stations were all similar in cross-
section and plan shape. The main structure of each is a secant piled box
20 m deep, 20 m wide and 60 m long. The internal structure to the box is
shown in cross-section in Fig. 9.10. The support to the box walls is provided
by the station roof and floor with a waling beam at approximately mid-height.
The ground conditions consisted of varying depths of well-compacted made
ground and over-consolidated sandy clay or clayey sand with bands of water-
bearing sands and gravels overlying limestone. The stations all extended into
the limestone which was heavily fractured and hard. The permeable sands
and gravels contained a secondary aquifer but the main aquifer was within
the limestone; the groundwater level was typically 2 m below ground level.
Secant piles were selected in favour of sheet piles (which would not
penetrate the limestone) and diaphragm walls (insufficient space on sites for
bentonite plant and reinforcement storage). The secant pile solution that
was used consisted of hard male piles, 1180 mm dia. with soft female piles
750 mm dia. installed only as far as the intact limestone. CFA and cased
techniques were used. The gaps between the male piles in the limestone
were sealed with grout.
Most of the stations were built using top-down methods to minimize
temporary works and settlements outside the box. The construction sequence
was:
(a) construct the secant piles and the station roof
(b) excavate to the waling level and construct the waling beam and the
permanent props
(c) hang the inner edge of the waling beam from the roof
(d ) excavate to base level and construct the base slab
(e) complete the internal elements, bottom-upwards
( f ) remove those hangers that are only required temporarily.
Geotechnical design in accordance with Eurocode EC7 (as detailed in Chapter
4) required three ultimate limits to be considered:
Cut-and-cover construction 413

Table 9.3 Soil parameters and


partial factors used in Soil parameter SLS Case B Case C
construction of the Copenhagen
Metro Made ground 0 30/1.0 30/1.0 30/1.25
Glacial till 0 32/1.0 32/1.0 32/1.25
c0 25/1.0 25/1.0 25/1.6
Limestone 0 40/1.0 40/1.0 40/1.25
c0 50/1.0 50/1.0 50/1.6

. Case A: deals with flotation


. Case B, wall design: deals with the strength of the structural members.
Ultimate load factor of 1.35, permanent unfavourable actions (forces)
. Case C, wall design: deals with the geotechnical design: ultimate load
factor of 1.0, permanent unfavourable actions (forces).
Soil parameters and partial factors applied to them were as detailed in Table
9.3.
Beadman and Bailey comment that the specified maximum crack width of
0.2 mm (rather than a maximum of 0.3 mm) was most onerous and required
substantial increased reinforcement quantities.
Examples of inclined secant pile walls for cut-and-cover works in the late
1960s on the Munich Metro were reported by Weinbold and Kleinlein10 .
The method replaced underpinning in hand-dug pits where space between
existing buildings was limited, and in some cases allowed the retention of
the existing facades which would otherwise have been rebuilt. Figure 9.11
shows typical cross-sections of the inclined pile walls on metros in Munich
and Frankfurt, both bored with a 128 inclination. These piles were installed
by Benoto rigs and were continuously supported by temporary casing
within the inclined bore at all stages of excavation and concreting. The
design of these inclined walls was based on elastic analysis after site experi-
mental verification of the modulus of subgrade reaction. The wall was
therefore assumed to be loaded by earth pressure at rest, groundwater
pressure and superimposed loads due to buildings, and was to be supported
by elastic embedment within the soil and by ground anchors and strutting.
Soil conditions, loading and sheeting moments and deflections at successive
excavation stages are shown for a typical section in Fig. 9.12.
Hana and Dina11 carried out a series of model tests on anchored inclined
walls. They concluded the following.
(a) The design of an inclined wall supported by rows of pre-stressed
anchors should incorporate a rectangular earth pressure envelope.
(b) Walls which tend to undercut the retained soil, inclined away from the
excavation (the same direction of inclination as shown in Figs. 9.11 and
9.12), experienced much larger soil subsidence than walls inclined
towards the excavation, at all stages of excavation from ground level
to final formation. Generally, for walls inclined away from the excava-
tion the largest lateral movements were at the top of the wall, whereas
for walls inclined in the opposite direction the largest lateral movements
occurred at the base of the wall. There were also vertical movements,
particularly for walls inclined away from the excavation. The bearing
capacity of the wall base was stated to be very important.
(c) Individual anchor loads changed as construction progressed. In
general, for walls inclined towards the excavation, anchor loads
reduced from initial values, whereas for walls inclined away from the
excavation initial anchor loads increased.
414 Deep excavations

Fig. 9.11. Inclined secant pile


construction for metro
construction: (a) cross-section
with temporary roadway, pile
construction for Benoto rig,
Munich Metro; (b) inclined
walls, Frankfurt Metro
Fig. 9.12. Inclined secant wall construction: (a) section; (b) soil profile; (c) lateral pressure on wall due to foundation pressure; (d) earth pressure at stage 1 excavation
to 3.1 m below street; (e) end-of-construction earth pressure; (f) moment in secant wall for construction stages; (g) deformation during construction stages10
Cut-and-cover construction
415
416 Deep excavations

(d ) The mechanics of inclined wall behaviour are similar to those of the


vertical wall and are controlled by wall and anchor.
Schnabel12 argued that earth pressure was reduced on walls sloping towards
the excavation. Site measurements for sloping walls were compared with
calculated pressures for vertical walls, with proposed reduction factors.

Diaphragm walls
Historic use and development
The earliest structural diaphragm walls were built in Italy by Icos in the 1950s.
The method was soon used to facilitate cut-and-cover construction for metros
in major cities. In Milan, Icos walls were used in a method developed by the
firm and shown in Fig. 9.13. The sequence of construction was designed to
minimize the disturbance to highway and traffic by early reinstatement of
the carriageway above the permanent cut-and-cover roof as excavation and
invert construction proceeded beneath it. This method, now familiar as top-
downwards construction in both basement and cut-and-cover construction,
became the basis of metro construction by Icos in many cities worldwide
and, over time, by their competitors. Innovations were introduced by Icos13
and later by others in Milan, on number 1 and 2 lines, when structural steel
column elements were lowered into barrettes, sections of Icos wall, below
tunnel invert level, as plunge columns to be used as structural support for
reinforced concrete mezzanine floor and roofworks to the tunnel. Details of
this construction are shown in Fig. 9.14.

Fig. 9.13. Construction


sequence used by Icos in
original cut-and-cover works for
Milan Metro (courtesy of Icos)
Cut-and-cover construction 417

Fig. 9.14. Milan Metro:


cross-section of fabricated steel
columns cast into reinforced
concrete base using diaphragm
wall techniques for vertical
load-bearing units (courtesy of
Icos)

Icos also introduced castellated sections of diaphragm walling on metro


construction in Milan (Fig. 9.15). The section, of greater breadth than the
straight wall, provides considerably enhanced strength in bending. More
recently, panel joints which can transmit vertical shear and tension from
one panel to its neighbour have allowed this efficient plan shape to be fully
exploited with a continuous wall section.
In the UK, diaphragm walls were introduced in 1962 for use on cut-and-cover
construction for a road underpass at Hyde Park Corner in London14 . The
engineer had decided that driven sheet piling could not be used because of
installation noise and vibration (a hospital was located nearby), so contiguous
bored reinforced concrete piles were specified. Icos diaphragm walls were
introduced by the main contractor but neither the contiguous piles of the
original scheme nor the alternative of diaphragm walls were considered as
part of the permanent subsoil support. This diaphragm wall scheme was also
successfully used to underpin the existing hospital walls where the underpass
diaphragm wall was built less than 1 m from the main hospital walls and more
418 Deep excavations

Fig. 9.15. Cut-and-cover construction using castellated plan shape diaphragm walls on one side: (a) plan; (b) cross-section
(courtesy of Icos)

than 8 m below it. These measures later became standard practice for such
locations. The use of short panels, increased wall reinforcement, pre-loaded
struts and reduced open lengths of main excavation limited horizontal and
vertical soil movements and wall movements to less than 3 mm.
The reluctance of designers outside Europe to use diaphragm walling as
a means of combined temporary and permanent soil support persisted
in the 1960s and 1970s. The cut-and-cover for the Calcutta Metro15 used
diaphragm walls only to resist buoyancy under permanent load conditions.
A factor of safety of 1.5 was used against flotation with full soil cohesive
strength being allowed in calculating wall adhesion to the clay subsoil.
In the UK, diaphragm walls were similarly used to resist buoyancy but with
no assumed contribution to flexural strength of the rectangular cut-and-cover
box structure housed between the walls. This reluctance to use the flexural
strength of the diaphragm wall after construction was evident in the second
tunnel crossing of the Thames at Dartford in 1972. By this time, diaphragm
wall construction had gained wide acceptance in the UK and the walls at
Dartford were extended to depths in excess of 30 m to minimize the length
of the driven tunnel. Nevertheless, the flexural strength of the diaphragm
walls was ignored for the permanent works design.
The 800 mm and 1 m thick diaphragm walls at Dartford were excavated by
kelly-mounted hydraulic grabs through soft alluvial silty clays and dense
gravels into hard chalk. Five frames of bracing were necessary to reduce
flexural stresses in the wall as bulk excavation proceeded to the deepest
sections at the junction with the bored tunnel, almost 30 m from ground
Cut-and-cover construction 419

level. These high flexural stresses were correctly anticipated by the wall
designers who appreciated the relatively large wall movements that would
be necessary to mobilize relatively small passive resistance in the soft clays
at formation level and immediately below it. Following the innovation used
by the contractor for the first Dartford Tunnel, hammerheaded concrete
struts were used throughout to brace diaphragm walls in the cut-and-cover
length, thus avoiding the need for separate walings.
Shortly after the Dartford Tunnel cut-and-cover works had been
constructed by diaphragm walling, the station at Heathrow for the Piccadilly
Line extension was built in cut-and-cover box, the diaphragm walls acting as
both temporary and permanent soil-retaining walls. Jobling and Lyons16 said
that cut-and-cover construction was chosen in preference to bored tunnelling
for three reasons:
(a) the use of station tunnels with space between escalator access tunnels
would have used more plan area than the cut-and-cover box and left
insufficient space for further station development for surface railways
(b) it was considered very costly to provide foundations for proposed
building development over driven tunnels
(c) subsoil strata, flood plain gravels overlying London clay, favoured box
construction in diaphragm walling.
The box shown in Fig. 9.16 was typical metro station size at that time, 131.5 m
long and 22 m wide with a depth to formation level of 17 m to keep the tunnel
drive below the flood plain gravels and within the London clay. The 1 m thick
diaphragm walls, propped by three frames, were designed using earth
pressures based on a value of Ka ¼ 0:25 for the gravel and K0 ¼ 0:75 for
the clay and a design groundwater level of 2 m below ground level. The
base to the box, between 1.9 and 2.575 m thick, was designed as a beam on

Fig. 9.16. Heathrow Central station, plan and vertical section16


420 Deep excavations

an elastic foundation using a modulus of elasticity of 107 MN/m2 for London


clay, and a modulus of subgrade reaction of 9055 kN/m2 . The design factor of
safety against flotation of the box was 1.2 on completion, but a rubble drain
beneath the base raft adjacent to the diaphragm walls restricted groundwater
pressure on the raft during construction. The raft was not designed to resist
hydraulic forces until loaded by the main internal columns.
The walls were temporarily braced by the frames of Rendex No. 6 struts at
floor level and by 300 mm  300 mm timber struts at platform level. A maxi-
mum deflection of only 5 mm at the top of the diaphragm walls had been
specified and pre-loading of the top frame and successive shimming of the
second frame was necessary to achieve this. Strut loads were monitored;
this showed that middle frame loads exceeded design values prior to the
lower frame being placed. This lower frame consisted of timber struts
spanning from the central raft section to the walls which were placed as a
soil berm 4 m  3 m in section progressively excavated from the face of the
walls. Excavation of these berms was uneconomical, being hindered by starter
steel from the raft and the strutting itself.
Metro construction in the Far East has utilized diaphragm walling for both
station and running tunnel construction in soil conditions generally more
demanding than those in Europe. The original contract for the construction
of the Hong Kong MTR in the late 1970s4 comprised running tunnels in
bored tunnel and cut-and-cover and the construction of twelve stations.
In general, the cut-and-cover tunnels were built within braced sheet pile
trenchworks, except in one section close to rather sensitive buildings where
a proprietary contiguous (PIP) piling system was used.
The contracts were let as design-and-construct contracts within specifi-
cations and layouts prepared by the engineer. Local ground conditions
comprised granites in various stages of decomposition, varying from strong
intact rock to stiff residual clays containing granite boulders. These ground
conditions impose many practical construction difficulties for bored tunnel,
large exavations and cut-and-cover work.
Table 9.2 compared those methods of station construction envisaged prior
to bidding and those adopted by the successful contractors. Generally, it had
been planned that the twelve stations would be built top-down using dia-
phragm walls, but with three main exceptions: bored tunnel construction
for Lok Fu station; top-down construction within sheet pile walls at Diamond
Hill; and bottom-upwards construction in open rock cut at Shek Kip Mei. In
the event, considerable changes were made as contractors’ alternatives for
both construction method and walling techniques were considered and then
adopted.
Choi Hung and Diamond Hill stations used a locally popular technique at
that time, the hand-dug caisson. At Choi Hung (Fig. 9.17) the caissons used
were interlocked to form the station walls, and at Diamond Hill (Fig. 9.18)
the system was modified for the semi-permanent walls by the use of anchored
plates spanning king piles installed within the caissons. Hand-dug caissons
were used at both stations to place permanent columns prior to bulk excava-
tion works. More recently the use of family caissons in Hong Kong has
declined significantly due to the unacceptable health and safety risks. Only
where all other methods are considered inapplicable are hand-dug caissons
used in Hong Kong, although at present some use is made of the method in
mainland China.
At Argyle station, Hong Kong, one of the largest, with a concourse level,
two track levels, 13 entrances and three ventilation shafts, the intensity of
street traffic and the proximity of tower blocks, some with piles founded
near station formation level, favoured the top-down construction method.
Cut-and-cover construction 421

Fig. 9.17. Choi Hung station,


Hong Kong MTR: interlocked
hand-dug caissons4

Fig. 9.18. Diamond Hill station,


Hong Kong MTR: jack arches
between soldier piles installed
in hand-dug caissons4

Due to difficult subsoil conditions with large granite boulders and the need to
penetrate bedrock, secant piles installed by Benoto oscillating rigs were
chosen by the contractor. The piles, 1.2 m in diameter and bored at 1 m
centres, were reinforced in both male and female piles by 914 mm  305 mm
universal beam sections. Figure 9.19 shows a plan and cross-section of the
station and Fig. 9.20 shows a typical cross-section during excavation for the
lower track. To reduce settlements due to dewatering and subsidence of
adjacent buildings, a bentonite cement and silicate grout curtain was made
below the toe of the structural wall to form a cut-off to the box from the
high water table where this could not be achieved by the Benoto rigs. This
grout curtain produced excellent results, restricting the abstraction rate
from the whole box to less than 0.0045 m3 /s under a differential head of
more than 20 m. Design of the walls during construction and in the permanent
case used active and at-rest earth pressures with plastic methods and limit
state checks. Station columns, heavily loaded, in some cases up to 15 MN,
comprised 1000 mm  800 mm steel boxes in 50B steel, 20 m long.
Figure 9.21 shows a shear shoe and plate arrangement used to transfer high
loads from slabs to walls and columns.
At Tsim Sha Tsui station, shown in Fig. 9.22, the bottom-up construction
sequence with PIP piles was used by the contractor in preference to the
original, pre-bid, top-down method using diaphragm walls, with the following
advantages4 .
422 Deep excavations

Fig. 9.19. Argyle station, Hong


Kong MTR: plan and vertical
section showing soil profile and
location of existing buildings4

(a) Where rock existed above formation level, difficult underpinning work
was avoided.
(b) Less noise and vibration was caused by PIP piling.
(c) The PIP wall was narrower than the diaphragm wall.
(d ) The PIP wall provided drier conditions in which to build the permanent
structural box.
(e) The PIP pile did not require wide, heavy reinforcement cages as used in
the diaphragm wall panels.
( f ) The work construction period was reduced.
In the event, the method was successful. Figure 9.23 shows the formation of
PIP walls, installed by a large continuous flight auger (CFA) with mortar
placement through the hollow stem, in a simlar way to CFA piles. The male
PIP piles (piles B), however, required a cement paste injection pressure
of 200 kg/cm2 to make a vertical mortar cut-off between adjacent female
Cut-and-cover construction 423

Fig. 9.20. Argyle station,


Hong Kong MTR: typical
cross-section showing
top-down construction for lower
track slab using secant pile
walls and ground treatment to
secure cut-off to bedrock4

Fig. 9.21. Argyle station, Hong Kong MTR: detail of shear shoe and plate to transfer high loads from slabs to secant pile walls:
(a) side view; (b) end view; (c) force transfer4
424 Deep excavations

Fig. 9.22. Tsim Sha Tsui


station, Hong Kong MTR: typical
cross-section of cut-and-cover
station constructed by bottom-
upwards method showing
temporary deck support4

Fig. 9.23. Tsim Sha Tsui


station: construction of wall
using PIP piles4
Cut-and-cover construction 425

Fig. 9.24. Tsim Sha Tsui


station: PIP construction
sequence: (a) augering;
(b) withdrawing auger and
injecting mortar; (c) completing
injection; (d) inserting
reinforcement cage or steel
column section; (e) augering;
(f) withdrawing auger then
injecting and jetting mortar;
(g) completing injection and
jetting; (h) inserting
reinforcement cage or steel
column section4

piles. A rock boring machine (typically a Koken N50 Big Man) was used to
bore through boulders or into bedrock. Figures 9.23 and 9.24 show the
sequence of piles used at Tsim Sha Tsui station: piles A were taken 500 mm
below formation level to support the traffic deck at road level; piles B were
taken 1.5 m into rock, and the pressure-injected sealing piles C were taken
to rockhead.
The following loading and permissible stress values were used in the
design:
. wall stiffness per metre width: EI ¼ 9:40  108 kN/cm2 per metre
. traffic load: 14.7 kN/m2
. building load: 353 kN/m2
. earth pressure: trapezoidal loading, active pressure
. water pressure from 1 m below ground level: fully hydrostatic
. design strength of pile mortar: 23.5 N/mm2 .
The design loading model for the cofferdam is shown in Fig. 9.25. A typical
cross-section of the temporary support and the permanent works in Fig.
9.22 shows the six frames of pre-loaded H steel strutting at centres of 2.1 m.
Considerable settlements to existing buildings resulted from the initial
Hong Kong MTR construction. Settlement was primarily due to dewatering,
diaphragm wall panel installation and bulk excavation, but more importantly
the combination of properties of the decomposed granite subsoil and the
groundwater regime, unusual for developed city centres, was conducive to
high installation deformations.
Davies and Henkel17 referred to the construction of Chater station and
settlements of the existing Courts of Justice building. A section of the con-
struction and soil profile is shown in Fig. 9.26. The permeability of marine
deposits was of the order of 107 m/s compared with 105 m/s for the under-
lying decomposed granite. Wide variations in drawdown were expected due to
local variations in the geological profile, and preliminary studies showed
unfavourable dewatering settlements could result. Pumping tests had shown
that for each 1 m of drawdown a settlement of 4 mm would result. A system
of groundwater recharge was used, however, both at the Courts and else-
where, with beneficial results.
426 Deep excavations

Fig. 9.25. Tsim Sha Tsui


station: earth pressure, water
pressure and horizontal
surcharge loading diagrams4

Fig. 9.26. Cross-section of


Chater station, Hong Kong,
showing soil profile and
location of existing structures17
Cut-and-cover construction 427

Fig. 9.27. Chater station,


settlement record of ground
adjacent to diaphragm wall
construction17

Settlements due to diaphragm wall panel excavation for Chater station


had unpredicted and serious consequences. Figure 9.27 shows the extent of
movements of three points, D, E and F, spaced 6 m, 15 m and 2 4m, respec-
tively, from the diaphragm wall. The progressive settlement of the points,
even as diaphragm wall installation was completed well away from the vicinity
of the points, is clear. This is a most unusual phenomenon. Measurements of
soil movement due to diaphragm wall installation before and since in widely
differing soils, show very small soil movements (a few millimetres) due to
panel excavation at points near the panel and very early reduction of soil
movement at distances of less than half the panel depth from the panel, as
428 Deep excavations

described in references by Farmer and Attewell18 , Uriel and Oteo19 , Clough


and O’Rourke20 , Thompson21 , Carder22 and Carder et al23 . Generally,
vertical settlement at the wall due to diaphragm wall installation in stiff clay
is of the order of 0.05% of wall depth and probably becomes negligible at a
distance of 1.5 times wall depth from the wall. The progressive movement
as panels were excavated at Chater station was therefore unprecedented and
unexpected. Morton et al.24 concluded that this cumulative movement
resulted from the combined effect of a relatively high permeability and a
medium compressibility of the completely weathered granite. A softened
zone of highly compressible soil occurred as the diaphragm panel was
excavated due to relaxation of horizontal soil pressures. After the panel was
concreted this softened zone compressed due to soil pressures and, in turn,
reduced arching action within the soil as further panels were excavated. It
may be summarized that such large soil movements only occur as a result
of panel excavation in expansive soil with a high permeability and access to
a groundwater supply.
In Singapore, metro construction did not proceed until the late 1980s and
many organizational and technical lessons learned on the Hong Kong MTR
were used to good effect. The works were let mainly on a design-and-construct
basis and Table 9.1 showed the wide range of station constructions chosen.
The ground conditions in Singapore3 are predominantly soft clays and
loose sands or, from an earlier period, stronger soils and rocks. The soft
clays and loose sands have been laid down in valleys eroded into the under-
lying rocks during periods of low sea level. Where the soft clays and loose
sands were virtually absent, cut slopes were used with improvement where
necessary by soil nailing, rock bolting or anchoring. Generally, cut-and-
cover structures were not supported on these clays; piles or diaphragm
walls were used to take loads down to a lower bearing stratum. Where
there was any significant depth of soft clay a continuous walling system was
used. Sheet piles were used for this purpose, except where existing buildings
were particularly vulnerable to soil movement, in which case diaphragm
walls were used. A typical cut-and-cover station 15 m deep would generate
a net active pressure in the marine clay until the shear strength exceeded
about 60 kPa. This strength would only occur at a depth of about 35 m,
where the clay persisted to this depth, and would require heavily reinforced
diaphragm walls and special construction measures to resist the very high
moments and deflections induced in the wall. A thick layer of weak soil
above or below formation level poses problems for the designer in coping
with these very high wall moments prior to placing the lowest bracing
frame.
At Bugis station, 1.2 m thick diaphragm walls were socketed into dense
cemented old alluvium up to 14 m to produce walls up to 54 m deep. Where
the depth of old alluvium support to the wall was greatest, the marine clay
above and below the final excavation level was strengthened with chemico-
lime piles. Figure 9.28 shows a cross-section of the station. Seven frame
levels were used, but even so, deflections of 150 mm were measured in the
diaphragm walls.

Diaphragm walling: recent developments in cut-and-cover works


In recent years, since the end of the 1990s, diaphragm walling has continued to
find extensive use in cut-and-cover works for deeper and more heavily loaded
walls. In turn, the technique has been improved and varied by competing
specialists. These improvements have been applied to excavation plant and
diaphragm walling site practice (as referred to in Chapter 8), but specific
changes have also been made because of the demands of the plan length
Cut-and-cover construction 429

Fig. 9.28. Bugis station,


Singapore MRT: cross-section
showing seven cofferdam
frames used to minimize wall
deformation3

and depth of cut-and-cover works, both in terms of design and production


improvements.

Semi top-down construction


The development of a variation on conventional top-down construction was
made, as explained previously in this chapter, on the N.E. line of Singapore
MRT system to reduce bracing and strutting, to expedite excavation and
still gain the advantages already secured by top-down construction, the
minimization of soil and wall deformation to reduce settlements and ground-
water leakage. A typical application, described by Mitchell et al.25 , was made
on two station boxes on Contract 705, after a tender stage comparison of
alternative methods; a summary is shown in Table 9.4, reproduced from
this paper. The first decision to use diaphragm walls instead of sheet piling
was made to avoid problems with nearby buildings and services. The next
decision, the choice of construction method was guided by the disadvantages
of large working space and the large props needed to build bottom-up (the
box widths were up to 60 m) and the difficulties with conventional top-
down due to a very heavy roof structure and a complex internal lay out
needed for station operation. These matters were solved by the use of large
construction openings in the roof to reduce loading whilst keeping the lateral
support of the structure. The roof openings are shown in Fig. 9.29, at Boon
Keng station.
The stiffness of this roof (the thickness of 2 m was required for civil defence
purposes) compared to the diaphragm walls of the box (not yet lined) meant
that a large moment was temporarily transferred to the walls at the junction
with the roof. From a crack width consideration, the necessity to comply with
a maximum width of 0.2 mm as specified caused considerable design effort to
show that in the long term the influence of creep in the diaphragm walls, the
influence of the 1 m thick R.C. lining walls and the re-application of water
pressure would limit cracks to the 0.2 mm width.
The semi-top-down method also had the advantage that it allowed the base
slab to act as a raft before all the weight of the superstructure is applied and
thereby reduced settlement.
430 Deep excavations

Table 9.4 N.E. line Singapore MRT; summary of alternative methods at tender stage25

Construction Main quantities Advantages Disadvantages


method (per station)

Sheet pile . Sheet pile: 18 000 m2 . Lowest cost . Difficult to install sheet piles in
(FSP 4 3420 t) . No influence of temp works on hard ground
. Steel struts and walers: station design . Noise and vibration
2900 t (total of 5 or 6 . Early start on site . Sensitive utilities and buildings
layers) very close
. Decking steel: 1700 t . Ground treatment under utility
crossing required
. Congestion of site with six layers
of struts
. Sequence of work affected by
strutting
. Removal of all temporary works
required
D-wall . Diaphragm wall: . High stiffness and water . Permanent works design
Bottom-up 16 000 m2 tightness of retaining wall approval needed to start D-wall
. Steel struts and walers: . D-wall can be installed below . Congestion of site with six layers
2500 t (total of 5 or 6 utility of struts
layers) . D-wall is part of permanent . Sequence of work affected by
. Decking steel: 1700 t works strutting
D-wall . Diaphragm wall: . High stiffness and water . Permanent works design
Top-down 16 000 m2 tightness of retaining wall approval needed to start D-wall
. Temp props: 1000 t . D-wall can be installed below . All walls have to be under-
. Steel strut and waler: utility pinned to soffit
400 t (1 layer) . D-wall is part of permanent . Large number of props to
works remove at end
. Minimal heavy duty falsework . All work under roof in confined
required space
. Fairly clear working area
D-wall . Diaphragm wall: . High stiffness and water . Permanent works design
Semi-top- 16 000 m2 tightness of retaining wall approval needed to start D-wall
down . Steel strut and waler: . D-wall can be installed below . Complex design
400 t (1 layer) utility . Some walls have to be
. Minimal temp props . D-wall is part of permanent underpinned to soffit
works . Large openings have to be closed
. Very clear working area later

A precedence network prepared for an example of semi-top-down construc-


tion is shown in Fig. 9.30, based on Mitchell et al. and prepared by Brian Bell
Associates.

Movement joints in box structures


The use of movement joints within the diaphragm walls of cut-and-cover
structures has, until now, not been deemed necessary by designers either on
the grounds of thermal movement or flexural movement. On the construction
of the Stratford station box, part of the Channel Tunnel Rail Link works in
London, movement joints have been incorporated in the 1070 m long box,
with its width of 50 m and a depth varying between 16 and 22 m. The structure
is permanently propped at each end by concrete props at 10 m centres and by
road and rail bridges in the middle of the structure. Temporary props are used
below the permanent props. The base slab is unreinforced concrete and a
Cut-and-cover construction 431

Fig. 9.29. Boon Keng station,


semi top-down construction of
the N.E. line, Singapore MRT
(courtesy of Benaim
Consultants)25

permanent dewatering scheme will operate by deep wells in the chalk stratum
below the box to relieve water pressure on the base slab during its 120-year
design life.
The movement joints, shown in Fig. 9.31, were designed to accommodate
differential deflections, settlements and in-plane movements. The design
requirements for the joint were: þ100 mm=0 mm for horizontal deflection,
þ10 mm=30 mm for horizontal in-plane movement and 10 mm for vertical
settlement with joint watertightness for the 120-year design life. The joints are
prefabricated off-site in 9 m lengths, in a steel–rubber sandwich construction
which is pre-compressed and pinned in the factory. After installation and
concreting of the wall each side of the joint the pin system is released following
initial concrete shrinkage in the wall, the joint remaining in compression.

Soft tunnel eyes


The Thames tunnel crossing for the Channel Tunnel Rail Link has recently seen
the use of a soft eye formed from glass fibre bar reinforced concrete within the
end diaphragm wall panels of the reception chamber for the tunnel boring
machine (TBM). This innovation was necessary as a replacement for the
normal slurry block on the outside of the chamber through which the TBM
travels without allowing a ‘blow’ of groundwater and soil into the reception
chamber. The composite cage, made from the conventionally steel reinforced
sections and the glass fibre bar section was lifted successfully in one piece
(Fig. 9.32). Wider application of ‘eyes’ of glass fibre reinforcement may be
expected to allow permanent access ways through structural diaphragm walls.

Cut-and-cover walls of varied plan form


Where working space allows, the straight walls of cut-and-cover sections can
be replaced by walls of varied plan shape to produce walls of improved stiff-
ness and flexural strength. The use of diaphragm walls built to T plan shape
and castellated plan shapes with shear joints between panels has been referred
to in Chapter 8. A further variation is the use of walls built as horizontal
arches with temporary tubular stop ends. The arch shapes span between
props or diaphragm cross walls as shown in Fig. 9.33. The improved stiffness
and flexural strength of these plan shapes allows reduced vertical propping at
increased centres with both cost and programme advantages where working
space permits.
432
Deep excavations

Fig. 9.30. Semi top-down construction: precedence network (after Mitchell et al.25 ) prepared by Brian Bell Association
Cut-and-cover construction 433

Fig. 9.31. Movement joint construction: Fig. 9.32. Soft tunnel eye construction:
Channel Tunnel Rail Link works, composite reinforcement cage35
Stratford station box, London

Observational techniques
The repetitive nature of cut-and-cover construction, in which wall panels are
progressively cast, bulk excavation made and bracing frames inserted, allows
any observed production or technical improvement in walling or strutting
to be introduced at an early stage as the work proceeds. The principles of
observational soil mechanics as described by Peck26 are particularly relevant
to cut-and-cover construction. An example of the successful use of this tech-
nique is the Limehouse Link highway tunnel in East London which was built
in the early 1990s. The original design of the top-downwards construction
required temporary 1350 mm dia. steel props between diaphragm walls on
each side of the cut-and-cover box below roof level. The props were lifted
into place using hoists supported from the soffit of the roof slab (Fig. 9.34).
Excavation then continued to formation level below the line of struts. This
excavation was slow and costly due to the presence of the struts.
The observational method was applied progressively in a number of stages.
Initially, props were destressed and removed one at a time as wall movements
were measured. Since wall displacements were small, a new section with ‘soft’
props was installed with a small gap allowed at the end of the strut prior to
load take-up. Since movement again proved to be very small as excavation

Fig. 9.33. Plan of diaphragm


wall construction using
horizontal arch plan shapes
434 Deep excavations

Fig. 9.34. Limehouse link,


London. Heavy props used in
top-down construction34

was taken below the props to formation level, the mid-height props were
omitted and excavation was made to full depth prior to the installation of a
strut at blinding level. The monitored wall movements were still very small
and this allowed the continuing omission of the centre struts. Eventually
the blinding struts were also omitted. Contingency struts were always kept
available, but were not needed. The trigger level for maximum wall movement
was defined as 70 mm but the maximum recorded movement was 11 mm, and
generally readings were less than 7 mm. Considerable savings resulted from
avoiding the use of these heavy props.
A similar application of the observational method to reduce the propping to
a cut-and-cover excavation was reported by Beadman et al.27 , for the excava-
tion support (by a secant pile wall) at Norreport station on the Copenhagen
Metro. A system of trigger levels at the various design sections along the
station defined risk and the need to implement contingency measures. The
design calculations were based on most probable soil parameters based on
back analysis of a previous station excavation. Deflection profiles (based on
analysis by the springs program, WALLAP) defined the trigger values. The
green limit was assessed as 70% of the design values, the amber limit used
the most probable soil parameters whilst the red limit, stipulating the stoppage
of excavation was defined as 120% of calculated horizontal deflection values.
This value of 120% of designed deflection still ensured the secant piles
remained within their ultimate capacity.

Precast diaphragm walls


Precast concrete panels were introduced into diaphragm wall works in France
by the firms Bachy and Soletanche during the early 1970s. Each company
Cut-and-cover construction 435

Fig. 9.35. Cairo Metro: junction


between precast and in situ
diaphragm walls

obtained patents for its particular technique. The innovation found early
application in cut-and-cover construction and was used in Paris for underpass
and metro construction, and in both Lille and Lyon for metro construction.
The technique has not been used in the UK, and appears to have found less
application in recent years in France. (More recently the two companies
have merged to trade internationally together as one.)
In the Far East, in Hong Kong and Thailand, thick, heavily reinforced in
situ diaphragm walls are preferred; in Japan there are only a few examples
of precast walls; and there are no known precast walls in the USA. This
lack of acceptance of a potentially attractive innovation is probably due to
a unit cost disadvantage between in situ and precast walls. The introduction
of the cutter machine and its use by the largest diaphragm wall contractors
may also have detracted from the popularity of precast walls; the reverse
circulation process cannot be economically applied when grout is used as
the stabilizing fluid during excavation.
The principal feature of the precast diaphragm wall is the absence of any
surface finishing subsequent to its exposure after bulk excavation. On bulk
excavation the cement–bentonite slurry strips away from the inside surface
of the wall to reveal the precast concrete surface, to true alignment. Figure
9.35 shows the junction between in situ and precast diaphragm wall construc-
tion on the Cairo Metro.
The use of prefabricated diaphragm walls for metro cut-and-cover con-
struction was described by Namy and Fenoux28 . They noted two fundamental
disadvantages of in situ reinforced concrete diaphragm walls:
(a) the surface finish and quality of excavation of the wall depends on
subsoil conditions
(b) the water resistance of the concrete and the joints may be inadequate.
The development of precast wall methods offers several advantages over in
situ diaphragm wall construction.
436 Deep excavations

(a) Site nuisance is reduced by more rapid execution. The sequence of panel
excavation is simplified by successive panel excavation, whereas in situ
diaphragms frequently use primary, secondary and intermediate panel
excavation sequences to allow hardening of concrete. Remedial works
in breaking down walls to level or to profile are largely unnecessary.
(b) Site concreting operations and stop end extraction are avoided.

Fig. 9.36. Paris Metro, St Denis,


cross-sections of precast
diaphragm wall construction:
(a) top of excavation supported
by hardened slurry and bracing;
(b) arrangement of precast
concrete elements used in
cut-and-cover walls (courtesy of
Soletanche–Bachy)
Cut-and-cover construction 437

(c) In the permanent phase, constructional thicknesses are reduced by the


improved concrete qualities brought by precasting (a 400 mm precast
wall can be equivalent to a 600 mm in situ wall panel). By incorporating
water bars into the precast panels, good wall finishes and better water
resistance are possible.

The use of precast wall panels enables prefabricated units to be made up


with soldier beams for temporary soil retention above the precast wall. An
example of prefabricated, precast diaphragm construction using the Panosol
system is shown in Fig. 9.36 in cross-section. The works, an extension of
the Paris Metro in the heart of St Denis, extended 500 m along a confined
route, 12 m wide wall-to-wall, bordered by old, delicate buildings.
The subsoils consisted of fill, gypsiferous marls and clayey greensands
overlying St Ouen limestone. The marls acted as an upper aquifer close to
street level, and the lower aquifer of limestone has its piezometric head near
the top of the greensand.
Refuge holes, cross-drainage holes and slab starter bars were incorporated
into the precast panels which transferred vertical load through the sealing
grout into the limestone. No lining to the precast walls was used in the finished
structure although an epoxy resin based treatment was applied to the joints on
the inside to provide long-term waterproofness. The wall works, including site
set-up, were completed in a period of seven weeks, with a daily output of 30
linear metres of walls complete.
A precast diaphragm wall was also used for metro cut-and-cover con-
struction on the line serving Charles de Gaulle Airport at Sevran. Again,
gypsiferous marls and greensands overlying St Ouen marls and limestone
contained separate aquifers (shown in Fig. 9.37). Due to the risk of large
voids and solution cavities, sometimes several metres deep in the marly lime-
stones, pre-grouting with bentonite–cement grout with a high sand content
was carried out under gravity. Smaller cavities in the greensands were injected
with bentonite–cement under low pressure. The insertion of the precast
panel and the water stop is illustrated in Fig. 9.38. At the base of the panel
a blade-shaped plate slides into the vertical groove within the adjacent
panel, which serves as a guide as the new unit is lowered. The water stop fol-
lows the blade during the lowering. The weight of the panels reached 37
tonnes; the standard panels were 9.50 m high by 3.35 m wide by 0.45 m
thick. Extra-thick panels were provided at areas of high loading and at
tunnel refuge holes. The site output averaged 30 m of precast panel placed
daily, the side walls of the 300 m cut-and-cover section being constructed in
less than five months.
New metro construction in Lyon and Lille during the late 1970s provided
an opportunity for the use of precast diaphragm walls at two stations,
Saxe-Gambetta and Gare de Lille, and on sections of running tunnel in cut-
and-over construction.
Saxe-Gambetta station (Fig. 9.39) was built at the junction of two lines in
Rhone alluvium 25 m thick underlain by a relatively impermeable sand. The
groundwater table was at a depth of only 3.5 m and the alluvium was very
permeable. The station incorporated a precast diaphragm wall to support
the soil temporarily and provide a cut-off into the sand substratum, together
with an in situ reinforced concrete tunnel section. A sandwich-type water-
proof membrane was applied to the inside face of the precast wall.
The precast wall was designed to support all loads – soil and groundwater
pressure and surcharge loads – during construction. In the permanent
condition, the load was divided between soil load on the Panosol wall and
water pressure on the reinforced concrete tunnel structure. In the temporary
438 Deep excavations

condition the permeable alluvium became impregnated with the cement–


bentonite slurry used in the wall excavation, with a resulting reduction in
short-term soil pressures and deformation.
The works were built in open trench (Fig. 9.39(b)) with the exception of one
section beneath Gambetta Road where traffic could not be diverted (Fig.
9.39(c)). In this latter section, intermediate supports were needed from barette
panels. The construction sequence for the works in the open was:
(a) relocate existing utilities
(b) install precast diaphragm wall from street level (economy in the use
of materials and panel weight was achieved by making a cut-off at
the level of the relocated utilities, an H-beam was set in the top of
each panel to allow cantilever support for a temporary roadway for
light vehicles)
(c) excavate to groundwater level
(d ) concrete between H-beams beneath temporary roadway to support
utilities
(e) install temporary ground anchors
( f ) complete excavation and station construction (Fig. 9.40).
Gare de Lille station, which was planned to connect with a future line, was
built in a larger box 230 m long, 27 m wide and between 14 and 16 m deep.
The ground conditions consisted of fill and alluvium overlying chalk with a

Fig. 9.37. Cross-section of


precast diaphragm wall
construction for the rail link to
Charles de Gaulle Airport, Paris
(courtesy of Soletanche–Bachy)
Cut-and-cover construction 439

groundwater table at the base of the surface fill material and just below per-
manent roof level of the tunnel section. As at Lyon Saxe-Gambetta station,
the same combination of Panosol precast wall and in situ reinforced concrete
tunnel section with a sandwiched waterproof membrane was used. The plan

Fig. 9.38. Rail link to Charles de


Gaulle Airport, Paris: (a) view of
assembly of precast units in
slurry trench; (b) view of
completed cut-and-cover tunnel
(courtesy of Soletanche–Bachy)
440 Deep excavations

and cross-section of the works is shown in Fig. 9.41. The rate of flow of
groundwater into the completed excavation was limited to about 50 m3 per
hour, demonstrating the effectiveness of the cut-off.
Precast units in both reinforced and pre-stressed concrete were used in
sections of the cut-and-cover for the running tunnels of the Lyon Metro,
depending on depth to formation. Typical sections are shown in Fig. 9.42,
illustrating the use of a grouted base within the walls below formation level
and at a depth to balance the groundwater pressure within the alluvium. In
some areas, to obtain cut-off within the underlying sandstone it proved
more economical to extend the depth of the self-hardening slurry wall
where the formation level was deeper.
Ingress of the self-hardening slurry, containing between 150 and 250 kg of
slag and cement per cubic metre, into the alluvium at the sides of the
excavation was high – estimated at between 1 and 1.5 m3 per square metre
of wall area. The assumed short-term strength properties of the alluvium
allowed for this loss and a value of 20 kN/m2 was used for cohesion in
the design of wall and strutting in the temporary condition. As the water
resistance of the permanent structure was achieved with the sandwiched
impermeable membrane, the panels were made of rectangular section with

Fig. 9.39. Lyon Metro, Saxe-


Gambetta station: (a) plan of
site area; (b) section A–A,
works in open trench; (c) section
B–B, at junction, constructed by
top-downwards method
(courtesy of Soletanche–Bachy)
Cut-and-cover construction 441

Fig. 9.40. Lyon Metro,


Saxe-Gambetta station:
completed station excavation
showing precast diaphragm
walls (courtesy of
Soletanche–Bachy)

Fig. 9.41. Gare de Lille station: (a) plan; (b) transverse sections (courtesy of Soletanche–Bachy)
442 Deep excavations

Fig. 9.42. Lyon Metro, typical


sections of running tunnel:
(a) deep section constructed
with precast reinforced
concrete units; (b) shallow
section using precast
pre-stressed concrete units
(courtesy of Soletanche–Bachy)

no special jointing devices, temporary waterproofing being obtained from the


self-hardening slurry. Figure 9.43 shows illustrations of the Panosol wall con-
struction applied to the Lyon Metro.
The versatility of precast diaphragm walls in cut-and-cover construction
is demonstrated in Figs 9.44 to 9.46. Bachy’s method was used to build a
culvert at Vitny. Figure 9.44 shows a cross-section of the completed works,
which were constructed to high standards of finish, alignment and water
Cut-and-cover construction 443

Fig. 9.43. Lyon Metro: two views of exposed precast Panosol panels in running tunnel
(courtesy of Soletanche–Bachy)

Fig. 9.44. Vitny culvert:


cross-section of completed
works (courtesy of
Soletanche–Bachy)

resistance. The culvert was located in a narrow commercial street in the centre
of the town. The finished culvert, of internal rectangular section, is 3.75 m high
and 3.5 m wide with approximately 3 m depth of cover from existing
carriageway levels, the roof being just below groundwater level. The sequence
of construction was as follows.
(a) Construct a 350 mm thick precast diaphragm wall within an excavated
slurry trench 600 mm wide. The top of the precast wall was carefully
levelled to the soffit level of the roof slab, and the cementitious slurry
within the trench above this level was reinforced with steel mesh.
(b) Excavate between the walls in a strutted excavation, the upper 3 m of
exposed cementitious slurry being protected by sheeting behind vertical
444 Deep excavations

Fig. 9.45. Vitny culvert:


cross-section showing
temporary soil support above
permanent walls (courtesy of
Soletanche–Bachy)

runners. The average depth of excavation to the underside of the culvert


base slab was 7.3 m from ground level, the precast walls and the slurry
securing a cut-off into the marl.
(c) Cast in situ reinforced concrete floor and roof slabs.
(d ) Complete waterproofing of joints.
(e) Backfill over roof slab and reinstate carriageway.
The culvert constructed under this contract was 700 m long, used 524 precast
concrete wall panels weighing approximately 15 tonnes each. An output of 7.5
linear metres of culvert structure was achieved per day.
Bachy’s patented continuous water bar system was used in this work. A
perspective view is shown in Fig. 9.47. Vertical sections of PVC water bar
are cast into a recess in the face of the prefabricated panel. These are sub-
sequently thermally welded to a third section of water bar in the horizontal
plane after exposure of the wall following the main excavation. The horizontal
water bar is cast into the in situ floor slab of the culvert. The vertical
recess between the panels is finally filled with mortar reinforced with steel
mesh.

Overall stability: Cut-and-cover works are frequently constructed in water-bearing soils and in
design for uplift such circumstances it is necessary to consider the risk of failure of the struc-
tural box by uplift pressures both during construction and during the
design life of the structure. The total downward self-weight of the structure
together with the frictional resistance of the external walls, anchors or tension
piles is required to exceed the upward hydrostatic force by an acceptable
factor of safety at each stage.
In particular, tidal conditions, should they exist, should be considered
pessimistically over the design life with allowance for inaccuracy in predicted
levels. It would be usual to consider the restoring force in this factor of safety
Cut-and-cover construction 445

Fig. 9.46. Vitny culvert, successive stages in construction: (a) excavation by grab;
(b) final stages of excavation and strutting; (c) culvert construction
(courtesy of Soletanche–Bachy)

to be based on:
(a) dead weight of structural elements based on minimal dimensions but
the displacement of the structure based on maximum overall
dimensions
(b) height of fill above the roof of the cut-and-cover to final finished levels
in permanent condition only
(c) frictional resistance due to piled walls or diaphragm walls based on
the inner and outer surface of the walls below the underside of the
base slab
446 Deep excavations

Fig. 9.47. Vitny culvert: water


bar system with continuity
between vertical and horizontal
water bars (courtesy of
Soletanche–Bachy)

(d ) total resistance from anchors or tension piles based on the ultimate


capacity of anchors or piles divided by 2.0, using conservative values
of soil or rock parameters, unless the results of pull-out tests are
available.
Usually, an overall factor of safety of at least 1.1 on dead weight of the struc-
ture and fill over is required. A minimum value of 1.4 is required when the
effects of friction and resistance due to anchors or tension piles are included;
a further check that upward force does not exceed
 
dead weight friction
þ
1:1 3:0

is prudent.
A draft Institution of Structural Engineers report on Basements and
Cut and Cover29 refers to design specifications for buoyancy and flooding
Cut-and-cover construction 447

Table 9.5 Typical safety factor requirements, uplift for cut-and-cover structures29

Condition Downward forces D Upward forces U

Partial safety factors Partial safety factor on water


density or displacement (f )

On weights (m ) On friction m , i.e. sides,


piles, anchors
During construction 1.01 2.0 1.01
In service 1.05 3.0 1.05
Extreme event (flooding to 1 m 1.03 2.5 1.03
above ground level)  
P D X
Criterion (for each condition): > U  f
m

of underground structures for railway clients in Hong Kong, Singapore and


London. Typical clauses included a summary of partial safety factors as
shown in Table 9.5. Accompanying such a table would be a list of
specified material densities. Water density variation of 2% between fresh
and sea water was noted, the appropriate density being used in the
calculation. These safety factors would be modified if the centre of
factored buoyancy did not reasonably correspond to the centre of factored
gravity.
The draft report29 also adds that these metro authorities prudently
require the threshold to their underground stations to be not less than 1.0 m
above local ground level with no apertures below this level to guard against
extensive flooding underground from one source. Similar requirements
apply to highway tunnels, particularly approaches to river crossings, in the
UK.

References 1. Megaw T.M. and Bartlett J.V. Tunnels: planning, design, construction. Ellis
Horwood, Chichester, 1981, Vol. 1.
2. Cook R. and Paterson J. Nam Cheong station, Hong Kong. Struct. Eng., 2002,
80, No. 11, 13–15.
3. Hulme T.W. et al. Singapore M.R.T. system: construction. Proc. Instn Civ.
Engrs, Part 1, 1989, 86, Aug., 709–770.
4. McIntosh D.F. et al. Hong Kong M.T.R. modified initial system: design and
construction of underground stations and cut and cover tunnels. Proc. Instn
Civ. Engrs, Part 1, 1980, 68, Nov., 599–626.
5. Ridley G. Blackwall Tunnel duplication. Proc. Instn Civ. Engrs, 1966, 35, Oct.,
253–274 (discussion 1967, 37, Jul., 537–555).
6. Morgan H.D. et al. Clyde Tunnel design, construction and tunnel services. Proc.
Instn Civ. Engrs, 1965, 30, Feb., 291–322.
7. Haxton A.F. and Whyte H.E. Clyde Tunnel construction problems. Proc. Instn
Civ. Engrs, 1965, 30, Feb., 323–346 (discussion 1967, 37, Jul., 511–535).
8. Megaw T.M. and Brown C.D. Mersey Kingsway Tunnel. Proc. Instn Civ. Engrs,
1972, 51, Mar., 479–502.
9. Beadman D.R. and Bailey R.P. Design and construction of the deep stations for
the Copenhagen Metro. Proc. Conf. on Deep Foundations, 2000, 375–387. Deep
Foundations Institute, New York, 2000. DFI, Englewood Cliffs, New Jersey,
2000.
10. Weinbold H. and Kleinlein H. Berechnung und Ausführung einer schrägen
Bohrpfahlwand als Gebäudesicherung. Der Bauingenieur, 1969, 44, Jan., 233–
239.
448 Deep excavations

11. Hana T.H. and Dina A.O. Anchored inclined walls, a study of behaviour. Ground
Engng, 1973, 6, Nov., 24–33.
12. Schnabel J.R. Sloped sheeting. ASCE J. Civil Engng, 1971, Feb., 48–50.
13. The Icos company in the underground works. Icos, Milan, 1968.
14. Granter E. Park Lane improvement. Proc. Instn Civ. Engrs, 1964, 29, 293–317
(discussion 1966, 33, 657–664).
15. Dasgupta K.N. et al. Calcutta Rapid Transit System and the Park Street under-
ground station. Proc. Instn Civ. Engrs, Part 1, 1979, 66, May, 261–275 (discussion
1980, 68, 127–129).
16. Jobling D.G. and Lyons A.C. Heathrow Station – Piccadilly Line. Proc. Instn
Civ. Engrs, 1976, 60, 212–217.
17. Davies R. and Henkel D.J. Geotechnical problems associated with construction
of Chater Station. Proc. Conf. Mass Transport in Asia, Hong Kong, 1980, paper
J3, 1–31. Concrete Society (HK) Ltd., 1980.
18. Farmer I.W. and Attewell P.B. Ground movements caused by a bentonite-
supported excavation in London clay. Ge´otechnique, 1973, 4, Dec., 576–581.
19. Uriel S. and Oteo C.S. Stress and strain beside a circular trench wall. Proc. 9th
Int. Conf. S.M.F.E., Tokyo, 1977, Vol. 1, 781–788. Japanese Society of Soil
Mechanics and Foundation Engineering, Tokyo, 1977.
20. Clough G.W. and O’Rourke T.D. Construction induced movements of in situ
walls. Proc. Design and Performance of Earth Retaining Structures, ASCE Special
Publication 15, 439–470. Cornell University, 1989.
21. Thompson P. A review of the retaining wall behaviour in overconsolidated clay
during early stages of construction. M.Sc. Thesis, Imperial College, 1991.
22. Carder D.R. Ground movements caused by different embedded retaining wall
construction techniques. Transport Research Laboratory, Crowthorne, 1995.
TRL Report 172.
23. Carder D.R. and Darley P. Long term performance of embedded retaining walls.
Transport Research Laboratory, Crowthorne, 1998. TRL Report 381.
24. Morton K. et al. Observed settlements of buildings adjacent to stations con-
structed for the modified initial system of the M.T.R. Hong Kong. 6th South
East Asian Conf. Soil Engng, Taipei, 1980, 415–429. Organising Committee,
Taipei, 1980.
25. Mitchell A., Izumi C., Bell B. and Brunton S. Semi top-down construction
method for Singapore MRT, NEL. Proc. Int. Conf on Tunnels and Underground
Structures, Singapore, 2000. Balkema, Rotterdam, 2000.
26. Peck R.B. Advantages and limitations of the observational method in applied soil
mechanics. 9th Rankine Lecture. Ge´otechnique, 1969, 19, Jun., 169–187.
27. Beadman D. et al. The Copenhagen Metro observational method at Norreport
station. Proc. ICE Geotech, Eng., 149(4), 231–236.
28. Namy D. and Fenoux G.Y. Tranchées couvertes en parois prefabriquees. Proc.
6th Euro. Conf. SMFE, Vienna, 1976, Vol. 1.1, 183–188. Austrian National
Committee, Vienna, 1976.
29. Institution of Structural Engineers. Design and construction of deep basements
including cut and cover structures. Draft, 2002. Institution of Structural Engi-
neers, London, 2002.
30. Tricoire, J. Le metro de Paris. Paris Museés, 1999.
31. Nam Cheong station, Hong Kong. Struct. Engr, 2002, 80, June 13–15.
32. Kell J. The Dartford Tunnel. Proc. Instn Civ. Engrs, 1963, 24, Mar. 359–372.
33. Bigey M. et al. Construction des metros. Regie Autonome des Transport Parisien,
Paris, 1973.
34. Powderham A.J. The observational method learning from projects. Proc. Instn
Civ. Engrs, Geotechnical Engineering, 2002, 155, Jan., 59–69.
35. News item. Ground Engng, Supplement, 2002.

Bibliography BS EN 1538: Diaphragm walls. British Standards Institution, London, 2000.


BS EN 1537: Ground anchors. British Standards Institution, London, 1999.
BS EN 1536: Bored piles. British Standards Institution, London, 1999.
BS EN 12063: Sheet piling. British Standards Institution, London, 1999.
Cut-and-cover construction 449

BS EN 12715: Grouting. British Standards Institution, London, 2000.


BS EN 12716: Jet grouting. British Standards Institution, London, 2001.
Pr EN 14490: Soil nailing. British Standards Institution, London, 2003.
Darling P. The Limehouse Link: two cut and cover techniques. Tunnels and Tunnel-
ling, 1991, 23, Jan., 16–18.
Roy T. Calcutta metro: contract section 16B: Proc. Instn Civ. Engrs, Part 1, 1983, 74,
Nov., 871–883.
Shafts and caissons:
10 construction and design

Shafts for civil Shaft construction deserves to be considered separately from both conven-
engineering purposes tional cofferdam and caisson works. Often of small diameter, shafts may be
rectangular in plan, or they can obtain maximum benefit from the arching
action of soil around the shaft by conforming to a circular or ellipsoid
shape. Inclined or vertical shafts may be built for permanent or temporary
works. In specialist works such as tunnel construction, a shaft may serve a
dual purpose – first, as a means of access to the tunnel drive during con-
struction, and, at a later stage, as a means of ventilation to the completed
tunnel.
A comparison of shaft and tunnel construction work shows the particular
difficulties that can arise in shaft construction:
(a) groundwater can accumulate on the shaft working face
(b) works and materials have to be transported from the shaft face to
ground surface
(c) excavated spoil has to be removed as a dead weight from the shaft
face
(d ) shaft linings have to be installed progressively downwards.
Although it is beyond the scope of this chapter, a study of installation tech-
niques for deep shafts for mining works can assist in design and construction
for shallower shafts for civil engineering purposes. Jones1 reviewed current
methods of shaft sinking and raise boring for mining work. Mechanical
mole tunnelling machines were then in use for vertical shaft construction
and possessed the twin virtues of increased safety and efficiency. Jones
described the purpose of comprehensive investigation prior to sinking to
determine the safest and most economical excavation method and best utiliza-
tion of freezing, grouting and lining techniques. The practice of collaring to
support the shaft near the ground surface using sheet piling, diaphragm
walling or caisson reconstruction was also described, together with details
of drilling jumbos, mucking out, and hoisting and lining methods. Excavation
methods using raise boring and raise climbing machines and conventional
long hole blasting were also described.
The installation of deep shafts, in the range 100 to 500 m deep was dealt
with by Grieves2 but remains outside the scope of this present volume.

Permanent shafts
In works associated with tunnelling for underground railways and road
tunnels, permanent shafts are used for lifts, escalators, staircase access and
for ventilation purposes. In sewage disposal schemes, shafts find use in pump-
ing station construction and drop shafts. Some of the deepest shafts are in
hydro-electric schemes and pumped storage works.
The vertical connection of tunnel drives to penetrate the sea bed
involves shaft work of a special nature. These works are typically necessary
for cooling water intakes to power stations and for sewage outfalls. In some
Shafts and caissons: construction and design 451

cases, offshore jack-up rigs have been used; the procedure is basically to
install a bulkhead within a sea bed excavation, to raise a shaft from the
tunnel to the underside of the bulkhead and then remove the bulkhead by
blasting.
At Wylfa power station3 a cooling intake was constructed on an exposed
coastline through 11 m of water at low tide with a tidal range of 6 m. The
headworks to the shaft were constructed from a pit blasted on the sea bed;
within the pit were installed a cylindrical shell and bulkhead, concreted into
the rock face. The lower section of the 1.2 m dia. shaft was then excavated
upwards from the tunnel to meet the bulkhead. This pilot shaft was enlarged
to 4 m diameter before removal of the bulkhead.
The Dublin outfall sewer at Howth4 was raised in sound rock from the end
of the tunnel works to just below the sea bed to avoid headworks in expected
high seas. After flooding the shaft and tunnel so constructed, the remaining
length of shaft was removed by underwater blasting.
The use of concreted shafts and drilled shafts to house steel columns to
withstand loads imposed from a superstructure overlaps with the subject of
piling. This is inevitable as the size and power of mechanical piling equipment
increases. Drilling using rock roller bits, large drag bits, mechanical augers
and large diameter down-the-hole hammers, and the support of soil and
rock by differential waterhead, drilling slurry or temporary casing, while
defined in North America as ‘caisson construction’, is beyond the scope of
this book.

Temporary shafts
Megaw and Bartlett5 described the use of temporary shafts for tunnel works
and stated the principal requirements of all-purpose working shafts for
tunnels as follows.
(a) They must be available from the earliest stage of construction until
tunnel completion.
(b) A shaft 4 to 6 m in diameter is typically needed to accommodate
hoisting equipment and provide access for workers. Note that in
pipe jacking works the shaft diameter may depend on dimensions
of precast pipe sections and access clearances; with shield-driven
tunnels the shaft dimensions will depend on the size required for
hoisting shield components and, unless a separate shield chamber is
used, the space needed at the bottom of the shaft to allow shield
fabrication. In addition, the shaft diameter, of at least 1.5 times the
tunnel drive diameter, must be sufficient to allow break-out from
the shaft bottom. For tunnels driven by TBMs the shaft dimensions
are directly dependent on TBM dimensions and clearances around
it; the break-out eye details and dimensions will depend upon shaft
wall construction.
Megaw and Bartlett referred to the spacing of shafts in tunnel works: much
earlier, the progress of tunnel works dug by hand was improved by a large
number of working shafts. In 1838, eighteen shafts were used to drive the
Kilsey Tunnel which was only 2.2 km long. Later, railway works needed a
shaft spacing of about 1 km for steam clearance purposes. The location of
shafts remains a compromise between optimum tunnel alignment and the
availability of adequate space at ground level. This problem was accentuated
in earlier times when shield-driven tunnels could only be made in a series of
straights. Nowadays, availability of adequate space in urban areas with the
requirements of muck-away and construction materials handling and tunnel
ventilation often determine shaft spacing.
452 Deep excavations

Sinking methods
The sinking techniques adopted depend on the shaft use, its diameter and
depth, soil and rock conditions, groundwater state, the proximity of other
structures and their sensitivity to settlement.

Hand-excavated shafts
Although hand excavation would appear to be expensive, where soil is suffi-
ciently stable to stand unsupported for small heights the conditions are rela-
tively dry and moderate labour rates prevail, the method can find economical
application. In China, for instance, ‘family caissons’ are taken down through
residual soils where there is risk of boulders that might impair excavation by
mechanical plant. These hand-excavated shafts are lined with in situ unrein-
forced concrete, typically 75 mm thick in small height lifts. Family caissons
often were used in Hong Kong until the 1990s but safety legislation since
that time has much reduced their use there.
A traditional method of pier construction for foundation support was
known as the ‘Chicago method’ since its introduction on the Chicago Stock
Exchange in 1894. Soil support was obtained from vertical poling boards
set on the pier periphery and held by steel rings. Hand excavation proceeded
in depth increments of 1 to 2 m, depending on conditions, and a further set of
boards and rings were placed. At full shaft depth a hand-belling operation was
carried out to increase the shaft diameter if soil conditions allowed.
Hand excavation may also be more economical in other circumstances; in
weak rocks, for instance, where belling operations may prove difficult for
mechanical augers. At Hartlepool Nuclear Power Station, 17 piers, each
2.3 m in diameter, were used as support to the reactor. The boring was
taken through 5 m of soft fill, clay and 30 m of glacial fill by rotary auger
using bentonite slurry for soil support. Weak bunter sandstone at this depth
was then hand-excavated a further 4 m from beneath a casing set into the
rockhead and backgrouted throughout its height. The base was belled out
by hand to 3.9 m diameter to reduce bearing pressures on the sandstone to
2.9 MN/m2 .

Mechanical excavation
Open shafts have traditionally been excavated by mechanical grabs suspended
from cranes or derricks, where space allows, assisted by mechanical loading
shovels at excavation level. More recently, hydraulic grabs mounted on
cranes or long dipper arms of hydraulic excavators have replaced mechanical
grabs and derricks. In rock, shallow shafts are drilled by hand-held rock drills;
in deeper works, a shaft jumbo is used with boom-mounted drills mounted on
a folding frame. It is usual to pull 1 to 2 m on each round. Beyond a depth of
30 m or so it has been usual to construct a temporary head frame and use
muck skips or kibbles to muck out the shaft.
The distinction between piling and shaft construction may only depend on
size where the shaft is to be backfilled to form a load-bearing member;
mechanical augers are used together with casing, either temporarily or perma-
nently as needed. Bentonite slurry may again be used as a method of soil
support; in the reverse circulation process, slurry is used for both soil retention
and as a means of transporting the excavated soil cuttings. A temporary top
casing is used in these instances to avoid soil disturbance at ground level, to
maximize the head of slurry and avoid contamination of concrete during
placing of soil fall-ins. Shafts formed in this way are typically 2 to 3 m in
diameter and up to 70 to 80 m deep, but much larger diameter and deeper
shafts for civil engineering works have been successfully completed using
purpose-made reverse circulation equipment.
Shafts and caissons: construction and design 453

Soil and rock support


The periphery of the shaft may require no lining in dry, sound rock, but other-
wise the following methods are available for soil or rock support:
. in soils – timbering; steel sheet piling; precast concrete, cast iron and
pressed, welded steel and segmental linings; diaphragm walling; secant
piling
. in rock – in situ concrete lining; shotcreting; rock dowels and rock
bolting.

Fig. 10.1. Details of


connections to one-pass shaft
linings (courtesy of Charcon)
454 Deep excavations

Each of these methods is now discussed in turn.


(a) Timbering. In soft ground, vertical poling boards are driven ahead of
excavation and secured by timber walings, strutting and diagonal
corner braces. Construction depths are limited with timbering, and
the works are labour-intensive by modern standards.
(b) Sheet piling. Deeper shafts may be sheet piled, the piles being secured by
walings in timber, steel or reinforced concrete. Ground anchors may be
used where space outside the shaft allows and space within is confined.
Sheet piling may be impeded and pile clutches broken in hard driving
caused by boulders or penetration into rockhead. Grouting may be
necessary in such cases to avoid ingress of groundwater into the shaft.
(c) Segmental linings. Where space and ground conditions allow, the first
ring is built a small depth below ground level supported, if the shaft
is large, by an external collar to avoid differential settlement around
the ring. In dry, sound soil, rings are built successively downwards in
an underpinning operation for each ring. After completion of the ring
the annular space behind the excavated soil face and the outer segment
face is grouted, thus avoiding later loss of ground and subsidence of the
ground surface. Where soil conditions are less favourable it may not be
possible to complete a whole ring without temporary support for each
segment from the central dumpling. In tunnel works segments are in
cast-iron or precast concrete. Solid reinforced concrete segments,
known as one-pass shaft linings, are available and avoid further in
situ lining works. These segments use stressed loop cross-joint connec-
tors (details in Fig. 10.1) and allow the introduction of precast corbels
for structural support for landings by bolting to the main lining (Fig.
10.2). Examples of conventional concrete segment and one pass linings
are shown in Figs. 10.3 and 10.4. Steel liner plates will themselves be
adequate in small diameter shafts, but in larger diameter shafts a
curved steel joist is set for every two or three courses of liner.
(d ) Diaphragm walling. In diaphragm wall shafts a segmental plan shape is
used and structural integrity is necessary at each panel joint to transfer
hoop compression between panels and retain structural stability.
Depths of the order of 30 to 40 m are not unusual for diaphragm wall
shafts, and the increasing use of reverse circulation cutter rigs, such

Fig. 10.2. Example of single-pass Fig. 10.3. Conventional bolted


shaft lining with corbels (courtesy precast concrete segment lining
of Charcon) (courtesy of Charcon)
Shafts and caissons: construction and design 455

Fig. 10.4. Single-pass lining to


line tunnel break-out shaft
(courtesy of Charcon)

as the Hydrofraise and Trenchcutter rigs, bring improved verticality


tolerance to the works and even greater depths. Figure 10.5 shows
the use of polygonal plan shaped shafts in Hong Kong for deep
access shafts through decomposed granite, constructed by Bachy-
Soletanche.
The construction of deep shafts using grab excavation methods was
originally made practicable by installing rotary bored piles at the
junction of each diaphragm wall panel. The verticality tolerance of
pile installation using special equipment was of the order of 1 in 300
or better. The piles were installed before the diaphragm wall panels
and acted as a guide to the panel grab, which was shaped to cut soil
from the curved pile face. The Icos company used this procedure of
interlocked circular units and rectangular panels to build a 4.6 m dia.
Fig. 10.5. Polygonal plan shape shaft through fine sands to the very considerable depth of 72 m for
shaft construction using mineworks at Speckholzerheide, Holland as long ago as 1954.
diaphragm walls, Eastern Harbour Improved excavation tolerances achievable with reverse circulation
Crossing, Hong Kong (courtesy of equipment such as the Hydrofraise and Trenchcutter rigs now allow
Soletanche–Bachy)
deep shaft construction without the need for bored piles at the junction
of each panel. Shafts in excess of 50 m deep can be constructed with this
type of equipment, the junction being achieved by cutting the end of the
primary panel during excavation of the adjacent secondary panel. Cut
joints obviate the need for temporary stop ends in larger diameter
shafts.
For a typical small-diameter shaft excavated by grab with an internal
diameter of 6 m or so one primary panel is installed and work proceeds
successively on two fronts to complete the circle as shown in Fig. 10.6.
Verticality of the face of the initial primary panel needs to be measured
as excavation proceeds and care should be taken to avoid poor vertical-
ity and overdig at the end of the panel which would allow concrete to
pass around the stop end tube. With shafts of larger internal diameter,
456 Deep excavations

Fig. 10.6. Small-diameter shaft:


a panel configuration

say 10 m or more, two primary starter panels are often preferred to


reduce rig standing time. These starters are located opposite each
other and it is necessary that only one panel should be open at any
one time to ensure stability. Figure 10.7 shows a typical shaft with
primary panels within the flat wall sections of the polygonal plan
shape. This arrangement avoids the difficulty of cleaning into the
curved concrete surface of the panel joint by a grab excavating the
adjacent panel.
It is essential to avoid chiselling in shaft panels as much as possible
in order to reduce overbreak. Where panels are founded on sloping
rockhead it may be practicable to use multiple reinforcement cages
with a stepped panel toe to avoid excessive chiselling into bedrock.
Individual cages for each bite become necessary with depths greater
than 30 to 40 m.
For deep panels reverse circulation cutter rigs prove their value,
especially in terms of verticality. Figure 10.8 shows a typical shaft con-
struction in plan using alternative primary and secondary panels with
cut joints to a depth of 60 m through marine clays with a high water
table onto granite bedrock. Each primary panel concrete volume
exceeded 400 m3 .
Some risk of non-verticality to panels exists during the cutting of
joints with cutter rigs when the plan deviation between adjacent
panels exceeds 12 to 158.
(e) Secant piling. The use of secant piling as a means of shaft construction is
limited to depths of the order of 20 m with CFA rigs. Deeper excavation
Shafts and caissons: construction and design 457

Fig. 10.7. 10 m dia. shaft: panel


configuration with two starter
panels, panels located within
flat wall sections

(of the order of 40 m or more) can be made by secant piling methods


using heavy rotary equipment, powerful casing oscillations and tempor-
ary, rigid twin-wall casings. The depth of excavation is limited by the
accuracy of installing the temporary casing to maintain the depth of
the secant cut with the male secondary piles. A rotary rig with casing
oscillator working immediately adjacent to an existing structure is
shown in Fig. 10.9.
( f ) In situ concrete lining. In rock excavation where the rock will stand
temporarily without support in dry conditions, an in situ lining of
mesh-reinforced concrete can be used. Excavation is by drilling and
blasting in successive increments of 2 m or more, with concrete pours
6 and 8 m high. Slip-forming methods can be applied in deep shafts.
The use of in situ lining overcomes any difficulty in varying amounts
of overbreak that occur when using segmental linings. Thermal
insulation is necessary when concrete is placed against frozen soil or
rock.
(g) Shotcreting. In shafts through rock which is relatively stable and dry,
sprayed concrete, or shotcrete, may be used to stabilize the face.
Applied in layers typically 50 to 60 mm thick, light mesh reinforcement
may be pinned to the rock face6;7 .
458 Deep excavations

Fig. 10.8. (a) Plan of deep shaft,


10 m minimum internal
diameter, two starter panels

(h) Rock dowels and bolts. Drilled radially from the shaft, dowels and bolts
are used as rock reinforcement, and mesh secured to the face of the
dowel or bolt may be used to secure the rock face. CIRIA report
1018 describes rock reinforcement in underground excavations. Rock
dowels may be cement- or resin-grouted or may consist of a hollow,
high-strength steel tube with a longitudinal slit which compresses
radially when driven into a drillhole exerting continuous outward
pressure. Another type of dowel uses high-pressure water to expand a
steel tube within a drilled hole. Rock bolts are tensioned elements of
ultimate capacity in the range 80 to 300 kN. Tensioning is by jack or
torque wrench. Several proprietary types are available: bolts in various
high-strength steels; more recently, in solid and tubular fibreglass bolts
have been developed. Grout may be cement or resin, pumped or
encapsulated. Mechanical anchors, also grouted for corrosion protec-
tion, are designed for use in hard rocks. The Sure type is an exception,
which was specifically developed for soft rocks. Mechanical types are
simply driven split-and-wedge types or wedge expansion types such
as the duplex anchor. Types of proprietary rock bolts are shown in
Fig. 10.10.
Shafts and caissons: construction and design 459

Fig. 10.8 (b) As-built survey of


inner face of shaft with cut joints
at 60 m depth. Note
displacement of panel P2

Fig. 10.9. Piled wall


construction using a rotary rig
with a casing oscillator in close
proximity to an existing
structure
460 Deep excavations

Fig. 10.10. Types of proprietary


rock bolts and anchors used in
shaft construction (courtesy of
MAI Systems)

Methods of design for reinforcement of openings in rock are reviewed by


Choquet and Hadjigeorgiou9 .

Groundwater control in shafts


Methods of groundwater control were described in detail in Chapter 2. In
summary, the methods used in shafts are:
. pumping from sumps in advance of the excavation
. pumping from wellpoints
. pumping from deep wells below or outside the shaft
Shafts and caissons: construction and design 461

. grouting of soils and fissures in rock with cement or chemical grouting


. freezing
. compressed air working.
For relatively small shafts the first five items are listed in approximately
increasing order of cost. Where compressed air working is needed to drive
water from the shaft, an air deck, man and muck lock and compressor instal-
lation will be needed. If planned for, this mobilization will fit into the works
programme, but if the decision to use air is made during shaft excavation it
could be expensive, causing delay and disruption. As described earlier, the
depth of compressed air working below water level or ground level is restricted
to maximum working air pressures. Groundwater lowering, often by deep
wells, may be used to effectively increase this depth. Where compressed air
working is necessary and segmental linings are used, all segment joints are
caulked. Where cast-iron segments are used, kentledge may be necessary
above the air deck to alleviate the low tensile strength of the cast-iron;
grouting behind the tubbing reduces risk of circumferential tension within
the iron.

Caissons The maximum practical depth of excavation within cofferdams is of the order
of 25 m. To reach founding levels at greater depths, previous generations of
engineers have used caissons – either open-well or pneumatic types. Histori-
cally, well caissons have been used for bridge foundations in India, Burma
and Egypt, using masonry or brick for the caisson walls. Prior to Victorian
times these wells were sunk by hand-excavation by divers. The working
depth limit was therefore of the order of 6 m or so below water level. It was
the British engineer of the late nineteenth century who introduced excavation
by grab and sand pump to allow well foundations to much greater depths. The
caisson for the Hardinge Bridge on the Lower Ganges reached depths between
32 and 36 m below river level in the late 1870s.
The pneumatic caisson was first used by John Wright in 1851 in foundations
to the bridge over the River Medway at Rochester. Brunel used the method on
the Saltash Bridge foundations some years later, and pneumatic caissons were
used by James B. Eads in 1869 on the St Louis Bridge over the Mississippi

Fig. 10.11. Station floor


construction, St Michel station,
Paris Metro10
462 Deep excavations

River. Eads sank two river piers under air to depths of 26 to 28 m below water
level.
Early use of open and pneumatic caissons for metro construction was made
at the beginning of the twentieth century for construction of line 4 of the Paris
Metro. The construction and sinking of the caissons for the crossing of the
Seine and two stations, Cité and St Michel was made by the contracting
firm, Chagnaud, in the years 1905 to 190710 .
The station caissons were much bigger than the pneumatic caissons used in
the river. St Michel station is under Rue Danton between Place Saint-Michel
and Place Saint-André des Arts. Its 118 m length was made up of three steel
caissons, a central unit 66 m long was curved with a radius of 300 m containing
the platforms, and one caisson on either side of it, each 26 m long, containing
the approaches.
The first construction operation was the excavation of a pit 3 m deep at the
station site into which the caisson was built. Figure 10.11 shows the construc-
tion of the station floor in the Rue Danton. Figure 10.12 shows the station
wall structure (concrete infill is placed between the steel uprights). Figure
10.13 shows the two completed caissons of the three in Saint Michel station.
The elliptical caisson sunk at St Michel is in the background. The second
elliptical caisson at Place Saint-André des Arts in the front cannot be seen.
The third photo gives a good idea of the impressive dimensions of this
structure which was subsequently sunk by hand-excavation, with great care
to avoid damage to neighbouring buildings.
The caissons at St Michel station are shown at progressive excavation
stages in Figs. 10.14 and 10.15. The first photograph shows the caissons,
some 11 000 tonnes of steel and concrete ready for sinking. The caissons
were sunk, one at a time, to a depth of 20 m below ground level, approxi-
mately 13 m below the Seine water level. The second photograph, Fig.
10.15, shows the Place Saint-Michel elliptical caisson and the central caisson
during sinking. The other elliptical caisson, still under construction, is in the
background.
Two other types of caisson, which are outside the scope of this book, are the
box caisson and caissons used to form buoyant foundations. The box caisson,
closed at the bottom and open at the top, has been used for quay and break-
water construction at several ports. Little11 described its use at Rotterdam,
where the box caissons, constructed ashore and towed to their location,
were founded on prepared sand beds. Similar construction has been used
extensively on the River Liffey at the Port of Dublin. These works were
described by O’Sullivan12 ; Bruun13 described caisson quay wall construction
in Gdynia, Poland, and at Sheibah, Kuwait, using similar techniques. The
method is susceptible to under-scour by current or the erosive action of
ship propellers; anti-scour mattresses are sometimes necessary.
A similar technique of floated-in box caissons has found economical use for
bridge foundations. The caissons, constructed in dry dock, are towed to site
and sunk on prepared foundation soils. The method was used in 1991 for
the Queen Elizabeth II Bridge at Dartford in the UK.
Caissons are used to form buoyant foundations to building structures
where ground conditions, such as very soft estuarine clays, are inadequate
to support the building loads without excessive settlements. An example of
this method, designed and built by Soil Mechanics Ltd for sugar mill and
preparation buildings in Guyana, was described by Golder14 . Precast post-
tensioned concrete walls were used for these caissons.
Reference must also be made to the use of large caisson structures for off-
shore foundations described by Hansen and Frode15 . Typical applications
include foundations for lighthouse structures. Caisson techniques also find
Shafts and caissons: construction and design
463

Fig. 10.12. Station wall structure, St Michel station, Paris Metro10


464
Deep excavations

Fig. 10.13. Two completed caissons at St Michel station, Paris Metro10


Shafts and caissons: construction and design

Fig. 10.14. Caissons at St Michel station, Paris Metro10


465
466
Deep excavations

Fig. 10.15. Caissons at St Michel station, during sinking10


Shafts and caissons: construction and design 467

application for off-shore oil exploration and storage structures. These


matters, however, also fall outside the scope of this book.

Open-well caissons
An open-well caisson consists, in simple terms, of a box open at both top and
bottom.
The cutting edge of the caisson is situated at the underside of the curb.
Where this is constructed in steelwork, the curb has vertical steel outer skin
plates and inner steel cant, or haunch, plates. After the beginning of sinking,
concrete is filled between the skin plates. This concrete is termed ‘steining’. At
the height where the caisson has gained sufficient strength from the composite
concrete–steelwork, the caisson walls are frequently cantilevered upwards in
reinforced concrete. More recently, reinforced concrete has become an attrac-
tive alternative to steelwork in caisson construction as methods of shuttering,
slip-forming and concrete placement have developed. The reinforced concrete
walls, canted internally, may be lined with steel plate for protection above the
steel cutting edge.
Details of steelwork and the cutting edges to the main caisson piers of the
Howrah Bridge, Calcutta, are shown in Fig. 10.16. The steel caissons, built in
1937–1938, were designed so that the working chambers within the shafts
could be temporarily enclosed by steel diaphragms to allow work under com-
pressed air if required. In the event, the main piers on the Howrah side were
sunk by open-well dredging, and only those on the Calcutta side required
compressed air to counter running sand.
The sinking of the Howrah caissons were described by Howarth and
Smith16 . The plan size of the main caisson was 55.3 m by 24.9 m, with 21
shafts each 6.25 m square. The two anchorage caissons were each 16.4 m by
8.2 m with two wells 4.9 m square. The Howrah side work site is shown in
Fig. 10.17. The caissons were sunk through soft river deposits to a stiff
yellow clay 26.5 m below ground level. After penetrating 2.1 m into this clay
all the shafts were plugged with concrete after individual dewatering with
some 5 m of backfilling in adjacent shafts.
The accuracy of sinking of these large caissons was exceptionally good,
within 50 to 75 mm of true plan position. Excavation was efficient, each
shaft being of ample size with long, straight cutting edges and thin walls
near the cutting edge to allow the grab to work close to it. In very soft soil
a number of shafts, symmetrical to the caisson axes, were left unexcavated
to allow strict control, while in very stiff clays a large number of the internal
walls were completely undercut allowing the whole weight of the caisson to
be carried by the outside skin friction and the bearing under the external
walls. Skin friction on the outside of the monolith walls was estimated at
29 kN/m2 while loads on the cutting edge in clay overlying the founding
stratum reached 100 tonnes/m.
General rules to help prevent initial tipping of large open caissons were
given by Purcell et al.17 as follows.
(a) Keep the weight of the caisson and its centre of gravity as low as
possible by keeping the interior concrete below the waterline during
the dangerous stages.
(b) Provide lateral support at as high a level as possible.
(c) Maintain symmetry at the plane of the cutting edge. The removal of
false bottoms and dredging in the wells should generally be such as to
preserve symmetrical conditions, particularly along the long axis.
(d ) Avoid sudden movement of material. Dredge so as to remove mud from
the centre, maintaining bearing on the cutting edges.
468 Deep excavations

Fig. 10.16. Steelwork and


cutting edges to main caisson
piers, Howrah Bridge,
Calcutta46
Shafts and caissons: construction and design 469

Fig. 10.17. Howrah Bridge work site16

(e) Do not dredge below the exterior cutting edges until the caisson is
founded. If the caisson will not settle satisfactorily, use jets on the
side to reduce skin friction.
( f ) In a large, rectangular caisson, sinking may be accomplished by rocking
slightly about the short axis, dredging first from one end and then the
other. This must never be done about the other axis.
(g) In unstable soils, avoid any action that might tend to make the soil
under the cutting edge quick, particularly the use of blasting and pump-
ing down the dredging wells. Jetting on the sides of the caisson, and in
the dredge wells if mud plugs tend to form, is usually safe and desirable.
Tomlinson18 pointed out that caisson proportions, while usually defined in
terms of plan shape by the superstructure, are often a compromise. On the
one hand, the walls are required to be thick to provide maximum weight for
sinking, and on the other, they should be thin to allow the grab to work
near the cutting edge. Lightness of weight is desirable to allow a shallow
draught during towage from casting yard to mid-river site but this, in turn,
is to the detriment of the caisson rigidity, which is essential during early
stages of sinking when the caisson may be unevenly supported.
Tomlinson pointed out that the size and layout of the dredging wells
depends mainly on the soil type. Dense sands and firm to stiff clays require
a minimum number of cross-walls and a minimum outer wall thickness
consistent with the weight requirements for sinking and rigidity against
distortion. The cross-walls need not extend to cutting edge depth. In sands
and soft silts, on the other hand, grabbing below cutting edge level causes
soil to move towards the centre of the shaft where the excavation is kept
low. Water and air/water jets may be used for excavation and lubrication
purposes. The use of bentonite slurry lubrication for caisson sinking will be
referred to later.
470 Deep excavations

The use of deep wells, open or pneumatic caissons of relatively small plan
size for bridge foundations in rivers is a traditional method in India and
Pakistan, and remains currently in use. The following extract from the
Indian standard specification for road bridges19 gives a useful review of pre-
cautions to be taken during the sinking of the small wells and basic design
criteria for walls built in concrete and brickwork.

710 WELL FOUNDATIONS


710.1 General
While selecting the shape, size and the type of wells for a bridge, the
size of pier to be accommodated, need for effecting streamline flow,
the possibility of the use of pneumatic sinking, the anticipated depth
of foundation, and the nature of strata to be penetrated should be
kept in view. Further for the type of well selected, the dredge hole
should be large enough to permit easy dredging, the minimum dimen-
sion being not less than 2 m. In case there is deep standing water,
properly designed floating caissons may have to be used.

710.2 Steining
710.2.1 General
The thickness of the steining shall be fixed from the following consid-
erations:
(i) It should be possible to sink the well without excessive kentledge.
(ii) The wells shall not get damaged during sinking.
(iii) If the well develops tilts and shifts during sinking, it should be
possible to rectify the tilts and shifts within permissible limits
without damaging the well.
(iv) The well should be able to resist safely earth pressure developed
during a sand blow or during other conditions like sudden drop
that may be experienced during sinking.
(v) Stresses at various levels of the steining should be within permis-
sible limits under all conditions for loads that may be transferred
to the well either during sinking or during service.

710.2.2 Design considerations


710.2.2.1 Use of cellular steining with two or more shells or use of
composite material in well steining shall not be permitted.
710.2.2.2 In case of plain and reinforced concrete single circular
walls, the external diameter of well shall not normally exceed 12 m.
Note: Wells of larger dimensions shall call for supplemental design
and construction specifications outside the purview of the code.
710.2.2.3 In case of plain cement concrete wells, the concrete mix for
the steining shall not be leaner than 1:3:6. However, in areas
subjected to marine or other similar conditions of adverse exposure,
the concrete in the steining shall not be leaner than M160 mix with
cement content not less than 350 kg per m3 of concrete and with
water cement ratio not more than 0.45.
710.2.2.4 The external diameter of brick masonry wells shall not
exceed 6 m. Brick masonry wells for depth greater than 20 m shall
not be permitted.
Shafts and caissons: construction and design 471

710.2.2.5 For brick masonry wells bricks not less than 70A shall be
used in cement mortar not leaner than 1:3 for the steining.

710.2.3 Thickness of well steining


710.2.3.1 The minimum thickness of well steining shall not be less
than 500 mm and satisfy the following relationship:
pffi
h ¼ Kd l
where
h ¼ Minimum thickness of steining in m.
d ¼ External diameter of circular well or dumb bell shaped well in
the case twin D wells, the smaller dimension in plan in metres.
l ¼ Depth of well in metres below L.W.L. or ground level whichever
is higher.
K ¼ a constant.
The value of K shall be as follows.
(i) Single Circular or dumb bell shaped well in Cement Concrete
K ¼ 0:030 for predominantly sandy strata.
K ¼ 0:033 for predominantly clayey strata.
(ii) Twin D wells in Cement Concrete
K ¼ 0:039 for predominantly sandy strata.
K ¼ 0:043 for predominantly clayey strata.
(iii) Single Circular or dumb bell shaped wells in Brick Masonry
K ¼ 0:047 for predominantly sandy strata.
K ¼ 0:052 for predominantly clayey strata.
(iv) Twin D wells in Brick Masonry
K ¼ 0:062 for predominantly sandy strata.
K ¼ 0:068 for predominantly clayey strata.
Note: (i) For boulder strata or for wells resting on rock where
blasting may be involved, higher thickness of steining,
better grade of concrete, higher reinforcement, use of
steel plates in the lower portions etc. may be adopted as
directed by the Engineer-In-Charge.
Note: (ii) For wells passing through very soft clayey strata the stein-
ing thickness may be reduced based on local experience
and in accordance with the decision of the Engineer-In-
Charge, to prevent the well penetrating by its own weight.
In such cases, the steining may be adequately reinforced
to get sufficient strength.
710.2.3.2 Where nominal steel is provided in the steining as per
clause 710.2.4.1 the same shall not be considered in the design for
strength.

710.2.4 Reinforcement in well steining


710.2.4.1 For plain concrete wells, vertical reinforcements (either
mild steel or deformed bars) in the steining shall not be less than
0.12% of gross sectional area of the actual thickness provided. This
shall be equally distributed on both faces of the steining (see Fig. 1).
The vertical reinforcements shall be tied up with hoop steel not less
than 0.04 per cent of the volume per unit length of the steining.
472 Deep excavations

Fig. 1.
Shafts and caissons: construction and design 473

710.2.4.2 In case where the well steining is designed as a reinforced


concrete element, it shall be considered as a column section sub-
jected to combined axial load and bending. However, the amount of
vertical reinforcement provided in the steining shall not be less than
0.2% (for either mild steel or deformed bars) of the actual gross
sectional area of the steining. On the inner face a minimum of
0.06% (of gross area) steel shall be provided. The transverse re-
inforcement in the steining shall be provided in accordance with the
provisions for a column but in no case shall be less than 0.04% of
the volume per unit length of the steining.
710.2.4.3 The vertical bond rods in brick masonry steining shall not
be less than 0.1% of the cross sectional area and shall be encased
into cement concrete of 1: M160 mix of size 150 mm  150 mm. These
rods shall be equally distributed along the circumference in the
middle of the steining and shall be tied up with hoop steel not less
than 0.04% of the volume per unit length of the steining. The hoop
steel shall be provided in a concrete band at spacing of 4 times the
thickness of the steining or 3 m, whichever is less. The horizontal
RCC bands shall not be less than 300 mm wide and 150 mm high,
reinforced with bars of diameter not less than 10 mm placed at the
corners and tied with 6 mm diameter stirrups at 300 mm centres (see
Fig. 2).

710.2.5 Stresses
710.2.5.1 Allowable stresses for different materials used in the
construction of wells shall conform to the relevant sections of the
IRC Code of Practice for Road Bridges.
710.2.5.2 For piers/abutment fully spanning the well, the steining
immediately below it, for a width equal to thickness of pier/abutment
after allowing for dispersion shall be checked for any tangential
stresses developing in the steining due to the concentration of the
load.
710.2.5.3 The stresses in well steining shall be checked at such
critical sections where tensile and compressive stresses are likely
to be maximum and also where there is change in the area of
reinforcement or in the concrete mix.

710.3 Stability of Well Foundations


710.3.1 The stability of well foundations shall be analysed under
the most critical combination of loads and forces as per clause 706
allowing for the soil resistance from the sides below scour level by
any rational method. The pressure on the foundations shall satisfy
the provisions of clause 708.
710.3.2 The stability of the well shall also be checked for the con-
struction stage for following conditions:
(i) In case of pier wells resting on rock without bottom plug and
carrying no superstructure, the well being subjected to full
pressure due to water current at full design scour.
(ii) Completed abutment well resting on rock and carrying no super-
structure. The well being subjected to full designed differential
earth pressure.
474 Deep excavations

Fig. 2.
Shafts and caissons: construction and design 475

Note:
(a) The seismic force shall not be considered in the above two cases.
(b) In case it is not stable against overturning and sliding, the safety
of the well shall be ensured by suitable methods.
710.3.3 Use of IRC:45 for the design of pier well foundations in
cohesionless soil is acceptable. For design of abutment wells in all
types of soils and pier wells in cohesive soil guidance may be taken
from Appendix 4.
710.3.4 The side earth resistance shall be ignored in case of well
foundations resting on rock.
Note: If allowable bearing pressure less than 1 MPa is assigned to the
rock strata, the side earth resistance may be taken into account
subject to the satisfaction of clause 708.2.1.
710.3.5 If the abutments are designed to retain earth without spillage
in front (solid abutment with solid returns) the foundations shall be
designed to withstand the earth pressure and horizontal forces for
the condition of maximum scour depth in front of 1.27 dsm with
approach retained and 2 dsm with scour all around.
710.3.6 However, where earth spilling from the approaches is
reliably protected in front, relief due to the spilling earth in front
may be considered, assuming that the designed scour extends up to
the toe of the protection after launching. However, no passive relief
shall be taken for the triangular portion of the earth in front of the
abutment up to bottom of well cap, except for its surcharge effect.

710.4 Tilts and Shifts


As far as possible the wells shall be sunk plumb without any tilts or
shifts. However, a tilt of 1 in 80 and a shift of 150 mm in a resultant
direction shall be considered in the design of well foundations.
Even after resorting to precautionary/remedial measures during
execution, if tilts or shifts or both exceed the above limits, their effects
on bearing pressure, steining stress, change in span length and shift in
the centre line of the bridge shall be examined individually or jointly.

710.5 Cutting Edge


The mild steel cutting edge shall be strong enough and not less than
40 kg/m to facilitate sinking of the well through the types of strata
expected to be encountered without suffering any damage. It shall
be properly anchored to the well kerb.
When there are two or more compartments in a well, the lower end of
the cutting edge of the middle stems of such wells shall be kept about
300 mm above that of the outer wells to prevent rocking as shown in
Fig. 1.

710.6 Well Curb


The well curb shall satisfy the following requirements:
(i) It should have a shape offering the minimum resistance while the
well is being sunk.
(ii) Be strong enough to be able to transmit super-imposed loads
from the steining to the bottom plug.
476 Deep excavations

To satisfy the above requirements, the shape and the outline


dimensions of the curb shall be as given in Fig. 1. The curb
shall invariably be reinforced concrete of mix not leaner than
M200: with minimum reinforcement of 72 kg per cu.m. excluding
bond rods. This quantity of steel shall be suitably arranged as
shown in Fig. 1 to prevent spreading and splitting of the curb
during sinking and in service.
(iii) In case pneumatic sinking is indicated, the internal angle of the
well curb shall be made steep enough to provide easy access
for the pneumatic tools.
(iv) In case blasting is anticipated, the outer faces of the curb shall be
protected with suitable steel plates of thickness not less than
6 mm up to half the height of the well curb on the outside and
on the inner face not less than 10 mm thick up to top of well
curb, suitably reduced to 6 mm to a height of 3 m above the top
of the curb. The steel plates shall be properly anchored to the
curb and steining. The curb in such a case shall be provided
with additional hoop reinforcement of 10 dia. mild steel or
deformed bars at 150 mm centres. The latter reinforcement
shall also extend up to a height of 3 m into the well steining
above the curb, in which portion the mix of concrete in the well
steining shall not be leaner than 1:1:3.

710.7 Bottom Plug


The bottom plug shall be provided in all wells and the top shall be kept
not lower than 300 mm in the centre above the top of the curb (see
Fig. 1). A suitable sump shall be below the level of the culting edge.
Before concreting the bottom plug it shall be ensured that its inside
faces have been cleaned thoroughly.
The concrete mix used in bottom plug shall have a minimum cement
content of 330 kg/m3 and a slump of about 150 mm to permit easy
flow of concrete through tremie to fill up all cavities. Concrete shall
be laid in one continuous operation till dredge hole is filled to required
height. The concrete shall be placed gently by tremie or skip boxes
under still water conditions.
In case grouted concrete e.g. colcrete is used, the grout mix shall not
be leaner than 1:2 and it shall be ensured by suitable means such as
controlling the rate of pumping that the grout fills up all inter-stices up
to the top of the plug.

710.8 Filling the Well


The filling of the well, if considered necessary, above the bottom plug
shall be done with sand or excavated material free from organic matter.

710.9 Plug over Filling


A 500 mm thick plug of 1:3:6 cement concrete shall be provided over
the filling.

710.10 Well Cap


710.10.1 The bottom of well cap shall preferably be laid as low as
possible taking into account the L.W.L.
710.10.2 As many longitudinal bars as possible coming from the well
steining shall be anchored into the well cap.
Shafts and caissons: construction and design 477

710.10.3 The design of the well cap shall be based on any accepted
rational method, considering the worst combination of loads and
forces as per clause 706.

710.11 Floating Caissons


710.11.1 Floating caissons may be of steel, reinforced concrete or
any suitable material. They should have at least 1.5 m free board
above the water level and increased, if considered necessary, in
case there is a possibility of caissons sinking suddenly owing to
reasons such as scour likely to result from lowering of caissons,
effect of waves, sinking in very soft strata etc.
710.11.2 They should be checked for stability against overturning
and capsizing while being towed and during sinking for the action of
water current, wave pressure, wind, etc.
710.11.3 The floating caisson shall not be considered as part of
foundation unless proper shear transfer at the interface is assured.

710.12 Sinking of Wells


The well shall as far as possible be sunk true and vertical through all
types of soils. Sinking should not be started till the steining has been
cured for at least 48 hours. A complete record of sinking operations
including tilt and shifts, kentledge, dewatering, blasting etc. done
during sinking shall be maintained at site. For safe sinking of wells
necessary guidance may be taken from the precautions as given in
Appendix 5.

710.13 Pneumatic Sinking of Wells


710.13.1 Where sub-surface data indicate the need for pneumatic
sinking, it will be necessary to decide the method and location of
pneumatic equipment and its supporting adaptor.
710.13.2 In cases where concrete steining is provided, it shall be
rendered air tight by restricting the tension in concrete which shall not
exceed 3/8 of the modulus of rupture. For the circular wells, the tension
in steining may be evaluated by assuming it to be a thick walled cylinder.
710.13.3 The steining shall be checked at different sections for any
possible rupture against the uplift force and, if necessary, shall be
adequately strengthened.

710.14 Sinking of Wells by Resorting to Blasting

PRECAUTIONS TO BE TAKEN DURING SINKING OF WELLS


1. Construction of Well Curb and Steining
1.1 Cutting edge and the top of the well curb shall be placed truly
horizontal.
1.2 The methods adopted for placing of the well curb shall depend on
the site conditions, and the cutting edge shall be placed on dry
bed.
1.3 Well steining shall be built in lifts and the first lift shall be laid
after sinking the curb at least partially for stability.
1.4 The steining shall be built in one straight line from bottom to top
and shall always be at right angle to the plane of the curb. In no
478 Deep excavations

case it shall be built plumb at intermediate stages when the well


is tilted.
1.5 In soft strata prone to settlement/creep the construction of
the abutment wells shall be taken up after the approach embank-
ment for a sufficient distance near the abutment has been
completed.

2. Sinking
2.1 A sinking history record be maintained at site.
2.2 Efforts shall be made to sink wells true to position and in plumb.
2.3 Sumps made by dredging below cutting edge shall preferably not
be more than half the internal diameter.
2.4 Boring chart shall be referred to constantly during sinking for
taking adequate care while piercing different types of strata by
keeping the boring chart at the site and plotting the soil as
obtained for the well steining and comparing it with earlier
bore data to take prompt decisions.
2.5 When the wells have to be sunk close to each other and the
clear distance is less than the diameter of the wells, they shall
normally be sunk in such a manner that the difference in the
levels of the sump and the cutting edge in the two wells do not
exceed half the clear gap between them (Fig. 3).
2.6 When groups of wells are near each other, special care is needed
that they do not foul in the course of sinking and also do not cause
disturbance to wells already sunk. The minimum clearance
between the wells shall be half the external diameter. Simulta-
neous and even dredging shall be carried out in the dredging
holes of all the wells in the group and plugging of all the wells
be done together.
2.7 During construction partially sunk wells shall be taken to a safe
depth below the anticipated scour levels to ensure their safety
during ensuing floods.
2.8 Dredged material shall not be deposited unevenly around the
well.

Fig. 3.
Shafts and caissons: construction and design 479

3. Use of Kentledge
3.1 Where a well is loaded with kentledge to provide additional
sinking effort, such load shall be placed evenly on the loading
platform, leaving sufficient space in the middle to remove exca-
vated material.
3.2 Where tilts are present or there is a danger of well developing a
tilt, the position of the load shall be regulated in such a manner as
to provide greater sinking effort on the higher side of the well.

4. Sand Blows in Wells


4.1 Dewatering shall be avoided if sand blows are expected. Any
equipment and men working inside the well shall be brought
out of the well as soon as there are any indications of a sand
blow.
4.2 Sand blowing in wells can often be minimised by keeping the
level of water inside the well higher than the water table and
also by adding heavy kentledge.

5. Sinking of Wells with Use of Divers


5.1 Use of divers may be made in well sinking both for sinking pur-
poses like removal of obstructions, rock blasting etc. as also
for inspection. All safety precautions shall be taken as per any
acceptable safety code for sinking with divers or any statutory
regulations in force.
5.2 Only persons trained for the diving operation shall be employed.
They shall work under expert supervision. The diving and other
equipments shall be of an acceptable standard. It shall be well
maintained for safe use.
5.3 Arrangement for ample supply of low pressure clean cool air
shall be ensured through an armoured flexible hose pipe.
Stand by compressor plant will have to be provided in case of
breakdown.
5.4 Separate high pressure connection for use of pneumatic tools
shall be made. Electric lights, where provided, shall be at 50
volts (maximum). The raising of the diver from the bottom of
wells shall be controlled so that the decompression rate for
divers conforms to the appropriate rate as laid down in the
regulation.
5.5 All men employed for diving purposes shall be certified to be fit
for diving by an approved doctor.

6. Blasting
6.1 Only light charges shall be used under ordinary circumstances
and should be fired under water well below the cutting edge so
that there is no chance of the curb being damaged.
6.2 There shall be no equipment inside the well nor shall there be
any labour in the close vicinity of the well at the time of exploding
the charges.
6.3 All safety precautions shall be taken as per IS: 4081 ‘Safety Code
for Blasting and related Drilling Operations’, to the extent applic-
able, whenever blasting is resorted to. Use of large charges,
480 Deep excavations

0.7 kg or above, may not be allowed except under expert direc-


tion and with permission from Engineer-In-Charge. Suitable
pattern of charges may be arranged with delay detonators to
reduce the number of charges fired at a time. The burden of
the charge may be limited to 1 m and the spacing of holes may
normally be kept at 0.5 and 0.6 m.
6.4 If rock blasting is to be done for seating of the well, the damage
caused by the flying debris should be minimised by provisions of
rubber mats covered over the blasting holes before blasting.
6.5 After blasting, the steining shall be examined for any cracks and
corrective measures shall be taken immediately.

7. Pneumatic Sinking
7.1 The pneumatic sinking plant and other allied machinery shall not
only be of proper design and make, but also be worked by com-
petent and well trained personnel. Every part of the machinery
and its fixtures shall be minutely examined before installation
and use. Appropriate spares, standbys, safety of personnel as
recommended in the IS Code 4188 for working in compressed
air must be kept at site. Safety code for working in compressed
air and other labour laws and practices prevalent in the country,
as specified to provide safe, efficient and expeditious sinking
shall be followed.
7.2 Inflammable materials shall not be taken into air locks and
smoking shall be prohibited.
7.3 Whenever gases are suspected to be issuing out of dredge
hole, the same shall be analysed by trained personnel and
necessary precautions adopted to avoid hazard to life and
equipment.
7.4 Where blasting is resorted to, it shall be carefully controlled and
all precautions regarding blasting shall be observed. Workers
shall be allowed inside after blasting only when a competent
and qualified person has examined the chamber and steining
thoroughly.
7.5 The weight of pneumatic platform and that of steining and
kentledge, if any, shall be sufficient to resist the uplift from air
inside, skin friction being neglected in this case.
7.6 If at any section the total weight acting downwards is less than
the uplift pressure of air inside, additional kentledge shall be
placed on the well.
7.7 If it is not possible to make the well heavy enough during excava-
tion, ‘blowing down’ may be used. The men should be withdrawn
and the air pressure reduced. The well should then begin to
move with a small reduction in air pressure. ‘Blowing down’
should only be used where the ground is such that it will not
heave up inside the chamber when the pressure is reduced.
When the well does not move with a reduction in air pressure,
kentledge should be added. Blowing down should be in short
stages and the drop should not exceed 0.5 m of any stage. To
control sinking during blowing down, use of packs or packings
may be made.
Shafts and caissons: construction and design 481

8. Tilts and Shifts of Wells


8.1 Tilts and shifts shall be carefully checked and recorded regularly
during sinking operations. For the purposes of measuring the tilts
along and perpendicular to the axis of the bridge, level marks at
regular intervals shall be painted on the surface of the steining of
the well.
8.2 Whenever any tilt is noticed adequate preventative measures
like putting eccentric kentledge pulling, strutting, anchoring or
dredging unevenly and depositing dredge material unequally,
putting obstacles below cutting edge, after jetting etc., shall be
adopted before any further sinking. After correction the dredged
material placed unevenly shall be spread evenly.
8.3 A pair of wells close to each other have a tendency to come
closer while sinking. Timber struts may be introduced in between
the steining of these wells to prevent tilting.
8.4 Tilts occurring in a well during sinking in dipping rocky strata can
be safeguarded by suitably supporting the kerb.

9. Sand Island
9.1 Sand island where provided shall be protected against scour and
the top level shall be sufficiently above the prevailing water level
so that it is safe against wave action.
9.2 The dimension of the sand island shall not be less than three
times the dimension in plan of the well or caisson.

The practical difficulties of sinking small caissons, 12.5 m in diameter in a fast


flowing river for the second Bassein Creek bridge at Mumbai, India were
described by Deshpande and Patel.20 The new bridge is supported on nine
caissons, three on land and the remaining six were floating caissons sunk
through creek water to basalt rock founding stratum. The site conditions
are difficult for caisson construction as there are two daily low and high
tides with a variation of 4.25 m and an average velocity of 2.4 m/s. The
depth of water under the navigation span is 20 m. Occasional wind storms
cause high water currents with a velocity of 3 m/s and during the monsoon
heavy floods occur.
The floating caissons were initially fabricated at a temporary yard con-
structed on the creek bank. The cutting edge and steel liners were concreted
and fabricated to a height of 8.15 m before launching at high tide.
The anchoring in the creek was difficult due to rapid underwater and
surface currents accompanied by eddy formation due to existing piers. At
low tide the caisson was positioned and grounded by pouring concrete into
the steining. The sinking of the caisson was then continued by excavating
the river bed soft marine clay by grab. Severe tilt occurred in sinking caisson
P6 due to an obstruction under one side. The tilt was of the order of 1:5 and
kentledge and grabbing on one side was needed to correct the tilt to the
maximum permissible of 1:60. Generally, tilt in other river caissons was
corrected by:
(a) grabbing on the opposite side
(b) removal of boulder obstructions by divers with hand tools or with water
jets
(c) use of drop chisel to loosen hard strata and thence by grabbing
(d ) light blasting
482 Deep excavations

(e) water jetting from the external face on its high side – reducing friction
on that side and thereby increasing sinking on that side
( f ) applying kentledge on the high side.
A 1 m layer of boulders at caisson P4 had to be removed by drop chisel and
water jetting and divers. The penetration of the caisson through the boulder
layer took 1 month. At times, it was difficult to sink the caisson through the
weathered basalt to founding level. The caisson was moved by filling it with
water and making a controlled blast to give it a jerk. The alternative was to
fill the caisson with water and then to pump it out fast enough to mobilize
and thence reduce the skin friction.
Traditional deep wells, each with one shaft, were specified on two bridges
built across the River Ravi in Pakistan as part of the Indus Basin Settlement

Fig. 10.18. Abdul Hakim railway


bridge: well detail during
sinking and section of typical
pier21
Shafts and caissons: construction and design 483

Scheme in the 1960s.21 One of the new bridges replaced a railway bridge where
the existing foundations were not deep enough for scour protection with the
planned river discharge of 4300 cumecs and modification was not possible.
The wells, 36.6 m below river bed level, were sunk by a method developed in
France in 1927 by Cacot. The method reduced side friction and minimized the
need for kentledge. Figure 10.18 shows the well dimensions for the railway
bridge and the hemispherical shape of the wide well base and the position
of airlift pumps, or ‘emulsifiers’, used to excavate the soil on the well periph-
ery. The reinforced concrete cutting shoe was cast on a form made from
rendered dry-stacked bricks. Reinforced brick steining in 3:1 cement
mortar was built over the shoe to a height that allowed the initial sinking to
be made by crane and clamshell. The following stage of brick steining was
then completed and the excavation was undertaken with central air lift.
Sinking rates varied from small fractions of a metre to more than 3.5 m per
day.
Positional tolerances could only be achieved by sinking at a rake or, more
effectively when caught early, by dredging outside the well and surcharging
one side of it. In three instances correction could only be made at a considerable
depth by divers using water jets cutting below the cutting edge on one side. After
completion of sinking, the spherical shoe was plugged with intrusion-grouted
aggregate and the well was filled with sand. Savage and Carpenter21 commented
that while the technique was an improvement when kentledge was not available,
they considered that well techniques would fall into disuse and be replaced by
large-diameter piles as the cost of masonry work increased.
Large caissons for river piers require support in position during lowering.
Various methods are used depending on river water depth and soil strength
below the river bed and include forming an enclosure, or corral, from pre-
driven piles, sinking on a sand island, sinking from within a cofferdam and
tethering to anchors on the river bed.
A piled enclosure or corral was formed to support the 46 m by 26.8 m
caisson for pier II construction of the Mississippi River Bridge at New Orleans
in the mid 1950s. The 6 m high caisson curb, fabricated at Pittsburgh and
towed 1900 miles on the Ohio and Mississippi rivers to the job site, was floated
into the U-shaped corral formed from 39 90 cm dia. pipe piles, gravel filled and
driven 18.3 m into the river bed. Figure 10.19 shows the open-well caisson
during fabrication and at pier II reading for sinking. It took one month to
sink the caisson to river bed level through 21.3 m of water. The wells were
then dredged and the caisson sunk a further 33 m into hard clay when a
tremie concrete plug, 6 m thick, was placed. The caisson is shown in plan
and section in Fig. 10.20. A total of 27 000 m3 of concrete was used in the
construction of the caisson to pier II.
Corrals were also used to retain the caissons for the tower piers to the
Mackinac Bridge22 . Anchors and steel cables were not used because of limited
manoeuvring space for floating craft, with risks of fouling and cutting the
cables. Four towers, piled with steel tubes into the river bed, were spaced
around the caisson perimeter. Each tower was connected horizontally by
box-type trusses. The circular caisson was built in steelwork with double
walls, with outer diameter 30.5 m and inner diameter 26.2 m. Grouting of
discharged aggregate by the Intrusion Prepakt technique was used for all
the concrete within the caisson and permitted very high rates of underwater
concreting. From a single floating plant, placing reached almost 80 000 m3
in 30 days in caissons and cofferdams to the pier foundations.
Yang23 described the caisson construction procedure for the Brooklyn
pier to the New York City Narrows Bridge. The sand island technique was
used for construction of the caisson curb, the island being formed within a
484 Deep excavations

(a)

Fig. 10.19. Mississippi River


Bridge, New Orleans, open-well
caisson: (a) during fabrication;
(b) about to be launched11 (b)

continuous chain of steel sheet piled cellular cofferdams. The plan and
sectional elevation of the completed caisson is shown in Fig. 10.21. The
sand fill within the cofferdam cells was taken to level þ3.04 m and the
island itself was sand-filled within water to a finished level of 3.65 m.
Dewatering was by open pumping aided by a single-stage wellpoint system
within the cofferdam enclosure. A clay blanket was placed outside the coffer-
dam and the enclosed sand island dewatered to 4.88 m. The rate of pumping
reduced to 3000 gallons per minute after placing of the clay blanket.
The caisson steel cutting edge, 2.1 m high and 39.3 m by 69.8 m in plan, had
internal walls 0.9 m thick and external walls 1.5 m thick. Sixty-six dredging
Shafts and caissons: construction and design 485

Fig. 10.20. Mississippi River Bridge, New Orleans: caisson within corral, plan and vertical sections11

wells, each 3.2 m in diameter, were formed in concrete above the cutting edge,
being poured in four 3 m lifts in each of the stages of caisson extension. The
effective weight of the caisson was 37 000 tonnes at the end of the first stage
extension, and 42 000 tonnes was added at each subsequent extension.
Excavation within the shafts started at the caisson centre, working
progressively to the edge cells. The excavation, taken through sands and
fine gravels into stiff clay, tended to make the caisson sag and the caisson
profile was constantly checked to avoid overstressing the structure. The
average rate of sinking was 60 cm per day for the first two sinkings stages,
the rate depending only on the rate of sand excavation from the cells. The
actual volume of excavation barely exceeded the theoretical volume during
the early stages. Dredging efficiency was reduced at lower depths into the
clay, due in part to the depth effect itself and also because of loss of soil
from grabs. At the lowest depths, the daily rate of sinking reduced to 36 cm
per day and the actual excavation volume exceeded the theoretical volume
by some 27%.
The soil at the cutting edge became more silty and less able to resist
excessive pore-water pressure. A differential head between 1 and 2.4 m was
maintained between cell water level and mean tide level but several cave-ins
occurred. The daily sinking record and the mass diagram during sinking of
the Brooklyn caisson are shown in Fig. 10.22.
In the 1970s the Barton tower of the Humber Suspension Bridge was also
founded on caissons built on a temporary sand island. The figure-of-eight
486 Deep excavations

Fig. 10.21. New York City


Narrows Bridge, Brooklyn pier:
caisson plan and sectional
elevation23

cofferdam containing the sand fill was subject to considerable scour and some
12 000 tonnes of chalk was needed to remedy this hazard, with much of the
cofferdam piling requiring extension and redriving.
The two 24 m diameter caissons consisted of two concentric reinforced con-
crete cylinders joined by six radial walls to form seven cells. The caissons,
extended in 3 m lifts, were excavated by 2 and 4 m3 grabs, and water jets
mounted on jigs and lowered into the cells were used to soften the soil beneath
the cutting edge, first at low pressure and later at pressures between 20 and
40 N/mm2 . The excavation was underwater to avoid unloading the Kimmer-
idge clay. The west caisson penetrated a pocket of ground containing water
under artesian pressure during sinking; as a result the bentonite slurry lubri-
cating skin around the caisson was largely flushed away and skin friction
greatly increased. The caisson was only sunk to final founding depth with
an additional 3000 tonnes of temporary steel kentledge and 4000 tonnes of
Shafts and caissons: construction and design 487

Fig. 10.22. New York City


Narrows Bridge: (a) caisson
daily sinking record;
(b) mass diagram of sinking
caisson23

permanent concrete to form extended outer walls to the caisson. Views of the
caissons during construction are shown in Fig. 10.23.
Mitchell24 referred to claims for reduction of up to 40% friction with the use
of bentonite slurry. This slurry was injected from closely spaced nozzles
connected to a header tube cast into the caisson walls, and the injections
were regular and in sufficient quantity to keep the annular space next to the
488 Deep excavations

Fig. 10.23. Humber Bridge: two


views of caissons during
construction. Each of the 24 m
diameter caissons had two
concentric walls connected by
cross-walls (courtesy of AMEC)

caisson full of slurry. Blaine25 , referring to the Baton Rouge Bridge, con-
cluded that the use of water jets from similar nozzles within the caisson wall
were, as designed, little value in clay and very dangerous in sands, causing
run-ins. At Baton Rouge the two-tier jetting arrangement (Fig. 10.24) was
unsuccessful. Fine sand had offered the greatest resistance in caisson sinking,
and this situation is common elsewhere. Blaine also concluded that, in sand,
blasting was a method of desperation which seldom did any good and
caused run-ins.
It should be noted that the obstruction to river flow by sand island con-
struction may well cause excessive bed erosion. Tomlinson18 noted river bed
scour at sand islands formed within steel shells at the Baton Rouge Bridge.
The scour, 12 m deep despite the use of woven board anti-scour mattresses,
Shafts and caissons: construction and design 489

was caused by constriction of the 730 m river width to 97 m wide waterways


between the sand islands. The fast flowing Mississippi River removed the
whole sand filling to one 37 m diameter island in two to three minutes and
caused severe tilt to the partly sunk caisson.
Blaine25 stated that, irrespective of the detailing of erosion protection at
Baton Rouge Bridge, the essential lesson was that the sand island method
of caisson sinking is potentially a major erosion hazard. The damage to the
caisson referred to by Tomlinson occupied the contractor’s entire organiza-
tion for three months following the accident, in providing further scour
protection and plumbing and deepening the displaced caisson.
The south pier to the Forth Road Bridge, built in the early 1960s near the
famous railway bridge, was founded on caissons sunk from the bottom of a
river cofferdam. The Forth Bridge foundation works were described by
Anderson26 . The south pier foundation was originally designed as two rectan-
Fig. 10.24. Baton Rouge
Bridge: two-tier jetting
gular caissons, floated into position and sunk under compressed air through
arrangement25 the boulder clay. The contractor, John Howard and Co. Ltd, produced an
alternative design using two circular caissons, which were intended to be
driven through the clay as open caissons in the dry. The contractor was con-
fident that the boulder clay would allow sheet piling for the cofferdam to be
driven, and also that it would form an adequate seal for caisson excavation
in the dry. The construction sequence is shown in Fig. 10.25 with a view of
the access to the south pier caisson prior to sinking shown in Fig. 10.26.
The site of the pier, a quarter of a mile from the shore, had a steel and
timber access jetty which ended with staging enclosing the pier site on three
sides.
The cofferdam for the pier, described previously, included a lower ring
waling formed successfully underwater by intrusion grouting. After
dewatering the cofferdam, the caisson was erected on a 100 mm concrete

Fig. 10.25. Forth Road Bridge: south pier construction sequence26


490 Deep excavations

Fig. 10.26. Forth Road Bridge,


south pier caisson: view of river
crossing showing extent of
access works (courtesy of
AMEC)

blinding. The caisson curb consisted of a heavy steel cutting edge mounted on
skin plating. The caisson walls extended to a height of 8.5 m in reinforced con-
crete before sinking. During the erection work the caissons were supported
around their internal periphery by concrete blocks at 1.5 m centres, and the
space between the blocks was then filled to form a reinforced concrete ring
beam of wedge section. After sinking through the gravel, the intermediate
sections were removed by blasting and, as the caisson reached the boulder
clay, the original concrete stooling was blasted away to allow the caisson to
sink on the cutting edge alone.
Although air ducts and locks were ready, the method succeeded in the dry,
in ‘free air’, as the caisson continued downwards at an average rate of 30 cm
each day. Mechanical excavation on the caisson floor achieved an average of
just less than 100 m3 per day in the solid measure. The downstream caisson
was kept 4.5 m lower than its neighbour to ensure that any heavy inflow
below the cutting edge did not flood both caissons. There was some seepage
but bentonite grouting remedied the situation and, additionally, reduced
skin friction during caisson sinking.
The vertical walls extended to a total height of 22.5 m above the cutting
edge and the caisson was founded on bedrock. Figure 10.27 shows sections
of the caisson and the finished pier. At the end of the sinking both caissons
were vertical and less than 150 mm out of position.
The construction of caissons on the Lower Zambezi Bridge was described
by Howarth27 . On this contract both open and pneumatic caissons were
used. It was estimated that 16 of the main wells could be sunk from sand
banks and from the exposed river bed in the dry season. It was thought
that sinking of the remaining wells would likely to be through varying
depths of water and it was decided to use floating caisson sinking sets.
The arrangement adopted for the open wells to piers 28 to 32 in the 1932 dry
season is shown in Fig. 10.28. This technique minimized the obstruction to
river flow and, in turn, the extent of river bed scour. Wire ropes from the
craft carrying the caisson curb and its attendant cranes were secured to
Shafts and caissons: construction and design 491

Fig. 10.27. Forth Road Bridge,


south pier construction: half-
sections showing caissons
during sinking and completed
pier26

heavy concrete or cast-iron anchors. Two 10 t sinkers were laid out upstream
and four 5 t sinkers were used as breast moorings. The breast mooring lines
from the port-side craft were moored to starboard sinkers and vice versa.
One 5 t sinker was laid out downstream.
Caisson buoyancy can be improved by the use of false bottoms to the wells,
the temporary bottoms only being removed after the caisson has sunk a small
depth into the river bed so that the caisson weight is then supported by outer
friction and its cutting edge. In the ‘floatation caisson’ method devised and
patented by Daniel E. Moran for the west bay caissons on the San
Francisco–Oakland Bay Bridge in the early 1930s, compressed air was used
to regulate the amount of draught and improve control of the caisson

Fig. 10.28. Lower Zambezi Bridge: floating well sinking set27


492 Deep excavations

Fig. 10.29. San Francisco–Oakland Bay Bridge, West Bay pier caisson with domed wells, during: (a) floating out; (b) sinking;
(c) dredging11

during sinking. The technique used a domed-end wall as shown in Fig. 10.29.
Figure 10.30 shows the caisson for pier W4 at its final depth. The caisson,
described by Little11 , was 28 by 60 m in plan and the 55 cells, each 46 m in dia-
meter, were carried to a height of 23.6 m above the cutting edge before towing
to site. The caisson had a draught of 6 m and was positioned with lines to 125 t
anchors. Substantial fendering around the caisson was independently
anchored. The caisson was sunk by releasing the compressed air from five
of the cylindrical shafts, decreasing the buoyancy and concreting the caisson
structure progressively from 23.6 m to 35.0 m when the draught was 19.2 m
with the cutting edge a small distance above river bed. The air pressure in
the cylinders at this stage was 1.5 bar. The air was released from all cylinders,
losing 6000 tonnes of buoyancy, and the caisson started to sink into the sandy
clay below the river bed under its own weight. The air domes were burnt off to
allow dredging to proceed, and finally the caisson was founded on a sloping
rockhead 33.5 m below the river bed.
Construction of the Tagus River Bridge in Portugal in the early 1960s
similarly used compressed air and steel domes to stabilize caissons for two
river piers, the first time this had been done in Europe. Riggs28 described
the methods used by Tudor Engineers, who were retained to design the
piers and develop construction procedures for founding caissons on sloping
basalt at considerable depth below the river bed. At pier 4, this rockhead
slope was defined by no less than 31 borings and an even greater number of
jet probings. A sloping rockhead was catered for by constructing the caisson’s
cutting edge to the same gradient (Fig. 10.31). This created problems in con-
structing and floating the caisson which were overcome by the use of the air
domes. By varying the air pressure in the wells the caisson list was corrected
after launching and during sinking.
The open caisson therefore allows permanent pier construction to be taken
down through sands, gravels and clays to adequate founding soils or rock at
very considerable depths. These depths are often necessary because of very
large projections of scour risk in fast flowing rivers, but are not without
limit. Mitchell24 referred to caissons being sunk to the exceptional depth of
Shafts and caissons: construction and design 493

Fig. 10.30. San Francisco–


Oakland Bay Bridge: plan and
cross-section of caisson for pier
W4 at full depth11

105 m below the Jamuna River in Bangladesh in the early 1980s. In relatively
small caissons, in the dry, excavation may proceed efficiently, but in the
majority of cases excavation is only possible underwater and obstructions
at shallow depth, such as tree trunks, may only be removed by a diver using
small charges; obstructions at greater depths, such as boulders below the
cutting edge may produce unavoidable delays in underwater excavation.
Inflow of material beneath the cutting edge may occur due to artesian pressure
in groundwater, causing subsidence outside the caisson and loss of bentonite
seal placed to lubricate the outer skin. As the caisson is founded, sloping
bedrock may cause difficulties in achieving uniform bearing for the pier.
Unless the caisson is at depths where compressed air working can be used,
underwater excavation and diving work may prolong construction.
Hang-ups in caisson sinking sometimes occur unexpectedly and early
estimates of skin friction may prove optimistic. Water jetting, air and water
jetting, bentonite grouting, and even small charges fired in the contained
water, may be used in an attempt to get the caisson moving. The use of
extra kentledge is the final resort. It is highly desirable, therefore, that all
caissons should be designed to allow extreme weighting by extra kentledge
should difficult hang-ups occur.
494 Deep excavations

Fig. 10.31. Tagus River Bridge,


tower pier 4: plan and
cross-section showing sloping
rockhead and shaped cutting
edge28

Difficulties in sinking a large caisson on the Cairo waste water scheme in


1986 were described by Grimes et al.29 The caisson, for a large pumping station
at Ameria, had an external diameter of 45 m and was 37 m deep. The ground
conditions were 8 m deep silty clay and made ground from ground level over-
lying dense to very dense sands and gravels together with fine thin hard grey
silty clay layers at 10 and 13 below datum. Although the pre-contract
design alternatives included the use of a diaphragm wall with a frozen soil
base, the two lowest-bidding contractors both opted for a caisson solution.
After sinking the caisson in an open well in the dry at shallow depth, wet
caisson sinking continued below groundwater level. Dredging equipment
operated from a floating platform which was used subsequently as the perma-
nent base to the dry well of the pumping station. Figure 10.32 shows the
sequence of operations prior to casting an underwater plug to the caisson
on which the dry well base would be sited. Bentonite was used to assist pene-
tration of the caisson to a total depth of 28 m, but ground loss failures caused
considerable difficulty. It became necessary to excavate by grab an annular
space outside the caisson walls which was filled with bentonite slurry to
achieve final levels.
After a study of the history of open caissons, Gerwick30 concluded:
(a) most cases of initial tipping are caused by attempts to rush the work
Shafts and caissons: construction and design 495

Fig. 10.32. Ameria pumping station, Cairo waste water project: sequence of construction stages 1 to 7 and
second-stage excavating equipment29
496 Deep excavations

(b) most cases of serious tipping occur when the steps taken to correct the
minor initial tipping are too radical
(c) false-bottom caissons are the lowest in cost but are the most prone to tip
(d ) dome caissons, such as used on the San Francisco–Oakland Bay Bridge,
are very expensive and in practice do not give assurance against tipping
(e) false-bottom caissons which permit removal of the false bottom are of
major value in preventing tipping
( f ) the double-wall caisson, as used on the Mackinac Straits Bridge, is
expensive but is unlikely to tip seriously because of its low centre of
gravity and high centre of buoyancy
(g) any caisson may act, at times, as a false-bottom type due to soil plugs
forming in the wells.

Open caissons in building construction


The use of caissons is not limited to bridge foundations; the technique has
found application as a construction method for basements to buildings and
for substructures to industrial structures such as pumphouses. In some
instances, simultaneous superstructure erection has assisted caisson sinking.
Since the 1970s, when slurry trench techniques provided an alternative
method of deep basement construction in constricted urban areas, the use
of caissons for building foundations has proved progressively less attractive.
The use of the method in Tokyo in 1950 to construct a four-storey basement
of very large plan size is shown in Fig. 10.33. Three basement storeys for the

Fig. 10.33. Land caisson for the


Nikkatsu building, Tokyo.
Bearing plates shown in plan
and section were used to
control verticality and plan
position of the caisson during
sinking31
Shafts and caissons: construction and design 497

Nikkatsu building were constructed above ground. Temporary timber bearing


plates beneath internal columns and outer walls were used to control move-
ment as the caisson was sunk. The triangular-shaped caisson had a plan
area of about 0.4 hectare and weighed approximately 25 000 tonnes.
After the caisson was founded at its final level, excavation was made in
gravel to cutting edge level and a raft constructed to form a fourth basement31 .

Fig. 10.34. Whickham caisson:


cross-section and plan32
498 Deep excavations

It was claimed that the cost of temporary diagonal bracing to stiffen the
caisson structure during sinking was more than offset by the elimination of
large quantities of shoring, strutting and sheet piling. It was noted that foot-
ways adjacent to the outside caisson wall subsided uniformly 150 to 175 mm;
at a distance of 6 m, street subsidence was about 12 mm.
In the early 1970s a method of caisson construction due to Fehlmann and
Lorenz was introduced into the UK. The patented method was based on a
caisson curb of special design to allow the caisson to sink under its own
weight, the soil adjacent to the cutting edge progressively failing under
shear. Bentonite slurry was used as a lubricant. A number of contracts were
completed at that time, and 200 contracts had been completed in Europe by
1972.
A 45 m dia. open caisson was sunk using this method through stiff silty clays
for a three-storey basement below a 30-storey tower block at Whickham,
County Durham, UK. The cross-section and plan of the caisson are shown
in Fig. 10.34. The caisson was divided into 16 sections by radial walls, and
a floor in the form of a helix was cast prior to sinking to provide horizontal
stiffness and act as a permanent access to car parking spaces in the finished
basement32 .
At Worthing the same method was used to sink a 26.5 m dia. caisson, 17 m
deep, for use as a foul and surface water pumping station; the reinforced
Fig. 10.35. Reinforced concrete concrete curb is shown in Fig. 10.35.
curb to caisson for land A larger caisson 64 m long and 27 m wide with an overall height of 19 m,
pumping station, Worthing sunk on land as an open well, was used for the substructure to a circulating
water pumphouse in the Huntly power project33 . The subsoil conditions con-
sisted of 2 m thick silt overlying a depth of 50 m of pumiceous sand of high
permeability interspersed with silt bands, and at 50 m a layer of hard silt of
low permeability. Sheet pile cut-off walls were driven around the caisson
periphery, which was built to full height before sinking. The sheet piling
reduced excavation volume and improved crane access. Water jets cast into
the walls were successfully used with sand pumps and air lifts to control
excavation. The caisson was supported permanently on bearing piles driven
to siltstone to avoid risk of horizontal movement due to liquefaction of the
pumiceous sands under seismic conditions.

Pneumatic caissons
The pneumatic caisson, like the open caisson, is a four-sided box in steel and
concrete but with the addition of an air deck so that the caisson bottom is like
a diving bell. Compressed air excludes water from this bottom section and
enables work under air to carry out the excavation in the dry. The air pressure
balances the pore-water pressure at the cutting edge. The maximum depth
to which pneumatic caissons can be sunk is therefore directly controlled by
the maximum air pressure at which work can proceed. Some labour health
laws permit a maximum pressure of 3.4 bar. This pressure balances an
external piezometric head of 35 m and, unless a reliable means of reducing
the groundwater head is used, is the absolute limit for the pneumatic caisson
working34 .
The use of high pressure limits the working hours in the caisson, and both
labour and insurance costs become very high. From a health and safety point
of view, the whole subject of working under compressed air for sustained
periods has been of increasing concern in recent years. The use of pneumatic
caissons has therefore declined for reasons other than just cost.
Although there have been experiments with helium/oxygen mixtures, an
ordinary oxygen/nitrogen mixture is used in caissons. Air is deliberately
allowed to escape beneath the cutting edge, and this is usually sufficient to
Shafts and caissons: construction and design 499

Fig. 10.36. Typical


arrangement of a pneumatic
caisson36

ensure adequate ventilation, although exhaust pipes and valves are also
provided.
In the open, a caisson may be sunk as deep as possible, the air locks put on,
air pumped in, and excavation and sinking continued.
Figure 10.36 shows a typical arrangement of a traditional pneumatic caisson.
The caisson has four essential parts: the working chamber; the access shafts; the
air locks which allow transfer from working pressure inside the caisson to
atmospheric pressure outside, and vice versa; and the air compressor plant.
The working chamber, of the order of 2.5 m high, accommodates between
six and eight workers except in the very smallest caissons. Excavation is by
hand with the assistance of clay spades and similar compressed air tools,
water jets and pumps. In loose sand and silt, a blow pipe may be used to
evacuate the material vertically out of the caisson. The blow pipe consists
simply of a pipe with a valve at its lower end leading to a T-junction at its
upper end, the air pressure in the working platform forcing water and soil
up after the valve is opened. Impressive rates of evacuation can be achieved
with this method in sands, the soil being excavated from beneath the
cutting edge by hand and placed in loose heaps for removal by the blow
pipe. The abrasive action of the air/water/sand mixture on steel valves and
ends should be noted. More recently, the use of high-pressure water jets
with sand pumps has proved advantageous, particularly for dense materials.
Water jets with nozzle pressures in the range 20 to 40 N/mm2 can be used
to augment the work of hydraulic bursters to break boulders.
500 Deep excavations

Fig. 10.37. Cairo waste water


scheme: cross-section of
full-depth caisson showing
personnel, mud locks and
excavation35

Figure 10.37 shows a cross-section through a caisson of later design used on


the Cairo waste water system construction in the early 1990s35 . Excavation, by
a Smalley hydraulic excavator into 2.5 m3 muck skips, took six to eight weeks
for each shaft from application of compressed air to pouring the concrete
plug. On contract 4 of the waste water scheme these shafts were sunk through
typical Cairo subsoil conditions of fill overlying silts and sands with a high
groundwater table, the excavation finishing in sand or gravel layers. The
diameter of the finished structure varied from 6 to 10 m. The caissons were
only used for the first 10 m after which a concrete segment-lined underpinning
shaft, later lined with in situ reinforced concrete, was used. Air pressure
varied with depth of the shafts, from 1.6 to 2.3 bar, and needed between 42
to 340 m3 /min free air delivery in the coarser soils.
The shafts are often in a figure-of-eight plan configuration with the muck
shaft, usually 1 m in diameter, being placed next to the main access shaft.
The shaft sections are usually 3 m lengths to facilitate extension heights to
the caisson as sinking proceeds.
For safety reasons, separate air locks for workers and materials, except in
the smallest caissons, are sited above maximum water level to allow workers
to escape should air pressure fall and the caisson flood. For medium-sized
caissons, say up to 100 m2 in plan, it is usual to provide two muck locks
and one man lock. For high pressures with long lock occupancy times, two
man locks may be necessary for medium-sized caissons. Figure 10.38 shows
Shafts and caissons: construction and design 501

Fig. 10.38. Typical


Gowring-type air lock36

the Gowring-type air lock with double man locks accommodating twelve men
and a muck lock designed to hold a 1 m3 muck skip.
Compressed air is usually supplied from stationary compressors often
driven by variable-speed motors which can increase the supply as caisson sink-
ing proceeds. Compressors must be in duplicate with separate power sources.
The compressors are connected to twin air receivers fitted with relief valves.
Compressor tools in the caisson are separately supplied with air at operating
tool pressure plus working chamber pressure.
The quantity of air supplied and the operation of man locks is under the
jurisdiction of local or national health and safety regulations. The air quantity
depends on five factors: the caisson size; the number of workers; air leakage
from supply lines, stuffing boxes etc.; air loss from under the cutting edge;
and air loss through the use of the locks.
Great care is essential in maintaining the true position and verticality of the
caisson in the initial 5 to 8 m of sinking. Corrections to caisson position
become increasingly difficult with depth. The position and levelling of the cais-
son should be checked at almost every movement. Howarth27 commented that
after a well had been sunk to a depth exceeding about one and a half times its
axis diameter it was a mistake to try to correct position in the plan of that axis;
instead, attention should be focused on keeping the well plumb. Small correc-
tion to tilt could still be made by careful undercutting.
Where the caisson is a ‘floater’, that is, handled from floating river craft,
some positional error in placing the caisson on the river bed is inevitable.
Corrective measures can be taken in the first few metres of sinking; slight
502 Deep excavations

tilting of the caisson for a small depth of sinking may bring the caisson back to
its true position, after which vertical sinking can be resumed. Alternatively,
timber cribbing may be successful, or steel wedge plates can be used under
the cutting edge to move the caisson bodily sideways as it sinks. Round cais-
sons, desirable in other respects, such as minimization of overall frictional
resistance, are more difficult to control than rectangular plan shapes.
At the start of excavation within the working chamber, in loose granular
soils a berm with side slopes of 1:2 horizontal to vertical is left around the
periphery of the chamber. As the caisson sinks and the centre area is progres-
sively excavated, the berm drops inwards as the cutting edge penetrates. The
berm in the corners is excavated last of all. Where the caisson refuses to move
because of high skin friction, ‘blowing down’ may be used. After the workers
have left the working chamber, the air pressure is reduced and the extra effec-
tive weight of the caisson allows it to drop. Considerable care is needed to
avoid large drops. Run-ins may occur in loose sands and silts, with risk of
caisson tilt or movement. The procedure may succeed in hard ground such
as stiff clays, but the presence of boulders may cause difficulties.
During sinking in a homogeneous clay, the air pressure is often maintained
somewhat less than the theoretical balance, with a reduced air loss below the
cutting edge. If such a caisson then penetrates water-bearing granular soil, the
air pressure must be increased to achieve a balance against the water head and
maintain a stable bottom to the excavation.
As the caisson nears rockhead, forward probing will establish the rockhead
shape. Rock may be shot and excavated from the centre of the chamber to
allow rock near the cutting edge to be broken out. Control of the drop is essen-
tial and 1 m deep holes at 0.5 m centres are drilled to pull about 60 cm depth.
Rock pillars are left in the corners of the working chamber and are removed
when the material below the remainder of the cutting edge has been cleared
away.
When a caisson approaches founding level in soil and is moving without
‘blowing down’, it is necessary to cast blocks of concrete in small pits at the
corners of the chamber on which to seat the caisson. A concreting and grout-
ing procedure seals the caisson at its final level, filling all cavities in the
exposed formation and the whole of the working chamber.
After the excavated surface of soil or rock has been levelled and cleaned off,
a 60 cm thick pour of fairly workable concrete is made across the chamber
floor and worked well below the cutting edge. Concrete of drier consistency
can be used to pack the haunches of the curb. Care is needed at this stage
because concrete placed in this way forms a seal and prevents air loss below
the cutting edge, reducing ventilation and increasing pressure.
Concrete of high workability is then placed by dropping through the
bottom door of the muck lock under air pressure to fill the working chamber.
Bleeder pipes taken vertically above the chamber show that the filling is
complete when grout moves up the pipes.
The whole is then grouted with 1:1 grout placed to fill the lowest section of
the air shaft and kept under air pressure of 0.34 to 0.68 bar for 48 hours. The
permanent length of the air shaft is then concreted. Figure 10.39 shows typical
concreting works in sealing a pneumatic caisson.
The construction details of the caisson, and in particular the caisson curb,
deserve comment. Figure 10.40 shows typical pneumatic caisson curbs or
shoes. The angle of the haunch above the cutting edge depends on the type
of soils to be penetrated. The cutting edge, made from stiffened steel plate,
is vulnerable to buckling during caisson sinking and many problems can
result. Wilson and Sully34;36 recorded that the cutting edge shown in
Fig. 10.40(c) was used through sand and on to rock on a pumphouse caisson
Shafts and caissons: construction and design 503

Fig. 10.39. Typical steel caisson for river pier showing deposition of concrete in working chamber36
504 Deep excavations

Fig. 10.39. Continued

at a steelworks. Considerable excavation was needed under the cutting edge,


and pressure from rock not cleared sufficiently from the back of the cutting
edge caused the steel plate to be bent inwards as the caisson dropped, causing
the concrete to spall on the inside of the haunch and to damage the haunch
reinforcement steel.
The cutting edge and haunch detail shown in Fig. 10.40(b) was designed for
use in boulder clay, the slight outside slope of the haunch being provided to
avoid pincer action in any uneven sinking. In the event, the detail proved
less than successful; soft silt overlying the boulder clay and within the small
angle of the haunching allowed the caisson to sink quickly and unevenly.
Wilson and Sully described troublesome blows caused by loose soil falling
and becoming lodged between the outside curb face and the excavated soil
surface during sinking. This loose soil formed an inefficient seal and any
excess air pressure within the chamber led to rapid air loss and water entry
below the cutting edge.
The grain silo caisson curb detail shown in Fig. 10.40(a) was used for sink-
ing predominantly through silty sand with some clay at deeper levels. Note the
sloping inner face to the steel cutting edge instead of the usual horizontal
lower face. The caisson sinking operation proved quite successful.
Gerwick30 pointed out that the cutting edge must meet the following
requirements.
(a) It must be strong and rugged to resist extremely high localized pressure,
such as might be caused by a boulder.
(b) It must be designed to resist twisting, shearing, crushing and particu-
larly the tendency to spread outward because of its sloping inner
surface.
(c) Its plates must be adequately stiffened and outside plates must be
heavy.
(d ) Connection and splice details must be rugged and strong.
(e) There must be provision for ease of concrete placement inside the
cutting edge, and provision for seal placement to avoid voids.
( f ) There must be sufficient vertical diaphragms to make the cutting edge
cross-section act as a whole.
Shafts and caissons: construction and design 505

Fig. 10.40. Typical cutting edge and caisson curbs for use in: (a) silty sand with some clay at depth; (b) boulder clay;
(c) sand on to rock; (d) soft clay36 ; (e) river deposits, boulder clay, sands, gravels and coal measures, progressively with
depth24

(g) All cutting edges must be tied together in a rigid frame to resist distor-
tion.
(h) The shoe must be designed for excavation under the cutting edge in
greater depths of water.
Wilson and Sully34 recommended steel caisson curbs in preference to those in
reinforced concrete. More recently, however, and particularly for small- to
medium-size caissons, the reinforced concrete curb is generally used. Wilson
506 Deep excavations

and Sully argued that the lower weight of the steel caisson had advantages of
draught during towage and control during the initial sinking when concrete
infilling could be adjusted to allow for the presence of soft soils. Wilson and
Sully also advised that caisson construction is most economical with the
haunches built in trench and the caisson roof shuttering supported from the
existing ground.
Above the cutting shoes the caisson walls should be set-in a distance of 25 to
50 mm to assist sinking. The wall thickness is itself dependent on the need to
achieve rigidity and resist stresses due to uneven sinking and on the weight
requirement to overcome skin friction between caisson and soil during
sinking. Adequate vertical reinforcement from the curb to the top of the
caisson is necessary to resist any tension in the walls caused by ‘hanging up’
during sinking.
The plan shape of caissons may be circular, ellipsoid or rectangular. On the
Cairo waste water scheme35 , a triple full-depth caisson was constructed to
allow the bentonite tunnelling machines for the same working site to break
out in opposite directions in close succession. Figure 10.41 shows tunnelling

Fig. 10.41. Cairo waste water


scheme: triple-cell caisson35
Shafts and caissons: construction and design 507

Fig. 10.42. Sinking of caisson in


a narrow city street, Cairo
waste water scheme35

from the triple-cell caisson. The caisson was originally sunk in free air to 8.5 m
below ground level and sinking continued under compressed air to a depth of
24.5 m using a maximum pressure of 2.07 bar, requiring a maximum air con-
sumption of approximately 113 m3 per minute. Construction time for sinking
was 20 weeks.
A considerable number of caissons were sunk for working shafts during the
several phases of the Cairo waste water scheme. In some instances caissons
were sunk in very narrow working sites in streets in the residential quarter
of the city. Figure 10.42 shows a particular example of a caisson site immedi-
ately adjacent to existing buildings. Originally designed, pre-contract, as
underpinned caissons of limited depth, on some sections contractors resorted
to full-depth caissons (Fig. 10.43). Flint et al.35 concluded that the lack of any
distress to very close existing buildings, some of which were very old and
probably fragile, was a commendation for the full-depth caisson method in
such circumstances.
These caissons were designed on the basis that kentledge would be needed
to overcome both skin friction on the outside of the caisson works and the
uplift due to compressed air. Resistance was generally greater and more
kentledge was used than was anticipated in most cases. Small (75 mm) dia-
meter pipework within the caisson walls introduced bentonite to reduce
friction during caisson sinking, but one or two caissons had to be ‘blown
down’ by a sudden reduction of air pressure of up to 25% lasting 30 seconds.
The sinking effort applied to the caisson, that is the effective weight of the
caisson plus any kentledge, must be sufficient to overcome the resistance to
penetration of the cutting edge and the skin friction between the soil and
the outside wall. Stated in these words the equation is somewhat over-
simplified; the effective caisson weight is much increased during blowing
down when the working chamber pressure is reduced, and resistance at the
cutting edge is also much reduced by undercutting. In addition, the use of
508 Deep excavations

Fig. 10.43. Cairo waste water


scheme, alternative methods of
shaft construction:
(a) Engineer’s original design;
(b) Contractor’s design35

bentonite slurry to lubricate the outer wall–soil friction is widely practised. On


the other hand, distortion or bulges in the outside surface of the caisson wall
can substantially increase resistance to penetration. Nevertheless, it is
essential that skin friction is never underestimated in considering the weight
necessary to sink the caisson nor overestimated in calculating the permanent
safe load capacity of the caisson foundation.
Table 10.1 Values of skin Terzaghi and Peck37 gave indicative values of skin friction (Table 10.1).
friction during sinking of Values observed during sinking of both pneumatic and open-well caissons
caissons37 in differing soil conditions are quoted in Table 10.2. Comparing these tables
Type of soil Skin friction fs leads to the conclusion that there is considerable scatter of skin friction in
(kN/m2 ) similarly described soils, and this makes accurate estimation of skin friction
Silt and soft clay 7.2–28.7
before sinking more difficult. Tomlinson18 pointed out that disturbance of
Very stiff clay 47.9–191.6 stiff clays by undercutting at the cutting edge is likely to reduce efficient con-
Loose sand 12.0–33.5 tact between the outer caisson wall surface and soil, and when sinking through
Dense sand 33.5–67.0
Dense gravel 47.9–95.8
soft, sensitive clays and silts, friction generally has been found to be less than
the remoulded strength, possibly two-thirds of this value when the caisson is
sunk rapidly and continuously. Additionally, in dense sands, grabbing and
dragdown during caisson sinking may cause soil disturbance and reduce
friction. Nevertheless, a conservative approach is essential and friction
should not be underestimated when calculating the minimum effective caisson
weight necessary at each increment of depth.
After reviewing eight caisson friction values, Tomlinson concluded that
these values were very erratic, ranging from 9.6 to 29 kN/m2 , with very few
higher values, and appeared to vary little with various soil types. Wide
variations were due to four factors:
. the presence of boulders
. effects of air escaping below the cutting edge
. ‘freezing’ effect caused by a stoppage in caisson sinking operations
. the shape of the caisson.
Handman38 stated that the deeper the designed founding level of a caisson the
higher the average sinking effort to bring it to its final level. Wiley39 observed
that the skin friction of the caisson lower section increased directly with depth
Shafts and caissons: construction and design 509

Table 10.2 Observed values of skin friction in caissons18

Site Type of Approximate Soil conditions Observed Remark


caisson plan size skin friction
(m) (kN/m2 )

Howrah Bridge, Open-well and 55  25 Soft clays and silty sands 29 Measured value
Calcutta pneumatic from pneumatic
caissons
Baton Rouge Open-well Caisson 1: 38:3
Bridge, River Stiff clay.
Mississippi Tight clay grading to very 40:7
sandy clay.
Tight clay grading to very 31:0
sandy clay with lubricating jets
Caisson 3:
9.5 m sand grading to gravel;
4:6 m, clay and sand; 12.2 m
stiff clay; 3 m clay and sand;
2 m sand 40:5
Caisson 4:
14:3 m sand and clay; 16:1 m 35:2
sand, sand and clay
Caisson 5:
8:5 m silt; 29:9 m fine sand 53:8
Lower Zambezi Open-well and 11  6 Mainly sand 22:9
Bridge pneumatic
Uskmouth Pneumatic 50  33:5 12:2 m soft clay 55:0
Power Station
Grangemouth Open-well 13  13 Very soft clay 4:75
19:5  10 Very soft clay 5:75–10:0
Kafr-el-Zayat Pneumatic 15:5  5:5 Sand and silt 18:7–26:3
Grand Tower Open-well 19  8:5 Medium fine sand and silt Above Dewatering from
Mississippi River W.L.: 51 wells shown to
Below effect friction
W.L.: 29:7
Verrazano Open-well 69:5  39 Medium dense to dense sand 84:75–95:4 At lowest stage of
Narrows and fine gravel sinking to 40 m
Gowtami Open-well 96 9:1 m sand
13:7 m stiff clay
7:6 m sand 12:6
New Redheugh Pneumatic 11 m 5:5 m depth below river bed. 33 North caisson.
Bridge, diameter Hard boulder clay and dense Reduction of
Newcastle sand at shoe friction of 20%
10 m depth. Dense sandy 38 assumed due to
gravel and cobbles at shoe bentonite above
12 m depth. Top of shoe in 36 point of injection
gravel, bottom in mudstone
King Edward Pneumatic 34:6  10:7 South caisson. 31:6
VII Bridge, Sand and gravel overlying
Newcastle shale.
Centre caisson 26:8
4:9 m sand and gravel
overlying 3 m shale.
North caisson 35:5
sand and gravel overlying
shale
510 Deep excavations

Fig. 10.44. Uskmouth Power Station, stages in erection and caisson sinking: (a) caisson shoe in erection position on shore;
(b) caisson shoe supported on hydraulic jacks prior to removal of staging and subsequent lowering;
(c) caisson during sinking with tower derricks and air locks in position; (d) caisson after sinking to final level;
(e) completed pump house36
Shafts and caissons: construction and design 511

but the passage of the caisson progressively reduced the friction at any given
elevation by lubrication.
Yang23 described measurements of sinking force made on large open
caissons sunk through sands, fine gravel and silty clay for the foundations
to the New York City Narrows Bridge. The lower part of the sinking was
made with water jet lubrication and, more successfully, with air-water jet
lubrication. He concluded the following.
(a) The effective weight of the caisson should always exceed the anticipated
resistance without excessive use of a lubricating system.
(b) The bearing capacity of the silty clays at the cutting edge was deter-
mined to be 5.9 to 6.5 times the shear strength, which is in good agree-
ment with the theoretical value quoted by Terzaghi and Peck37 .
(c) The skin friction on the caisson could be estimated using Coulomb’s
active earth pressure. The skin friction on the caisson was a maximum
at the very beginning of sinking and reduced to 45% of average over-
burden pressure. The lower range of skin friction depended on the
method and manner of applying water jets and compressed air, but a
factor of 40% of average overburden pressure should be considered
the practical minimum where a ‘built-in’ jetting system was installed.
One of the largest caissons sunk in the UK was at the Uskmouth Power Station
near Newport, South Wales, in the early 1950s. The construction stages are
shown in Fig. 10.4436 . The caisson, weighing 42 000 t when complete, had a
plan size of 50 m by 33.5 m and was 24.4 m deep. The caisson shoe had been
fabricated on the river bank and was rolled to its sinking position over a trestle
framework. After lowering by hydraulic jacks, the shoe, which weighed 510 t,
was weighted further with concrete, allowing it to sink through the river water
and into the made up ground. Concreting was continuous and over 5000 m3
was placed in a month, the caisson sinking under its own weight a further
depth of 2.4 m without excavation. Compressed air was then applied and the
excavated pneumatic caisson sank a total depth of 19.4 m.
The caisson was divided into three working chambers with two air shafts to
each chamber. A maximum of 30 men split into five gangs worked in these
chambers. Although it was expected that the caisson would tend to drift
towards the centre of the river during sinking, the extent of this drift was
1.45 m, further than predicted.
The effects of dewatering in sand from deep wells adjacent to a pneumatic
caisson on the Grand Tower pipe bridge spanning the Mississippi River were
recorded by Newall40 . A pneumatic caisson was designed to allow under-
pinning work from within the land caisson into uneven bedrock. Four deep
wells were installed at an early stage to reduce the groundwater table to
allow maximum air pressures of 3.4 bar within the caisson. The caisson,
built from reinforced concrete with a welded steel cutting edge, is shown in
Fig. 10.45. Newall stated that the pumping operation affected the friction
on the outer skin; as the well pumps were turned off the caisson began to
sink, but during pumping the caisson was effectively held. It was also observed
that when pumping was reliably continuous the caisson maintained its
theoretical position, but when pumping was discontinued through break-
down, the caisson wandered from position. Newall concluded that it is
possible to use dewatering to control the position of land caissons during
sinking.
Pneumatic caissons were used successfully for three bridges constructed
across the River Tyne at Newcastle, over a period of 80 years24;41;42 . The
plan area of the caissons reduced progressively as each bridge was built, but
perhaps the most evident difference between the works is the improved
512 Deep excavations

Fig. 10.45. Pneumatic caisson,


Grand Tower Pipe Bridge, River
Mississippi: (a) cut-away view;
(b) horizontal cross-section;
(c) arrangement of
underpinning40
Shafts and caissons: construction and design 513

Table 10.3 Comparison between open well and pneumatic caissons43

Open-well caissons Pneumatic caissons

‘Normal’ heavy excavating plant only required. No Special compressed air plant required. All excavated
restriction on excavation material has to pass through air locks
No unusual precautions for personnel and normal rates Special medical precautions, e.g. examination and
of pay rejection of unsuitable men — special care during and
after decompression — provision of medical locks.
Higher rates of pay
No limit of depth so far reached: 240 ft below water Practical limit of depth, 120 ft below water level. In
level at San Francisco Bay bridge through 200 ft of some countries restrictions on maximum air pressure
mud and soft deposits
No men working inside caisson except occasional Men work inside caisson during sinking
diver’s inspection
Obstacles may hold up sinking and may have to be Obstacles clearly seen and the best method of dealing
removed ‘blind’ by blasting or other means with them assured
A zone around the inside perimeter of the well Whole of the area beneath the caisson accessible to the
(say 2 ft wide) not accessible to the grabs excavators
(this is not serious except for small wells)
Final cleaning up before concreting must be done Foundation can be thoroughly cleaned and inspected
underwater by grabs. If the depth is not too great, before concrete is placed
a diver can inspect
Excavation is best suited to soft materials Excavation can be done in any type of material

health and safety provision for workers in compressed air over the period. The
most recent works used the Medical Code of Practice for Work in Compressed
Air (1982), in draft form at the time of construction, and the Blackpool
Decompression Tables. Examinations, including with X-rays, were regular,
in particular to detect a septic necrosis of bone.
The use of pneumatic caissons is reducing in the same way as the use of
compressed air to exclude groundwater and stabilize excavation faces is
reducing in civil engineering works as a whole. The detrimental physiological
effects of working under air have prompted the development of bentonite and
earth pressure balanced shields for tunnel works, for example. With greater
cost and greater risk in both financial and health terms, the application of
compressed air to caisson work is likely to reduce further. Wilson and
Smith43 compared open-well and pneumatic caissons before either financial
or health disadvantages became so evident (Table 10.3).
A useful review of German contractors’ experience of caisson design and
construction in the 1990s was given by Arz et al.44 The following summary
refers to the design aspects within this text, for both open-well and pneumatic
caissons.

Plan shape of caissons


Caisson plan shapes which are the most statically efficient are shown in Fig.
10.46. The circular plan shape provides both the minimum cutting edge
length and the minimum wall area in contact with the soil in terms of caisson
plan area. For this reason large caissons are often circular in plan or nearly so.

Fig. 10.46. Caisson plan


shapes recommended by Arz
et al. as statically efficient44
514 Deep excavations

Fig. 10.47. Typical cutting edge


profiles44

In general, the caisson cross-section is governed by the caisson use and


whether it is an open or pneumatic caisson. If the plan shape is rectangular
it is often advantageous to use longitudinal or cross-walls to divide the caisson
into approximately square plan areas for ease of construction. To reduce
the risk of tilt during sinking the ratio of length to width in plan should not
exceed 2:1.

Design of walls
The outer surface of the exterior walls is generally set in from the vertical
surface of the cutting edge. This is shown in Fig. 10.47. Even a set in distance
of 3 cm is sufficient to reduce at-rest earth pressures to active values against the
outer wall surface. With open caissons this set-in distance can be larger but
certainly with compressed air caissons it needs to be less than 10 cm to
avoid excessive air loss.
The thickness of the external walls is determined not only by the applied
loads but also by the requirements of minimum self weight. For larger caissons
an external wall thickness of 1 m is usually sufficient. The lower level of internal
walls may be raised above the cutting edge level to improve sinking (see Fig.
10.48). The thickness of the internal walls is usually 0.6 to 0.8 m. Because of
the volume of concrete being placed in these thick external walls it may be
necessary to install cooling water pipes to reduce early age thermal cracking
in the concrete above the cutting edge.

Design of the cutting edge


The sectional profile of cutting edges varies considerably due to subsoil con-
ditions and the value of the load acting on the cutting edge. The profile must
be such that the downward load induces a bearing capacity failure below the
cutting edge in order to allow the caisson to sink, although this must be
controlled and the rate of sinking must not be excessive. It may be necessary
Shafts and caissons: construction and design 515

Fig. 10.48. View of outer cutting


edge to caisson at Bordeaux;
showing cutting edge to inner
walls at higher level44

to build a corbel above the cutting edge in order to arrest the sinking by
temporary vertical propping laid on the excavated surface (Fig. 10.49).
For pneumatic caissons there is a minimum embedment depth of the cutting
edge to avoid excessive air loss. The height of the working chamber must be
dimensioned to allow sufficient headroom for the safe working of operatives
and plant.
In open caissons the cutting edge is generally comparatively slim and the
splay angle of the cutting edge is acute. The base of the cutting edge requires
protection and steel sections can be used to achieve this. Figure 10.50 shows
typical methods of protection.
Fig. 10.49. Temporary corbel Deep caissons will usually need bentonite slurry to assist sinking. Alterna-
above cutting edge used for tive means of supplying bentonite, by ring main or through vertical down
temporary propping44 pipes at 1 m centres cast into the walls are shown in Fig. 10.51.

Fig. 10.50. Protection of cutting


edge with steel section44

Fig. 10.51. Alternative slurry


ring main locations44
516 Deep excavations

Fig. 10.52. Forces acting on an


open caisson during sinking
(see Fig. 10.53 for explanation
of parameters)44

Fig. 10.53. Forces acting on a


pneumatic caisson during
sinking44

Caisson calculation
The caisson installation phase will generally determine caisson design. For an
open caisson, the forces during sinking are shown in Fig. 10.52. The forces
acting on a pneumatic caisson during this phase are shown in Fig. 10.53.
The design assumptions regarding wall friction resisting caisson sinking are
shown in Fig. 10.54. If bentonite slurry is used a deduction of 58 can be made
from the friction angles. The design must consider:
(a) the air pressure at each stage
(b) the weight of the caisson at each stage
(c) the timing and amount of ballasting if required
(d ) concreting at each stage.
During construction a minimum of four load cases must be considered:
(a) the caisson must stand unsupported before sinking without air pressure
(b) the stage during sinking when maximum moment occurs on the caisson
walls causing tension on the inner face of the external walls
(c) condition at full depth with minimum moment on the external walls
Fig. 10.54. Wall friction
resisting caisson sinking – (d ) catastrophic case at full depth with complete loss of air pressure (not for
design assumptions44 open caisson).
Shafts and caissons: construction and design 517

Fig. 10.55. Calculation of net


vertical forces at final depth for
a pneumatic caisson44

Figure 10.55 shows the calculation of net vertical forces at final depth for a
pneumatic caisson. Assuming least favourable parameters, a net vertical
load of 50 kN/m is likely on the cutting edge although much higher values
will occur at the earliest stages of sinking. These values may reach 800 kN/
m. The actual values and distribution of this load on the cutting edge
cannot be verified by measurement. Klockner45 gives average bearing pressure
at which cutting edges will come to rest. These may be used in the absence of
more accurate data: for gravels, 1.2 to 1.6 MPa; for sands, 0.9 to 1.3 MPa. A
check using alternative pressure distributions shown in Fig. 10.55 should be
made. A lower acceptable factor of safety would apply to these distributions.
A further condition that should be checked occurs during surges in the rate
of sinking and a pendulum effect as correction is made to verticality when
increased stresses are induced in the external walls. The resulting indicative
earth pressure distribution is shown in Fig. 10.55.

Table 10.4 Open and pneumatic caisson construction in the 1980s–1990s due to Arz et al.44

Site Year of Caisson plan Wall thickness Ground Founding


construction size (m) (m) conditions depth (m)

Compressed air caissons


Bridge Piers, Kehl 1958–1989 7:00  22:00 – – 18.7
Well shaft, Sturzelberg 1971 Diameter 6.20 – – 24.5
Four caissons for bridges, Fray-Bentos 1974 Diameter 10.00 – Silt, medium– 5–7
Uruguay fine sand
Contiguous caissons for Baulos 1974–1975 10:00  33:00– – Silt, sand 14.0
Wibaustraat, U-Bahn, Amsterdam 66.00
Pylon foundation, Rotterdam 1977–1979 9 00  53:50
: – Clay, sand 12.7
Pumping station, Mannheim 1980 17:80  35:40 – Silt, sand, gravel 11.4
Stormwater pumping station, 1987 8:90  9:80 – Fine–coarse sand 11.5
Berlin-Spandau
Pumping station, Bremen-Oslebshausen 1988–1990 Diameter 26.90 – Silt, fine– 27.4
medium sand
Open caissons
Pipeline shaft, Verbindungskanal, 1982 6:5  10:5 1.00 Fine–coarse 16.00
Berlin-Charlottenburg sand, rock
Pumping station, Yenikapi, Istanbul 1985–1986 26:80  36:50 – Fill, sand, clay 13.35
Ameria Pumping Station, Cairo 1985–1988 Diameter 44.20 2.50 Clay, hard 28.00
with depth
Parking garage, Konstanz 1988–1991 Diameter 52.30 0.65 Marine clay, silt, 16.00
sand, gravel
518 Deep excavations

In the final condition, after construction, earth at-rest pressures should be


used to check external wall stresses and bearing pressures beneath the caisson
(unless there is a more onerous loading).
Typical approximate quantities of reinforcement for preliminary estimating
purposes are given as follows:
. in the cutting edge, 250 kg/m3 of concrete
. in the working chamber roof, 170 kg/m3 of concrete.
Table 10.4 gives a list of caisson construction examples, open and pneumatic,
in the 1980s and 1990s.

References Shafts
1. Jones M. The ins and outs of uppers and downers. Tunnels and Tunnelling, 1984,
16, Sept., 31–37.
2. Grieves M. Deep shafts. Tunnelling engineering handbook. Ed. J. O’Bickell,
Chapman and Hall, New York, 1996.
3. Chapman E.J. et al. Cooling water intakes at Wylfa power station. Proc. Instn
Civ. Engrs, 1969, 42, Feb., 193–216.
4. O’Shee S.F. The construction of the Howth Tunnel. Trans. Institution of Civil
Engineers of Ireland, Vol. 84, 1957–58, 49–110.
5. Megaw T.M. and Bartlett J.V. Tunnels: planning, design, construction. Ellis
Horwood, Chichester, Vol. 1, 1981.
6. Guide to shotcrete. Report by ACI Committee 506. American Concrete Institute,
New York, 1985.
7. Proc. Conf. Shotcrete for Ground Support. American Concrete Institute, New
York, 1972, SP-54.
8. CIRIA. A guide to the use of rock reinforcement in underground excavation.
CIRIA, London, 1983, Report 101.
9. Choquet P. and Hadjigeorgiou J. The design of support for underground excava-
tions in comprehensive rock engineering. Ed. J.A. Hudson. Pergamon Press,
Oxford, 1993, ch. 4, 313–348.

Caissons
10. Tricoire J. Le Metro de Paris. Paris Museés, 1999.
11. Little A.L. Foundations. Arnold, London, 1961.
12. O’Sullivan. Recent developments in Dublin Port. Proc. Instn Civ. Engrs, 1970,
suppl., 153–189.
13. Bruun P. Port engineering. Gulf Publishing, Houston, TX, 1976.
14. Golder H.Q. Floating foundations, Foundation engineering handbook. (Ed.
H.F. Winterkorn and H.Y. Fang, Van Nostrand Reinhold, New York, 1974,
537–555.
15. Hansen K. and Frode J. Proc. Symp. Concrete Sea Structures, London, 1972,
214–223. F.I.P., London, 1972.
16. Howarth G.E. and Smith S.J. The New Howrah Bridge, Calcutta, construction.
J. Instn Civ. Engrs, 1947, 28, May, 211–257.
17. Purcell C.H. et al. Difficult problems overcome in sinking deep caissons. Engn
News Record, 1963, 12 Dec.
18. Tomlinson M.J. Foundation design and construction. Longman, Harlow, 1986.
19. Indian Roads Congress. Standard specifications and code of practice for road
bridges: section VII foundations and substructure. 1988.
20. Deshpande D. and Patel D. Problems encountered for deep caissons. Proc. Deep
Foundations Institute Conf., Nice, 2002. DFI, Englewood Cliffs, New Jersey,
2002.
21. Savage C.D. and Carpenter T.G. Indus Basin settlement scheme. TSMB link
canal scheme. Proc. Instn Civ. Engrs, 1965, 32, Dec., 549–571 (discussion 1966,
35, Sept., 184–204).
22. Boynton R.M. Mackinac Bridge. Civ. Engng, 1965, 45, May, 44–49.
Shafts and caissons: construction and design 519

23. Yang N.C. Conditions of large caissons during construction. Highway Res. Rec.,
No. 74, 1965, 68–84.
24. Mitchell A.J. Caissons. Handbook of structural concrete. Ed. F.K. Kong, et al.
Pitman, London, 1983.
25. Blaine E.S. Practical lessons in caisson sinking from the Baton Rouge Bridge.
Engng News Record, 1947, 6 Feb., 213–215.
26. Anderson J.K. Forth Road Bridge. Proc. Instn Civ. Engrs, 1965, 32, Nov. 321–
512.
27. Howarth G.E. Construction of the Lower Zambezi Bridge. J. Instn Civ. Engrs,
1937, 4, Jan., 369–422.
28. Riggs L.W. Tagus River Bridge tower piers. Civ. Engng, 1965, Feb., 41–45.
29. Grimes J.F. et al. Greater Cairo waste water project. Proc. Instn Civ. Engrs, 1993,
97, special issue, 34–37.
30. Gerwick B.C. Handbook of heavy construction. Eds. J.A. Havers and F.W.
Stubbs, McGraw-Hill, New York, 1971.
31. Mason A.C. Open caisson method used to erect Tokyo office building. Civ.
Engng, 1952, Nov., 46–49.
32. Nisbet R.F. Whickham tower block. Structural Engr, 1973, 51, No. 7, 225–231.
33. Brown A.S. and Cox D.D. Design and sinking of a large concrete caisson at the
Huntly power project. New Zealand Engng, 1980, 35, No. 2, 28–32.
34. Wilson W.S. and Sully F.W. Compressed air caisson foundations. Institution of
Civil Engineers, London, works construction paper No. 13, 3–30, 1949.
35. Flint G.R. et al. Greater Cairo waste water project. Proc. Instn Civ. Engrs, 1993,
97, special issue, 18–33.
36. Wilson W.S. and Sully F.W. The construction of the caisson forming the founda-
tions to the circulating water pumphouse for the Uskmouth Generating Station.
Proc. Instn Civ. Engrs, Part 3, 1952, 1, 335–356.
37. Terzaghi K. and Peck R.B. Soil mechanics in engineering practice. Wiley, New
York, 1967.
38. Handman F.W. Lower Zambezi Bridge. J. Instn Civ. Engrs, 1937, 4, Jan., 325–
368.
39. Wiley H.L. Sinking of the piers for the Grand Trunk Pacific Bridge. Trans ASCE,
1909, 62, 132–133.
40. Newall J.N. Pneumatic caisson pier. Civ. Engng, 1956, May, 51–55.
41. Davis F.W. and Kirkpatrick G.R. The King Edward VII Bridge, Newcastle on
Tyne. Min. Proc. Instn Civ. Engrs, 1908, 174, Apr., 158–221.
42. Anderson D.A. Tyne Bridge, Newcastle. Proc. Instn Civ. Engrs, Part 3, 1929–
1930, 230, 167–183.
43. Wilson W.S. and Smith H.S. Relative methods of sinking bridge foundations.
Institution of Civil Engineers, London, 1950.
44. Arz P. et al., Grundbau. Ernst and Son, Berlin, 1994.
45. Klockner, W. Grundungen BK 1982 Teil II. Ernst and Son, Berlin, 1982.
46. Ward A.M. and Bateson E. The New Howrah Bridge, Calcutta, construction.
J. Instn Civ. Engrs, 1947, 28, May, 167–210.

Bibliography Shafts
Auld F.A. Design of concrete shaft linings. Proc. Instn Civ. Engrs, Part 2, 1979, 67,
Sept., 817–832.
Bills R.F. Developments in shotcrete equipment. Proc. Engng Foundations Conf.
American Concrete Institute, 1977, SP-54, 201–210. American Concrete Institute,
Farmington Hills, Michigan, 1977.
Collins S.P. and Deacon W.G. Shaft sinking by ground freezing. Ely Ouse Essex
scheme. Proc. Instn Civ. Engrs, 1972, suppl. 7, 129–156.
Design and construction of tunnels and shafts. Proc. Australian Conf. on Tunnelling,
Australian Institute of Mining and Metallurgy, Melbourne, 1976.
Harding P.G. Fresh air for Frejus: vent shaft excavation in Europe’s top tunnel.
Tunnels and Tunnelling, 1980, 12, No. 4, May, 25–27.
Lancaster-Jones P.F. Problems of shaft sinking. Tunnels and Tunnelling, 1975, 7, July
Aug., 26–28.
520 Deep excavations

Lyons A.C. and Reed A.J. Modern cast iron tunnel and shaft linings. Proc. 2nd
Rapid Excavation and Tunnelling Conf., San Francisco, 1974, Vol. 1, 669–689.
American Institute of Mining, Metallurgical and Petroleum Engineers, Baltimore,
Maryland.
Prater E.G. Examination of some theories of earth pressure on shaft linings. Canadian
Geotech J., 1977, 14, No. 1, Feb., 91–106.
Provost A.G. and Griswold G.G. Shaft sinking considerations and problems. Proc.
2nd Rapid Excavations and Tunnelling Conf., San Francisco, 1974, Vol. 2, 1095–
1113. American Institute of Mining, Metallurgical and Petroleum Engineers,
Baltimore, Maryland.
Wilson N.E. Designing access shafts for tunnel construction. Engng J., Montreal,
1980, 63, Oct., 12–15.

Caissons
Ameria Pumping Station. Construct. Ind., Int., 1989, 15, No. 5.
Arz P. et al. Grundbau Abschnitt B des Beton Kalenders. Teil II. Ernst and Son, Berlin,
1991.
Chandler J.A. et al. Jamuna River 230kV crossing Bangladesh. Construction of
foundations. Proc. Instn Civ. Engineers, 1984, 76, Nov., 965–984.
Gales R.R. Hardinge Bridge over the Lower Ganges at Sara. Min. Proc. Instn Civ.
Engineers, 1917, paper 4200, 18–99.
Hayward D. Humber Bridge. New Civ. Engnr, 1975, 10 April, 22–24.
Hayward D. Stubborn caisson sinks under 6500 tons. New Civ. Engnr, 1975, 17 July,
16–17.
Hinch L.W. et al. Jamuna River 230kV crossing Bangladesh. Design of foundations.
Proc. Instn Civ. Engrs, 1984, 76, Nov., 927–949.
Hyatt K.E. and Morley G.W. The construction of Kafr-el-Zayat railway bridge.
Institution of Civil Engineers, London, 1952, paper No. 19 works construction.
Krauss F.E. Pneumatic technique for buoyant caissons. J. ASCE Waterways
Harbours Coastal Engng Div., 1973, 99, Feb., 19–26.
Lingenfelster H. Senkkasten Grundbau Taschenbuch Teil 3, 4. Ernst and Son, Berlin,
1992.
Meldner V. Pneumatic caisson work for the bridge across Lillebaelt. Baumasch,
Bautech., 1971, Jul., 289–295.
Pike C.W. and Saurin B.F. Buoyant foundations in soft clay for oil refinery structures
at Grangemouth. Proc. Instn Civ. Engnrs, Part 3, 1952, 1, Dec., 301–334.
Porter T.G. et al. Shaft excavation in soft clay by caisson construction. Proc. Conf.
Retaining Walls. Institution of Civil Engineers, London, 1992, 281–290.
Schwald R. and Schneider H. Tiefgarage Obere Augustinergasse in Konstanz.
Mitteilungsheft der G Baresel, Stuttgart, 1991.
Schwald R. and Schneider H. Gestarte Absenkung eines offener Vortrage der
Baugrundtagung. DGEG, Dresden Heraasg, 1992.
Soil movement due to deep excavations
11
Introduction In earlier times the task of the temporary works engineer was to design the
peripheral soil support to a deep basement excavation to provide an adequate,
but not over-generous, factor of safety against collapse. Risk was identified
in terms of the adequacy of the structural strength of strutting, shoring,
anchoring, sheeting or walling. Addressing the risk of excessive deformation
of sheeting and bracing was frequently not a high priority. Now this has
changed and the provision of deep basement accommodation on urban sites
has raised to a new importance the serviceability design conditions of accep-
table horizontal and vertical soil movement around and below the excavation.
As basements are built to greater depths and building developments occupy
greater plan areas, the problems of subsidence, heave and horizontal soil
movement themselves become priorities. Insurers are no longer prepared to
cover risks of property damage which can be recognized, from previous
experience, as inevitable. This chapter addresses those factors which cause
soil movement around an excavation, typically a large deep basement excava-
tion, the measures which can be taken to alleviate soil movement, and the
methods available to the designer to predict movement.
A recent review by Long1 was made of some 300 case histories of wall and
ground movements due to deep excavations worldwide. Generally this data
base ignored geographical boundaries, and variations in local standards of
specification and workmanship and its limitations in this respect must be
acknowledged. Broadly, the collected information was grouped into four
categories: predominantly stiff to medium-dense soils; predominantly stiff to
medium-dense soils with embedment into a stiff stratum; predominantly
stiff to medium-dense soils with a low safety factor against base heave; and
cantilever work. Further subdivision was made for internally propped walls,
anchored walls, top-down construction and soil strengths. Comparison was
made with the charts of Clough et al.2 and Peck3 and the regional studies in
Oslo (Karlsrud4 ), Taipei, Taiwan (Ou et al.5 ), Singapore (Wong et al.6 ) and
the UK (Carder7 , Fernie and Suckling8 ).
The conclusions reached by Long may include the disadvantages of a wide
sweep of published data but cannot be disregarded bearing in mind the very
large data base collected. Long concluded the following.
For retaining walls in stiff clays with a large safety factor against excavation
base heave:
(a) Normalized maximum lateral movement values h max are frequently
between 0.05%H and 0.25%H where H is the excavation depth.
(b) Normalized maximum vertical settlement values v max are usually
lower, at values frequently between 0 and 0.20%H.
(c) There is no discernible difference in the performance of propped,
anchored or top-down systems.
(d ) The values recorded are somewhat less than would be expected from the
charts produced by Clough and O’Rourke9 , possibly because the soils
in the data base are on average stiffer.
522 Deep excavations

(e) They seem relatively independent of system stiffness and are perhaps
controlled by excavation base heave and limited by arching effects in
these stiff soils.
( f ) The data indicate that less stiff walls may perform adequately in
many instances and worldwide design practice may be somewhat
conservative.
For retaining walls that retain a significant thickness of soft material (greater
than 0.6 of excavation depth) with stiff material at dredge level and where
there is a large safety factor against base heave:
(a) The h max and v max values increase significantly from the stiff soil cases.
(b) The values are close on average to those predicted by Clough and
O’Rourke9 .
(c) There is some promise in the use by Addenbrooke10 of the flexibility
number for the analysis of the collected data.
For retaining walls embedded in a stiff stratum that retain a significant thick-
ness of soft material (greater than 0.6 of excavation depth) and have soft
material at dredge level but where there is a large safety factor against base
heave (determined intuitively):
(a) The h max and v max values increase significantly from the situation
where stiff soil exists at dredge level.
(b) The Clough et al. charts considerably underestimate movements.
In cases where there is a low safety factor against base heave, large movements
(h max to 3.2% of excavation depth) have been recorded. The data mostly fall
within the limited values suggested by Mana and Clough11 and it is suggested
that the relationship between movement, system stiffness and safety factor
proposed by Clough et al. form a good starting point for preliminary esti-
mates of system performance.
For cantilever walls, the normalized maximum lateral movements:
(a) are relatively modest and average 0.36% of excavated depth
(b) are surprisingly independent of excavation depth and system stiffness
(c) suggest that less stiff walls would perform adequately in many cases.
To relate these findings to a site-specific assessment of wall displacement may
prove difficult or indeed, impossible. The findings by Long represent his
conclusions from a wide data base but may not accurately represent risk of
wall displacement on a particular site.

Factors that influence The principal factors that determine the extent of soil deformation have been
soil movement listed for conditions in Hong Kong12 . For wider geographical application the
list of factors influencing soil deformation around a deep basement excavation
is slightly longer:
(a) effects of stress changes within the subsoil
(b) dimensions of the excavation
(c) soil properties
(d ) initial horizontal stresses within the soil
(e) groundwater conditions and changes to them
(f) stiffness of the sheeting and bracing system
(g) effects of pre-load in bracing and anchoring
(h) construction methods
(i ) construction workmanship.
Soil movement due to deep excavations 523

This list is not given in any order of priority since the importance of each
factor varies from job to job. The list is now examined in more detail.

Effects of stress changes within the subsoil


The complexity of stress changes in four elements in an over-consolidated clay
which is supported by a diaphragm wall during excavation are shown in
Fig. 11.1 (after Gaba et al.)13 . The locations of the four elements are:
. Element A: immediately behind the wall
. Element B: immediately in front of the wall
. Element C: beneath the centre of the excavation, some distance from the
wall
. Element D: behind the wall and remote from it.
The initial pore pressures, prior to installation of the wall are hydrostatic
below an in situ groundwater level.
The changes in the short term during construction and in the long term as
steady seepage is established are summarized in Table 11.1. In more detail the
progressive changes in stress and pore-water pressure for elements A and B are
as follows.
At soil element A
. Over-consolidation of soil in geological time, following deposition and
removal of overburden, no further deposition of recent deposits: stress
path 00 to 0. (K0 greater than 1.)
. Wall excavation below slurry: reduced lateral total stress, pore-water
pressure reduces: stress path 0 to 1.
. Wall concreting: increase in lateral total stress, pore-water pressure
increases to approximate in situ values: stress path 1 to 2.
. Excavation in front of wall: wall moves forward, horizontal total stress
reduced with reduction in pore-water pressure. Following yield, exces-
sive negative pore-water pressure occurs: stress path 2 to 3.
. Steady-state seepage develops with time (as the permeability of the soil
fabric allows). Long-term steady-state seepage pore-water pressure is less
than the initial hydrostatic value but probably greater than pore-water

Fig. 11.1. Stress changes in four elements in an over-consolidated clay during excavation13
524 Deep excavations

Table 11.1 Changes in the short term during excavation and construction and in the long term as steady state seepage is
established13

Element A Element B Element C Element Da

Vertical total stress during excavation Constant Decreases Decreases Unchanged


Horizontal total stress during excavation Decreases Decreases due to Decreases Unchanged
unloading.
Increases due to
wall movement
Pore-water pressure during excavation Decreases Decreases Decreases See noteb
Pore-water pressure in the long term Probably Increases Decreases See noteb
increases
Undrained shear strength in the long term Probably Decreases Decreases Unchanged
decreases
Strain during excavation Vertical Vertical Vertical Unchanged
compression extension extension
Strain in the long term Vertical Vertical Vertical Unchanged
compression extension extension
Notes
a
Assumed to be located sufficiently remotely from the wall so as not to be affected by changes in soil stress due to excavation in front
of the wall.
b
Depends on ground permeability.

pressure immediately after excavation. Pore pressure increasing in the


long term as steady-state seepage is established: stress path 3 to 4.
Soil element A experiences an overall increase in vertical effective stress and a
decrease in horizontal effective stress to bring element A into an active state.
At soil element B
. Excavation makes a large reduction in vertical total stress with a large
reduction of pore pressure. Groundwater at formation level.
. Movement of the wall below formation level towards the soil in front of
the wall increases horizontal total stress and is likely to result overall
in an increase in horizontal effective stress and a reduction in vertical
effective stress during excavation.
. Steady-state seepage develops with time, pore pressures increase
reducing vertical and horizontal effective stresses: stress path 3 to 4.
The vital factor influencing the horizontal movement of soil below formation
level, and therefore the magnitude and extent of vertical settlement is the
proximity of the unloading stress path of element B to the failure envelope.
If the stress path 2 to 4 is well within the passive failure envelope, this
shows that the yield is small and both heave and the resulting horizontal
soil movement will also be small. Conversely, if the effective stress points
for element B are close to the failure envelope this indicates risk of excessive
yield, local passive failure and high lateral movements.
A summary of the stress changes that occur at soil elements A, B, C and D
(from reference13 ) is given in Table 11.1.

Dimensions of the excavation


The plan shape, the plan area and the excavation depth all critically influence
the extent and distribution of soil movement around and below a basement
excavation in given soil conditions. The depth obviously affects movement;
Tomlinson14 referred to unavoidable inward movement in normally strutted
Soil movement due to deep excavations 525

or anchored excavations of the order of 0.25% of excavation depth in soft


clays and 0.05% in dense granular soils or stiff clays. It is usual to assume
that the volume of horizontal soil movement at the sheeting within a unit
length of excavation support is approximately equal to the volume of vertical
soil movement at ground level over the same unit length. As a rule of thumb,
horizontal soil movements are likely to extend to a maximum lateral dimen-
sion of two to three times the excavation depth. The deformed soil profile
therefore begins to take shape, although changes with time (due to pore-
water pressure dissipation) and the effects of irregular plan shape complicate
a simple assessment of settlement risk.

Soil properties
Soil properties were summarized by Peck3 . Figure 11.2, after Peck, shows
smaller wall movements and ground settlements in stiffer soils (such as
granular soils and stiff clays) than in softer soils (e.g. soft and medium clays
and loose silts).
As reported by Long, soil movements due to excavations in soft clays may
prove to be embarrassingly large, particularly where the clays have been
assumed incorrectly to be isotropic. Clough et al.15 and Mana and Clough11
showed the rate and magnitude of lateral wall movement both increase rapidly
as the risk of base heave increases and the factor of safety against base failure
reaches unity.
Overall deformation in terms of heave below the excavation and vertical
settlement around it will depend on many factors including soil stiffness
and, in weaker soils, soil strength. In weak clays and loose silts, yield in soil
zones may result in providing passive resistance to peripheral sheeting or
walling, with large movements resulting. From a practical viewpoint, in
loose cohesionless soils with high piezometric pressure due to groundwater,
excavation conditions may be close to quick conditions with risk of vertical
soil subsidence and loss of ground between timbering, sheet piles or dia-
phragm wall joints. Soil and groundwater conditions, therefore, pre-empt

Fig. 11.2. Observed


settlements behind
excavations3
526 Deep excavations

all other factors as the prime critical risk of soil movement around deep
excavations.

Initial horizontal stresses within the soil


Where high, locked-in horizontal stresses exist within soils, typically within
over-consolidated clays, soil deformations surrounding excavations increase,
even at relatively shallow depths. For soils with comparatively low values for
coefficient of earth pressure at rest K0 , deformations are much less16 .

Groundwater conditions
The effects of groundwater on soil settlement are varied and occur at different
stages of excavation. Where sheeting penetrates a cohesionless stratum but
does not achieve a cut-off at depth, a steady groundwater seepage condition
will develop whereby flow is established beneath the sheeting and upwards
to the formation level of the excavation. This flow causes a decrease in
groundwater pressure, an increase in effective stress and settlement outside
the periphery of the excavation. At the same time passive resistance reduces
due to the upward flow on the inside of the sheeting, and further horizontal
movement occurs as sufficient passive resistance is mobilized. The establish-
ment of a steady-state groundwater regime therefore causes both vertical
and horizontal soil movement.
Where dewatering of sheeted excavations causes drawdown to the exterior
groundwater table, again where the sheeters to the excavation do not make an
adequate cut-off at full penetration, effective vertical soil pressure increases,
resulting in vertical settlement. Since the drawdown is greatest near the
excavation and reduces progressively with increasing distance from it, this
settlement profile will be similar in shape to that due to relief of overburden
by the excavation itself.

Stiffness of the support system


Parametric studies using Winkler spring or finite element soil–structure inter-
active programs and observations made on site show that the exterior ground
settlement profile surrounding a sheeted excavation reduces as the stiffness of
the sheeting and the bracing supporting it increase. The elastic stiffness of the
bracing system appears to be most important. The vertical embedment of the
sheeting beneath formation level will also materially alter the effective stiffness
of the sheeting and influence external soil movement, both vertically and
laterally.
A study of the effects of wall stiffness, bracing stiffness, vertical spacing of
supports and embedment was reported by Goldberg et al.17 ; a summary of the
results is shown in Fig. 11.3 in which the stability number is plotted against the
stiffness parameter. The data presented also suggest that sheeting stiffness and
support spacing effectively influence external soil movements.
Experience over some years in temporary works design using a Winkler
spring program has confirmed site observation that increasing strut stiffness
decreases external soil movements, although less effectively at very high
values of stiffness. These findings regarding the practical importance of
sheeting and strutting stiffness are not confirmed by Clough and Davidson18
nor by Tomlinson14 . These authors stated that the amount of yielding for
any given depth of excavation is a function of the characteristics of the
supported soil and not of the stiffness of the supports. Tomlinson referred
to steel structural members, even of heavy section, as being insufficiently
stiff to reduce yielding by any significant amount. Reinforced concrete
diaphragm walls, he noted, deflect by amounts similar to those experienced
with sheet pile walls. This similarity has not been the experience of the
Soil movement due to deep excavations 527

Fig. 11.3. Effects of wall


stiffness and support spacing
on lateral wall movements17

Author and, although it is agreed that soil stiffness should be regarded an


important factor, it is considered that both sheeting and strutting stiffness
contribute significantly to the extent that soil behind the wall is allowed to
move.
In summary, the cross-sectional area of the soil vertical settlement trough
outside the basement walls is approximately equal to the cross-sectional
area of the horizontal soil deformation curve at the wall relative to the original
wall line. The deformed shape of the walls, and of the soil immediately
adjacent to it, is made up of deformation between prop levels (and wall defor-
mation prior to the insertion of props) together with inward deformation due
to the compression of the props. Prop compression occurs as all forces on the
inside and outside of the wall reach equilibrium, earth pressures on the outside
face reducing from at-rest earth pressures, and passive pressures on the inside
face below dredge level being progressively mobilized as horizontal deforma-
tion occurs. The balancing forces mobilized in the props and in the soil below
dredge level cause horizontal movement the amount of which depends on the
stiffness of the props and the stiffness of the soil in compression below dredge
level.

Effects of pre-loading
Within the Author’s experience the pre-loading of bracing to deep excavations
has been beneficial in reducing settlements outside the excavation in a variety
of granular soils and medium–stiff to stiff clays using relatively stiff sheeting
and walling such as steel sheet piling and reinforced concrete diaphragm wall-
ing. The pre-load tightens the bracing system and thereby reduces one cause of
horizontal movement and vertical settlement. Apart from this practical
improvement, movement is further reduced due to improving soil stiffness
caused by the hysteresis effect on the soil stress–strain curve as the soil is
progressively unloaded in shear at each frame level (due to pre-loading) and
reloaded for the next bracing frame.
The advantages of reduced soil settlement by pre-loading are not fully
accepted by some temporary works designers, although O’Rourke19 summar-
ized his views by stating that in most instances of cross-lot bracing pre-loaded
to 50% of its design load, further movement would be prevented at that frame
level, and overstress of the bracing at that frame was unlikely if the pre-load
was limited to 50% of the design value.
528 Deep excavations

The pre-loading of bracing can be compared with the pre-stressing of


ground anchor tendons. The Author’s practice over a long period has been
to pre-load anchors to the design load to the completed excavation, calculated
on the basis of the trapezoidal strut load envelope recommended by Peck3
and described in Chapter 7. These loads are applied successively by post-
tensioning jacks at each frame level as the excavation reaches that level. A
10% overload is applied at each anchor to allow for slippage and creep
within the tendon. Winkler spring analysis is then used to check sheeting
stresses and deformations with the applied anchor loads for each stage of
the excavation. No excessive movement or overstress has been experienced
within this procedure.
The efficiency of pre-loading tendons in reducing vertical settlements out-
side the periphery of the excavation is related to anchor length. Where the
fixed length of the anchor is located within the zone of soil movement
caused by the bulk excavation, only a limited reduction in settlement may
be expected. This effect is most likely in cohesive strata, while in sands and
gravels pre-loaded anchors with a fixed grouted length at least beyond the
theoretical Coulomb wedge (at an angle of 458 þ =2 to the horizontal)
may be expected to effectively limit horizontal soil movement and vertical
settlement.

Construction methods
The choice of overall construction method for the basement, either top-down-
wards or bottom-upwards, the technique used for walling or sheeting the base-
ment periphery, and the period taken for each excavation stage, all influence
the extent of soil movement around the excavation for given groundwater
conditions and basement dimensions.
The top-downwards method, using basement floors to successively prop
peripheral diaphragm or piled walls, is frequently used to restrict vertical
soil settlements. No comparative soil measurements are available to substanti-
ate this view, although certain features of the method would appear to mini-
mize soil movement. These include the regular propping to the exterior wall
afforded by the floor at each storey height, the considerable elastic stiffness
of this prop, and the avoidance of movement involved in repropping. (The
exterior wall is repropped when conventional cross-lot bracing is replaced
by temporary supports between the exterior sheeting or walls and the sub-
structure as the permanent works progress in bottom-upwards construction.)
Closer examination of a particular site may, however, reveal a less satisfactory
situation. The regularity of support provided by the floors to the exterior
walling at each storey height would not necessarily provide support at the
optimum levels, especially where external surcharge loads are applied from,
say, existing foundations to adjacent structures or where the height from
the penultimate support to formation level should be minimized, say where
soft or weak strata exist immediately below final formation. Again, some
contractors prefer to construct the exterior walls and excavate to first base-
ment floor level without casting the ground floor. This procedure loses
some advantage of the method in restricting soil movement. Potts et al.20
showed the results of numerical analysis, concluding that the use of temporary
soil berms to support cantilevered external walls from ground floor to
first basement floor level only partly reduces the extra settlement caused
by excavation to first basement level prior to commencement of floor
construction.
The choice of walling or sheeting and its method of installation also
influence the extent of vertical soil settlement. With some methods, such as
walls from vertical soldier piles and horizontal laggings, the loss of ground
Soil movement due to deep excavations 529

caused by the need to have an open face of excavation as the laggings


are placed is likely to cause greater settlements than those due to other walling
methods. Similarly, where sheet piles are used in granular soils and
heavy vibratory or percussive installation methods are chosen in preference
to, say, assistance by water jetting, the resulting soil settlement may be
significant.

Construction workmanship
Published records of the influence of inadequate construction standards,
usually workmanship standards, are numerous and only serve to confirm
the common sense knowledge of site staff that short cuts and sloppy attitudes
towards workmanship in timbering and excavation support works inevitably
lead to support movement, soil subsidence and even local failure and pro-
gressive collapse. The designer must bear some responsibility with regard to
standards of workmanship. For example, the materials chosen for construct-
ing excavation support should be the best available at site, the chosen method
of walling or sheeting should comply with the experience of the supervisors
and operatives that are to build it, and construction details such as site-
welded connections, reinforcement fixing and stressing works should all be
related to the available site skills.
Many causes of additional movements (and failure) of excavations due to
bad site practices appear in reference 1. These include late installation of
supports, over-excavation, poor pile driving and caisson construction, loss
of water through holes for tie-backs and joints or sheet pile interlocks and
diaphragm wall joints leading to loss of ground, remoulding and undercutting
of clay berms, and excessive surcharge loads from spoil heaps and construc-
tion equipment. Many more items of inadequate workmanship or supervision
standards that can cause movement, subsidence or collapse can be added to
the list. In particular, the lack of rigidity and tightness of shores and braces
are important causes of wall and soil movement. Failure to provide or tighten
wedges between walling and walings is a significant cause of movement and
subsidence. Similarly, with king post walls, failure to efficiently wedge hori-
zontal laggings to vertical soldiers and ensure good uniform contact between
soil and laggings is a direct cause of soil subsidence behind the wall. Peck3
pointed out that the choice of detail of lagging connection to soldier could
cause settlements adjacent to the excavation to vary widely; settlements
adjacent to walls using the detail in Fig. 4.6(b) were three times those using
the detail in Fig. 4.6(a).

Measuring techniques The measurement of small displacements and angular rotations of surfaces
and their accuracy and existing structures has required the development of existing surveying
techniques and the use of electronic techniques. A general review of such
methods was given by Dunnicliff21 . More recently, a review of the methods
of field measurements made on greenfield sites and existing structures on
the extension to the Jubilee Line metro extension was made by Standing
et al.22 Examples of the best accuracies for the different measuring systems
were given in this paper and are reproduced in Table 11.2 as a guide to the
best expectation of accuracy of measurement for such techniques elsewhere.

Measures to reduce To ensure minimum soil movement horizontally and vertically, around and
soil movement at the below a deep excavation of given dimensions in given soil conditions, several
curtilage of a deep measures are necessary. Not all may prove to be financially worthwhile, but
excavation they are:
530 Deep excavations

Table 11.2 Examples of the best accuracies for measuring systems22

Instrument type Building example Resolution Precision Accuracy


(monitoring method)

Precise level (NA 3003) Treasury, Palace of Westminster 0.01 mm 0.1 mm 0.2 mm
Total station (TC 2002) Ritz: (vertical displacement) 0.1 mm 0.5 mm 0.5 mm
Ritz: (horizontal displacement) 0.1 mm 1 mm 1 mm
(angular displacement) 0.1 arcsec 2 arcsec 5 arcsec
Photogrammetry Elizabeth House 1 mm 1 mm 2 mm
Tape extensometer Elizabeth House 0.01 mm 0.03 mm 0.2 mm
Demec gauge Palace of Westminster 0.001 mm 0.01 mm 0.01 mm
Rod extensometer Elizabeth House 0.001 mm 0.01 mm 0.2 mm
Electrolevel Elizabeth House 2 arcsec 10 arcsec 10 arcsec

(a) provide a wall support which provides both temporary and permanent
soil support
(b) make the sheeting or wall support flexurally stiff
(c) avoid installation vibration or other causes of loss of ground
(d ) ensure the wall has adequate embedment in a stiff stratum
(e) ensure the wall receives support at frequent vertical centres and reduce
these centres progressively with depth
( f ) locate the lowest support near formation level
(g) make the bracing stiff in compression
(h) pre-load the bracing or pre-tension the ground anchors
(i) avoid delays in construction of either walling or bracing, avoid keeping
diaphragm wall panels open for long periods and avoid delays in
bracing works or anchor installation at each support level
( j) avoid any loss of ground by over-excavation or removal of fines during
pumping
(k) avoid drawdown caused by dewatering outside the basement
(l ) in weak soils, improve ground conditions below formation level to
ensure adequate passive resistance inside the sheeting from soil with
sufficient strength and high stiffness (such improvement could be
made by localized jet grouting, pin piles, mix-in-place piles or vibro-
replacement).
For deep basements, where soil conditions permit a cut-off against ground-
water ingress, the top-downwards method of construction may prove attrac-
tive in meeting some of these criteria to reduce external soil settlement. The
method has disadvantages, however, including high excavation costs to
remove soil from below basement floor construction, the risk of overall
delay caused by any local hold-up in a sequence of interdependent construc-
tion activities, and the problems in terms of space and access of several specia-
list firms working on site at the same time.
In shallower basements the use of top-downwards construction may be
prohibitively expensive. In such circumstances, the risk of excessive settle-
ments around the site will be minimized by the above methods. In particular,
cantilevered walls and excessively high sheeting are a frequent cause of
excessive soil movement outside the excavation and should be avoided
where possible by propping the sheeting from temporary bases or from a
previously constructed raft at the centre of the basement plan shape. Where
walls are cantilevered at any stage individual piles or diaphragm panels
Soil movement due to deep excavations 531

Fig. 11.4. Relationship between


maximum ground settlement
and stability number at the end
of construction after berm
excavation for berms of varying
batter: (a) fully penetrating,
fixed-end MZ-27 sheet pile wall;
(b) partially penetrating,
free-end MZ-27 sheet pile wall23

within the basement wall should be connected by a stiff capping beam in


reinforced concrete.
The use of soil berms to minimize lateral movements of walls or sheeters at
the periphery of a deep excavation should be noted. The general consensus on
the use of berms23;24 is that the increase in vertical stress using a relatively
small volume of soil is often sufficient to reduce lateral movements to walls
or sheeters by 50% while the berm is left in place. If the berm is removed in
short lengths while rakers or struts are placed, the final lateral movement,
and thence the vertical settlement of soil outside the excavation, can be use-
fully reduced. Numerical studies by Clough and Denby23 on an excavation
in soft to medium clay showed a theoretical relationship (Fig. 11.4) between
settlements behind sheet piles with berms and the stability number H=cub ,
where cub is the undrained shear strength at the base of the excavation for
the condition after the berm has been removed and the rakers installed. The
reduction in ground settlement increases as the stability number increases,
and at high stability numbers increasing berm size leads to larger reduction
in settlements. This apparent improvement may not be produced, however,
where deep-seated movements occur at high-stability numbers with low-
strength clays.
Burland et al.24 showed the effect of a soil berm at one stage of a 16 m deep
basement excavation in London. The peripheral soils were supported by a
diaphragm wall with a depth of embedment 3 m below final formation level.
The initial excavation was 10 m deep, at which depth a thick waling slab
532 Deep excavations

Fig. 11.5. Observations of wall


lateral movement; effect of soil
berm24

was cast on the exposed surface of the London clay. Excavation was then
made to the full depth leaving a soil berm below the waling slab, as shown
in Fig. 11.5. After the basement raft had been concreted in the central area
the berm was removed in short lengths and the raft completed to support
the wall. Observations of lateral movement of the wall shown in Fig. 11.5
indicate the efficiency of the berm in reducing wall movement before its
removal. Burland et al. considered that the weight of berms could be very
effective in controlling the softening of the clay, and hence the movements
and stability of the toe of the wall. Burland et al. also gave the results of a
numerical analysis showing the use of berms to reduce movements while
excavating the lowest and fourth basement floor spaces to an excavation
within London clay. The berm, removed in short lengths and replaced by
the propping action of the basement floor, contributes an important shear
connection with the diaphragm wall and adds significant surcharge to the
soil below final formation level, providing passive resistance to the toe of
the wall.
Two recent papers describing strut load measurements, at Mayfair car
park25a and Canary Wharf underground station25b , London have highlighted
the influence of corner stiffness of walling in rectilinear excavations and
the discrepancy between 2D and 3D analyses. The effect of stiffness of
corner panels to relieve load from corner brace struts and reduce lateral
wall displacements within the plan length of the walling has been long
accepted by designers, but field measurements of strut loads and 3D finite
element analyses now confirm this. The calculated reduction in displacement
of walls due to three-dimensional analysis is also discussed by Simic and
French26 .
Where soil movements at the outside of a deep excavation are a vital con-
sideration and space allows a wall thickness substantially greater than usual
but the proximity of nearby structures prohibits the use of anchors, a cellular
type of wall may be considered. Although not cost-effective in terms of area of
diaphragm wall per linear metre of completed wall, the cellular wall, utilizing
the self-weight of the enclosed soil for stabilization, is remarkably efficient in
limiting horizontal soil movements and vertical settlement behind the wall.
Plan forms of a T-panel wall and a double-flange cellular wall were shown
in Fig. 8.44.
In Medinah, Saudi Arabia, a cellular diaphragm wall was built to avoid soil
deformation at the rear of a 17.5 m deep excavation which was required not to
be obstructed by temporary raking shores. The excavation was 100 m wide
Soil movement due to deep excavations 533

and more than 1.5 km long. Cross-lot strutting was therefore not practical and
ground anchor capacities were uneconomical in the conditions. The
excavation was through silty clays and silty sands overlying bedrock up to
55 m deep in close proximity to precious religious shrines. The maximum
acceptable lateral soil movement immediately behind the wall, either in the
short term or prior to placing permanent props (with basement floors) some
18 months after excavation, was 25 mm. The maximum acceptable lateral
soil movement in the proximity of piles to existing substructures within
50 m of the perimeter of the deep excavation was 5 mm. The cellular dia-
phragm wall, shown in Fig. 11.6, was analysed using finite element methods
and measured small strain modulus values with a non-linear finite elastic–
plastic numerical model. The combined immediate and drained soil move-
ments were shown to be less than these maximum values. The movement of
the wall as-built was significantly less than that predicted, as shown in Fig.
11.7. Difficulties were experienced modelling the effect of the shear stiffness
of the cross-walls in the two-dimensional model, and only when this stiffness
was reduced was a deformed shape produced by the analysis typical of a
cantilever wall.
The proximity of nearby buildings required major temporary works of a
different kind during the substructure construction of a large arts centre at
the Barbican in London in the 1970s (Fig. 11.8)27 . Nearby tower blocks,
although piled, were susceptible to tilt caused by soil movement during bulk
excavation for the theatre, but even more importantly, analyses predicted
excessive shear stresses in these piles if significant soil movement were allowed
between the piles due to the excavation. Even though cellular diaphragm walls
of considerable stiffness were designed at each side of the theatre basement,
the predicted horizontal inward movement of these walls below formation
level exceeded the maximum that the pile shear could withstand. To prevent
this movement below formation level, two props were constructed in tunnels
between the cellular walls and pre-loaded with thrusts up to 10 000 kN. The
length of the north wall of the theatre exceeds 60 m, with a minimum height
of 14 m. The wall was designed to be propped apart by the two diaphragm
walls at the east and west sides of the theatre basement and at low level by
the two pre-loaded tunnel props, as shown in Fig. 11.8. A horizontal waling
was formed by the arch slab at the 6 m level spanning across the whole base-
ment width. The wall itself spans horizontally across the low-level supports
and vertically between the arch slab and the prestressed concrete beam
within the wall at lower level. Measurements during construction showed
that the north theatre wall was moved northwards by a maximum of 10 mm
and the south wall moved southwards by 5 mm. The jacks were maintained
in an operable state for one year after the basement excavation and were
then stabilized by the exchange of hydraulic fluid with epoxy resin grout with-
out loss of pressure. The jacks had therefore fulfilled their purpose and instead
of soil movement towards the theatre excavation, the pre-loaded tunnel props
caused small movement in the opposite direction. The essential point in this
basement design was the risk, avoided by the use of the pre-loaded tunnel
props, of progressive wall/soil movement below the excavated level of the
theatre basement as it was dug out, the stiffness of the peripheral cellular
walls below excavated level being insufficient to reduce it to acceptable
levels without the action of the pre-loaded props.
A later, but similar, use of tunnelled struts was made for basement construc-
tion at Westminster station, London. Stringent settlement criteria were neces-
sary to avoid damage to the Big Ben clock tower, the adjacent metro tunnels,
trunk sewer and nearby parliamentary buildings. Top-downwards construc-
tion was used for basement construction but risk of soil deformation below
534 Deep excavations

Fig. 11.6. Medinah car park: (a) plan of cellular wall construction; (b) design soil profile; (c) design cross-section; (d) key plan
(dimensions are in m)
Soil movement due to deep excavations 535

Fig. 11.7. Medinah car park: (a) key plan; (b) finite element mesh; (c) predicted horizontal soil movement due to panel
excavation; (d) predicted horizontal soil movement due to bulk excavation; (e) observed horizontal soil movement at rear of
wall by inclinometer E
536 Deep excavations

Fig. 11.8. Barbican Arts Centre: site plan and cross-section of site during excavation showing use of pre-loaded tunnel
props27

the final excavation level required structural support to the diaphragm wall
box below that level. A cross-section is shown in Fig. 11.9. Both low-level
diaphragm cross-walls and tunnel struts were considered as alternative
means of strutting. The low-level tunnel struts were finally adopted because
of the risk of poor contact between the outer box diaphragm wall and the dia-
phragm cross-wall together with the difficulty of installing jacking equipment
under 40 m head of bentonite slurry. Three hand-dug tunnel struts, 1770 mm
dia. were used, lined with precast concrete segments and filled with reinforced
concrete. Hammerhead walings, 1800 mm deep were constructed at the ends
of the struts. Access for tunnelling was gained from two 3 m diameter lined
pile shafts. At each strut a jacking chamber, 2440 mm dia., was constructed
with a jack capacity of 38 000 kN and a stroke of up to 50 mm.
Soil movement due to deep excavations 537

Fig. 11.9. Westminster underground station, London. Use of tunnel struts in top-down construction33
(continued p 538)

Methods of predicting Soil movement behind a supported excavation can be predicted empirically,
soil movement semi-empirically, by finite element or finite difference methods, or by other
methods such as velocity fields.

Empirical methods
The risk of settlements in the vicinity of proposed deep excavations can be
assessed, in broad terms, from published data from sites in similar soil condi-
tions. The most useful records include those published by Peck3 , O’Rourke
et al.28 and others2931 .
Peck’s work is summarized in Fig. 11.10(a) showing vertical settlement (as a
percentage of excavation depth) against distance from the excavation (plotted
non-dimensionally as a ratio of excavation depth). Peck used this plot to draw
attention to the distances from the cut at which settlement occurs, and to the
experience that settlements in plastic clays were likely to be greater than in
cohesive soils and stiff clays. Both immediate and consolidation settlements
are included in the settlement data in Fig. 11.10. It should be noted that in
very soft to soft clays, settlements as great as 0.2% of the excavation depth
can occur at distances of three or four times the depth. The critical influence
of excavation depth on vertical settlement in shown in Fig. 11.10(b) for base-
ments in Chicago soils, generally supported by sheet piling with small embed-
ment and cross-lot strutting, or more usually with rakers. The upper 5 m of
soil in downtown Chicago consists of fill and sand underlain by a soft clay,
becoming stiffer with depth until hardpan is met at 23 m. The single-storey
basements shown therefore do not penetrate into the soft clays and the
recorded settlements were probably caused by the caisson construction on
which the basements were founded rather than by the basement excavation.
The care required in extrapolating data obtained from one set of soil condi-
tions to another site with an inexact match of soil conditions is self-evident.
538 Deep excavations

Fig. 11.9. Continued


Soil movement due to deep excavations 539

Fig. 11.10. Summary of settlements adjacent to open cuts in various soils as a function of distance from edge of excavation:
(b) settlement associated with foundation construction in Chicago: summary of results and settlements as a function of
excavation depth3

The data published by Peck serve only to show the order of settlement and the
extent to which such settlements are likely to occur in soft clays.
O’Rourke et al.28 published settlement data for excavations supported by
soldier piles and horizontal laggings with cross-lot strutting, in dense sand
and interbedded clays in Washington, DC (Fig. 11.11). In these conditions,
maximum settlements of the order of 0.3% of excavation depth were recorded
immediately to the rear of the sheeting and extended up to twice the excava-
tion depth laterally from the rear of the excavation. O’Rourke et al. also pub-
lished records of settlement readings from Chicago which, while similar to the
earlier data by Peck, showed three zones of settlement related to the salient
construction characteristics. These plots are shown in Fig. 11.12.
Using these data, O’Rourke19 published further work on settlement due to
braced excavations. He considered the pattern of stress in three stages of exca-
vation and strutting (see Fig. 11.13).
(a) Stage 1, initial excavation before strutting. The cantilevered sheeting
deforms with horizontal soil strains and strain contours of triangular
shape, decreasing with depth and distance from the sheeting.
(b) Stage 2, excavation to formation level. After installation of the top frame
support, lateral movement is prevented at that level, but further inward
540 Deep excavations

Fig. 11.11. Summary of


measured settlements adjacent
to strutted excavations in
Washington, DC28

wall movement at lower levels caused by further excavations produces


tensile strains at approximately 458 to the vertical.
(c) Stage 3, replacement of temporary supports. As the lower struts are pro-
gressively removed to build the permanent structure there is further
inward movement of the wall. As the upper supports are removed,
the lower part of the sheeting is held by re-strutting to the permanent
structure and movement associated with a cantilever wall deflection

Fig. 11.12. Summary of settlements adjacent to strutted excavations in Chicago28


Soil movement due to deep excavations 541

Fig. 11.13. Horizontal strains as measured at successive stages of strutted excavations: (a) stage 1, initial excavation;
(b) stage 2, excavation to subgrade; (c) stage 3, removal of struts; (d) principal components of wall movement (bold arrows
show movement vectors)19

occurs in the upper parts of the wall. The strain contours in this stage
are therefore those from a combination of the inward movement of
the sheeting at depth and to cantilever action at the higher levels.
O’Rourke19 developed a relationship between the deformed wall slope and
the vertical settlement of the ground at the rear of the wall and defined a
coefficient of deformation as the ratio of maximum movement due to the
cantilever action of the wall to the maximum total lateral movement of
the wall, including elastic bulging of the wall (as shown in Fig. 11.14).
These data referred to walls founded in stiff strata and where lateral soil
movement below the sheeting towards the excavation was minimal. This
relationship could therefore be used to estimate maximum vertical settlement
from calculated values of horizontal deformation of the wall. However, no
vertical settlement profile results from this computation, and the maximum
value of settlement, at the rear of the sheeting, is of limited use on its
own.
Clough et al.32 and Mana and Clough11 examined data from sheet pile walls
and king post walls in clays supported by cross-lot struttings with either free-
end or fixed-end support. The results (in Fig. 11.15) show the relationship
between maximum lateral wall movement and the factor of safety against
basal failure by heave. Lateral movements of sheeting are shown to increase
very rapidly below a factor of safety of 2.
542 Deep excavations

Fig. 11.14. Ratio of horizontal to


vertical soil movement as
function of the coefficient of
deformation19

Fig. 11.15. Empirical


relationship between the factor
of safety against basal heave
and non-dimensional maximum
lateral wall movement15
Soil movement due to deep excavations 543

Fig. 11.16. Empirical


relationship between maximum
ground settlement and
maximum lateral wall
movements11

Mana and Clough11 produced an empirical relationship between maximum


ground settlement and maximum wall movement from data in varied overall
ground conditions in clays in San Francisco, Oslo and Chicago (Fig. 11.16).
Perhaps the limited conclusion to be drawn from this plot is that maximum
vertical settlements appear most likely to be equivalent to maximum hori-
zontal displacements in clays.
Clough had earlier summarized empirical data on anchored sheeting and
walls (Fig. 11.17) for subsoils varying from sands and silts to stiff clays and
shales to soft clays. Most values of maximum movement remain below 1%
of excavated depth and no significant variation is shown with soil type.
Clough31 suggested that the maximum reduction in soil movement using
pre-stressed anchors was achieved with pre-stress forces obtained from
ground pressures slightly greater than those advised by Terzaghi and Peck34 .

Fig. 11.17. Observed


movements of anchored wall
systems for varying soil types31
544 Deep excavations

Fig. 11.18. Semi-empirical


method to estimate settlement
in sands: (a) ground settlement
adjacent to wall; (b) variation of
settlement ratio with soil
properties37

Semi-empirical methods
Several methods have been devised which enable the settlement profile at the
rear of the wall supporting a deep excavation to be calculated from empirical
relationships determined for the lateral movement of the wall. Caspe35
published a method of analysis which related the settlement profile to the
deflected shape of the wall. In this method:
(a) there is a surface behind the wall which defines the limit of soil deforma-
tion due to the excavation
(b) a variation in horizontal strain in the soil between this no-strain surface
and the wall is assumed
(c) at all locations, vertical strain is assumed to be related to horizontal
strain by Poisson’s ratio.
Others have commented that for plane strain conditions this last assumption
is incorrect and the relationship between vertical and horizontal strain should
be expressed by the ratio v=ð1  vÞ. Caspe’s method was altered by Bowles36
to take this into account, with reasonable agreement between the calculated
settlement profile and site measurement.
A further semi-empirical method devised by Bauer37 is shown in Fig. 11.18.
This method, applicable to excavation in sands, was claimed to show reason-
able fit of settlement profiles with site movements, although the calculated
width of settlement influence appears to limit the lateral extent of this zone
to less than the excavated depth for practical values of .

Finite element and finite difference methods


Numerical methods using finite elements or finite differences allow a soil–
structure analysis. As mathematical tools these methods provide convenient
two-dimensional plane strain solutions (three-dimensional soil–structure
solutions are increasingly becoming available) and use commercially available
programs (such as PLAXIS and CRISP) or in-house programs developed by
academic or professional organizations (such as ICFEP from Imperial
Soil movement due to deep excavations 545

College, London). The methods attempt to address all theoretical require-


ments with boundary conditions that realistically model the site problem
and incorporate, for instance, a stage-by-stage simulation as the excavation
progresses, including time-related aspects such as dissipation of excess pore
pressure. Displacements are the primary unknown solved by the methods,
so prediction of horizontal displacement and vertical settlements fall con-
veniently to this solution.
Finite element packages generally offer the user a choice of constitutive
models, ranging from simple elastic models to sophisticated non-linear
elastic–plastic models. The final choice of model will depend on the
accuracy required of the prediction and the availability of appropriate
input data, particularly with regard to soil parameters. Some of the issues
facing the designer in the choice of constitutive model were raised by
Woods and Clayton38 and included two items related to soil stiffness:
linearity and small-strain behaviour. Although the solution, using the
simple linear model, has been available for many years, it has always been
appreciated that most natural soils are of non-linear nature, even at the
very low-strain values that occur in wall deformation and settlement profile
prediction. In addition, the use of finite element programs to predict
movement around excavations has been shown generally to exaggerate
deformation unless soil stiffness at very small strain volumes is used in the
analysis. To obtain these values, specific measurement procedures have
been designed for use in the triaxial test. Even so, choice of a suitable average
operational strain level is necessary, particularly for soils from previously
undeveloped areas. Where the excavation is near previous sites where meas-
ures have been made, back analysis will provide appropriate soil stiffness
parameters, providing the excavation and subsoil conditions are similar.
Good agreement between the use of small-strain non-linearity to predict
settlement behind a strutted excavation and field behaviour was described
by Jardine et al.39
The Author’s experience of use of the PLAXIS model is that soil deforma-
tions tend to be overestimated when using the Mohr–Coulomb soil model;
more accurate deformation prediction is obtained by use of the PLAXIS
soil hardening model, especially in stiff clays. This model is an elastic–plastic
type of hyperbolic model handling soil stiffness in terms of stress level and
stress path.
In addition to the use of appropriate soil stiffness parameters, the quality
of prediction will of course depend on the selection of accurate K0 values
for the particular site. The method of back analysis on its own may not
prove sufficiently dependable to obtain these values because of the relatively
large variations in at-rest pressures within relatively small distances.
The finite element method was developed by Mana and Clough11 to formu-
late a design method for estimating wall deformation and the settlement
trough for a strutted excavation in soft to medium clays without resorting
to use of a finite element program for a particular design problem. Their
procedure was as follows.

(a) At each construction stage where prediction of movement is needed,


calculate the minimum factor of safety against basal heave using Terza-
ghi’s method.
(b) Estimate the maximum wall movement hmax from the relationship
between factor of safety against basal heave and maximum wall move-
ment shown in Fig. 11.19. Approximate ground movement vmax can
be estimated by assuming that vmax lies within the range 0.6 hmax
to 1.0hmax .
546 Deep excavations

Fig. 11.19. Analytical


relationship between maximum
lateral wall movement and
factor of safety against basal
heave11

Fig. 11.20. Effect of wall


stiffness on maximum lateral
wall movement and maximum
ground settlement11

(c) Based on the wall stiffness factor, the strut stiffness factor, the depth to
a firm soil layer and the excavation width B, determine the influence
coefficients w , s , D , B , using Figs. 11.20–11.23.
(d ) Determine the influence coefficient for the design strut pre-loading p
using Fig. 11.24.
(e) Determine the modulus multiplier influence coefficient m from Fig.
11.25.
( f ) Using the value of hmax from step (b) and the influence coefficients
determined in stages (c)–(e), calculate a revised value for the maximum
lateral movement from hmax ¼ hmax w s D p m .
(g) Revise the estimate of hmax using the relationship vmax ¼ 0:6hmax
to 1:0hmax .
(h) Plot the ground settlement profile using the calculated value vmax and
the profile shown in Fig. 11.26.
This method can be used for walls supported by anchors provided the anchors
themselves are embedded in a mass of soil or rock which is materially beyond
the movement zone.

Other predictive methods


A method originally developed by Roscoe40 and developed by James et al.41
and Serrano42 uses stress and strain fields for increments of structural
Soil movement due to deep excavations 547

Fig. 11.21. Effect of strut


stiffness on maximum lateral
wall movement and maximum
ground settlement11

Fig. 11.22. Effect of depth to


firm layer on maximum lateral
wall movement and maximum
ground settlement11

Fig. 11.23. Effect of excavation


width on maximum lateral wall
movement and maximum
ground settlement11
548 Deep excavations

Fig. 11.24. Effect of strut pre-


load on maximum lateral wall
movement and maximum
ground settlement11

Fig. 11.25. Effect of modulus


multiplier on maximum lateral
wall movement and maximum
ground settlement11

Fig. 11.26. Envelopes to


normalize ground settlement
profiles11
Soil movement due to deep excavations 549

deflection and load. In a series of iterative steps, stress and strain fields are
produced which comply with all the parametric values for a particular
problem. More recently Maruoka et al.43 presented a predictive method for
vertical settlement in cohesive soils based on the deformed shape of the wall
and patterns of zero extension lines in the adjacent soil. They concluded
that strain fields consisting of straight lines and circular areas could be used
with a rigid body spring model and finite element analysis to give reasonably
accurate settlement profile predictions. This method, however, requires the
initial prediction of the deformed shape of the wall due to the excavation,
and while a solution based on a Winkler spring model could be used to do
this, any inaccuracy in this prediction would presumably be reflected in the
accuracy of the final settlement profile. A simpler solution based on the
kinematics of a mechanism involving soil and wall mass is referred to in
Eurocode 744 .

Soil movement during diaphragm wall and bored pile wall


installation
The soil deformations considered in this chapter so far have been due to
unloading of the soil surrounding the deep substructure during bulk excava-
tion. The soil structure, however, undergoes several stress changes during
installation of the peripheral walling prior to bulk excavation. These changes
may be due to dynamic stresses set up by the driving of sheet piles, or the relief
of stress due to augering of piles or the excavation of diaphragm wall panels.
Each in situ soil stress change has an associated volume change and a resulting
vertical settlement. The stress changes and settlements which result from the
installation of diaphragm wall panels and bored piles at the periphery of a
deep excavation deserve special comment. Soil movements associated with
diaphragm wall installation are generally small and limited in lateral extent,
but experience with the initial sections of the Island Metro Line in Hong
Kong indicated otherwise, and the causes for this difference should be noted.
Stress changes near a diaphragm wall panel excavation occur as a result of
unloading due to panel excavation, recharging with bentonite slurry and the
subsequent fluid pressures from liquid concrete. (The in situ stresses from
the concreting operation were measured in tests by Reynaud45 , which empha-
sized the relatively high pressures caused by concreting relative to soil values
of K0 .) The soil stress changes due to diaphragm wall panel installation are
therefore relatively complex and do not stem only from the excavation opera-
tion. Similar stress changes occur during bored pile installation.
The extent of soil movement due to excavation of diaphragm wall panels
depends on soil properties, groundwater levels, panel width and the length
of time between excavation and concreting. There are limited published
records of in situ measurements of soil movement due to panel installation:
those due to Uriel and Oteo46 in Seville, Spain; Farmer and Attewell47 in
London; Symons and Carder48 in London clay; and Humpheson et al.49 ,
Davis and Henkel50 , Morton et al.51 and Stroud and Sweeney52 in Hong
Kong, should be referred to. Plots of maximum movement due to the installa-
tion of diaphragm walls and bored piles due to Thompson75 and Carder7 are
reproduced in Figs. 11.27 and 11.28.
Early measurements on the sites in London and Seville within stiff over-
consolidated clays confirmed the general opinion of diaphragm wall specialist
firms that soil movements are generally small and reduce rapidly at short
distances from the panel, say equal to the panel length. Measurements
made by the Author at panel excavations in lightly over-consolidated silty
clay and silty sand washdown soils in Medinah, Saudi Arabia, also confirm
this view. The movements caused by panel excavation in Hong Kong, as
550 Deep excavations

Fig. 11.27. Maximum


movement due to installation
of planar and counterfort
diaphragm walls in stiff clays75

described in Chapter 9, were substantial. Measurements made at these sites


are compared in Table 11.3. The soil conditions in Hong Kong are materially
different from those in London, Seville and Medinah. In the Chater station
excavation in Hong Kong, fill and marine deposits to a total depth of 15 m
overlie decomposed granite, and the groundwater is only approximately
2.5 m below ground level. Measurements were made during panel excavations
up to 35 m deep at three points, 6 m, 15 m and 24 m from the diaphragm wall.
Soil movement due to deep excavations 551

Fig. 11.28. Maximum


movement due to contiguous
and secant bored pile wall
installation in stiff clays7

The vertical settlements at these points reached 50 mm, 40 mm and 25 mm,


respectively, increasing progressively as the adjacent panels were excavated
and concreted.
The following conditions caused these movements in Hong Kong,
partly due to the breaking down of soil arch structures spanning panel
excavations:
(a) a high groundwater table providing a ready supply of moisture
(b) a relatively high soil permeability (105 m/s in the completely decom-
posed granite) which allowed the ready transmission of moisture
552 Deep excavations

Table 11.3 Observed soil deformation due to diaphragm wall panel excavation

Site and reference Ground conditions Panel size Maximum horizontal soil Remarks
deformation at stated
horizontal distance from
panel centre line

1. London London clay: over- 15 m deep  At 1 m: 16 mm Observations made on


(Farmer and Attwell47 ) consolidated fissured stiff 6.1 m long  At 2.5 m: 6 mm one panel only, kept
silty clay 0.8 m wide At 4.5 m: 2.6 mm open 7 days before
No movement at 6.1 m concreting
from panel
2. Seville Sandy silt up to 13 m 34 m deep  At 2.4 m: 7.5 mm. Excavation for circular
(Uriel and Oteo46 ) overlying Quaternary 3.4 m long  During panel excavation wall, 25 m dia.
gravels up to 13 m 0.8 m wide maximum horizontal measurements made
thickness overlying plastic deformation 7.5 mm and 2.4 m from two panels,
fissured clay 7 m depth, 2.4 m from diametrically opposite
panel centre line each other
3. Hong Kong Chater Rd/Jackson Rd. 36 m deep  At 1.4 m: 30 mm at 20 m Trial panel
(a) (Stroud & Sweeney52 ) 7 m of fill overlying 4 m 6.1 m long depth.
of marine deposits At 6.4 m: less than 10 mm
overlying completely at 20 m depth
decomposed granite
(b) (Davis and Henkel50 ) Chater station 17.4 m deep Maximum movement of
about 0.15 to 0.2% depth
of excavation movement,
some movement at least
for horizontal distance
equal to panel depths
4. Medinah Car Park Layered silty sand and 30 m deep  At 15 m: 8 mm at ground Trial panel. Readings
sandy silt overlying basalt 6.8 m long  level, reducing linearly made over 11 day period
bedrock at 30 m depth 0.8 m wide with depth to zero at 30 m after excavation. No
subsequent movement

(c) soil with high swell potential which required moisture to cause signifi-
cant volume change.
The lesson to be learnt from the Hong Kong diaphragm wall excavations is
that where these three subsoil conditions occur together, large soil settlements
result from diaphragm wall panel installation in advance of bulk excavation in
addition to settlements related to bulk excavation.
Mention should also be made of the effects of poor slurry quality control.
Where bentonite is subject to prolonged use, vital properties (particularly
viscosity) become ‘tired’ and contamination and pH are inadequately con-
trolled, high fluid loss results. In turn, this leads to less effective lateral support
and greater movement.
Researchers have given considerable attention to the installation effects of
diaphragm wall panels. Although soil deformation effects may be of limited
practical effect (with the exception of the described conditions in Hong
Kong) there is likely to be some reduction in in situ earth pressures close to
the wall as panel excavation is made. These pressure reductions, likely to be
small and only near the wall, may be of the order of 20% for a diaphragm
wall panel and about 10% for a bored pile wall. The published analytical
work on panel and pile installation includes papers by Tedd et al.53 (finite
element data from Bell Common wall), Powrie54 (elastic stress analysis),
Gunn et al.55 (finite element analysis), Ng et al.56 (finite element analyses,
Lion Yard Cambridge), Page57 (centrifuge model tests), de Moor58 (finite
element analyses of a number of panels in sequence), Ng and Yan59 (3D
Soil movement due to deep excavations 553

finite element analysis), Gourvenec and Powrie60 (3D finite element analysis of
panels in sequence), Ng and Yan (3D finite difference analysis of panels in
sequence)61 , Powrie and Batten62 (axisymmetric analysis of a single bored
pile), and Cowland and Thorley63 (Building settlements in Hong Kong due
to panel excavation). Lings et al.64 examined the pressure effect due to wet
concrete.

Building response to The earliest published work on the effects of settlement on structures was
ground displacement directed to the tolerance of buildings to settlement and their own weight.
Papers by Skempton and MacDonald65 (1957), Meyerhoff 66;67 (1953, 1956),
Polshin and Tokar68 (1957) and others were based on angular distortion of
the structure, or deflection ratio, and horizontal strain was not a prime con-
sideration. Burland and Wroth69 (1974) and others showed how tensile strains
when considered with simple elastic beams could be used to give deflection
criteria for the onset of structural damage. Later, in 1977, Burland et al.70
reproduced ‘critical tensile strain’ with the concept of ‘limiting tensile
strain’, which could be varied to take account of the properties of differing
structural materials and serviceability limit states.
Boscardin and Cording71 (1989) further developed the relationship between
building damage and angular distortion and horizontal strain. A study of
eighteen case histories of buildings affected by excavation, either braced
excavations or tunnel workings, examined the effects of angular distortion
and strain. The classification of visible damage used by Boscardin and
Cording is reproduced in Table 11.4. Using this data together with the
analysis of a deep beam model they proposed a correlation between angular
distortion, horizontal strain and degree of cracking damage as shown in
Fig. 11.29. Burland72 in 2001 progressed the matter further by adopting

Table 11.4 Classification of visible damage to buildings71

Class of Description of damagea Approximate widthb


damage of cracks (mm)

Negligible Hairline cracks <0.1


Very slight Fine cracks easily treated during normal redecoration. Perhaps isolated <1
slight fracture in building. Cracks in exterior brickwork visible upon close
inspection
Slight Cracks easily filled. Redecoration probably required. Several slight fractures <5
inside building. Exterior cracks visible, some repointing may be required for
weathertightness. Doors and windows may stick slightly
Moderate Cracks may require cutting out and patching. Recurrent cracks can be 5 to 15 or several
masked by suitable linings. Tuck-pointing and possibly replacement of a cracks >3 mm
small amount of exterior brickwork may be required. Doors and windows
sticking. Utility service may be interrupted. Weathertightness often impaired
Severe Extensive repair involving removal and replacement of sections of walls, 15 to 25 also
especially over doors and windows required. Windows and door frames depends on number
distorted, floor slopes noticeably. Walls lean or bulge noticeably, some loss of cracks
of bearing in beams. Utility service disrupted
Very severe Major repair required involving partial or complete reconstruction. Beams usually >25 depends
lose bearing, walls lean badly and require shoring. Windows broken by on number of cracks
distortion. Danger of instability
a
Location of damage in the building or structure must be considered when classifying degree of damage.
b
Crack width is only one aspect of damage and should not be used alone as a direct measure of it.
Note. Modified from Burland et al.70
554 Deep excavations

Fig. 11.29. Relationship of


damage to angular distortion
and horizontal tensile strain71

limiting values of tensile strain associated with the Boscardin and Cording
categories of damage and showed the derivation of an interaction diagram
similar to Fig. 11.29 but for a particular value of the ratio of building
length to building height. Burland points out that there are wide gaps in
knowledge in these predictions of likely damage, including paucity of observa-
tions relating building stiffness to slope and magnitude of structural damage,
the effect of piling to these observations, and the effects of time-related soil
movement.
The conclusions to Boscardin and Cording71 are worthy of reiteration.
(a) Buildings sited adjacent to excavations are generally less tolerant of
differential settlements due to excavation than similar structures settling
under their own weight. This is due to the lateral strains that develop in
response to most excavations.
(b) As a structure is subjected to increasing lateral strains, its tolerance to
differential settlement decreases. As a consequence, measures to miti-
gate excavation-related building damage should include provisions to
reduce the lateral strains sustained by the structure.
(c) Since the ground movements develop in the form of a travelling
wave gradually impinging on a structure, the ratio of the length of
the portion of the structure affected by the excavation induced
ground movements to height of the structure will initially be very
small and grow gradually. Thus, the tolerance of a building to initial
deformation will be governed by its tolerance to shearing deformation
and horizontal extension.
(d ) Increasing the ratio of longitudinal stiffness to shear stiffness, E=G, as
openings in a wall are assumed to cause, increases the range of
deformed length to height ratio, L=H, in which the diagonal strain
capacity or shearing resistance is the limiting factor.
(e) Points (c) and (d) suggest that angular distortion, a measure of shearing
strain, would be an appropriate parameter to correlate with observed
response.
( f ) If reasonable estimates of "crit and L=H can be made for a structure in
the vicinity of proposed underground construction, the limiting deflec-
tion ratio and angular distortion for that structure can be estimated and
compared to the estimated ground movement to permit engineers to
assess the potential for damage and suitability of possible remedial
measures.
(g) Frame-type structures, depending on size and other details, can often
resist some of the ground movements. Increasing the number of stories
Soil movement due to deep excavations 555

creates a structure stiffer in shear that would tend to tilt rather than dis-
tort. Conversely, increasing the number of bays typically increases the
length of a structure and causes it to distort more to accommodate
the ground movements caused by the excavation.
(h) Horizontal ties in the form of reinforced grade beams or similar items
are effective means of controlling the strains and distortions in both
bearing wall and frame structures adjacent to excavation.
(i) This study only examined cases where the lateral strains are tensile in
nature. However, compressive strains may create a critical condition
for structures sited over the centre-line of a tunnel or mine. The effects
of compressive lateral strain were examined by the National Coal
Board73 .

Measures to alleviate Whilst use of construction methods such as the top-down system or pre-
the effects of settlement loading of temporary struts may achieve reductions in settlements below
nearby buildings, these measures may not reduce settlements to acceptable
levels. Further measures may be needed to reduce risk of cracking or tilt. In
summary, these are as follows.
(a) Strengthening the ground by means of grout injection, cement or
chemical, or by mix-in-place or pin piles, in order to increase soil stiff-
ness. In extreme cases, freezing of the subsoil may prove an effective
solution in granular, water-bearing soils.
(b) Strengthening the affected building by the insertion of improved
vertical support by underpinning and horizontal ties to resist horizontal
tensile strains imposed by the soil deformation. Shear stiffness of the
building can be improved by temporarily filling window and door open-
ings in facades and cross-walls with brickwork or blockwork of
requisite strength.
(c) Structural jacking applied progressively as the deep excavation is made
to counteract vertical settlement, possibly with improvements of
temporary strengthening to the structure.
(d ) Compensation and fracture grouting applied progressively as deep
excavation is made. Compaction grouting applied to both granular
and cohesive subsoils can provide a means of lifting structures to
counteract the effects of vertical settlements. The injection of a stiff
viscous spherical bulb of grout leads to some compaction and
permeation in granular soils, which can be controlled by grout design
and control of applied pressures. In cohesive soils neither compaction
nor permeation will occur materially and control of grout stiffness and
applied pressure will determine the extent of volume of the grout bulb
and vertical soil displacement.
Fracture grouting is generally applied to cohesive soils at high
pressure, to create thin layers of grout, often of considerable lateral
extent. The grout exploits existing planes of weakness at fissures and
joints within the soil structure and the extent of this intrusive plane
of grout is controlled by grout volume and pressure. Successive injec-
tions of compensation and fracture grouting can be made from tubes
à manchette, sleeved grout tubes, drilled in arrays from positions
both inside and outside the affected structure.
Recently published data and experience associated mainly with
tunnelling for the Jubilee Line Extension in London74 are valuable in
the consideration of compensation and fracture grouting as a means
to alleviate settlements caused by deep excavations.
556 Deep excavations

References 1. Long M. Database for retaining wall and ground movements due to deep
excavations. ASCE J. Geotech. Eng. 2001, Mar., 203–224.
2. Clough G., Smith E. and Sweeney B. Movement control of excavation systems by
iterative design. Proc. ASCE Geotech. Eng., 1989, 2, 869–884.
3. Peck R.B. Deep excavations and tunneling in soft ground. Proc. 7th Int. Conf.
S.M.F.E., Mexico City, 1969, State of the art volume, 225–290. Sociedad de
Mexicana de Mecanica de Suelos, Mexico City, 1969.
4. Karlsrud K. Performance monitoring in deep supported excavations in soft clay.
Proc. 4th Int. Geo. Symp. Field Instrumentation, Naryang Inst., Singapore, 187–
202.
5. Ou C., Hsieh P. and Chiou D. Characteristics of ground surface settlement. Can.
Geotech. J. Ottawa, 30, 758–767, 1993.
6. Wong I., Poh T. and Chesah H. Performance of excavation for depressed
expressway in Singapore. ASCE J. Geotech. Eng., 1997, 123, No. 7, 617–625.
7. Carder D. Ground movements caused by different embedded retaining wall con-
struction techniques. Transport Research Laboratory, Crowthorne, 1995. TRL
report 172.
8. Fernie R. and Suckling T. Simplified approach for estimating lateral movement
of embedded walls in UK ground. Proc. Int. Symp. Geo. Aspects of Underground
Construction in Soft Ground. City University, London, 131–136, 1996.
9. Clough G.W. and O’Rourke T.D. Construction induced movements of in situ
walls. Proc. Design and Performance of Earth Retaining Structures, ASCE Special
Publication 15, 439–470. Cornell University, 1989.
10. Addenbrooke T.I. A flexibility number for the displacement controlled design of
multi propped retaining walls. Ground Eng. 1994, 41–45.
11. Mana A.I. and Clough G.W. Prediction of movements for braced cuts in clay.
ASCE J. Geotech. Engng, 1981, 107, Jun., 759–777.
12. Review of design methods for excavations. Geotechnical Control Office, Hong
Kong, 1990.
13. CIRIA. Embedded retaining walls: guidance for economic design. Gaba A.R. et al.
CIRIA, London, 2002.
14. Tomlinson M.J. Foundation design and construction. Pitman, London, 4th edn.,
1980.
15. Clough G.W. et al. Prediction of support excavation movements under
marginal stability conditions in clay. Proc. 3rd Int. Conf. Numerical Methods
in Geomechanics, Aachen, 1979, Vol. 4, 1485–1502. Balkema, Rotterdam,
1979.
16. Potts D.M. and Fourie A.B. The behaviour of a propped retaining wall: results of
a numerical experiment. Ge´otechnique, 1984, 34, No. 3, 383–404 (discussion 1986,
35, No. 1, 119–121).
17. Goldberg D.T. et al. Vol. 1, Lateral support systems and underpinning; Vol. 2,
Design fundamentals; Vol. 3, Construction methods. Federal Highway
Administration, FHWA-RD-75–128, FHWA-RD-75–129, FHWA-RD-75–130.
(National Technical Information, 1976 PB 257210, PB 257211, PD 257212.)
18. Clough G.W. and Davidson R.R. Effects of construction on geotechnical perfor-
mance. Proc. 9th Int. Conf. S.M.F.E., Tokyo, 1977, 15–53. Japanese Society of
Soil Mechanics and Foundation Engineering, Tokyo, 1977.
19. O’Rourke T.D. Ground movements caused by braced excavations. ASCE J.
Geotech. Engng, 1981, 107, Sept., 1159–1178 (discussion 1983, 109, Mar., 485–
487).
20. Potts D.M. et al. The use of soil berms for temporary support of retaining walls.
Proc. Conf. Retaining Structures, Institution of Civil Engineers, London, 1992,
440–447.
21. Dunicliff J. Geotechnical instrumentation for monitoring field performance. J. Wiley,
New York, 1988.
22. Standing J.R., Withers A.D. and Nyren R.J. Measuring techniques and their
accuracy. Proc. Conf. Building Response to Tunnelling, Vol 1, 273–299. CIRIA.
Thomas Telford Ltd., London, 2001.
23. Clough G.W. and Denby G.M. Stabilizing berm design for temporary walls in
clay. ASCE J. Geotech. Engng, 1977, 103, Feb., 75–90.
Soil movement due to deep excavations 557

24. Burland J.B. et al. Movements around excavations in London clay. Proc. 7th
Euro. Conf. S.M.F.E., Brighton, 1979, Vol. 1, 13–29. British Geotechnical
Society, London, 1979.
25. a Richards D., Holmes G. and Beardman, D. Measurement of temporary prop
loads at Mayfair car park. Proc. ICE Geotech. Eng., July, 1999, 137, No. 3,
165–174.
25. b Batten M., and Powie W. Measurement and analysis of temporary prop leads at
Canary Wharf underground station, East London. Proc. ICE. Geotech. Eng.,
July, 2000, 151–164.
26. Simic M. and French, D. Three dimensional analysis of deep underground
stations. Proc. Conf. Value of Geotechnics, Institution of Civil Engineers,
London, 1998.
27. Stevens A. et al. Barbican Arts Centre. Structural Engr, 1977, 55, No. 11, Nov.,
473–485.
28. O’Rourke T.D. et al. The ground movements related to braced excavation and their
influence on adjacent buildings. US Department of Transport, 1976, DOT-TST76,
T-23.
29. Clough G.W. and Tsui Y. Performance of tied back support systems in clay.
ASCE J. Geotech. Engng, 1974, 100, Dec., 1259–1273 (discussion 1975, 101,
Aug., 833–836).
30. Clough G.W. Performance of tied back walls. Proc. ASCE Spec. Conf.
Performance of Earth and Earth Supported Structures, Lafayette, Indiana,
1972, Vol. 3, 259–264. ASCE, New York, 1972.
31. Clough G.W. Deep excavations and retaining structures. Proc. Symp. Analysis
and Design of Founds, Bethlehem, PA, 1975, 417–465.
32. Clough G.W. et al. Prediction of support excavation movements under marginal
stability conditions in clay. Proc. 3rd Int. Conf. Numerical Methods, Aachen,
1979, Vol. 4, 1485–1502. Balkema, Rotterdam, 1979.
33. Crawley J. and Glass P. Westminster Station, London. Proc. Deep Foundations
Conf., Vienna, 1998, 5.18.1–5.18.14. Deep Foundations Institute, Englewood
Cliffs, New Jersey, 1998.
34. Terzaghi K. and Peck R.B. Soil mechanics in engineering practice. Wiley, New
York, 2nd edition, 1967.
35. Caspe M.S. Surface settlement adjacent to braced open cuts. ASCE J. S.M.F.E.,
1966, 92, Jul., 51–59 (discussion 1966, 92, Nov., 255–256).
36. Bowles J.E. Foundation analysis and design. McGraw-Hill, New York, 4th
edition, 1988.
37. Bauer G.E. Movements associated with the construction of a deep excavation.
Proc. 3rd Int. Conf. Ground Movements and Structures, Cardiff, 1984, 694–706
(discussion 870–871 and 876). UWIST, Cardiff, 1984.
38. Woods R.I. and Clayton C.R.I. The application of the Crisp finite element
program to practical retaining wall problems. Proc. Conf. Retaining Structures,
Cambridge. Institution of Civil Engineers, London, 1992, 102–111.
39. Jardine R.J. et al. Studies of the influence of non-linear stress-strain characteris-
tics in soil-structure interaction. Ge´otechnique, 1986, 36, No. 3, 377–396.
40. Roscoe K.H. The influence of strains in soil mechanics. Ge´otechnique, 1970, 20,
No. 1, 129–170.
41. James R.G. et al. The prediction of stresses and deformation in a sand mass
adjacent to a retaining wall. Proc. 5th Euro. Conf. S.M.F.E., Madrid, 1972,
Vol. 1, 39–46. Sociedad Española de Mechanica del Suelo y Cimentationes,
Madrid, 1972.
42. Serrano A.A. The method of associated fields of stress and velocity and its appli-
cation to earth pressure problems. Proc. 5th Euro. Conf. S.M.F.E., Madrid, 1972,
Vol. 1, 77–84. Sociedad Española de Mechanica del Suelo y Cimentationes,
Madrid, 1972.
43. Maruoka M. et al. Ground movements caused by displacements of earth
retaining walls. Proc. Conf. Retaining Structures, Cambridge. Institution of
Civil Engineers, London, 1992, 121–130.
44. Eurocode 7. Foundations. British Standards Institution, sixth version, London,
1993.
558 Deep excavations

45. Reynaud P.Y. Mesure des pressions developpées dans une paroi moulée en cours
de bétonnage. Bull. de Liaison des Laboratoires des Ponts et Chausse´es, 1981,
May-June, 135–138.
46. Uriel S. and Oteo C.S. Stress and strain beside a circular trench wall. Proc. 9th
Int. Conf. S.M.F.E., Tokyo, 1977, Vol. 1, 781–788. Japanese Society of Soil
Mechanics and Foundation Engineering, Tokyo, 1977.
47. Farmer I.W. and Attewell P.B. Ground movements caused by a bentonite sup-
ported excavation in London clay. Ge´otechnique, 1973, 23, Dec., 576–581.
48. Symons I.F. and Carder D.R. Field measurement on embedded retaining walls.
Ge´otechnique, 1992, 42, No. 1, 117–126.
49. Humpheson C. et al. Basement and substructure for the Hong Kong and
Shanghai Bank, Hong Kong. Proc. Instn Civ. Engrs, 1986, 80, Aug., 851–883
(discussion 1987, 82, Aug., 831–858).
50. Davis R. and Henkel D.J. Geotechnical problems associated with construc-
tion of Chater Station. Proc. Conf. Mass Transport, Hong Kong, 1980,
paper J3.
51. Morton K. et al. Observed settlement of buildings for the modified initial system
of the mass transit railways, Hong Kong. Proc. 6th Asian Conf. S.M.F.E., Taipei,
1980, Vol. 1, 415–429. Organising Committee, Taipei, 1980.
52. Stroud M.A. and Sweeney D.J. Readings from diaphragm test panel, Chater
Road. A review of diagraphm walls and anchorages. Institution of Civil Engineers,
London, 1977, 142–148.
53. Tedd P. et al. Behavior of propped embedded retaining wall in Bell Common
Tunnel. Ge´otechnique, 1984, 34, No. 4, 513–532.
54. Powrie W. Discussion on performance of propped and cantilever rigid walls.
Ge´otechnique, 1985, 35, No. 4, 546–548.
55. Gunn M., Setkunanasthan A. and Clayton C. Finite element modelling of
installation effects. Proc. Conf. on Retaining Structures, Cambridge, 46–55.
Thomas Telford Ltd., London, 1993.
56. Ng C. et al. An approximate analysis of the three dimensional effects of
diaphragm wall installation. Ge´otechnique, 1995, 45, No. 3, 497–508.
57. Page J. Changes in lateral stress during slurry trench installation. Ph. D. Thesis,
Univ. of London, 1995.
58. de Moor E. An analysis of bored pile/diaphragm wall installation effects.
Ge´otechnique, 1994, 44, No. 2, 341–347.
59. Ng C. and Yan R. Stress transfer and deformation mechanisms around a
diaphragm wall panel. ASCE J. Geotech. Eng., 1241, No. 7, 638–648.
60. Gourvenec S. and Powrie W. Three dimensional finite element analysis of
diaphragm wall installation. Ge´otechnique, 1999, 49, No. 6, 801–823.
61. Ng C. and Yan R. Three dimensional modeling of a diaphragm wall construction
sequence. Ge´otechnique, 1999, 49, No. 6, 825–834.
62. Powrie W. and Batten M. Prop loads in large braced excavations. CIRIA,
London, 2000. Project Report 77.
63. Cowland J. and Thorley C. Ground and building settlement associated with
adjacent slurry trench excavation. Proc. 3rd Int. Conf. on Ground Movements,
Cardiff, 1984, 871–876. UWIST, 1984.
64. Lings M., Ng C. and Nash D. Lateral pressure of wet concrete in diaphragm
wall panels cast under bentonite. ICE Proc. Geotech. Eng., 1994, 107, No. 3,
163–172.
65. Skempton A.W. and MacDonald D.H. The allowable settlement of buildings.
Proc. ICE, 1957, III, No. 5, 727–784.
66. Meyerhoff G.G. Some recent foundation research and its application to design.
Struct. Eng., 1953, 31, 151–167.
67. Meyerhoff G.G. Discussion of Skempton and MacDonald’s paper. Proc. ICE,
1956, II, No. 5, 774.
68. Polshin D. and Tokar R. Maximum allowable non-uniform settlement of
structures. Proc. 4th Int. Conf. S.M.F.E., London, 1957, 402–405. Butterworth
Scientific Publications, London, 1957.
69. Burland J. and Wroth C. Settlement of buildings and associated damage. Proc.
Conf. on Settlement, 1974, 611–654. Cambridge, Pentech Press.
Soil movement due to deep excavations 559

70. Burland J. Broms B. and De Mello V. Behavior of foundations and structures.


Proc. Int. Conf. S.M.F.E., Tokyo, 1977, Vol. 2, 495–546. Japanese Society of
Soil Mechanics and Foundation Engineering, Tokyo, 1977.
71. Boscardin M. and Cording E. Building settlement due to excavation induced
settlement. ASCE J. Geotech. Eng., 1989, 115, No. 1, 1–21.
72. CIRIA. Building response to tunnelling. Ed. J. Burland, J. Standing and F. Jardine,
Thomas Telford Ltd., London, 2001, 23–41.
73. National Coal Board. Subsidence engineers’ handbook. National Coal Board
Production Dept., London, 1975.
74. CIRIA. Building response to tunnelling. Ed. J. Burland, J. Standing and F. Jardine,
Thomas Telford Ltd., London, 2001, Vols. 1 and 2.
75. Thompson P. A review of retaining wall behaviour in overconsolidated clay during
the early stages of construction. M.Sc. Thesis, Imperial College, 1991.
76. Marino G.G. Subsidence damage to homes over room and pillar mines in Illinois.
PhD Thesis, University of Illinois, 1985.

Bibliography Chang C.Y. and Duncan J.M. Analysis of soil movement around a deep excavation.
ASCE J. S.M.F.E., 1970, 96, Sept., 1655–1681.
Cole K.W. and Burland J.B. Observation of retaining wall movements associated
with a large excavation. Proc. 5th Euro. Conf. S.M.F.E., Madrid, 1972, Vol. 1,
445–453. Sociedad Española de Mechanica del Suelo y Cimentationes, Madrid,
1972.
Cowland J.W. and Thorley C.B. Ground and building settlement associated
with adjacent slurry trench excavation. Proc. 3rd Int. Conf. Ground Movement
and Structures, Cardiff, 1984, 723–738 (discussion 871–876). UWIST, Cardiff,
1984.
Creed M.J. et al. Back analysis of the behaviour of a diaphragm wall supported
excavation in London clay. Proc. Int. Conf. Ground Movements and Structures,
Cardiff, 1981, 743–758. UWIST, Cardiff, 1984.
Gumbel J.E. and Wilson J. Observations of ground movement around a trench
excavated in London clay. Proc. Int. Conf. Ground Movements and Structures,
Cardiff, 1981, 841–856.
Jardine R.J. et al. Some practical applications of a non-linear ground model. Proc.
10th Euro. Conf. S.M.F.E., Florence, 1991, Vol. 1, 223–228. Balkema, Rotterdam,
1991.
Milligan G.W. Soil deformation behind retaining walls. Proc. 3rd Int. Conf. Ground
Movement and Structures, Cardiff, 1984, 702–722. UWIST, Cardiff, 1984.
Potts D.M. and Burland J.B. A numerical investigation of the retaining walls of the Bell
Common Tunnel. Transport Research Laboratory, Crowthorne, 1983, Supplemen-
tary Report 783.
Simpson B. et al. A computer model for the behaviour of clay. Ge´otechnique, 1979, 29,
No. 2, 149–175 (discussion 1980, 30, No. 3, 336–339).
Tedd P. et al. Behaviour of a propped embedded retaining wall in stiff clay at Bell
Common Tunnel. Ge´otechnique, 1984, 34, No. 4, 513–532.
Wood L.A. and Perrin A.J. Observations of a strutted diaphragm wall in London
clay: a preliminary assessment. Ge´otechnique, 1984, 34, No. 4, 563–579.
Appendix: Selection of typical soil
parameters and correlations for
initial design purposes

Typical values of soil and rock density

Density when drained Density when submerged


above ground water below ground water level
level (mg/m3 ) (mg/m3 )

Gravel 1.60–2.00 0.90–1.25


Hoggin (gravel–sand–clay) 2.00–2.25 1.00–1.35
Coarse to medium sands 1.70–2.10 0.90–1.25
Fine and silty sands 1.75–2.15 0.90–1.25
Loose fine sands Down to 1.50 Down to 1.90
Stiff clay 1.80–2.15 0.90–1.20
Soft clay 1.60–1.90 0.65–0.95
Peat 1.05–1.40 0.05–0.40
Granite .
2 50 (not crushed —
or broken)
Sandstone 2.20 (not crushed —
or broken)
Shale 2.15–2.30 1.20–1.35
Stiff to hard marl 1.90–2.30 1.00–1.35
Chalk 0.95–2.00 0.30–1.00

Atterberg limits

wLL (%) wPL (%) Ip (%)

Sand Non-plastic
Silt 30–40 20–25 10–15
Clay 40–150 25–50 15–100
Clay examples London clay 73 25 48
Gosport 80 30 50
Shellhaven 97 32 65

Coefficient of permeability ranges of soils

Soil type k (cm/s)

Clean gravel >1.0


Clean sands, clean sands and gravel mixtures 1.0 to 103
Fine sands, silts, mixtures of sands, silts and clays 103 to 107
Homogeneous clays <107
Appendix 561

Typical values of 0 , drained shear strength c0 and undrained shear strength cu

K (m=s  106 ) 0 c0 (kN/m2 ) cu (kN/m2 )

Sand loose, round 308 — —


Sand loose, angular 32.58 — —
Sand medium dense, round 32.58 — —
Sand medium dense, angular 358 — —
Gravel without sand 37.58 — —
Coarse gravel, sharp edged 408 — —
Soft normally-consolidated clays 158–208 5–10 10–25
Stiff over-consolidated clays 258–288 20–25 75–150
Silt 258–308 — 10–50
Peat 108–158 0–5 —

Typical values of Poisson’s ratio

Soil type Description 0

Clay Soft 0.35–0.4


Medium 0.3–0.35
Stiff 0.2–0.3
Sand Loose 0.15–0.25
Medium 0.25–0.3
Dense 0.25–0.35

Typical values of Young’s modulus ðEÞ and shear modulus ðGÞ. (These values for E are
secant values at peak deviatic stress for dense and stiff soils and at maximum deviatic
stress for loose, medium and soft soils.)

Soil type Description E (MPa) G (MPa)

Clay Soft 1–15 0.4–5


Medium 15–30 5–11
Stiff 30–100 11–38
Sand Loose 10–20 4–8
Medium 20–40 8–16
Dense 40–80 16–32

You might also like