Temporary Works - Principles of Design and Construction-ICE Publishing (2012)
Temporary Works - Principles of Design and Construction-ICE Publishing (2012)
Edited by
Murray Grant and Peter F. Pallett
06 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Control of groundwater 79
Neil Smith
6.1. Introduction 79
6.2. Techniques 80
6.3. Investigation for dewatering 84
6.4. Analysis and design 85
References 88
Further reading 89
vi
12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Trenching 145
Ray Filip
12.1. Introduction 145
12.2. Techniques 149
12.3. Design to CIRIA 97 trenching practice 155
12.4. Controlling water 157
References 158
Further reading 158
vii
viii
22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Falsework 275
Peter F. Pallett and Godfrey Bowring
22.1. Introduction 275
22.2. Materials and components 277
22.3. Loads on falsework 279
22.4. Falsework design 282
22.5. Providing a stable structure 286
22.6. Workmanship and inspections 287
References 288
23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Formwork 289
Laurie York and Peter F. Pallett
23.1. Introduction 289
23.2. Vertical formwork 289
23.3. Economy 290
23.4. Specifications and finishes 291
23.5. Tolerances/deviations 292
23.6. Formwork materials 292
23.7. Concrete pressure calculation 295
23.8. Wall formwork design 300
23.9. Column formwork design 303
23.10. Striking vertical formwork 304
23.11. Workmanship/checking 304
References 305
ix
Index 361
xi
Contributors
A. Bell Chief Engineer, Cementation Skanska Limited
Andrew Bell is a geotechnical engineer and Chartered Civil
Engineer with over 15 years experience, gained principally in the
design and construction of piled foundations and deep
basements. Andrew is currently Chief Engineer at Cementation
Skanska.
P. Boddy MEng CEng MICE, Head of Temporary Works for
Interserve Construction Limited
Currently Head of Temporary Works for Interserve
Construction Limited, Paul Boddy has also worked for British
Waterways and has a broad experience (both practical and
technical) of floating plant from large to small.
xiii
xiv
xv
xvi
xvii
xviii
Chapter 1
Safety, statutory and contractual
obligations
Andrew Rattray HM Principal Specialist Inspector, Construction Engineering Specialist Team,
Health and Safety Executive
Peter F. Pallett Pallett TemporaryWorks Ltd
Safety of temporary works is paramount to protect not only site workers, but also the
public and others who may be affected by the work. Temporary works should be
‘engineered’ and given the same degree of care and consideration as the permanent
works. It is also important that temporary works are designed to be robust enough to
withstand the rigours of site use and detailed to ensure that local or single component
failure does not lead to progressive collapse. Knowledge gained from past collapses/
incidents and research has informed current temporary works guidance and standards.
BS 5975 is an industry consensus view on good practice and provides recommendations
for procedures for the design, construction, use and dismantling of all types of temporary
works. Statutory legislation has a direct effect on the design and operation of temporary
works, imposing duties and obligations on many of the parties involved. In addition, all
users of temporary works will have contractual obligations under which they work, often
imposing specific requirements for the works.
1.1. Introduction
‘Temporary works’ is a widely used expression in the construction industry and is defined in
BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures and the permis-
sible stress design of falsework (BSI, 2011) as ‘parts of the works that allow or enable
construction of, protect, support or provide access to, the permanent works and which
might or might not remain in place at the completion of the works’. The construction of
most types of permanent works will require the use of some form of temporary works.
Temporary works should be an ‘engineered solution’ that is used to support or protect an
existing structure or the permanent works during construction, or to support an item of
plant or equipment or the vertical sides or side-slopes of an excavation or to provide
access. It is imperative that the same degree of consideration and care is given to the
design and construction of temporary works as to the design and construction of the perma-
nent works. The management of temporary works is discussed in Chapter 2 on management.
Examples of temporary works include (but are not limited to) the following.
In order to ensure the strength and stability of any temporary works structure, there are
three fundamental aspects that need to be considered, which can be simplified as follows.
g Foundations: the ability of the ground to carry the loads transmitted from the
temporary works structure without failure or excessive deformation or settlement.
g Structural integrity: the ability of the temporary works structure itself to carry and
transmit loads to the ground via the foundations without failure of the structural
elements, including fixings and connections (e.g. by buckling, bending, shear,
tension, torsion) and without excessive deflection.
g Stability: the ability of the temporary works structure to withstand horizontal or
lateral loading without sway, overturning or sliding failure (stability may be
inherent in the temporary works structure itself or provided by the permanent
works).
Failure to adequately design, construct and maintain temporary works can lead to
The code of practice on falsework (BS 5975) was revised in 2008 to provide recom-
mendations and guidance on the procedural controls to be applied to all aspects of
temporary works in the construction industry, as well as specific guidance on the
design, specification, construction, use and dismantling of falsework. BS 5975 (BSI,
2011) describes procedures as well as technical aspects since the success of falsework
and temporary works is closely linked to good management.
1.2. Background
The first report on falsework was by the Joint Committee of the Institution of Structural
Engineers and the Concrete Society (1971) and introduced classes of falsework. The
British Standard Code of Practice committee was established and, ironically, a fortnight
later a major collapse occurred in the UK by the River Loddon near London. This collapse
and other significant falsework collapses in the 1970s, together with an apparent lack of
authoritative guidance, led to the government setting up the Advisory Committee on
Falsework which produced the Bragg Report (Bragg, 1974, 1975) named after the
committee chairman. Industry produced the first code of practice (in compliance with
one of the recommendations of the Bragg Report) as BS 5975 in 1982 which was informed
by the recommendations of the Bragg Report and the earlier Joint Committee report.
The Bragg Report made some very pertinent comments about falsework which apply
equally to all temporary works and which still hold true today, including:
‘Falsework requires the same skill and attention to detail as the design of permanent
structures of like complexity, and indeed falsework should always be regarded as a
structure in its own right, the stability of which at all stages of construction is
paramount for safety.’
g competency
g design procedures
g design responsibilities
g communication and coordination
g inspection and supervision
g lateral stability.
BS 5975 codified all relevant aspects that should be considered when preparing a permis-
sible stress design for falsework and included recommendations for materials, design and
work on site. Because the success of falsework is closely tied up with its management, the
There have been significant changes to the construction industry since the mid 1970s
which have affected how falsework, and more generally temporary works, are dealt
with. A Standing Committee on Structural Safety paper (SCOSS, 2002), updated in
2010, identified the principal changes which included the following.
g Few main contractors now have their own temporary works departments, whereas
in the 1970s almost all would design temporary works in-house; the responsibility
for temporary works now often falls to a specialist contractor/supplier which can
result in a lengthy supply chain.
g In the 1970s, most falsework (and temporary works) were constructed from
scaffold tube and fitting components whereas proprietary systems now dominate
the market; the design skills and knowledge of the performance of the systems
therefore now tends to lie within the specialist organisations.
g There has been a gradual but inexorable loss of traditional skills within the
construction industry; in practical terms, this means that the site foreman with a
lifetime’s experience of ‘what works’ has largely been lost.
g Procurement routes are now largely chosen to maximise commercial benefit and have
little regard to considerations of the flow of information; the difficulties caused by
long supply chains are further exacerbated when design and erection responsibility
are split and when design/supply briefs do not include site visits/inspections.
Health and Safety Executive research (HSE, 2001) into various aspects of falsework
produced some worrying findings that were equally applicable to temporary works in
general; these included the following.
g sufficiency of information
g adequacy of supervision
g role of the Temporary Works Coordinator
g competency of those erecting falsework/temporary works.
The actions to deal with these concerns are straightforward and require no more than the
application of the good practice given in BS 5975 (BSI, 2011). They also fit well with the
aspirations of the Construction (Design and Management) Regulations 2007 (HSE, 2007)
in respect of their aim of improving the overall coordination and management of health
and safety throughout all stages of a construction project. The European limit state
design codes, such as BS EN 12811-1 for scaffolding (BSI, 2003) and BS EN 12812 for
falsework (BSI, 2008), have little (if any) reference to procedures or site practice. This
means that the procedural items in BS 5975 (BSI, 2011) take on a significant importance.
The duties of all parties under CDM 2007 (Client, CDM Coordinator, Designer,
Principal Contractor, Contractor) are set out in the associated Approved Code of
Practice (CDM ACOP) (HSE, 2007) and need to be read and understood by all those
involved in the procurement and use of temporary works. CDM 2007 and the CDM
ACOP (HSE, 2007) make several specific references to the design and management of
temporary works; these are considered below.
1.4.1 Designers
The definition of ‘structure’ in CDM 2007 (HSE, 2007) includes ‘any formwork, falsework,
scaffold or other structure designed or used to provide support or means of access during
construction work’. In the CDM ACOP (HSE, 2007), designers are deemed to include
‘temporary works engineers, including those designing auxiliary structures, such as form-
work, falsework, façade retention schemes, scaffolding, and sheet piling’. Specific guidance
for all designers is given in CDM07/4 Industry Guidance for Designers (CITB, 2007). It is
therefore clear that, under CDM 2007, temporary works designers have exactly the same
designer duties as Permanent Works Designers, including the following.
g Ensuring that they are competent in their specific field of temporary works design
(and able to address the particular relevant health and safety issues).
g Avoiding foreseeable risks, so far as is reasonably practicable, to those involved in
the construction, use and dismantling of the temporary works and providing
adequate information on the remaining significant risks, for example, by following
the principles of the Eliminate, Reduce, Inform, Control (ERIC) model (CITB,
2007) or any other similar risk reduction technique as part of the design process.
g Coordinating and cooperating with others, for example, liaising with the
Permanent Works Designers to ensure that designs are compatible and that the
permanent works can accommodate any assumed loadings from the temporary
works, such as the lateral restraint of falsework.
The equal care and consideration required by the designers of temporary works and the
permanent works reflects comments made in the Bragg Report (Bragg, 1975) more than
30 years previously.
The WAHR (2005) impose duties on employers and the self-employed to assess the risks
from work at height and to organise and plan the work so it is carried out safely. The
objective is to make sure that work at height is properly planned, including the selection
of relevant equipment, appropriately supervised and carried out in a safe manner. The
WAHR set out a simple hierarchy for managing and selecting equipment for work at
height; duty holders must avoid work at height where they can; use work equipment
or other measures to prevent falls where they cannot avoid working at height; and,
where they cannot eliminate the risk of a fall, use work equipment or other measures
to minimise the distance and consequences of a fall should one occur. Priority should
be given to collective protection measures over personal protective measures. When
working under the control of another person, all employees and the self-employed
have a duty to report any activity or equipment which is defective.
In relation to temporary works, the WAHR place duties on clients and designers to
ensure that strength and stability calculations for scaffolding are carried out unless
calculations are already available or the scaffold is assembled in conformity with a
generally recognised standard configuration (HSE, 2007; schedule 3.2). A recognised
standard configuration could be the National Access and Scaffolding Confederation
(NASC) Technical Guidance TG20 for tube and fitting scaffolds (NASC, 2008) or the
manufacturer’s guidance for system scaffolds. In such cases calculations have already
been prepared and, provided the scaffolding is erected to the stated rules for the solution
adopted, further calculations are not required. The WAHR require that, depending on
the complexity of a scaffold, an assembly, use and dismantling plan shall be drawn up
by a competent person; such a plan should describe the sequence and methods to be
adopted when erecting, dismantling and altering the scaffold, if this is not covered by
the published guidance referred to above.
The WAHR state certain requirements for all working platforms. Where there is a risk
of falling, there are requirements for guardrails, toe boards, barriers and similar means
of protection; the top guardrail should be at least 950 mm high, with intermediate
guardrail(s) and toe boards(s) positioned to give a maximum unprotected gap of
470 mm. Where the platform is for a sloping workplace at an angle greater than 108,
the guardrail requirements are more onerous. Although no minimum width of a
working platform is stated, the width values given in good practice documents such as
BS EN 12811 (BSI, 2004) or TG20 (NASC, 2008) are used.
In the UK, there is a separate judicial system for handling the Civil Procedures involved
in disputes both under contract law and the law of tort. Under contract law, both parties
accept and agree to liabilities and duties whereas tort is a violation of a duty established
by law. Liability in tort usually arises from a breach of duty established in law, and
includes the tort of negligence. However, although the subject of Contract and Tort
Law is outside the scope of this book (see ICE, 2011), its implication for temporary
works needs to be considered.
Contracts may be based on existing industry formats such as the ICE Conditions of
Contract (ICE, 1999) and the Engineering and Construction Contract (ECC), previously
the New Engineering Contract (NEC3) (ICE, 2005). Contracts may also require compli-
ance with detailed specifications. Typical specifications are Highways Agency (HA,
2006), the National Building Specification (NBS) (published annually), Civil Engineering
Specification for the Water Industry (CESWI, 2011) and the National Structural Concrete
Specification for Building Construction (NSCS) (CSG, 2010). The contract will place an
obligation on the contractor to comply. This contractual obligation will frequently be
passed on to subcontractors and suppliers, therefore obligating their compliance. The
contract may also include lists of standards deemed to be included, thus making a parti-
cular British Standard a contractual requirement. Hence, although not legally enforceable
as a mandatory statutory requirement, use of a particular British Standard would become
obligatory under the contract. Failure to adopt the BS recommendations would therefore
risk a court action for damages. This could affect the temporary works design and such
obligations should be included in the temporary works design brief (see Chapter 2 on
management).
contractor has to carry out permanent works design in order to verify the construction
sequence and method.
The professional advisor is responsible to the client for ensuring that the contractor’s
temporary works will produce a finished job which complies with the contract
documents, in particular that it is not detrimental to the permanent works. This
means that the professional advisor does not have a contractual duty to verify the
contractor’s temporary works calculations; hence the importance of the independent
temporary works design checks recommended by BS 5975 (BSI, 2011) and discussed
in Chapter 2. It is always important to establish the responsibilities under the various
subcontracts. Phrases in the contract such as ‘in accordance with recognised codes
of practice’ etc. will make the use of certain codes of good practice a contractual
requirement.
1.8. Robustness
Temporary works often comprise a structure with many components, junctions and
connections and, unlike the permanent works, are often reused a number of times and
moved from site to site. Temporary works should therefore be robust enough to
withstand the rigours of site use. Careful attention should be paid to the way in which
components, connections and junctions are detailed to reduce the dependence on
workmanship. For example, the design of a particular joint or component may be
justified by calculations using a minimum thickness of section, whereas engineering
judgement would dictate that using a thicker/larger section would reduce the risk of
damage during transport and when in use, giving a more robust and consequently
safer structure. Any critical component or connection should be inherently robust in
its own right. Detailing of the temporary works should be such that local or single
component failure does not lead to the progressive collapse of the whole structure.
This does not imply that the design of temporary works should be over-conservative,
but that due consideration should be given to providing alternative load paths so that
in the event of the failure of one member or component the load may be redistributed
through others. An appreciation of the way the temporary works structure behaves
will lead to safe temporary works that do not progressively collapse.
10
In particular, children do not have the ability to perceive danger in the same way as adults
do and may see construction sites as potential playgrounds. While the numbers of children
being killed or injured on construction sites has reduced, there is no room for complacency.
Each year, two or three children die after gaining access to construction sites but many
more are injured. Other members of the public have been seriously injured by
All construction sites require measures in place to manage access to the site through well-
defined site boundaries and to exclude unauthorised persons such as members of the
public. The site boundary should be physically defined, where necessary, by suitable
fencing. In populated areas, this will typically mean a 2 m high small mesh fence or
hoarding around the site or work area. Consideration must also be given to the provision
of protection from any work activity taking place outside of the site boundary, for
example, erection/dismantling of the site fencing/hoarding, utilities excavations, scaffold
erection/dismantling and delivery and storage of materials.
Many hazards have the potential to injure members of the public. In particular, the
following need to be considered whenever temporary works are being planned or
carried out.
g Falling objects: ensure that objects cannot fall outside the site boundary.
g Excavations and openings: barriers or covers are required.
g Delivery and site vehicles: ensure that pedestrians cannot be struck by vehicles
entering/leaving site; do not obstruct pavements so forcing pedestrians into the road.
g Scaffolding and other access equipment: prevent people outside the site boundary
being struck during the erection, dismantling and use of scaffolding and other
access equipment.
g Slips, trips and falls within pedestrian areas: inadequate protection of holes, uneven
surfaces, poor reinstatement, trailing cables and spillage of materials are common
causes.
g Storing and stacking materials: keep all materials within the site boundary if
possible, or provide protection.
More comprehensive guidance on this subject is given in the HSE publication Protecting
the public – Your next move (HSE, 2009).
11
g Key issues for the safety and stability of temporary works are adequate
foundations, structural integrity and lateral stability.
g Coordination and cooperation is required between the temporary and the
Permanent Works Designers to ensure that designs are compatible and that the
permanent works can accommodate any assumed loadings from the temporary
works, for example, the lateral restraint of falsework.
g The majority of construction phase plans prepared by Principal Contractors
should contain the management arrangements for controlling the risks associated
with temporary works, for example, suitable temporary works procedures.
g The appointment of a competent Temporary Works Coordinator is an essential
step for the safe management of substantive temporary works structures.
REFERENCES
Bragg SL (1974) Interim Report of the Advisory Committee on Falsework. Department of
Employment and Department of the Environment. HMSO, London.
Bragg SL (1975) Final Report of the Advisory Committee on Falsework. Department of
Employment and Department of the Environment. HMSO, London.
BSI (2004) BS EN 12811-1: Temporary works equipment. Part 1: Scaffolds – Performance
requirements and general design. BSI, London.
BSI (2008) BS EN 12812: Falsework – Performance requirements and general design. BSI,
London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
CESWI (Civil Engineering Specification for the Water Industry) (2011) Civil Engineering
Specification for the Water Industry, 7th edition. WRc plc, Swindon.
CITB (Construction Industry Training Board) (2007) The Construction (Design and
Management) Regulations. Industry Guidance for Designers. Construction Skills,
Norfolk.
Concrete Society (1971) Falsework. Report of the Joint Committee of The Concrete
Society and the Institution of Structural Engineers. Technical Report TRCS 4. Concrete
Society, Crowthorne.
CSG (Concrete Structures Group) (2010) National structural concrete specification for
building construction, 4th edition. Publication CCIP-050, The Concrete Centre, Camberley.
HA (Highways Agency) (2006) Specification for Highway Works. Manual of Contract
Documents for Highway Works. Highways Agency, London.
HSE (Health and Safety Executive) (2001) Investigation into aspects of falsework. HSE
Contract Research Report 394/2001. HSE Books, Sudbury. Available at www.hse.gov.
uk/research/crr_htm/2001/crr01394.htm.
HSE (2007) Managing health and safety in construction. Approved code of practice, Publi-
cation L 144, HSE Books, Sudbury.
HSE (2009) Protecting the public – Your next move. Publication HSG 151, HSE Books,
Sudbury.
HSOA (Health and Safety (Offences) Act) (2008) The Health and Safety (Offences) Act.
HMSO, London.
ICE (Institution of Civil Engineers) (1986–1999) ICE Conditions of Contract, 5th and 6th
editions. ICE Publishing Ltd, London.
12
ICE (2005) ECC Engineering and Construction Contract. ICE Publishing Ltd, London.
ICE (2011) ICE Manual of Construction Law. Thomas Telford Publishing Ltd, London.
NASC (National Access & Scaffolding Confederation) (2008) TG20: Guide to good prac-
tice for scaffolding with tube and fittings. NASC, London (including TG20 Supplement,
February 2011).
NBS (National Building Specification) (annually) Formed finishes, Section E20. National
Building Specification Ltd, Newcastle-upon-Tyne.
SCOSS (Standing Committee on Structural Safety) (2002) Falsework: full circle? SCOSS
Topic Paper SC/T/02/01. SCOSS, London.
WAHR (Work at Height Regulations) (2005) SI 2005/735: The Work at Height Regula-
tions. HMSO, London (and Work at Height (Amendment) Regulations 2007).
FURTHER READING
Burrows M, Clark L, Pallett P, Ward R and Thomas D (2005) Falsework verticality:
leaning towards danger? Proceedings of ICE, Civil Engineering, February 2005: 41–48.
ICE (Institution of Civil Engineers) (2010) ICE Manual of Health and Safety in Construc-
tion. Thomas Telford Publishing Ltd, London.
NASC (National Access & Scaffolding Confederation) (2010) SG25: Access and Egress
from Scaffolds. NASC, London.
Smith NJ (2006) Managing Risk in Construction Projects, 2nd edition. Blackwell Publishing,
Oxford.
13
Chapter 2
Management of temporary works
Godfrey Bowring Consultant
This chapter examines why and how temporary works and the interface between the
permanent works and temporary works designers are managed and controlled. The
importance and independence of the Temporary Works Coordinator (TWC) and
support upon which the TWC depends from the construction team are discussed. The
categories of design check to protect the team (including the operatives) from errors,
omissions, misunderstandings and perhaps even the pressures of contract, programme
and cost are reviewed.
2.1. Introduction
Temporary works are traditionally defined in the Institution of Civil Engineers (ICE)
Conditions of Contract in very general terms, no doubt to preclude contractors from
trying to argue that an aspect of temporary works, not specifically listed, would in
some way give rise to a variation.
The Conditions of Contract (ICE, 1986), known as the ICE 5th Edition, defines temporary
works, permanent works and the works as follows:
Traditionally, temporary works were the responsibility of the contractor with the
Engineer or the Permanent Works Designer (PWD) only required to provide specific
details of the permanent works design to allow the contractor to design the temporary
works. With this one exception, there was negligible liaison between the permanent
15
and temporary works designers; a situation which is inconsistent with the Construction
Design and Management (CDM) Regulations (HSE, 2007). See Chapter 1 on the
contractual and legal aspects. Strangely, the mindset of this traditional division of
responsibilities can still be found on notes on permanent works drawings to the
effect that the temporary works or temporary stability of the permanent works are the
responsibility of the contractor. The guidance to the CDM Regulations, Industry
Guidance for Designers (CITB, 2007), is clear, stating that ‘all designers should take
full account of the temporary works, no matter who is to develop those works’. This
therefore requires that both permanent works and temporary works designers consider
the temporary works.
While it is often tempting to provide a list of typical temporary works such as formwork,
falsework, excavation support, temporary access, scaffolding, façade retention, temporary
slopes, hoardings, cofferdams, trenching, temporary jetties and crane foundations, a list
can never be either exhaustive or exclusive and is something that the ICE contractual
definition attempts to avoid.
Temporary works and permanent works, and the operatives and public who are entitled
to depend upon their safe execution and use, are identical insofar as their being
dependent upon the application of the same engineering principles. It therefore follows
that the temporary works and permanent works deserve the same rigour and respect.
In the late 1960s and early 1970s, there was unfortunately a series of significant collapses
of falsework and of permanent works in a temporary condition together with associated
fatalities. The number and scale of these collapses was sufficient to prompt the govern-
ment to commission a report into the failures. The Government Advisory Report known
as the Bragg Report (Bragg, 1975) was an industry milestone, which is still held in
respect. It identified a series of causes and made numerous recommendations, the
most significant of which were the adoption of procedural controls, coordination
between designers of permanent and temporary works, proper management of the
process from inception through to loading and unloading and the appointment of a
falsework coordinator to manage and be responsible for the process. Industry wisely
realised that the Bragg recommendations for falsework applied equally to all temporary
works and adopted the more general term Temporary Works Coordinator (TWC).
The core Bragg recommendations, which were subsequently incorporated into BS 5975
which was first published in 1986, remain as valid today as when first written. The
adoption of the recommended procedural controls for temporary works in general has
served industry well and has been instrumental in preventing further major collapses
and failures and in promoting safe practice.
Nevertheless, research commissioned by the Health & Safety Executive (HSE, 2001;
Burrows et al., 2005) showed that the basic concepts were being forgotten. This was
further supported by the Standing Committee for Structural Safety (ICE/SCOSS,
2002) stating that there was ‘a lack of understanding of the fundamentals of stability,
at all levels of industry’ and ‘a lack of adequate checking and a worrying lack of
16
design expertise’. All this serves to reinforce the Bragg recommendations and to caution
against complacency.
In 2008, the procedures recommended for falsework were formally introduced for all
temporary works by publication of an updated edition of BS 5975: 2008 þ A1: 2011
(BSI, 2011). This British Standard also includes the permissible stress design of false-
work; see Chapter 22 on falsework for more information.
Although the use of the procedures in British Standards is not mandatory, every organi-
sation has a legal duty to ensure that it is operating a safe method of work. This is usually
embodied in its company procedures, written and specifically adapted for the operations
carried out by the particular company/organisation. In many cases, the procedures
recommended by BS 5975 (BSI, 2011) are incorporated into Quality Manuals, Best
Practice Guidance etc., and should be followed. Company procedures are often made
binding on subcontractors to ensure they follow the relevant company procedures/forms.
This chapter reflects UK practice, though much of what is written will apply elsewhere. It
follows the procedural controls embedded in BS 5975 (BSI, 2011), but makes no attempt
to either list or replicate every aspect.
17
simple terms, the Permanent Works Designer should consider the buildability of any
temporary works required and communicate clearly the residual risks including design
assumptions to the designer of the temporary works. It is no longer acceptable for a
designer to attempt to abdicate responsibility by the provision of a note on the
drawing stating that the contractor is responsible for all checks of temporary conditions
and/or loading of the permanent works. The CDM Coordinator has a duty to ensure
cooperation between the permanent and temporary works designers and ensure that
the permanent works can support any loadings from temporary works (HSE, 2007;
para 104). There is an argument that the coordination roles of the CDMC and the
TWC overlap; both parties should therefore ensure that their coordination does not
lead to omissions or ambiguity.
If all the good advice is followed, the implementation of temporary works should be
relatively risk free. Regrettably, that is not always the case. The principle factors
which continue to put temporary works at risk are listed in Table 2.1.
18
Factor Effect
Package management and Loss of ownership and therefore of responsibility for the
subcontracting temporary works by the principle contractor. Design
interfaces and responsibilities established by commercial
framework, with insufficient attention being paid to the
engineering risks
Supplier design The advice received will be partial in that the design will
inevitably incorporate only the supplier’s own
equipment and there will be aspects which will either
make fundamental assumptions or require separate
design. Falsework foundations and top restraint are
typical examples
Lack of understanding of temporary Temporary works design will be made more difficult and
works and buildability by permanent therefore with a higher risk. Interface between and
works designers of permanent works interaction of temporary and permanent works not fully
addressed
The procedural controls within The controls, if complied with, reduce the risk of failure
BS 5975 are not adhered to to an acceptably low level. Non-compliance, either in
part or in the whole, significantly increases the risk of
failure
when the TWC is absent for any reason. The key point is that this deputy has both the
authority and responsibility to act as the TWC when and if the TWC is absent from site
for any reason.
BS 5975 makes it clear that, for a given site, the principal contractor should appoint a
TWC.
It is important to realise that the TWC manages the process. It is only on the TWC’s
authority that the temporary works are either loaded or unloaded. The Project
Manager (PM) therefore has to be aware that the TWC’s role is critical to the progress
of the works and is there to protect, in terms of safe working practice, the operations and
operatives on site. The PM must therefore support, and moreover be seen to support, the
TWC.
Interestingly, it is the PM’s responsibility to manage the erection of the temporary works
in the same way that it is the PM’s responsibility to manage the permanent works. The
19
TWC’s role of technical independence and responsibility for permits to load is therefore
deliberately separated from the role and pressures of production, therefore minimising
the risk of being compromised. In the event that the TWC believes that the support is
lacking, that the role is being compromised in some way or that other duties are
preventing the proper execution of the role, it is essential that the DI is informed. This
confirms the earlier point that the DI has to be someone in the organisation with suffi-
cient authority and experience to ensure that the temporary works are not compromised.
g on small contracts, the appointed TWC may not have the resources to be
everywhere since he is responsible for several sites or for a large complex site
g subcontractors may have responsibility for elements of temporary works within
their work package.
In both cases, the Principal Contractor may find it advantageous for one or more TWSs
to be nominated to act as a point of reference and to handle the day-to-day temporary
works. It is therefore feasible for a subcontractor to nominate a TWS to be technically
responsible for that subcontractor’s temporary works within a given site, while remem-
bering that the TWC (as a single point of reference and authority for all temporary works
on that site) has the overall responsibility. Any TWS is therefore technically responsible
to the TWC on all relevant matters; although the TWS may on certain sites be permitted
to sign permits, this should only occur when specifically authorised.
20
throughout a contract. This ensures that the temporary works are properly identified and
managed, whatever the method of procurement. The register may include items that,
through methods of working, prove not to be required; it is infinitely preferable to
have a register that has a number of items which are not required rather than one
from which items are missing.
It should be a live register which is regularly distributed by the TWC. Moreover, it can be
used to demonstrate to interested parties (including HSE inspectors and QA auditors)
that the site and the temporary works are being managed safely and properly.
The brief should include the programme, any materials that are available, preferred
methods of working and the key information from the designer of the permanent
works (e.g. relevant borehole information and design risks and assumptions). It
should also define the limits of responsibilities. This is best exemplified by proprietary
supplier’s designs for falsework which will, almost without exception, exclude the
necessary foundation design and make fundamental assumptions about the ability of
the permanent works to provide top restraint (see Table 2.1). In both cases, it is essential
that the TWC understands the engineering principles involved and ensures that the brief
identifies and resolves these and similar points fully. Note that although the TWC role
is to manage the temporary works, the appointed person has to have the technical
competence to understand the issues involved relevant to the nature of the work.
It should be noted that TWCs and site personnel who make decisions about methods of
working or materials to be used are taking on the role and responsibilities of the designer
under the CDM regulations. They may have a greater potential to affect the safety in use of
a particular temporary works scheme than the designer who proves, in analytical terms, a
concept that has already been decided and included in the brief. This principle is reinforced
by a legal maxim in respect of design and build, which predates and even predicted the
CDM Regulations to the effect that ‘he who decides, designs’. This maxim is salutary.
Finally, the brief should set down the required level of information (or output) to be
provided by the designer, which should comprise appropriate layouts (such as sketches
and/or clear working drawings) with the particular design risks and design assumptions
clearly communicated to those who need to know. This principle applies equally to a
standard solution, for which the information source should be available and the limits
21
of use and the design risks of such a solution should be clearly communicated to the
TWC. Interestingly, the design risks for a standard solution will in all probability be
more extensive and restrictive than for a specific bespoke design of similar character.
While all this may appear obvious, there are still proprietary suppliers whose design
output is limited to a computer printout. It is therefore worth reflecting on the following
points.
g Designs of temporary works and permanent works, and the operatives and public
who depend upon their safe execution, depend upon the same engineering
principles.
g The Project Manager and the site team expect to be provided with working
drawings and specification for the permanent works. Is it not reasonable for the
same team to be provided with equivalent drawings/sketches and specifications for
the temporary works?
g The person undertaking the design check is required to check from first principles,
given that the check is not a simple arithmetical check or a check of calculations
(ICE/SCOSS, 2002). This therefore presumes that the checker should be working
from drawings/sketches.
g The TWC is expected to ensure that the temporary works is inspected before
loading. In so doing, it is not unreasonable to assume that the TWC requires the
relevant layout and information; this may be from sketches, brochures or a
number of drawings (which have been subject to a design check) against which the
check prior to loading is carried out.
All temporary works designs require to be independently checked. The degree of checking
should be related to the scale of the temporary works; a simple scheme may be checked by
someone in the same office, whereas a more complex scheme might have to be checked by an
outside organisation. The level of check may also depend on the location and the adequacy
of the team members. For example, a major contractor, with many years’ experience and
competent operatives, would consider a 4 m trench as routine work and relatively low
risk in terms of the team’s competence and compliance with good practice and procedures.
A small house builder, rarely digging deeper than 1 m and whose team has limited experi-
ence of significant temporary works, would, however, consider 4 m as high risk.
The four recommended categories of design checks for temporary works in BS 5975 (BSI,
2011) are listed in simplified form in Table 2.2.
It is important that the principle of a ‘higher’ level category of check being required when
the temporary works are more complex is not misunderstood. Any design check should
22
Note: top-restrained falsework is the method by which the temporary structure is stabilised for lateral movement
by connection to external restraints at its head (e.g. to the permanent works of columns or adjacent walls)
provided that these elements have been designed to provide the required restraint
be carried out, by the checker, to a degree of rigour which the checker considers
appropriate to enable the design check certificate to be signed. The degree of check
should not vary depending on whether it is designated Category 0, 1, 2 or 3. The checking
process may be more straightforward for items which have a lower category, but the
checker is only ever carrying out a check that is sufficient to enable him to sign, whatever
the category. The significance of the various check categories is only that they determine
the degree of independence that the checker has in relation to the designer. Some items
of temporary works design therefore require greater independence than others for the
design check, whether because of the need to involve the thinking or experience of a
second organisation or because of contractual requirements or constraints. The risk is
that some checkers may consider that a lesser check is acceptable for lower category
items.
Other forms of contract may have other requirements for certification of the temporary
works, particularly those from the Highways Agency (HA). Temporary works on HA
contracts which have a public interface will be subject to an Approval in Principle
(AIP or equivalent) followed by the associated design and check certificates.
On rail contracts, a ‘Form C’ is required for all temporary works that affect the safety of
the railways. There is also a Network Rail (NR) form covering the design, design check,
NR approval and, if applicable, issue by subcontractor and approval by the zone civil
engineer.
It should always be remembered that checks that are required by, or undertaken by,
other organisations should in no way be considered as an alternative to or as a reason
to reduce or omit any of the checking stages that the organisation responsible for the
temporary works is required to undertake.
23
The TWC should provide the checker with the design brief, the relevant layouts and
information (drawings/sketches) and the residual design risks identified by the designer
of the permanent works; calculations are not normally provided. The checker’s role is to
carry out an independent check of the concept, working from first principles (which is
why calculations are not provided). With the increasing use of computers and other
design aids, the importance of simple rule-of-thumb checks should not be overlooked.
The check is therefore not an arithmetic check of the original designer’s calculations,
due to the risk that a fundamental error may be repeated by the checker. The checker
must issue a check certificate, which lists all the documents and drawings as necessary
with their revision status. Indeed, without such a list, the design check certificate is of
questionable value.
Although BS 5975 (BSI, 2011) recommends the four categories shown in Table 2.2,
individual companies may have different views on the scale of works and ‘simple’
versus ‘complex’ or ‘minor’ versus ‘major’ can have different meanings. Often the
categories of temporary works are defined in other terms, but the essential philosophy
remains that all temporary works designs are checked from first principles. Note that
use of ‘unclassified’ can never be used as a scale because, fundamentally, all items of
temporary works have to be given a category (even if it is only ‘Category 0’).
Finally, and only when the TWC is in all respects satisfied, the TWC (or TWS where
authorised) signs and issues a permit to load to the Project Manager, which will generally
be limited in time. If the TWS is authorised to sign the permit, the TWC must also be
included in the process as the single point of authority and responsibility. It would be
24
incorrect to issue a permit to load falsework some weeks in advance of a pour date,
given the inclination of site teams to use misplaced initiative and ‘borrow’ key
components for another element of temporary works. Falsework permits are therefore
normally issued the day before and are valid for the following day only. The actions
of anyone on the site team who wilfully loaded temporary works in full knowledge
that the permit had been withheld or had not been issued would normally be treated
as a disciplinary matter.
The TWC can inevitably come under immense commercial pressure from the site team to
sign and avoid delay and cost. It is in these situations that the TWC has to have sufficient
strength of character and the confidence and full support of the PM. Given the
consequence of failure, the presumption is that the TWC has to withhold permission if
there is sufficient doubt.
Checklists for some temporary works have been issued; refer to available guidance on
falsework (Concrete Society, 1999), formwork (Concrete Society, 2001) and trenching
(CIRIA, 2001).
For concrete structures, the value and method of assessing the concrete strength at the
time of striking should be agreed beforehand with the PWD and carefully controlled
on site. Modern methods of strength assessment are now available; see Chapter 24 on
soffit formwork.
Structures that combine temporary works with prestressing of the permanent works or
multi-storey floor slab construction with a requirement for back-propping need careful
planning and control. The exact order of removal of supports can affect the load transfer,
so the permit to unload should state any required sequence or procedures for striking or
unloading the structure.
2.5. Summary
This chapter describes the background to and the reasons why and how temporary works
are managed and controlled. It has not sought to repeat fully the detailed procedural
points in BS 5975 (BSI, 2011). The TWC’s role is to protect the team, including the
operatives, from errors, omissions, misunderstandings and perhaps even the pressures
of programme and cost. The TWC can be assisted by a TWS, but remains the responsible
person.
25
Failures of temporary works can all too easily cause serious injury or fatalities with the
associated distress to families and, in all probability, legal action. Contracts will incur
delay and cost that is difficult to recover. The industry relies upon and therefore owes
it to those who work with, within, on, or under temporary works to follow the estab-
lished procedures. In the event of failure and enquiry or legal action, individuals must
be able to demonstrate that established practice rules have been followed.
Finally, while it can never be an absolute rule, failures are generally caused by a series
or combination of factors. It follows that such failures can best be prevented by the
consistent application of a similar series of checks and balances; this is why temporary
works are controlled.
REFERENCES
Bragg SL (1975) Final Report of the Advisory Committee on Falsework. Department of
Employment and Department of the Environment. HMSO, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
Burrows M, Clark L, Pallett P, Ward R and Thomas D (2005) Falsework verticality:
leaning towards danger? Proceedings of ICE, Civil Engineering, February 2005: 41–48.
CIRIA (Construction Industry Research and Information Association) (2001) Trenching
Practice, 2nd edition. CIRIA Report 97, London.
CITB (Construction Industry Training Board) (2007) The Construction (Design and
Management) Regulations. Industry Guidance for Designers. Construction Skills,
Norfolk.
Concrete Society (1999) Checklist for Erecting and Dismantling Falsework. Ref CS123.
Concrete Society, Crowthorne.
Concrete Society (2003) Checklist for the Assembly, Use and Striking of Formwork. Ref
CS144. Concrete Society, Crowthorne.
HSE (Health and Safety Executive) (2001) Investigation into aspects of falsework. HSE
Contract Research Report 394/2001, HSE Books, Sudbury. Available at www.hse.gov.
uk/research/crr_htm/2001/crr01394.htm.
HSE (2007) Managing Health and Safety in Construction, Approved code of practice.
Publication L 144, HSE Books, Sudbury (including Construction (Design and Manage-
ment) Regulations 2007).
ICE (Institution of Civil Engineers) (1986) ICE Conditions of Contract, 5th edition.
Thomas Telford Publishing Ltd, London.
ICE/SCOSS (Institution of Civil Engineers/Standing Committee on Structural Safety)
(2002) Falsework: Full Circle? Report on Temporary Works, Topic Paper SC/T/02/01,
London.
26
Chapter 3
Site compounds and set-up
Geoff Miles Quality Manager, Costain
Typically, a construction project manager may be responsible for completing the design
and build of a new £40M facility over a timescale of 2.5 years. In terms of his role and
responsibilities, that makes him the equivalent of the CEO of a small-to-medium
‘start-up’ enterprise, with revenue of £16M in the first year. In most other industries,
manufacturers already have their permanent, sophisticated, production facilities fully
operational before they take on new orders for clients. Upon award of a new contract,
and within just a few weeks, Builders and Civil Engineers have to first create their
‘temporary factory and welfare facilities’ on a distant site before they can start their
client’s work. They establish (or set-up) this often concurrently with commencement
of delivery of the permanent works. Set-up requires management of a very large
number of trades. This chapter provides advice on the range of items that may need
to be designed, costed and included in the project preliminaries cost plan.
3.1. Introduction
Apart from the obvious operational needs to provide production facilities, the following
legislation imposes obligations on all parties involved when working in the UK: The
Health and Safety at Work Act (1974); The Management of Health and Safety at
Work Regulations (1999); The Construction (Design and Management) Regulations
(2007); The Workplace (Health, Safety and Welfare) Regulations (1992).
Under the CDM regulations (2007), Clients, Designers, Principal Contractors and
Contractors all have specific duties.
The Client must make sure that the construction phase does not start unless there are
suitable welfare facilities and a construction phase plan in place. The Principal
Contractor must plan, manage and monitor the construction phase in liaison with
contractors; prepare, develop and implement a written plan and site rules (the initial
plan must be completed before the construction phase begins); ensure that suitable
welfare facilities are provided from the start and maintained throughout the construction
phase; and secure the site.
The Contractor must plan, manage and monitor their work and that of their workers and
ensure there are adequate welfare facilities for their workers. Where contractors are
involved in design work, including for temporary works, they also have duties as
designers. When working overseas or in other jurisdictions, advice should be sought
27
regarding any local requirements and obligations. Otherwise, the UK standard is a good
model to use to provide a safe and healthy working environment for employees in any
location, at home or overseas.
The best time to plan and quantify the detailed compound, offices and welfare
requirements is during the Early Contractor Involvement or Tender Planning stage. It
is important to have a detailed solution in place in order to estimate the preliminary
cost budget accurately.
Table 3.1 provides a generic list of what may need to be planned, designed and costed.
This should be carefully considered and edited to exclude any unnecessary items.
Equally, any missing site-specific items missing should be added.
A detailed description of issues including land and access; communications, energy, clean
water supply and waste water disposal; office and welfare accommodation space plan-
ning; and materials (unloading, distribution, fabrication, handling, storage and
testing) is provided in the following sections.
28
Access Public access, diversions Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Traffic light controls Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Rail branch/siding/platform Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Highways licences Application form and letter of submission
Electricity Metered electricity supply from Design, layout drawing 1 : 100, details 1 : 20,
existing network 1 : 10 and specification
Generator and diesel tanks, Design, specification, 1 : 100 plant layout
refuelling logistics (emergency or drawing and base (civils) details
permanent)
Electrical distribution power and Design, layout drawing 1 : 100, details 1 : 20,
lighting including 110 V, major 1 : 10 and specification
plant, etc.
Substation/switchgear building Design, specification, 1 : 100 layout
drawing, building details and plant
drawings
Trenches and ductwork for buried Design, layout drawing 1 : 100, details 1 : 20,
services 1 : 10 and specification
Construction site electrical Specification and method statement
regulation testing regime
29
Gas Metered gas supply from existing Design, layout drawing 1 : 100, details 1 : 20,
network 1 : 10 and specification
Bulk LPG tank Design, specification, 1 : 100 plant layout
drawing and base (civils) details
Gas distribution pipework Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Gas valve and meter building Design, specification, 1 : 100 plant layout
drawing and building details
Trench and ductwork for buried Design, layout drawing 1 : 100, details 1 : 20,
services 1 : 10 and specification
Construction site gas regulation Specification and method statement
testing regime
Environmental impact assessment Report
(for gas leakage, etc.)
Security fencing re gas equipment Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Automatic safety and shutoff Design, layout drawing 1 : 100, details 1 : 20,
systems 1 : 10 and specification
Voice and Aerial/dish mast base Design, layout drawing 1 : 100, details 1 : 20,
data 1 : 10 and specification
Satellite receiver switched link Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Cooper wire, switched link Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Fibre-optic, switched link Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Data and voice wiring network Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Cable ducts and draw-pit layout Design, layout drawing 1 : 100, details 1 : 20,
1 : 10 and specification
Voice and data equipment Schedule
schedule
Internal communication system – Schedule
radios
Security Fire alarm system Fire alarm system schematic drawing 1 : 100
and specification
30
Welfare Facilities for all site personnel General arrangement layout drawing 1 : 100
Predicted manpower plan Labour histogram produced from bar-chart
programme
Canteen Detailed room drawings 1 : 50
Drying room Detailed room drawings 1 : 50
Changing room Detailed room drawings 1 : 50
M and F toilets Detailed room drawings 1 : 50
Showers and emergency showers Detailed room drawings 1 : 50
Living accommodation on site Detailed room drawings 1 : 50
Health and first-aid Detailed room drawings 1 : 50
Remote welfare facilities about site Detailed room drawings 1 : 50
Offices Facilities for all management staff General arrangement layout drawing 1 : 100
Management organogram Organisation chart
Subcontractors offices Detailed room drawings 1 : 50
Main contractors offices Detailed room drawings 1 : 50
Office furniture layout Detailed workstation and furniture layout
drawing 1 : 50
IT office equipment Detailed IT, printing and office equipment
layout drawing 1 : 50
Filing system/room Detailed room drawings 1 : 50
Stationery store Detailed room drawings 1 : 50
Cleaners store Detailed room drawings 1 : 50
M and F toilets Detailed room drawings 1 : 50
Print room Detailed room drawings 1 : 50
Drawings room Detailed room drawings 1 : 50
31
Gantry For congested sites (usually inner Structural steelwork design layout and
city) fabrication drawings
Highways licence Application form and letter of submission
Compound Temporary soil heaps Location and design drawing 1 : 100, and
specification
Hard-standings Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Fencing Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Hoarding Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Noise Screens Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Gates Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Vehicle entry barrier Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
32
Waste Skip standing and segregation Design, drawing 1 : 100, details 1 : 20, 1 : 10
management area and specification
Concrete plant ‘wash out’ area Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Food refuse bins and storage Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Incinerator Design, drawing 1 : 100, details 1 : 20, 1 : 10
and specification
Plant Refuelling station and fuel storage Design, drawing 1 : 100, details 1 : 20, 1 : 10
maintenance and specification
Recharging station; battery- Design, drawing 1 : 100, details 1 : 20, 1 : 10
powered vehicles and specification
Vehicle repairs and maintenance Design, drawing 1 : 100, details 1 : 20, 1 : 10
bay and specification
33
Access routes should also be walked, especially in remote areas away from major roads or
in congested town centres where many restrictions can apply (e.g. limited access in length,
width, height and axle-weight). Look for overhead cables, noise and time limitations.
To prepare for a site visit, contact all relevant authorities and request copies of their
drawings showing all existing infrastructure, i.e. roads, sewers, water mains, electricity
cables, gas mains and communications cables. During the site visit make a point of
finding physical proof of items such as manholes, draw-pits, valve access covers, fire
hydrants, substations and cable pylons or poles shown on the drawings provided.
Also identify all potential locations for the site compound(s).
Finding that location requires detailed analysis and sketch modelling of the drawings and
specifications provided for the permanent works. The timing of the availability of such
information depends very much on the two principal methods of procurement, either
g traditional: the Client provides a complete, fully detailed design, specification and
bill of quantities for the main contractor/specialist subcontractors to price or
g design and build: the Client provides a schematic design and performance
specification requiring the main contractor/specialist subcontractors to complete
the detailed design, specification, materials take-off and cost plan.
34
In cases where the detailed design will not be completed pre-tender or pre-guaranteed
maximum price, then the following options must be considered
g the use of a piece of land or an existing building close to the site (this removes the
risk of the temporary set-up being in the way of the permanent works), or
g include for relocating the set-up in the tender cost budget.
Lease negotiations for land and premises can take a considerable time. This will involve
securing the option to lease contracts and obtaining outline planning approval during the
tender or ‘early contractor involvement’ period. Without these, it would be irresponsible
to include assumptions about their availability for use in any tender submissions.
See Figures 3.1 and 3.2 for typical compound layouts for a road project and a small site.
35
Position of compound
and high value storage area
70 m × 40 m
18 × 11 m 1.8 m
Close board timber fencing
e
on
2.4 m Chainlink fencing
gz
itin
wa
icle
h
Ve
cti er
tru tt
3 No. 6 m
on
ns hu
1 No. pedestrian
access gate
Sto
ati a
dic re
)
ve
(In t 1 A
Lo
ati a
dic re
g the proximity of utilities to connect up to, i.e. sewers, water mains, electricity
mains, gas mains, telecommunications cable network, and
g access between the road network and site for cars and heavy goods vehicles.
Established roads and utilities infrastructure may not exist in remote or under-developed
areas and so energy and water sources, water treatment, communications and transport
links have to be considered from first principles. Provision for designing and constructing
the following should be included.
36
IT systems, software and support are high cost overheads. Profit margins in construction
are notoriously low, generally between 1 and 2%. Budgets for IT investment are there-
fore very low compared to other, more profitable industries. As a result, only basic
services and systems are affordable.
In the UK, the mobilisation lead time for a new cable internet link, measured from
placing an order to commissioning, is on average 20 weeks. However, there is rarely
more than 4–6 weeks between signing contracts and being expected to make a start
and have a management team based on site.
1 During the first 20 weeks, use mobile 3G or satellite receiver systems. These
provide an instant email and database service. Performance is variable and often
frustratingly slow, but is better than no service at all.
2 In the period leading up to week 20, the local area cable network (LAN),
switchgear and fibre-optic/copper-wired link will be installed and ‘go live’. The
LAN will be used from week 20 for the remaining duration of the project, linking
users to shared equipment such as printers, plotters, scanners, internet and
telephones. Access to the project database and documents will be available at
acceptable upload/download speeds.
Before deciding on a system and placing supplier orders, two steps are required. For the
first step, a briefing document must be drawn up to pass to the IT system designer. This is
comprised of three parts: as follows.
1 A questionnaire asking for site address and project duration with the start and
completion dates; number of users; software required and the type of web-based
37
document sharing system; and types and number of shared equipment such as
printers, plotters, scanners, internet, telephones, projectors, etc.
2 A compound layout drawing and a detailed office plan layout drawing showing
the positions of communications hub room, work stations, printers, plotters,
scanners, telephones and projectors. Positions of data socket outlets should be
drawn on the plan, including some in the meeting room(s). It is prudent to have a
few extra socket outlets included for visiting ‘hot desk’ users.
3 A site location map.
The brief is passed to the IT systems designer in the second step so that a fully costed IT
system specification and design can be produced. This should always be checked and
explained to ensure that the proposals are a correct interpretation of the brief.
3.3.2 Energy
Electricity is the most popular and flexible energy to use. Reliable sources include either:
national grid or local distribution cable networks; or hired electricity generators
(commonly diesel engine powered). Generators can also be driven by gas engine or
turbine or even jet engine. The choice of fuel will depend on safety risk assessment, avail-
ability and cost. Isolated wind power and hydro-electric alternatives are not reliable.
g In offices: power socket outlets, light fittings, space heaters, water heaters,
domestic appliances, office equipment, IT and communications switchgear, fire
alarm, intruder alarm, access control, trace heating.
g In welfare facilities: power socket outlets, light fittings, space heaters, water
heaters, domestic appliances, kitchen equipment, fire alarm, intruder alarm, access
control, trace heating.
g Compound: external lighting, CCTV system, access control, vehicle entry barrier,
weighbridge, wheel-wash, battery charging plug-in sockets for large battery-
powered moving plant.
g Materials storage shed: power socket outlets, light fittings, space heating, fire
alarm, intruder alarm, access control, dehumidifier.
g Materials testing laboratory: power socket outlets, light fittings, space heating, fire
alarm, intruder alarm, access control, trace heating, cube curing tank, special test
equipment.
g Maintenance workshop: power socket outlets, light fittings, space heating, fire alarm,
intruder alarm, access control, welding equipment, hoist and vehicle inspection ramp.
g Fabrication workshop: power socket outlets, light fittings, space heating, fire
alarm, intruder alarm, access control, overhead gantry crane, manufacturing
plant, welding equipment.
38
g Items of major plant located in the compound or elsewhere on site, for example,
tower cranes, concrete batching/mixing plant, mortar silo mixers, materials or
passenger hoists, etc.
g Site distribution to transformers on all floor levels to provide power outlets for
110 V power tools, welding points in plant rooms, etc.
g Site distribution to all floor levels for temporary lighting, fire alarm system, etc.
Using the schematic described above as well as utilisation factors, the total electricity
demand can be calculated and used to pick the most economical source of supply.
The second step is to prepare a detailed distribution layout plan showing cable sizes,
routing and switchgear locations. In order to do this, the electrical engineer will need
the completed detailed compound layout plan and temporary buildings layout plans.
Distribution of electricity on site is covered by BS 7375 (BSI, 2010).
Once the total quantity and flow rate required has been calculated, the next step is to
decide upon a water source. Depending upon the location, the choices of source will
be one of the following.
g Connect to the existing water mains supply. Obtain an application form from the
relevant water supply company and apply for a new (temporary) metered water
supply pipe. They provide a proposal and quotation for the supply. Once they
have received payment for the quoted sum, they then install their pipework from
their mains (including a water meter) up to a stop valve located at the site
boundary. Water usage is paid for on a rate per cubic metre.
g Import water by bulk tanker. Again, obtain an application form from the relevant
water supply company and apply for them to supply water in bulk tankers. They
provide a proposal and quotation for the supply. Once they have received
payment for the quoted sum, they will commence deliveries. Water usage is paid
for on a rate per cubic metre. A tanker hard-standing area is required with
suitable valve chambers to allow the tanker discharge pipes to be connected to the
site network.
g Extract water from a river, a borehole or the sea. Water extraction licences and
borehole drilling licences have to be applied for and granted before proceeding
39
with this option. Water pumps, purification plant and clean water ‘buffer’ storage
tanks are required. When using seawater, desalination plant is also needed.
g Build a dam and form a reservoir. This option would only be the choice of last
resort. Obtaining local planning permission will be necessary before proceeding.
Water pumps, purification plant and clean water ‘buffer’ storage tanks are
required.
Similarly, surface water from roofs, guttering, hard-standings, car parks and roads has to
be collected by a system of pipework, manholes, petrol interceptors and disposed of by
one of the following methods
g From ground surface run-off during rainstorms. A system of ditches, land drains
and silt traps linked to a soak-away should be designed.
g From groundwater seeping into excavations. This is normally dealt with by
excavating sump holes to collect the water and then pump it into a settlement
tank, before pumping it to a soakaway or settlement pond.
g From the de-watering of deep substrata to allow excavation below the water
table. A typical dewatering scheme will comprise a series of well points connected
via a common header collection pipe to the suction side of a large pump. The
pump runs constantly and the discharge water is passed into a segmented
settlement tank. The water is then discharged into a soakaway or settlement
pond. Water quality should be monitored by daily sampling. Consideration
should be given to include piezotubes in the de-watering design to allow water
sampling at depth.
40
Once these quantities are known, the following must then be considered.
g Segregation or integration: the need to separate people relative to their job roles,
work groups, different employers and for reasons of privacy. Bringing multi-
disciplinary teams together in an open-plan environment has become common
practice. There will be a requirement to provide different-sized meeting rooms for
privacy and confidentiality.
g Productivity and quality: open-plan offices should not be assumed. Construction
site offices are notoriously noisy. Many technical and managerial job roles require
privacy, relative quietness, wall space to display bar chart programmes, table-top
space to layout two AO or A1 size drawings side by side, filing cabinet, cupboard
and a desk. Providing individual offices where needed will improve productivity,
quality and staff motivation. Regular team meetings should be used to open up
lines of communication.
g Communication and motivation: operatives need a good-quality canteen and
welfare facilities. Break times are short and time spent away from the workface
must be minimised. Design seating and table layouts that allow sufficient places
for all to be served in ‘one sitting’. Employees should not have to sit in a van to
eat due to lack of space. Equally important are changing, clothes drying and
shower rooms. Workers should not be expected to travel to work in their overalls
or dirty clothes. The induction/training room needs to be a dedicated space and
large enough to accommodate daily new arrivals for Safety, Health and
Environment (SHE) inductions and to accommodate all operatives for shift
briefings.
g Storage and records: separate and secure rooms with ventilation and temperature
control for IT switchgear ‘hub’ cabinet and document filing.
g Modular or volumetric accommodation system: there are numerous suppliers and
systems to choose from. Cost comparisons will need to take into account any
foundations required, setting up and dismantling, craneage and transport and
connectivity (i.e. prewired and only needing to be ‘plugged in’).
Deciding upon and drawing up room layout and furniture plan is a fairly straight-
forward exercise. Most suppliers of modular accommodation have standard generic
drawings available to use or modify into site-specific layouts. In cases where
existing buildings are used as temporary facilities, obtain floor plans from the
landlord or carry out a dimension survey of all rooms and draw the floor plan.
The completed furniture layouts are also used to show small power, data and
voice socket outlet positions. Figure 3.3 shows a typical layout of office and welfare
facilities.
41
42
43
When considering the choices for deciding on the load size, the method of delivery onto
site, storage and movement to the workface, principal considerations including the
following.
Alternative methods and designs for providing a solution for each of these have to be
drawn up and fully costed so that they can be compared and allow the most practical,
efficient and cost-effective combination to be determined. Main Contractors usually
rely upon specialist suppliers and subcontractors to provide details of their requirements.
REFERENCES
BSI (2010) BS 7375: Distribution of electricity on construction and demolition sites. Code
of practice. BSI, London.
Construction (Design and Management) Regulations (2007) HMSO, London.
Health and Safety at Work Act (1974) HMSO, London.
Management of Health and Safety at Work Regulations (1999) HMSO, London.
Workplace (Health, Safety and Welfare) Regulations (1992) HMSO, London.
FURTHER READING
Hall F and Greeno R (2001) Building Services Handbook, 5th edition. Butterworth Heine-
mann, Oxford.
44
Chapter 4
Tower crane bases
Laurie York Consultant
Tower cranes are widely used in the construction industry, especially for building construc-
tion. They are normally supplied on hire and the customer, usually the Principal Contractor,
is responsible for the design and construction of the base upon which the crane will be
erected. Details of loadings imposed on the base by the chosen crane are provided by the
crane supplier and the customer designs the base as temporary works. Loadings are normally
provided for ‘in-service’ conditions (when the crane is working but wind speeds are restricted
to ‘working wind’ speed) and ‘out-of-service’ conditions (when the crane is not working but
is free to weather-vane, and maximum wind speeds may occur). A further set of loadings is
sometimes also provided, related to the erection of the crane. The temporary works base
design process will include checking that the loading details are appropriate for the location
in which the crane is to be erected (particularly with regard to the wind speed used by the
crane supplier in deriving the loadings) and will also entail liaison with piling contractors
(if the base is to be piled) and the Permanent Works Designer (if the crane base is to
impose temporary loading on any part of the permanent works during the construction
operations). Similarly, if the tower crane mast is to be supported off an adjacent structure,
the temporary works designer must liaise with those responsible for that structure.
4.1. Introduction
The process of foundation design for tower cranes which are to be used on construction
projects starts during the tender preparation. A particular crane type and configuration
will be identified, for example, saddle-back or luffing, with decisions made concerning
whether the crane is to be static or mobile and assessments made concerning the require-
ments for the maximum loads to be lifted, the height required under the hook and the
maximum radius required about the mast for handling the loads. Other factors which
influence the choice of crane type and configuration include the number of cranes
needed, the proximity of hazards and any restrictions on oversailing which may have
been imposed under the contract conditions.
Figure 4.1 shows typical tower crane coverage of a site using a mobile tower crane. The
track upon which the crane travels may be curved in plan (subject to minimum radius of
curvature conditions, as advised by the crane manufacturer) if this is more suitable for
the area to be serviced by the crane.
Similar considerations are involved when static tower cranes are to be utilised. Figure 4.2
shows typical tower crane coverage of a site, using two static tower cranes. Care is
45
Site boundary
Extent of area
Working radius covered by crane
of crane
Oversailing restrictions
to be investigated
where coverage
extends beyond Site
Foundation
for rail track
Working radius
of crane
Figure 4.2 Typical coverage of site using two static tower cranes
Site boundary
Maximum working
radius of crane
Mast height of each crane
to be set to ensure that
crane jibs do not clash
where coverage ‘overlaps’
Oversailing restrictions
to be investigated
where coverage
extends beyond Site
Extent of area
covered by cranes
Maximum working
radius of crane
46
required if more than one crane is to be used to ensure that, where the coverage areas of
two (or more) cranes overlap, the jib lengths and mast heights are set for each crane so
that they cannot clash with each other.
The location for each tower crane base will be determined by taking account of the
proposed permanent works to be constructed. Sufficient temporary works design for
the foundation will be undertaken at tender stage to enable financial provision for the
construction (and sometimes subsequent removal) of the base to be included in the
tender, and for the crane erection to be planned within the construction programme.
Following award of the contract, full design and detailing is undertaken by the
temporary works designer. CIRIA Report C654 Tower crane stability (Skinner et al.,
2006) provides guidance to designers on all aspects of tower crane foundation design.
Details of overturning moments, vertical and horizontal loads and slewing torques
which will be imposed on the foundation are provided by the crane supplier. This
information is normally supplied in the units of kilonewtons (kN) and metres (m), and
the foundation designer must take care to ensure that corresponding units are used
throughout the design process.
All overturning moments on the crane (from the load on the crane hook, crane counter-
weights and wind loading) are converted into vertical loads acting, through the bogies, on
to the foundation. Details will be provided by the crane supplier. The temporary works
designer must then design a suitable foundation for each of the rails, incorporating
either structural fill (ballast as for railway tracks) or reinforced concrete beams. Loads
provided by the crane supplier will normally be unfactored working loads, but this must
be checked by the foundation designer before any design is commenced. The crane supplier
should also give full details of suitable rail sections, the fixings used to connect the rails to
any foundation and the spacing of ties required between the two beam foundations.
47
compacted structural fill. The foundation will then be similar in principle to a railway
track foundation, with the loads from the bogies spread through the layers of track
support and at 458 through the structural fill and the various soil strata below. The
bearing pressures on each stratum are then compared with allowable bearing pressures.
The temporary works designer should obtain advice from a suitably qualified geotech-
nical engineer with regard to the load-spreading ability of the fill material. Additional
supervision may also be necessary during the time that the crane is in operation on
the site, to monitor any settlement of the foundation and to apply any remedial measures
as necessary.
Extreme conditions involving large loads from the bogies and poor ground may result in
the need for piled foundations. In such cases (which are rare), the foundation is designed
as a continuous beam supported by piles and loaded by the two bogies on the rail. Care is
needed to ensure that all necessary design cases are considered to provide the worst-case
sagging and hogging moments, shear and pile loadings.
P þ Wbase
LB
48
x Bearing pressure on
ground from u/s of base
L
This bearing pressure is then compared with the allowable bearing pressure. This process
can be applied iteratively to achieve the best solution for the beam section.
Having obtained the design loadings for in-service and out-of-service, they are then
compared and the more onerous case used for the beam design. This design loading,
from the bogie, is considered to be spread evenly over the full width and length of the
beam adopted for the design (Figure 4.5). The reinforced concrete section is then
designed, both longitudinally and transversely.
Bearing pressure on
ground from u/s of base
B
49
Table 4.1 Partial (load) safety factors for design of foundation beam
Data taken from CIRIA C654
The maximum bending moment in the beam occurs below the centres of the bogie
wheels; its value is
P Lx 2
0:5
L 2
The maximum shear in the beam occurs at the same location but, in accordance with
BS 8110, recognition of the enhanced shear strength of beams near their supports
allows the beam to be designed to suit the shear at a distance, equal to the effective
depth of the beam d, away from the centres of the bogie wheels. The value of the
shear at these points is:
P ðL xÞ
d
L 2
In the transverse direction, a 1 m length of the beam is considered ‘supported’ by the rail,
with upwards loading as for the longitudinal direction. For simplicity, the width of the
rail is neglected (Figure 4.6).
50
and the shear (calculated in the same way as for the longitudinal direction) is
P B
1:0 d
LB 2
This design process, using BS 8110 (section 3.4) and incorporating the appropriate
partial material factors m, will produce reinforcement requirements for the bottom of
the beam and for shear links. It is unlikely that shear links will be required in the
transverse direction but, if the calculations indicate that shear reinforcement is
required in that direction, consideration should be given to modifying the chosen
section dimensions in order to remove this requirement, as fixing shear links in both
directions within a beam can be difficult. It is also prudent to detail minimum
reinforcement content, in accordance with the recommendations of BS 8110, for the
top of the beam. This will resist any tension in the top of the beam caused by
moments resulting from the beam ‘spanning’ between the bogies, and will assist in the
fixing of the shear links.
On completion of the design of the reinforced concrete section, the reinforcement should
be detailed similarly throughout the full length of each of the rail foundations.
When it is necessary to adopt a piled foundation, the beam supporting the crane rail is
designed to suit the various combinations of loads from the crane bogies at all positions
along the rail. The loads are factored, using the appropriate values of f noted above, and
the beam designed with reference to BS 8110 as before.
51
specially fabricated by the crane manufacturer and which are cast into a reinforced
concrete base. The reinforced concrete base is designed by the temporary works designer,
either as a ground-bearing pad or as a piled foundation, to withstand the loads and over-
turning moments as advised by the crane supplier. The minimum depth requirement for
the ground-bearing pad or for the pilecap will be advised by the crane supplier, ensuring
that the embedment depth of the crane mast into the concrete is sufficient to be capable
of transmitting the loads from the crane into the foundation.
The choice of whether to adopt either a crane which utilises kentledge at the base of
the mast to provide stability or a crane which has its mast cast into the foundation is
probably dictated by space available on the site. A tower crane with kentledge around
the mast can take up a significant area whereas a crane with a cast-in mast base
section only occupies a limited space, particularly if the foundation is constructed with
its top at (or below) existing ground level.
52
D
G
Moments taken
about this edge
L
The chosen size for the concrete block is taken and its self-weight calculated. Depending
on the crane type and configuration, base sizes are often in the range 1–1.5 m deep and
4.5 m square–6 m square. The (unfactored) vertical load from the crane (as provided by
the crane supplier) is then added to the unfactored self-weight of the foundation.
Moments are then taken about one of the bottom edges of the concrete block, using
unfactored values (Figure 4.8).
L L D 24
The overturning moments imposed by the crane on to the foundation (MO) are
M A þ ðH DÞ
L
ð V þ GÞ
2
M S 5 1:67 M O
This process may be applied iteratively to achieve the best base size for stability.
53
L – 2ey
ey
L
L
At this stage, it may be appropriate (if possible) to review the overall dimensions of the
concrete block to reduce the imposed bearing pressure on the soil without reducing the
overall stability factor of safety below 1.67.
54
Table 4.2 Partial (load) safety factors for design of block foundation
Data taken from CIRIA C654
In-service Out-of-service
condition condition
If a concrete block which will provide adequate stability without overstressing the soil
below cannot be designed using the approach described above, it will be necessary to
design a piled foundation.
A similar design process to that used to determine the bearing pressure on the ground is
then followed for both in-service conditions and for out-of-service conditions, using the
appropriate partial load factors and the partial material factors m in accordance with
the guidance in BS 8110.
The transformed uniform pressure derived less the factored bearing pressure due to the
self-weight of the block ( f ¼ 1.0) (as noted in Table 4.2) is then used to design the foun-
dation block reinforced concrete section (Figures 4.10 and 4.11). An example of this
calculation can be found in CIRIA Report C654 (Skinner et al., 2006).
55
A A
L – 2ey
ey
L
L
legs, in plan, so that congestion of reinforcement at the pile top does not create problems
for positioning the mast base section. The pilecap should extend a reasonable distance in
plan beyond the piles in order to avoid problems associated with punching shear around
the pile perimeter. If piles are being installed elsewhere on the site, it is recommended that
the tower crane foundation is designed using similar piles. It is often possible to design
foundations incorporating four piles. This is recommended and any foundation should,
if possible, be square with a symmetrical arrangement of piles.
The depth of the pilecap is chosen to suit the embedment requirements of the tower crane
mast cast-in section and plan dimensions chosen, assuming four piles. In a similar
manner to that adopted when carrying out the design of a foundation without piles,
initial analysis uses unfactored loads and moments and is carried out for both in-
service and out-of-service conditions,with the overturning moment imposed by the
Transformed uniform
bearing pressure
Section A–A
56
Mo
L
s
x
s
L
crane applied about a diagonal axis (Figure 4.12). An example of this calculation can be
found in CIRIA Report C654 (Skinner et al., 2006).
The final working load combinations of vertical and horizontal loads, as calculated, must
be checked with the pile designer to ensure that the piles will be suitable for the loading
predicted.
A similar design process to that used to determine the pile working loads is then followed
for both in-service conditions and for out-of-service conditions, using the appropriate
partial load factors (Figure 4.13). Having derived maximum and minimum ultimate
pile loads, a comparison with the equivalent working pile loads will normally show
Table 4.3 Partial (load) safety factors for checking capacity of piles to provide stability to base
Data taken from CIRIA C654
57
Mo
L
s
x
s
L
that piles designed to suit working loads will meet the requirements for stability; normal
pile design is based on the working pile loads and incorporates a factor of safety of
at least 2 in compression and 3 in tension. If, however, the ultimate pile loads for
stability are greater than the pile design loads, the pile design must be reviewed by the
pile designer.
Loading on the beams within the pilecap is assumed to be imposed through the tower
crane mast legs. These loads are derived from the in-service and out-of-service load
combinations advised by the crane supplier, converted into loads in the individual
mast legs.
For the beams shown, four load cases are considered: with the overturning moment from
the crane applied about the axis parallel to each of the sides of the pilecap, and about
each diagonal axis. The guidance given in CIRIA Report C654 can be adapted to give
the appropriate values for the partial load factors f for use in the design, as listed in
Table 4.4.
58
Mo Mo
L
s
Mo
s
L
Maximum sagging and hogging bending moments and maximum shear can be deter-
mined for each beam and the reinforced concrete section designed in accordance with
BS 8110 using the appropriate values for m. It is often the case that, due to the depth
of the pilecap necessary to provide the embedment for the tower crane mast legs, the
reinforcement content in both the top and the bottom of the beams will be determined
by the minimum content required in accordance with BS 8110. On completion of the
design of the four beams, if the pilecap is square and there is a symmetrical arrangement
of the tower crane mast and piles, the arrangement of reinforced concrete beams may be
repeated for the other direction across the pilecap with no further design. Finally, top
and bottom reinforcement, similar to that detailed for the beams, is incorporated in
the remaining areas of the pilecap. As with foundations not incorporating piles, it is
not normally necessary to check the section for punching and pull-out shear around
the tower crane mast legs, as the legs will have been designed by the crane manufacturer
to suit the embedment requirements which they stipulate. Checks on punching shear at
the pile tops should, however, be undertaken.
In-service Out-of-service
condition condition
59
REFERENCES
BSI (1997) BS 8110 Part 1: Structural use of concrete. BSI, London.
Meyerhof GG (1953) The bearing capacity of foundations under eccentric and inclined loads.
Proceedings of 3rd International Conference on Soil Mechanics and Foundation Engineering,
Zurich, 1: 440–445.
Skinner H, Watson T, Dunkley B and Blackmore P (2006) Tower Crane Stability. C654,
CIRIA, London. ISBN 978-0-86017-654-1.
60
Chapter 5
Site roads and working platforms
Christopher Tate Chief Engineer, VolkerFitzpatrick
Except on very compact construction sites, vehicles are likely to be used to transport
materials from place to place. Site traffic may comprise road-going vans and delivery
lorries, or construction plant ranging from dumpers weighing a few tonnes to tipper
trucks of thirty-five tonnes or more. In order for these vehicles to move safely and
efficiently across terrain that might be rough and undulating with poor load-carrying
capacity, temporary site roads are needed; these should be distinguished from ‘haul
routes’ used by heavy earthmoving equipment. Site roads generally fulfil only the most
basic function of a permanent carriageway and will often be designed for a relatively
short working life; they are generally constructed cheaply using unbound granular
material or stabilised soil and must be capable of being simply and rapidly repaired
should localised failure or excessive rutting occur during service. Purpose-built
working platforms are needed to support heavy construction plant operating in
defined locations; for example, tracked piling rigs and cranes and mobile cranes with
outriggers. Platforms must be strong enough to ensure that the ground underneath is
not overstressed during peak loading and they should be relatively flat and level so
there is no risk of plant toppling during manoeuvring.
5.1. Introduction
Provision for the efficient movement of delivery vehicles, muck-away lorries and
construction plant across a site is vital, yet may be given little forethought and a low
or non-existent budget. Ground that appears hard and able to support traffic in a hot
summer season can quickly be transformed into a quagmire as the autumn and winter
rain arrives; production is curtailed and minds become focused on the problem of
transporting materials from A to B. Temporary site roads capable of functioning all
year round should be central to project planning. If designed appropriately and built
before ground conditions are allowed to deteriorate, site roads are cost effective; they
will of course have a tangible cost but should more than prove their worth by keeping
supply lines open (Figure 5.1).
Working platforms are considered by some to be a luxury. Cases of mobile cranes tilting
or toppling during a heavy lift or piling rigs unable to extract casing as a result of
poor ground support suggest otherwise. Correctly designed platforms enable lifting or
piling operations to be undertaken with confidence at any time of year and whatever
the foreseen ground conditions, and have a place on most projects.
61
Figure 5.1 The need for a properly designed and constructed site road
Courtesy of Tensar International Ltd
This chapter deals primarily with the design and construction of temporary roads and
working platforms for use by site vehicles and plant. The temporary diversion of
public roads within the site is a special case, which is covered briefly.
A minimum road width of 3.5 m will accommodate normal site traffic in single file, with
local widening (typically 32/R) on short radius (R) bends to allow for the passage of long
or articulated lorries. Passing places 15 m long with 1 : 5 tapers should be provided at
suitable intervals and at blind bends; the length may be reduced to 12 m if site traffic
is not expected to include articulated lorries. For two-way working, a 6.5 m wide road
will avoid passing vehicles running too close to an edge and causing excessive haunch
damage.
62
The damage inflicted on a road by the passage of a vehicle is related to the axle loading.
Current knowledge is that there is a fourth power relationship; i.e. doubling the vehicle
axle loading results in 16 times the damage. This effect is taken into account by assigning
each type of vehicle likely to be using the site roads a number of ‘standard’ axles (a
standard axle is 8160 kg or 80 kN). Table 5.1 lists the numbers of standard axles
associated with various types of road vehicle and construction plant operating at full
payload (BritPave, 2007). It is clear that a single fully laden lorry will have as much
influence on the structural design of the road as a fleet of vans.
In the absence of data for a particular item of wheeled construction plant, its number of
standard axles (sa) can be estimated from the manufacturer’s technical specification
Table 5.1 Equivalent standard axles for typical vehicle types (fully laden)
Data taken from BritPave (2007) courtesy of Alex Lake
63
Tracked plant is designed to exert a relatively low ground pressure while travelling and
its effect on the road design compared to say, a fully laden four-axle tipper lorry, will be
insignificant.
On large civil engineering projects where heavy dump trucks or motorised scrapers are
employed for earthmoving, it is not normal practice for haul routes to be hardened
since the pattern of plant movements may be subject to continual change. Earthworks
operations by their nature generally take place in the late spring to early autumn
period; at this time of year the ground itself is able to provide the necessary support
for the vehicles involved (albeit with continual re-profiling of an often well-rutted
surface). Heavy earthmoving plant would therefore not be routinely included in the
traffic-loading assessment for site roads, although in special circumstances some
account may need to be taken (e.g. should haul routes cross site roads).
The product of the anticipated number of movements (n1, n2, n3 . . .) of each fully
laden type of vehicle (V1, V2, V3 . . .) and the number of standard axles associated with
that vehicle (sa1, sa2, sa3 . . .) should be summed for all the expected vehicle types:
This will give the total number of standard axles to be carried by each branch of the
site road network over its lifetime. Because site roads are likely to have a relatively
uneven surface, there will be an element of dynamic loading not usually considered in
the design of public roads; to take account of this it is suggested that the calculated
total number of standard axles be doubled. The resulting number of standard axles is
the design traffic loading:
sa(design) ¼ 2sa(total)
In cases where the site road is merely the foundation layer for a permanent road to be
built at a later date, a more simplistic approach to assessing design traffic loading can
be adopted. There are published data relating the number of standard axles expected
during construction to both
g the length of permanent road under construction (Powell et al., 1984) and
g the size of the development served by the permanent road (BSI, 2001).
If a comprehensive ground investigation report is available for the site it may well
contain some or all of the information needed. Such information includes
64
Should information be lacking it will be necessary to excavate trial pits at intervals along
the proposed route to sample and test the ground to obtain the required data. If time is
short, in situ testing may substitute for some of the laboratory tests. For example, a
Mexe-probe or Transport Research Laboratory (TRL) dynamic cone penetrometer
(DCP) could be used to estimate current CBR (HA, 2009) and a hand vane to
measure undrained shear strength. For a clay soil, shear strength can be converted to
CBR using the approximate relationship:
where y ¼ 23 for fill (Black and Lister, 1978) and y ¼ 30 for undisturbed ground.
Basic soakage tests (BRE, 1991) performed in the trial pits would help decide if the site
roads need to be drained or whether natural percolation into the subgrade will suffice.
If the trial pits expose weak ground that would benefit from stabilisation, or soil
stabilisation is being considered in lieu of granular material for the road construction,
it is essential that bulk samples be taken to a specialist laboratory to determine the
feasibility and most economic method of stabilising (i.e. which type of binder is most
suitable and the necessary percentage addition). Chemical tests and swelling tests must
also be undertaken to ensure there are not elevated levels of undesirable or potentially
disruptive substances in the soil (HA, 2007).
65
(in mm) of unbound granular material that will limit rut depth to 75 mm for a given
traffic loading (sadesign) and subgrade CBR (%):
The equation was first presented by Giroud and Noiray (1981) and relates to low
permeability subgrade soils (clay/silt); the granular material itself should have a
CBR > 80%.
For example, for traffic loading of 1000 standard axles (approximately 100, 4-axle
tippers with 20 tonne payloads and dynamic loading allowance) and a subgrade CBR
of 2.5% (firm clay), the thickness of unbound granular material (well-graded crushed
rock) needed for the temporary road would be:
While a rut depth of 75 mm is acceptable for a temporary site road, it is too much if the
granular layer is to become the foundation for a permanent road. For the latter case,
Powell et al. (1984) present a second equation (rearranged below) where the unbound
granular material is Type 1 subbase and the rut depth is limited to 40 mm:
Using the above example, the thickness of unbound granular material increases to:
The use of both design equations is generally restricted to sadesign 410 000. The subgrade
CBR used for design purposes should be the lesser of
g the value measured in the trial pits at the time of the ground investigation
g the equilibrium value based on soil type and plasticity index (Powell et al., 1984)
g the value estimated from Mexe-probe/DCP/hand vane testing at the time of
construction.
Having calculated a suitable thickness for the granular road, the following rationalisa-
tion should be adopted.
g If the subgrade is weak and the design CBR falls below 2.5%, formulate an
alternative design incorporating geogrid reinforcement; as well as saving on the
thickness of granular material, the geogrid should give better support for
construction plant during placing and compaction.
g For subgrades with a design CBR of 2.5% or more, the calculated thickness of
unbound granular material may be adopted, subject to the following provisos.
(i) Where the design CBR does not exceed 15%, a minimum road thickness of
225 mm would be prudent.
66
Figure 5.2 Summer construction on chalk subgrade; temporary site road not required
(ii) On non-clay subgrades with a design CBR in the range >15–30%, the road
thickness should not be less than 150 mm.
(iii) On subgrades of gravel where the CBR is >30%, it may not be necessary to
provide a road structure at all if the ground is relatively free-draining (soil
infiltration rate >104 m/s).
(iv) Subgrades of chalk can be troublesome. Even though low-density chalk can
generally be trafficked directly during the drier seasons (Figure 5.2), it may be
impassable during winter (particularly if damaged by frost) when the CBR
can drop below 2.5%. The site road design should be assessed accordingly.
It should be noted that if the site road is also the foundation layer for a permanent public
road there may be special requirements in respect of layer thickness and the quality of
materials used; the views of the supervising highway authority should be sought.
A geotextile covered with compacted unbound granular material to form a road needs
the subgrade to deform significantly under wheel loading in order to function as a
67
Figure 5.3 Interlock of aggregate and geogrid increases lateral restraint (confinement)
Courtesy of Tensar International Ltd
‘tension membrane’ (i.e. a state in which it both confines the granular material and
reduces the pressure applied to the subgrade soil). Therefore, a rut must develop
before the full contribution of the geotextile occurs; typically the rut depth would be
75 mm, which limits the effectiveness of geotextiles as reinforcement in permanent
works. However, a geotextile could act as reinforcement in the special case of a
narrow temporary site road where traffic is channelised and a regime for maintenance
of rutting is implemented. While a geotextile may not in most circumstances permit a
reduction in the thickness of the unbound granular layer, it may confer other benefits.
For example, it could prevent the intermixing of the aggregate with a soft cohesive
subgrade soil (separation) and filter any soil water rising up into the granular layer
on wet sites (filtration). Guidance on designing with geotextiles is available from
manufacturers (e.g. Terram, 2010).
Design software for geogrid-reinforced site roads (and working platforms) is available
from a leading manufacturer (Tensar, 2010). Based on the output from this software,
Figure 5.5 compares the thicknesses of granular material needed to accommodate a
range of traffic loading with and without geogrid reinforcement. A geogrid can be
supplied with a factory-attached geotextile (i.e. as a geocomposite) to provide the
combined benefits of confinement, separation and filtration.
Geocells are an extension of the geogrid principle, effectively increasing the depth of the
‘ribs’ to the full thickness of the granular layer. The quality of aggregate used to fill the
68
Figure 5.4 Increased restraint (confinement) at base of unbound layer leads to better load
spreading and improved subgrade bearing capacity
Courtesy of Tensar International Ltd
cells may be inferior to that needed with geotextiles or geogrids, thereby recouping some
of the extra cost of the geocell structure. Reinforcement of unpaved granular roads is a
relatively new application for geocells and empirical design guidance needs to be sought
from manufacturers.
69
Figure 5.5 Comparison of unbound granular material thicknesses for site roads with and without
geogrid reinforcement (40 mm rut depth)
1300
1200
1100
1000
900
800
Thickness: mm
700
600
500
400
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5 5.5 6 6.5 7
Subgrade CBR: %
For lightly loaded platforms used for storage or supporting falsework on firm ground, a
thin concrete blinding layer could suffice (in conjunction with timber sleepers below
70
Figure 5.6 Geogrid-reinforced platforms for tracked sheet piling rig and mobile crane outriggers
concentrated loads such as scaffolding standards); a spread of load at 458 through the
support is normally assumed for design purposes. At the other extreme, piled founda-
tions might be needed for heavy gantries where the ground is marshy with negligible
bearing capacity. Traditional soil mechanics methods can be used for the design of
these types of platforms.
The frequently used Building Research Establishment design method (BRE, 2004)
determines the thickness of platform needed to resist punching shear under the tracks
or bearing plates of the operating plant. The method is only applicable where the
working platform is significantly stronger than the subgrade and is not appropriate
for very soft cohesive subgrades (undrained cohesion <20 kPa). While the BRE
method can be used for the thickness design of a platform to support the outriggers of
mobile cranes, trafficking by the wheeled vehicles themselves should be assessed in
accordance with Section 5.2.
For a platform on a very soft cohesive subgrade (undrained cohesion <20 kPa) geogrid
reinforcement is considered essential, not least to facilitate construction. Geogrids can
71
Figure 5.7 Comparison of unbound granular material thicknesses for typical piling platform with
and without geogrid reinforcement
1300
Soilmec R625 rig (1500 mm gear)
Ø'FILL = 45º
1200
1100
1000
900
800
BR470 design, unreinforced
Thickness: mm
(punching shear)
700
600
500
200
100
0
0 10 20 30 40 50 60 70 80 90 100
Subgrade strength (undrained cohesion): kPa
also reduce the overall thickness of granular material needed on stronger subgrades
(Figure 5.7). The relevant design method (Tensar, 2010) assumes a spread of load
through the reinforced granular layer of 458, the platform being made sufficiently
thick to limit the pressure on the subgrade to less than its safe bearing capacity. As
with site roads, the geogrid placed at the base of the granular layer serves to restrain
outward displacement of aggregate subjected to a vertical stress, thereby improving
load-spreading ability and increasing subgrade bearing capacity. For platforms more
than about 400 mm thick, a second geogrid may be needed within the body of the
granular layer to extend the zone of confinement.
72
outlined for site roads should be undertaken, with emphasis on undrained shear
strength rather than CBR as the fundamental design property for cohesive subgrades.
The investigation would need to be expanded to include standard penetration testing
(to estimate the angle of shearing resistance) should the subgrade prove to be of a
granular nature.
It is generally recommended that the gradient of platforms used by tracked cranes and
piling rigs with raised masts be restricted to 10% (which includes access ramps), although
a level surface is preferable. At the base of a ramp it is good practice to increase the
platform thickness by 50%.
Road alignment, marking and signing will need to be approved by the local highway
authority; temporary lighting may be stipulated at potentially hazardous locations
(e.g. short radius bends and junctions).
The thickness design of the temporary road pavement and its foundation can be based on
well-established principles relating to traffic loading (number of standard axles) and
subgrade strength. However, the intended short life means that the current design
charts for highway pavements (HA, 2006) will not necessarily give an economic
design, since the minimum thickness limit for the bound layer(s) has been set relatively
high.
Reference could be made to earlier design guides for bituminous roads (Powell et al.,
1984) and concrete roads (Mayhew and Harding, 1987) where data are provided for
traffic loading as low as 100 000 sa.
For even lower traffic loadings, Figure 5.3 in Powell et al. (1984) could be extrapolated
back to 1000 sa using the following equation to estimate the required thickness of
bitumen-bound material, hDBM100 (in mm):
where satemp is the traffic loading (standard axles) expected over the lifetime of the
completed road. The bitumen-bound material should comprise a binder course and a
surface course; the surface course should have adequate skidding resistance.
The thickness h (in mm) of Type 1 subbase needed as the road foundation may be derived
from Equation 5.2 where sadesign is a measure of the traffic (standard axles) employed in
the building of the road.
73
For example, consider a temporary public road diversion 200 m long built on a subgrade
with CBR 6% to carry 5000 standard axles after construction:
Because temporary roads can be relatively thin they may suffer winter damage when a
prolonged period of freezing is experienced and the subgrade is highly frost-susceptible
(e.g. chalk). There is no way to mitigate this risk cheaply and the consequences of prema-
ture failure during a harsh winter need to be assessed.
5.5. Construction
5.5.1 Materials and maintenance
Granular material for building temporary roads/platforms may be derived from site if
suitable deposits of gravel or moderately strong rock occur naturally in the ground;
on brownfield sites it should be possible to process demolition arisings to create
acceptable granular material. If site-won granular material is unavailable, a design
that includes geogrid reinforcement would reduce the volume of import needed and
should prove more economic; in these circumstances, stabilisation of the subgrade also
becomes an attractive option.
For convenience, granular materials are classified in accordance with the Specification
for highway works (HA, 1998) and should be compacted accordingly. The standard
designs for roads (Equations 5.1 and 5.2) require the use of well-graded crushed rock
gravel or Type 1 subbase. For lower-quality material, for example, Type 2 subbase or
Class 6F2 capping, an increase in the design thickness is needed. This could be as
much as 100% based on material equivalency (BSI, 2001), although realistically a
50% increase is considered more appropriate providing the lower-quality materials
have negligible plasticity and are relatively free-draining. No such increase in the
design thickness is considered necessary for geogrid-reinforced granular roads, although
it is important that the maximum particle size of the aggregate remains compatible with
the selected geogrid mesh size. The use of granular material with a rounded particle shape
should be avoided (at least near the surface of the road) since it may lack stability under
wheel contact pressure.
Where secondary (recycled) aggregates are imported (these will normally be cheaper than
primary aggregates), they must have been recovered under an acceptable quality
protocol (e.g. WRAP, 2010).
74
For working platforms where the design criteria are different, the angle of shearing
resistance (;0 ) of the granular material must be known. This can be determined by a
large shear box test in the laboratory but, since the maximum particle size that can be
included in the test specimen is 20 mm, the result may underestimate the potential of the
material. It is the author’s experience that for 75 mm maximum size, well-graded, moder-
ately strong crushed rock or crushed concrete, ;0 ¼ 458 may be safely presumed, provided
that the material is non-plastic and relatively free-draining. At the other extreme, naturally
occurring sand/gravel deposits with slight plasticity complying with the requirements for
Type 2 subbase (commonly referred to as ‘hoggin’) will have ;0 358.
Site roads and working platforms constructed from unbound granular material must be
kept clear of mud and slurry otherwise the desirable free-draining properties will be
adversely affected. Any localised deep rutting or haunch damage should be reinstated
with fresh material at the earliest opportunity, or progressive failure will result.
Soil stabilisation is a specialist operation and should not be attempted by the inexperienced
(refer to Chapter 7). A strong material (equivalent to or stiffer than unbound granular
material) can be formed from most soil types by using an appropriate binder. The binder
may be Portland cement, quicklime, fly ash, ground granulated blast furnace slag or combi-
nations of these. It is important for a stabilised road/platform to be protected against
surface abrasion by site traffic, standing water and frost. To this end, a topping of granular
material should be provided except perhaps in cases where the stabilised soil is compressive
strength class C3/4 or higher. One disadvantage of stabilised soil construction is that,
should premature failure occur, it is not usually practicable to replace the damaged layer
with like material. The only option is to dig out the affected area and substitute compacted
granular material, which might result in a recurrent ‘soft-spot’.
On small sites, or for short-term and one-off applications, the placing of proprietary
mats could be considered for traffic or plant access over soft ground. Access mats are
made variously from heavy-duty flexible plastic composites, aluminium extrusions/
plates or timber and can be bought, or more usually hired, from specialist outlets.
Some types are lightweight and capable of being man-handled into position; they can
be used to good effect where damage to existing surfacing (e.g. turf ) must be avoided.
Access mats function by increasing the load contact area; the restraint offered by
adjoining linked mats serves to restrict deformation. However, they can prove to be
prohibitively expensive for long-term or widespread use.
Inevitably, traffic leaving a construction site will have picked up mud from the site roads
or piling platforms. The provision of a wheel-wash facility (or at least a rumble strip)
close to the exit will help minimise the spread of mud to the adjoining public roads.
On larger sites where bulk excavation is taking place, a purpose-made wheel/vehicle
wash unit is essential (models are available that recycle the wash water and extract/
dewater the suspended solids).
It is important that plant operators and drivers using the site roads and working
platforms are aware of the layout and boundaries. Edges should be defined by fencing
75
or timber sleepers; in circumstances where there is an abrupt drop over the edge, a
physical restraint should be provided, which might comprise an earth bank or pre-
fabricated barrier units.
5.5.2 Drainage
During inclement weather rainwater could accumulate in unbound granular con-
struction layers, resulting in softening of the underlying stabilised soil (where used) or
cohesive subgrade (particularly where the soil infiltration rate is <106 m/s). Potentially,
this could lead to early failure of the site road or working platform.
For site roads, both the surface of the granular layer and the underlying stabilised soil or
subgrade should be shaped to a crossfall (5% if feasible, but 2.5% minimum) to shed
water, the run-off being channelled away from the edges of the road in drains or
temporary shallow open ditches (‘grips’).
Working platforms may be much wider than site roads and providing a crossfall is not
necessarily practicable. It is also desirable that the surface of the platform be relatively
level to avoid construction plant tilting. Shaping the underlying stabilised soil or
subgrade to fall in two, or even four, directions towards the edges of the platform
may be a solution, in conjunction with perimeter drains or grips. Alternatively, for
very wide platforms constructed entirely of unbound granular material, shallow land
drains could be installed in a herringbone pattern below the platform to channel any
water reaching the subgrade to a common outfall; these drains need not be piped, but
may simply be slots in the subgrade filled with a free-draining granular drainage medium.
When disposing of water drained from site roads and working platforms it is essential to
avoid pollution of aquifers and watercourses. Recourse to silt traps and oil interceptors
may be needed and Environment Agency guidance should be sought.
Drainage of surface water from a paved temporary public road would normally be to a
filter drain on the low side of the carriageway with an outfall into the existing highway
drainage system. In built-up areas, kerbs could be used to direct surface water back to the
main highway.
REFERENCES
Black W and Lister NW (1978) The strength of clay fill subgrades; its prediction and
relation to road performance. In: Clay Fills, Proceedings of the Conference held at the
Institution of Civil Engineers 14–15 November 1978. ICE, London.
BRE (Building Research Establishment) (1991) BRE Digest 363: Soakaway design. IHS
BRE Press, Watford. (Revised 2007.)
BRE (2004) BR 470: Working platforms for tracked plant. BRE Bookshop, Watford.
BritPave (2007) Concrete hardstanding design handbook: Guidelines for the design of con-
crete hardstandings. 2nd edition. British In-situ Concrete Paving Association, Camberley.
BSI (2001) BS 7533-1: Pavements constructed with clay, natural stone or concrete pavers
(Part 1: Guide for the structural design of heavy duty pavements constructed of clay
pavers or concrete paving blocks). BSI, London.
76
Chaddock BCJ and Atkinson VM (1997) Stabilised sub-bases in road foundations: struc-
tural assessment and benefits. TRL Report 248. Transport Research Laboratory,
Crowthorne.
Giroud JP and Noiray L (1981) Geotextile-reinforced unpaved road design. Proceedings of
the American Society of Civil Engineers, Journal of the Geotechnical Engineering Division,
107, No. GT9.
HA (Highways Agency) (1998) Specification for highway works. Manual of contract docu-
ments for highway works Vol. 1. HMSO, London. (Amended 2009.)
HA (2006) Pavement design. Design manual for roads and bridges Vol. 7, Section 2, Part 3.
HD 26/06. HMSO, London.
HA (2007) Treatment of fill and capping materials using either lime or cement or both.
Design manual for roads and bridges Vol. 4, Section 1, Part 6. HA 74/07. HMSO,
London.
HA (2009) Design guidance for road pavement foundations (Draft HD25). Interim Advice
Note 73/06 Revision 1 (2009). HMSO, London.
Mayhew HC and Harding HM (1987) Thickness design of concrete roads. Department of
Transport, TRRL Research Report 87. Transport and Road Research Laboratory,
Crowthorne.
Powell WD, Potter JF, Mayhew HC and Nunn ME (1984) The structural design of bitumi-
nous roads. Department of Transport, TRRL Report LR1132. Transport and Road
Research Laboratory, Crowthorne.
Tensar (2010) TensarPave. Available from Tensar International Limited, Blackburn.
Terram (2010) Geotextiles. Terram Limited, Pontypool. Available at http://www.terram.
com (accessed September 2010).
WRAP (Waste and Resources Action Programme) (2010) Guidance notes for the
producers’ compliance checklist. Banbury. Available at http://www.wrap.org.uk
(accessed September 2010).
FURTHER READING
BritPave (2007) HBM and Stabilisation 2: The design and specification of residential and
commercial road pavements. British In-situ Concrete Paving Association, Camberley.
Jewell RA (1996) Soil reinforcement with geotextiles. CIRIA SP123. London (Construction
Industry Research and Information Association).
WRAP (Waste and Resources Action Programme) (2006) Guidance on the use of HBM in
working platforms. Banbury.
77
Chapter 6
Control of groundwater
Neil Smith Director, Applied Geotechnical Engineering
6.1. Introduction
Where excavations are required to extend below the groundwater level, it is essential to
plan beforehand the means of dealing with the water. There are generally four possible
methods as follows.
1 Exclude the water by surrounding the area to be excavated with a cut-off wall
(there are various ways of doing this; see Chapters 8, 9, 11, 13 and 14).
2 Extract sufficient groundwater to depress the level below the excavation surface.
(Methods (1) and (2) may be used in combination, where a partial cut-off reduces
the volume of water to be extracted.)
3 Reduce the pore water pressure in the soil sufficiently to stabilise the ground
around and below the excavation.
4 Excavate below water (refer to Chapter 15 on Caissons and Shafts).
g sump pumping
g wellpointing
79
g deep wells
g ejectors.
The rarely-used technique of electro-osmosis is based on different principles and will not
be considered here.
6.2. Techniques
6.2.1 Selection of appropriate technique
Figure 6.1 illustrates the ranges of application of the various different pumped systems
for groundwater control. The blurred boundaries are deliberately chosen to emphasise
their approximate nature. The particle sizes shown below the permeability k axis are a
guide only. Note that where there are strata of widely differing permeabilities, it may
be necessary to use more than one technique.
Figure 6.1 Relations between pumped well systems, permeability and drawdown
Adapted from Roberts and Preene (1994a)
0
Sump pump
5
and may not be necessary
Dewatering not feasible
flows:
cut-off or
10
wet
excavation
Deep wells may be
necessary
15
20 –8
10 10–7 10–6 10–5 10–4 10–3 10–2 10–1
Permeability: m/s
Silt and interlaminated Sand and gravel mixtures Clean gravel
silt/sand/clay
80
Well screen
*A filter is essential in stratified soil. May not be needed in well-graded sandy gravel
reduce the stability of the soil close to the excavation surface. The sump can be fed by a
pattern of trenches or French drains leading from the perimeter of the excavation. Sump
intakes can be filtered by granular, geotextile wrapping or combined filters.
6.2.3 Deep wells, wellpoints and ejectors: extraction from below the
excavation
Wellpoints work by suction pumping and so the head which they can lift is limited by
atmospheric pressure and suction efficiency. Figure 6.2 illustrates the main parts of a
wellpoint.
Note that the header pipes, which connect a number of wellpoints to a pump, can
obstruct surface activities as can be seen from Figure 6.3 (in which the tops of wellpoints
and pumps can also be seen).
In practice, the achievable drawdown is limited to about 5 m but the use of a second stage
of wellpoints can increase the drawdown to around 9 m. In rare instances, multiple levels
of wellpoint can be used (Figure 6.4). Wellpoints can be used to enclose an excavation or
simply to dewater a length of trench, for example, to install services. To dewater long
trenches, sections of wellpoints which extend beyond both ends of the open trench are
activated progressively as the work proceeds.
81
Deep wells are not limited in the drawdown they can achieve because the water is sucked
into the bottom of the well and then pushed up the discharge pipe. Pumps operate from
within prepared wells which have had lining tubes and surrounding filters previously
installed.
Figure 6.4 Three stage well-point system. Note rise of water level below excavation; a similar rise
is present with deep wells
Note:
Sand filters are needed
in stratified deposits
82
Ejectors work on the principle of a nozzle and venturi pump. Water fed to the base of a
hole is forced to flow back up the hole through a nozzle and venturi to pull groundwater
in from below. The practical depth limit on their use is about 50 m. They tend to be used
for the control of pore pressure in fine-grained soil. (There is some similarity with the
airlift system in that fluid is introduced at the bottom of a discharge pipe (riser) to
cause an upward flow.)
6.2.3.1 Development
It is important to ‘develop’ deep wells. This is, effectively, a cleaning process to remove
drilling debris from the hole and also, in suitable ground, wash out fine particles from
joints or fissures, thus increasing the effective size of the well. Removal of debris
before installing the pump also reduces pump wear. In carbonate rocks such as limestone
or chalk, concentrated hydrochloric acid can be added to dissolve slurry and clean
fissures. This process releases carbon dioxide in large quantities and should only be
carried out by appropriately experienced personnel.
6.2.4 Filters
Whenever groundwater is pumped from a soil, seepage pressures are set up at the point of
entry of the water into the well. The seepage pressures can dislodge particles from the soil
near the borehole wall so that the flow carries them into the system. This can have
deleterious effects both on the system and the ground. Excessive sediment transport
will cause wear or clogging to the pump and pipework. Removal of fines from the
ground is a common source of settlement of ground in the vicinity of dewatering
schemes.
In order to prevent the removal of soil particles, filters are installed. Traditionally, these
were made from soil with a grading specified to meet criteria relating the particle size
distribution of the filter to that of the aquifer. In some cases, two layers of filter were
needed. Filter rules are described in detail in Preene et al. (2000).
6.2.5 Hazards
Hazards resulting from the application of the techniques described above include the
following.
g Settlement: when the groundwater level is reduced, the effective stress between soil
particles increases. At shallow depth and in compressible soil, this increase can be
significant, causing settlement of the ground in the drawdown zone surrounding
the system and consequential damage to structures.
g Removal of fines: if the filtration system does not function properly, the flow of
water can lead to the extraction of fines from the surrounding ground and the
creation of voids. In some circumstances, this can lead to catastrophic collapse of
the surface.
83
g Fouling and clogging: the effectiveness of wells, wellpoints and ejectors can
deteriorate over time because of fouling of the screen by bacterial growth and/or
encrustation by chemicals precipitated from the water.
g Corrosion: some groundwater (for example, saline water at sites near the sea) has
a chemical composition which may cause corrosion of pipework and pumps.
6.2.6 Recharge
Where settlement is considered to present a hazard, the effects of dewatering can be
ameliorated by pumping water back into the ground within the area of drawdown
outside the excavation. This will change the groundwater profile during system operation
and reduce the drawdown below the surrounding ground.
It may also be necessary to monitor water quality (both the chemistry and the concentration
of suspended solids), particularly for discharge or recharge purposes. Where the ground
conditions dictate, accurate settlement monitoring should be undertaken; this should begin
well before the dewatering is turned on. Note that in some areas close to tidal water, the
ground surface can move with the tide. Any natural variation in level could make it very diffi-
cult to interpret ground level data during operation of the groundwater lowering system.
6.2.8 Consents
It is normally necessary to obtain a license from the Environment Agency (EA) to extract
more than 20 m3/day of water from the ground. The abstracted water is classified in a
legal sense as trade effluent. As such, a discharge permit is required before the water
can be fed to surface water, the sea, water courses or recharged back into the ground.
Note that, even if nothing else is done to the abstracted water, exposure to the
atmosphere may cause chemical changes to occur. Advice about obtaining permits
from the EA can be found on their website.
84
Direct measurements of permeability can be made in a number of ways, but there are
significant drawbacks to most methods. Testing can be carried out in situ during drilling
of site investigation boreholes. However, the ground just below the bottom of the bore-
hole may have been disturbed and there may be problems in allowing enough time to
establish an accurate rest level. Additionally, sediment can be deposited at the bottom
of the borehole as a thin low-permeability layer, especially during inflow tests. If the
head is reduced too far for a rising head test, the ground around the base of the hole
may be disturbed by ‘blowing’. Where piezometers are installed, tests can be carried
out which are similar to the in-borehole tests. The rest level should have been determined
by repeated measurements and the risk of silting or blowing should be minimal.
However, tests in boreholes and piezometers suffer from the small volume of ground
tested. In most circumstances, it is necessary to carry out a number of tests in different
places at different depths.
Permeability tests can be carried out on samples in the laboratory, but sample
disturbance may be a problem. Tests which pass a vertical flow through a tube sample
may measure a misleadingly low permeability if the sample is anisotropic, as many
soils are. To obtain sufficient flow in low permeability soils, it is generally necessary
to apply a much higher hydraulic gradient to the specimen than would occur in practice;
this can lead to misleading results. The sample size may not be large enough to include
representative fabric (soil structure). Particle size distributions can be used to provide an
estimate of permeability in certain soils – uniform sands are most suited to this (see
Equation 6.2). In a layered soil, the different soil layers should be separated before
taking specimens for particle size distribution testing.
The best way to measure permeability at the scale required for a dewatering scheme
design is by a pumping test. This is effectively a small-scale trial of the dewatering
system. The test pumping well should be surrounded by a number of monitoring
wells so that the drawdown curve can be determined, ideally in two orthogonal
directions. These tests are relatively expensive and take between 1 and 2 weeks to
conduct. Their major advantage is that the mass permeability of the soil is measured
under conditions which are similar to those which will confront the actual dewatering
system.
85
g Detail the system (e.g. determine depths and spacings for wellpoints or deep
wells).
6.4.2 Permeability
One-dimensional flow in saturated soil is governed by Darcy’s Law which states that the
velocity of laminar fluid flow through a soil is proportional to the gradient of fluid
pressure (head) between the two ends of the element. Turbulent flow in soil is only
likely in coarse gravels with high flow. Darcy’s Law is defined:
ðho hi Þ
q ¼ k ð6:1Þ
l
where q is the velocity of flow (m/s), ho is the head (m) at the point where the fluid
emerges from the element, hi is the head (m) at the point where the fluid enters the
element, l is the length (m) of the element and k is the coefficient of proportionality
(m/s), referred to as the permeability.
The flow is positive in the direction of negative gradient. Between clay and coarse gravel
there is a huge range of permeability, c. 1010 to 101 m/s. (An equivalent range of
distance is from 1 m to 2.5 times the distance from the Earth to the Moon.) In highly
favourable conditions, the most accurate determination of permeability is within a
factor of c. 3.
The permeability is related to pore size which, in uniform soil, is related to the particle
size. Hazen’s formula can be reliable for some sands and gives a general indication of
the order of magnitude of k:
k ¼ 0:01D210 ð6:2Þ
where D10 is the sieve size (mm) which allows 10% of the soil by weight to pass through
and k is measured in m/s.
Permeability is not a constant, because the pore size decreases if the soil is compressed.
This can be important in soft soil (e.g. organic clay) where significant changes in
permeability can occur as a result of consolidation, which may be a result of the
groundwater lowering.
Soil is not homogeneous. Even in what might appear to be a uniform soil, the
permeability is almost always greater in the horizontal direction than the vertical.
Many soils are layered. Thin clay layers in a sand stratum can reduce the vertical
mass permeability dramatically. Glacial soils are very heterogeneous. Predominantly
clayey glacial soil can contain pockets, lenses or ribbons of coarse granular material
which may provide a conduit to a source of water.
Equation (6.1) shows that the rate of flow is also proportional to the difference in
head along the element. In a dewatering scheme this is equivalent to the drawdown,
86
that is, the difference between the equilibrium water level and the lower level required for
construction.
Figure 6.5 shows the two principal ‘ideal’ aquifer conditions (confined and unconfined)
with partially penetrating wells or slots. Fully penetrating wells, as the name suggests,
reach down to the top of the impermeable base layer.
Equations have been derived for volumetric flow rates to wells and slots for confined and
unconfined aquifers. As might be expected, the principal variables are the permeability
and the required drawdown. In determining the required drawdown, allowance must
be made for the increase in head with distance from the well. In the case of a group of
wells around an excavation, the head at the centre of the group is higher than the
head at the wells themselves (see Figure 6.3).
The expression for radial flow to a fully penetrating well in a confined aquifer is given by
the Theim equation:
2kDðH hw Þ
Q¼ ð6:3Þ
lnðRo =re Þ
where Q is in m3/s if k is in m/s and D, H and hw are in m.
There are similar relationships for a well in an unconfined aquifer and for planar flow to
slots in confined and unconfined aquifers. The equations, together with factors for flows
from partially penetrating wells and slots, are given in Preene et al. (2000).
The parameters are defined in Figure 6.4. Note that re is the equivalent well radius.
Where a group of wells is distributed around a rectangle a b, re may be taken as
(a þ b)/. If only a single well is being considered, re ¼ rw (i.e. the well radius).
87
Well Slot
Radius re
Width b
Ro Lo
Piezometric level Q
Confined
aquifer
P H
Aquifer hw D
permeability k
Impermeable
Ro Lo
Q
Water table
Unconfined
Aquifer P aquifer
H
permeability k
hw
Impermeable
Well Slot
However sophisticated the analysis, the design of the dewatering system must take
account of the uncertainties in both the actual values of permeability and in the structure
of the ground affected by the process. It is essential to monitor the effectiveness of the
system, ideally starting the process with enough time allowed in the programme to
adjust the design if the drawdown is insufficient. When starting a dewatering scheme,
it is wise to be aware of where and how quickly additional resources can be obtained
(if they are not actually held in reserve). As far as pump capacity is concerned, an
immediate need to increase capacity can often be met by use of a standby pump(s)
which should be on hand in case of breakdowns.
REFERENCES
Bond A (ed.) (1994) Validation and use of geotechnical software. Association of Geotech-
nical and Geoenvironmental Specialists, Kent.
Preene M, Roberts TOL, Powrie W and Dyer MR (2000) Groundwater control – design
and practice. CIRIA Report C515. CIRIA, London.
88
Roberts TOL and Preene M (1994) Range of application of groundwater control systems.
In Groundwater Problems in Urban Areas, Wilkinson WB (ed.) Thomas Telford,
London, 415–423.
FURTHER READING
Roberts TOL and Preene M (1994) The design of groundwater control systems using the
observational method. Ge´otechnique, 44(4): 727–734.
89
Chapter 7
Lime and cement stabilisation
Mike Brice Associate, Applied Geotechnical Engineering
Neil Smith Director, Applied Geotechnical Engineering
Weak soils can be improved in strength and workability by the addition of lime and/or
cement. The use of the process as a temporary measure is generally to provide a sound
working surface for construction (though the treated soil has been altered permanently).
Lime is usually applied to the soil in the form of quicklime. This has an immediate effect
in taking moisture from wet soils and is especially effective in soft, wet, plastic clay. The
lime also reduces the plasticity of the clay and so makes it both stronger and more workable.
If added in sufficient quantity, it produces a cementing effect with considerable additional
increase in strength over time. Cement also reduces the moisture content and plasticity,
although less effectively than lime. It does, however, have a greater cementitious effect. It
is more suitable for treating lower plasticity or non-plastic soils. The results of adding
lime or cement are very soil-specific, so preliminary testing is essential to ensure that the
desired effect(s) can be achieved and to find the optimum concentration and type of additive.
The effects vary with plasticity and other properties, so that a substantial test programme is
needed in variable soils such as glacial till. Blends of lime and cement can be used and other
cementitious agents, such as pulverised fuel ash (PFA), may be incorporated.
7.1. Introduction
This chapter deals with the use of lime and cement additives to improve the strength of
near-surface soils for temporary works construction. In a temporary works context, the
addition of lime or cement is most likely to be required in the context of haul road or
working platform improvement (see Chapter 5), and in the rendering of wet material
suitable for use as a general fill. In some circumstances, such as the enhancement of
thick soft deposits, deep methods such as lime mixing or lime columns may be beneficial.
These require highly specialised plant and are of a different character to superficial lime
and cement mixing. They are not considered further here.
The two additives differ in their actions: lime is of more benefit in the rapid drying of soils
and as an additive to improve ‘heavy’ plastic clay soils. In low plasticity clay soils
(PI < 10%) and in predominantly granular soils lime may be used as a drying agent,
but otherwise its effectiveness is limited and cement is of greater benefit.
91
commonly used for soil treatment, being significantly more effective in drying out a soil
by the heat of hydration than slaked lime, and is less dusty to handle. Only quicklime is
discussed below.
In the first instance, the addition of lime dries the soil both by absorbing water directly
(up to 32% of the dry lime weight) and by driving off water by the exothermic nature of
the hydration reaction. As a rule of thumb, the reduction in soil moisture content
amounts to approximately 1% per percentage point of added quicklime. This effect is
seen in all soils. In cohesive soils there is also a chemical reaction with clay minerals in
the soil; the lime will alter the nature of the clay minerals, greatly increasing the
plastic limit and thereby reducing the plasticity. Both of these reactions are immediate
and are not affected by ambient temperature. They can combine to transform a soft
plastic soil into a dry friable material, making it significantly stronger and easier to
work. Together they are termed soil modification. The effectiveness of the second
reaction is dependent upon the nature of the clay minerals within the soil (montmorillo-
nite being more affected than illite and chlorite for example), and the clay content and
mineralogy will also affect the quantity of lime required.
The proportion of lime required is small; between 1 and 3% (by dry weight) of lime to soil
is needed in order to effect soil modification. It has been noted that 0.5% lime addition
should be adequate for the majority of British plastic clays that have been classified as
unacceptable on the basis of excess moisture content (Perry et al., 1996a).
If a higher proportion of lime is added then the pH of the soil is elevated to pH > 12 and a
cementitious action of the lime on the clay minerals in the soil occurs, leading to an
increase in strength with time (typically weeks, although measurable strength increases
will still occur after several years). Above 5–8% lime (depending on the original
clay content of the soil), the effect is lost and long-term strength begins to decrease.
Such high proportions of lime may not be economical in any event. This cementitious
reaction is affected by soil temperature, and the reaction will be suspended for as long
as the temperature falls below 4–58C. It also requires the presence of water, therefore
wetting of treated fills may be required if the cementitious effect is to be relied upon.
The cementitious action is termed ‘soil stabilisation’.
A resting (‘mellowing’ or ‘maturing’) period between mixing and compacting the fill is
recommended where the stabilisation component of the lime treatment is important.
The resting period, which is typically 24–72 hours for lime-treated soils, allows the
lime to hydrate and migrate through the soil thereby fully effecting the changes in
plasticity index and general handling improvements. Following this resting period, the
soil is remixed and compacted. If the resting period is too protracted then carbonation
of the quicklime becomes appreciable, and less quicklime is then available for the
stabilisation reaction.
For high-strength end products (such as road-capping layers), compaction to less than
5% air voids is usually required and control of the moisture content is needed to
ensure this is achievable.
92
Figure 7.1 Typical relationships between strength of treated high plasticity clay, percentage lime
and time after mixing
Data taken from Bell (1988)
3000
t=0 +
t = 7 days
t = 28 days
t = 56 days +
+
+ t = 182 days
Unconfined compressive strength: kPa
+
2000
+
+
+
1000
+
+
0
0 2 4 6 8
Percentage lime
Figure 7.1 depics data for the strength of a lime-treated high plasticity clay from Bell
(1988). The flattening of the t ¼ 0 line at lime concentrations above 2% clearly shows
the limitation of the modifying effect. The subsequent strength gains largely reflect the
stabilisation process.
7.2.2 Cement
Cement is used predominantly for soil stabilisation, similar to the cementitious action
of excess lime. There is a modest immediate drying effect (soil modification) on first
mixing the cement with the soil. This amounts to a 0.3–0.5% reduction in overall
moisture content per percentage point of cement added, but this is usually viewed as a
secondary benefit. The cementitious action is far faster than the equivalent action in
the lime stabilised soil and is not dependent on the presence of clay minerals. The
reaction is again suspended below a soil temperature of 58C, but the period over
which the treated soil is susceptible to low temperature is so much shorter than with
lime treatment that cement additives might confidently be used closer to the winter
season. The faster ‘set’ of a cement additive means that there is far less leeway in the
duration of the resting period. Testing is required to establish the optimum duration
of the resting period.
93
Figure 7.2 Relations between lime/cement content, compressive strength and original clay
plasticity
Data taken from Christensen (1969)
3000
+ Lime
Cement
Clay – originally CL, PI 15%
Clay – originally CH, PI 30%
Clay – originally CH, PI 43%
Unconfined compressive strength: kPa
2000
+
+
+
1000 +
+
+
+
+
0
0 1 2 3 4 5
Percentage lime or cement
Figure 7.2 presents data from Christensen (1969) for lime- and cement-modified soils of
different plasticity, tested for strength at 7 days. It can be seen that there is little
difference between the effects of the lime and cement for the high plasticity soil, but a
very marked difference in the results for low plasticity clay.
7.2.3 Blends
In some cases a combined treatment with a lime/cement blend may be appropriate.
In any case it is imperative, in view of the significant costs of lime/cement treatment,
that pre-contract laboratory testing is carried out. The best additive to use for a given
soil and the proportions in which it is to be added in order to achieve the required
performance must be established.
Where it is only required to produce a modest increase in strength, for example, simply to
make a site subgrade traffickable, the proportions of additive needed may be very small.
There can be problems with the even distribution of such small proportions of additive,
so a dry ‘inert’ bulking medium such as pulverised fuel ash (PFA) can first be added.
94
(PFA is not inert and will also confer an additional cementitious element to the long-term
strength.)
These changes arise as the clay minerals flocculate, thereby effectively changing the grain
size of the clay minerals.
The stabilisation process is marked by an increase in strength and stiffness with time.
The changes brought about by the initial soil modification usually render the treated soil
frost susceptible. Lime-stabilised soil remains frost susceptible for around 3 months after
placement, and frost action during that time would probably disrupt the cementitious
bonds developed during the stabilisation process.
Cement stabilised soils are likely to be less frost susceptible than lime-stabilised, due
both to the different grain size distribution of the original soil and to the more rapid
development of strong cementitious bonds.
In stabilised soils, the lime can undergo an expansive chemical reaction with sulphate
chemicals in the soil, potentially leading to heave of the treated soil. Heave magnitudes
of 60% of treated layer thickness have been recorded. Such effects are the single main
cause of failure of lime- or cement-treated earthworks. The reaction involves the
formation of Ettringite and Thaumasite and is therefore directly equivalent to sulphate
attack on concrete. The rapid oxidation of soil sulphides to sulphates during the lime
mixing process makes it essential to determine both sulphide and sulphate contents of
the soil to be treated during the planning stage.
Organic material in the soil interferes with the stabilisation reaction, tending to reduce
the pH locally. An upper limit of 2% organic content for lime-treatable fills is often
quoted (HA, 1991). Above this limit, lime addition is taken to be ineffective although
it is recognised that the form of the organic matter is important; Oxford Clay with up
to 10% disseminated bituminous content has been successfully treated.
95
7.4. Testing
The proportion of additive (lime or cement) is expressed as a percentage of additive per
unit dry weight of soil. Testing soil to establish the suitability of lime/cement treatment is
very much dependent upon the end use of the treated product.
It is not appropriate to apply the results of testing of the original soil to the treated
product. The nature of the soil is altered by treatment and new earthworks control
criteria need to be derived. The Moisture Condition Value (MCV) test is recognised as
being appropriate for the site control of treated soils.
Before the commencement of treatment on site, laboratory testing should be carried out
in order to determine (for the treated soil) the following properties.
1 The sulphate, sulphide and organic contents of the soil to be treated (Longworth,
2004; HA, 2007). It is important to note that sulphides and sulphates are often
leached out of the top metre or so of UK soils. Hence, if soils from deeper than
this are to be treated, then it is important that they are adequately sampled and
tested. Groundwater should also be tested for sulphates. A limit of 1% total
sulphate is usually applied in UK highway works, but it is recognised that lower
sulphate concentrations can still result in unacceptable swelling of treated
materials and so site-specific testing (relating the sulphate content to heave in an
inundation test) should be carried out where soil stabilisation is required (Perry
et al., 1996b).
2 The relationships between moisture content, MCV and dry density as well as (if
applicable to end use) undrained strength and CBR for different proportions of
added lime/cement. The Initial Consumption of Lime (ICL) is the quantity of lime
required to bring the soil to a pH of 12.4, and marks the boundary between a soil
modification effect and soil stabilisation. Where soil stabilisation is required (as
opposed to simple modification), the ICL marks the lowest lime proportion that
needs to be tested. Increments of 0.5% lime are usually tested until the optimum
lime proportion is identified. It is important that if a resting period is to be used
on site then a similar resting period is allowed in the laboratory (Perry et al.,
1996b).
3 Heave on inundation, carried out over 14–28 days on a compacted sample in a
CBR mould, is important as an indicator of adverse sulphate reactions and should
be carried out where soil stabilisation is required.
4 Frost susceptibility tests may be appropriate, but these are expensive tests requiring
large samples and it may be more appropriate to ensure that treated soils are
protected from frost or are not stabilised during the season when frost is a risk.
Quality control testing of the end-product is necessary. The testing undertaken would
depend on the original intention of the stabilisation process. For example, if the intention
96
is to produce a surface with a given traffickability, then plate-bearing tests or in situ CBR
tests would be appropriate. The rate for such testing is commonly of the order of 1 test
per 1000 m2 of treated surface. If on the other hand the purpose was to produce a useable
bulk fill, then compaction/density testing (to ensure the treated product could be readily
compacted to achieve a given standard, for example, 95% Proctor maximum dry density)
would be appropriate. A suitable testing rate in this case might be 1 test per 250 m3 of
treated fill.
7.5. Plant
Lime/cement additives are usually mixed into the soil in situ, either for subsequent
compaction in place or for subsequent excavation and placement/compaction elsewhere.
There may be practical problems on site with dispersion through the soil of small propor-
tions of lime/cement. In this situation, the additive can be bulked up by the addition of
PFA to improve dispersion. This also imparts a cementitious component, leading to a
further increase in strength with time.
In addition to storage facilities for the additive, the plant requirements comprise
spreading and mixing equipment as well as the preliminary grading and subsequent
compaction plant. On economic grounds, agricultural machinery (drag-bars, ploughs,
rotovators, etc.) may be suitable for the spreading and mixing on minor works. For
best results, however, and to ensure a uniformity of product, specialist spreading/soil
pulverisation plant is required (Figure 7.3). More than one pass of the soil pulveriser
may be required to produce uniform results. Wetting of the fill may be required in
Figure 7.3 Addition of lime using an integrated spreader and mixer unit and a simple mixer unit
(working on pre-spread lime), both tractor drawn
Courtesy of Con-Form Contracting
97
A delay between mixing of the lime with the soil and the subsequent compaction allows
the quicklime to slake and, depending on the delay, to migrate through and modify the
clay soil, facilitating compaction. The duration of this delay, often between 1 hour and
72 hours depending on the soil and on site constraints, should be modelled during the
original testing phase. Following compaction, the cementitious reactions are likely to
lead to a gain in strength with time. The time over which the strength increases signifi-
cantly is termed the ‘curing’ period.
REFERENCES
Bell FG (1988) Stabilisation and treatment of clay soils with lime. Part 1 – Basic principles.
Ground Engineering 21(1): 10–15.
Christensen AP (1969) Cement modification of clay soils. RD002.01S. Portland Cement
Association, Illinois. Available from the Association website.
HA (Highways Agency) (1991) HA44/91: Earthworks. Design and preparation of contract
documents. Advice note published by the Highways Agency as part of the Design
Manual for Roads and Bridges, Volume 4, Geotechnics and drainage, Section 1 Earth-
works, Part 6. HMSO, London. (Amended 1995.)
HA (2007) HA74/07: Treatment of fill and capping materials using either lime or cement
or both. Advice note published by the Highways Agency as part of the Design Manual
for Roads and Bridges, Volume 4, Geotechnics and drainage, Section 1 Earthworks,
Part 6. HMSO, London.
Longworth I (2004) Assessment of sulphate-bearing ground for soil stabilisation for built
development. Ground Engineering 37(5): 30–34.
Perry J, MacNeil D and Wilson P (1996a) The uses of lime in ground engineering: a review
of work undertaken at the Transport Research Laboratory. In Lime Stabilisation.
Thomas Telford Publishing, London.
Perry J, Snowdon R and Wilson P (1996b) Site investigation for lime stabilisation of
highway works. In Advances in Site Investigation Practice. Thomas Telford Publishing,
London.
FURTHER READING
BRE (British Research Establishment) (2005) BRE Special Digest 1: Concrete in aggressive
ground. BRE, Watford.
BSI (2004) BS EN 13286 Parts 1–53 Testing stabilised materials. BSI, London. (This
replaces BS 1924: 1990.)
98
BSI (2004) BS EN 13286-47: Unbound and hydraulically bound mixtures. Test method for
the determination of California bearing ratio, immediate bearing index and linear swelling.
BSI, London.
BSI (2004) BS EN 14227-1: Unbound and hydraulically bound mixtures. BSI, London.
Mitchell J and Jardine FM (2002) A guide to ground treatment. CIRIA, London.
99
Chapter 8
Jet grouting
John Hislam Director, Applied Geotechnical Engineering
Neil Smith Director, Applied Geotechnical Engineering
Jet grouting is a process of increasing the strength and/or decreasing the permeability of
soil by wholly or partially eroding a volume of ground and replacing it with grout or a
mixture of grout and the disturbed soil. The erosion and grout placement are both
effected by high-pressure jetting from drillholes via a tube known as a ‘monitor’.
Zones of grout can either be cylinders created by full rotation of the monitor, segments
of cylinders created by limited rotation or panels created without rotation. Closely
spaced drillholes are used to make overlapping zones of treated ground to form the
desired shape. Any shape is possible provided that the drillholes can be suitably
oriented. A number of factors affect the strength of the treated ground, but unconfined
compressive strengths of up to 8 MPa or 2 MPa are quite possible in coarse or fine soil,
respectively, although a few weeks will be required for the maximum strength to be
reached. The permeability of the treated ground can be reduced to 10 7 m/s or less.
The technique can be used to form retaining structures around excavations or (with
subhorizontal drillholes) arches above tunnels under construction or groundwater
cut-offs. By treating only cylinders of limited height in a defined depth zone, bottom
cut-offs or props between retaining walls can be created. On a small scale, tunnel to
shaft break-ins or break-outs can be facilitated or leaky basements sealed.
8.1. Introduction
Jet grouting is a method of ground improvement achieved by replacing the in situ soil
with cement grout or by mixing soil and cement grout in situ. It can be used where
permeation grouting (filling the voids between soil particles) would be impossible or
ineffective. It is carried out from drillholes using radial fluid jets and forms generally
cylindrical zones of cemented ground co-axial with the drillhole. High-pressure fluid is
used to erode the soil and flush part or all of it to the surface. If only some of the soil
is removed, that which remains must be thoroughly mixed with the grout.
Figure 8.1 shows the basic systems of single, double and triple fluids. The pipework
which delivers the fluids into the ground is termed (confusingly) the ‘monitor’.
Figure 8.2 shows an example. The monitor is incorporated into the drillstring. During
drilling, the drill flush is emitted through the bit. Usually the jet grout mixture is used
as the flushing medium during drilling as this helps to stabilise the sides of the hole.
As for the drillhole cuttings, the excess spoil generated during the jetting process is
flushed to the surface via the annular gap between the monitor and the hole.
101
Grout Air
Air
Grout
Air Water
Drill bit Air
Grout
The technique is not restricted to the creation of vertical columns. It can be used from
drillholes of any inclination. The treatment can be used over a limited section of a drill-
hole to form a disc of improved ground. The treated ground has a relatively high
compressive strength and a low permeability. It can, therefore, be used to form arches,
gravity structures or tubes to create, for example
1 a water cut-off, either around the sides or below the floor of an excavation
2 a retaining structure to support the sides of an excavation
3 a horizontal low-level prop before excavation
4 a supporting arch to maintain ground stability during tunnel construction
5 on a small scale, it can be used to stop leakage through basement retaining walls
6 a region of low permeability ground to permit tunnel break-out from, or break-in
to, shafts.
Coupland (2010) and Wit et al. (2007a) provide case histories of two interesting projects;
many others can be found in the literature. Most applications entail the construction of
cylinders of ground, since the jets are rotated throughout the process of erosion and
filling. If rotation is limited or prevented, different geometric forms are possible. The
work is covered by an ‘execution of special geotechnical work’ attachment to EC7
Part 1 (BSI, 2001).
102
For the combined fluid jets, the compressed air is forced through an annular nozzle with
the nozzle delivering the water or grout at its centre. The flow of air therefore surrounds
the more viscous fluid as it leaves the jet. High jetting velocities are used (up to 300 m/s),
requiring fluid pressures of the order of 20–60 MPa and compressed air pressures of
c. 1.5 MPa. Triplex pumps are used to attain the high pressures required. Extreme
care must be observed when the monitor is above ground level as the jets, especially
because they carry particulate matter, can cause damage and injury.
The size of the cylinder which can be treated with this method depends to a significant
extent on the nature of the soil, as well as the method adopted. While jet grouting can
103
treat virtually all soils (and some very weakly cemented rocks), some materials are more
resistant to erosion than others so that the size can vary between strata. Stiff clay is much
more resistant than silt or fine sand. Very coarse soil particles require more force to move
them than fine grains. The drillhole would typically be about 150 mm diameter. Grouted
column diameters from 0.75 m (for the single-fluid method) to 2 m are commonly quoted
and up to 5 m diameter has been claimed. Cylinders for temporary works would
normally be used in groups and the selection of the spacing between drillholes must
therefore ensure sufficient overlap of adjacent cylinders. In theory, the depth which
can be treated is not limited but in practice the method has been used to about 50 m
below ground.
If the mixing process is not properly effective, lumps of unmixed soil can be trapped
within the column (see Stark et al., 2009).
The treatment is generally terminated a short distance below ground level because of the
risk of the jetted fluids breaking through to the surface. While the plant used does not
need to be large (so that restricted access working is possible), columns may have to
be constructed in stages where there is a height constraint. If this is the case, it is
obviously important to ensure good overlap between adjacent sections.
The strength varies with the construction technique, for example, the proportion of the in
situ soil which is mixed with the grout, how well it is mixed and the nature of the soil
itself. Time is also a factor which may influence the design if the work is on the critical
path. Figure 8.3 gives an indication of the strength values which may be achieved,
depending on soil type and time after mixing. It can be seen that the full strength of
the mixture will take several weeks or more to be reached.
Permeabilities of 10 7 m/s and lower have been reported from various sites. Preliminary
testing for strength and/or permeability is necessary to ensure that the design and the
achieved properties are compatible. Generally, samples of the treated ground are
recovered in cored drillholes and tested in the laboratory. Piezometers can be installed
in these holes to permit measurement of the in situ permeability.
104
10 Notes:
1. Actual strengths depend on many variables, including the
proportion of soil. Those shown are indicative only.
2. *The method may not be suitable for fibrous peat but may
work in amorphous peat.
8
Unconfined compressive strength: MPa
Dec
6
re
asin
g so
il gr
ain
4
size
0
0 20 40 60
Time since mixed: days
The second aspect of design involves the selection of the construction method, the deter-
mination of the column diameter and grout mix requirements and the monitor rotation
and lifting speeds. All of these are empirical; they depend partially on the particular
equipment and should therefore be the province of the specialist contractor.
105
If the location is critical, it may be appropriate to survey long drillholes before grouting.
Parameters to be monitored during construction should include the grout density, the
fluid pressures and flow rates, the rate of rotation and withdrawal. When a column
has just been completed, its diameter can be checked using callipers. Conventional
instrumentation such as inclinometers can be used to monitor the performance of
retaining structures. It may be prudent to deploy precise survey techniques and borehole
extensometer systems during the grouting process to measure and control heave or
settlement.
Preliminary trials can be used to verify the condition of the treated ground by
exhumation (for example, see Stark et al., 2009). Treated ground can be drilled to
recover cores for testing strength and permeability. Piezometers can be installed within
or around treated ground to confirm its effectiveness in controlling groundwater flow.
When used to stabilise very soft clay (having a consistency similar to toothpaste) for a
1 km length of tunnel for the Singapore Mass Rapid Transit scheme, heave of up to
550 mm was recorded (Berry et al., 1987). Average heave above the westbound tunnel
(at 15–25 m depth) was 140 mm and above the eastbound tunnel (at 10–15 m depth)
was 70 mm.
106
dirty and potentially unsafe site. Control issues regarding drillhole location become
important. Off-site disposal of this effluent adds a significant cost to the process.
REFERENCES
Berry GL, Shirlaw JN, Hayata K and Tan SH (1987) A review of grouting techniques
used for bored tunnelling with emphasis on the jet grouting method. Proceedings of the
Singapore Mass Rapid Transport Conference, Institution of Engineers, Singapore.
BSI (2001) BS EN 12671: Execution of special geotechnical works: Jet grouting. BSI, London.
Coupland J (2010) New York’s Fulton Street Transit Centre – Dey Street Structural Box.
Proceedings of the DFI/EFFC Conference on Geotechnical Challenges in Urban Regenera-
tion, London.
de Wit JCM, Bogaards PJ, Langhorst OS, Schat BJ, Essler RD, Maertens J, Obladen BKJ,
Bosma CF, Sleuwaegen JJ and Dekker J (2007a) Design and construction of a metro
station in Amsterdam, challenging the limits of jet grouting. Proceedings of 14th
European Conference on Soil Mechanics and Geotechnical Engineering, 24–27 September
2007, Madrid: 1061–1066.
de Wit JCM, Bogaards PJ, Langhorst OS, Schat BJ, Essler RD, Maertens J, Obladen BKJ,
Bosma CF, Sleuwaegen JJ and Dekker H (2007b) Design and validation of jet grouting
for the Central Station, Amsterdam. Proceedings of 14th European Conference on Soil
Mechanics and Geotechnical Engineering, 24–27 September 2007, Madrid: 1299–1305.
Stark TD, Axtell PJ, Lewis JR, Dillon JC, Empson WB, Topi JE and Walberg FC (2009)
Soil inclusions in jet grout columns. DFI Journal 3(1): 33–44.
FURTHER READING
Bell AL (ed.) (1994) Grouting in the Ground. Thomas Telford Limited, London.
107
Chapter 9
Artificial ground freezing
Neil Smith Director, Applied Geotechnical Engineering
Artificial ground freezing (AGF) is a method for temporarily increasing the strength and
decreasing the permeability of ground. Since AGF works in all ground types, it is
particularly useful in highly variable water-bearing ground and for deep shaft sinking.
The ground is frozen by passing a cold fluid through pipes buried in the ground. The
system creates overlapping cylinders to form a freeze-wall to stabilise the void and
exclude groundwater. The moisture content of the soil must normally be above 10%
and the rate of flow of groundwater less than about 2 m/day if brine is the coolant or
20 m/day for nitrogen. The shape of the frozen ground is limited only by the ability to
orientate the freeze tubes. Several weeks are required for the frozen ground to be
ready for use if using brine, but nitrogen is much quicker. The system design needs
specialist knowledge of the complex properties of frozen ground. The structural design
of vertical freeze-wall cylinders can be based on empirical formulae. For other applica-
tions, conventional analyses can be used to determine the required strength and
dimensions of the frozen ground. Numerical thermal modelling is required to predict
freeze-wall growth. Monitoring of the work throughout the entire process is essential.
9.1. Introduction
It is well known to everyone who lives in a climate where freezing temperatures occur
that, when frozen, the soil near the surface is generally stronger than when unfrozen.
It is also obvious that water ceases to flow when it is frozen. From these two facts, the
principal benefits of ground freezing are clear: the soil is strengthened and groundwater
flow is prevented. Artificial ground freezing (AGF) is therefore a specialist process which
is used for temporarily preventing groundwater ingress into excavations and, where the
ground is unstable, assists in strengthening the ground. Its use is not widespread due to
the relatively small number of opportunities where the process may be required and, as a
result of this, there are few organisations with real competence in its use. General opinion
appears to balk at the up-front cost of employing the process but experience has shown
that, in a number of cases, if a value, risk and cost-effectiveness exercise had been carried
out first, the adoption of AGF would have brought savings. The fact that AGF operates
by modifying the water within the soil pores means that it is much less affected by
variations in soil type than, for example, dewatering or grouting and hence there is
less chance that treatment will be ineffective.
The basis of ground freezing is to pass a cold fluid through a pipe buried in the ground
and thereby freeze the pore water. This solidifies a column of ground approximately
109
centred on the pipe. By combinations of suitably closely spaced pipes, it follows that the
shape of a mass of frozen ground is limited only by the ability to orientate boreholes. It is,
therefore, a very flexible technique of ground stabilisation. Provided the soil has a
sufficient moisture content and groundwater is not flowing too quickly, any soil can
be frozen so that heterogeneous mixtures of sand/clay layers or glacial clay with
gravel lenses can be treated effectively.
When construction has been completed, the fluid circulation or gas exhaust is stopped and
the ground begins to thaw. The ground does not necessarily return to its previous condition.
Considerable time is required for a brine-based system to bring the ground to a condition
in which it is sufficiently frozen for work to begin. The stages are drilling boreholes and
installation of pipework, chilling the brine and building the freeze-wall. This process may
take up to four months.
The distribution system would typically consist of a header pipe (or ring main) and freeze
tubes within the ground. The freeze tubes can be connected to the header in series or in
parallel but, for risk reduction, it is normal that there should be at least two parallel
circuits to ensure that if a pipe leaks the whole system does not have to be shut down.
Adjacent freeze tubes should be connected to different parallel circuits to ensure
that the worst spacing between pipes is double the installed spacing. This needs to be
considered in the design factor of safety.
More recently, the ability to deliver liquid nitrogen to site in reasonably large quantities
and to store it in vacuum-insulated tanks has provided an alternative to brine. The
system does not require refrigeration plant or pumps to circulate the fluid. The nitrogen
vaporises at –1968C at a pressure sufficient to drive the fluid through the circuit without
the aid of pumps. At the end of the process, it is vented to the atmosphere rather than
re-circulated. The very low temperature of the fluid reduces the freeze-wall formation
110
Pump Pump
Compressor
Condenser
time from several weeks to a few days. The high temperature differentials make it more
appropriate to connect the freeze tubes in parallel rather than in series.
The loss of product and consequent maintenance costs may make the use of nitrogen
uneconomic if the ground must be kept frozen for an extended period. It is, however,
possible to use nitrogen initially to produce the freeze-wall rapidly, and then to
replace it with a brine system for longer-term freezing.
Schematic diagrams of the two systems are shown in Figures 9.1 and 9.2. Figure 9.3
shows a system installed in preparation for shaft construction.
Vacuum
insulated tank
storing liquid
nitrogen Vent to
atmosphere
Freeze
tubes
(connected in parallel)
111
Figure 9.3 Freezing system for shaft construction through 30 m of superficial deposits to bedrock
at Belmont, NSW
Courtesy of the British Drilling and Freezing Company Limited
significant advantage of freezing is that, given suitable conditions, the frozen ground will
encircle service pipes, making it possible to reduce the disruption and re-routing of
services often required by other forms of ground treatment. It is possible to use the
technique on a small scale, for example, to plug leaks in basements while a permanent
seal is formed.
g Calculations of the thermal behaviour of the ground are needed to ensure that the
columns of frozen ground around each freeze tube meet to provide a continuous
freeze-wall which is sufficiently thick. This is a specialist subject which requires the
knowledge and experience of the specialist contractor’s staff. The text in the
112
Time is a further factor to be considered in design. If the time available for freeze-wall
formation is short, then nitrogen may be preferred to brine or the spacing between
freeze tubes may be reduced. Over-capacity should be built into the design as a protec-
tion against the malfunction of any elements of the system, or simply a less effective
system than anticipated.
For thermal design, the heat capacity is measured per unit volume (not mass) of
material and is the quantity of heat transfer required for a unit change in temperature
of a unit volume of the ground. It is denoted by C and has units of J/m3K where K is
degrees Kelvin. The range of C is also small for a range of soils, roughly from 1.8 to
2.5 MJ/m3K. It decreases with decreasing temperature, tending towards zero at a
temperature of absolute zero.
The thermal diffusivity () is the ratio K/C and is the equivalent in heat flow to the
coefficient of consolidation in water flow and has the same units (L2/T). It relates to
the rate at which the temperature of the material can change. A high value of
implies a high rate of change of temperature. When water changes to ice, its value
increases roughly eightfold from 1.45 10 7 to 11.5 10 7 m2/sec. Hence, if other
factors are unchanged, the rate of temperature drop increases when the soil becomes
frozen.
113
particles. It also contains unfrozen water and air so the composite material is very
complex, as is its behaviour. Ice itself exhibits significant creep characteristics at low
shear stress levels; this is manifested in the flow of glacier ice. In soil, the creep of
the ice transfers stress to the intergranular (effective) stress between particles; if this
eventually exceeds the frictional capacity of the soil, failure will occur.
As examples of the variability of the strength of frozen soil, low strain peak shear
strengths of frozen Ottawa sand have been found to range from c. 5 to 8 MPa, increasing
with confining pressures which ranged from about 1.4 to 6.9 MPa. Strength is also
heavily dependent on strain rate, however. The strength of frozen silt has been shown
to vary by a factor of 3 for strain rates varying from 10 6 to 10 3 s 1. While the varia-
tions in strength can be seen to be very significant, it is also clear that the strength of the
frozen material is very much higher than strength values associated with unfrozen soil.
The two aspects of design must proceed in tandem. The engineer designing the freezing
system must be provided with the strength and stiffness requirements for the frozen
ground. It is also important that the designer knows the duration for which the
ground is to retain its enhanced strength and stiffness.
Other applications include the construction of temporary arch supports for tunnel
construction in unstable ground and stabilisation of localised zones for tunnel break-in
to (or break-out from) shafts. A relatively frequent application of the technique is to
‘rescue’ tunnelling projects which have run into difficulty. Freezing has been used in
instances where dewatering and/or grouting has been ineffective. Another infrequent use
is to form a temporarily strengthened roadway for heavy plant to cross weak ground. In
uses other than to facilitate shaft construction, the project engineer may use a conventional
analytical approach to determine the required properties and dimensions of the frozen zone.
This then provides the criteria for the design of the freezing system and freeze tube layout.
When water freezes and turns to ice, its volume increases by approximately 9%. In some
soils, however, especially low plasticity clays and silts, much larger ground movements
may be caused by frost heave. This occurs when lenses of frozen water are formed by
114
water which migrates in response to suctions which occur in the ground during freezing.
The localised increases in volume can also cause significant increases in pressure on
adjacent structures such as foundations or retaining walls.
Ground settlement or deformation can occur on thawing, with the potential to create
voids around the newly completed structure. In such circumstances, it may be necessary
to use compensation or ‘back-wall’ grouting to fill the voids and ensure the long-term
integrity of the structure.
Various researchers have investigated the effect of the freeze-thaw cycle on soil strength.
Some have found that the strength of clay increases following the cycle, where other
studies report a decrease in strength. If the post-freezing strength of the soil is a
matter of concern, it is advisable to carry out testing in advance of the works.
The thickness of the freeze-wall will continue to increase until the rate of heat extraction
is balanced by the supply from the ground. The void excavated within the freeze-wall will
generally be maintained at a temperature in which men and machines can work effec-
tively. There will therefore be some water generated in the void by melting. In addition,
where the void has to be ventilated and the air drawn from the surface is warm and
humid, substantial condensation can take place.
9.5. Monitoring
It may be necessary to survey the drilled holes prior to installation of the freeze tubes.
Significant deviations from the design hole location may result in gaps in a freeze-wall.
Accurate hole-drilling is especially important for deep shafts and when precision is
required (e.g. the rescue of a tunnelling machine).
Once the freezing process has been started, the ground temperature should be monitored
at suitable locations throughout the length of dedicated freeze tubes to ensure that the
plant is functioning correctly and to check that the ground is freezing as predicted.
Temperature monitoring must continue through the whole period of freezing until the
ground has thawed.
Ground levels must be monitored to check for heave and post-thaw settlement. For
schemes where the frozen zone is required to control groundwater, piezometers are
required to measure groundwater conditions.
REFERENCES
Haasnoot J (2010) Large scale ground freezing in the Netherlands. Proceedings of the DFI/
EFFC Conference on Geotechnical Challenges in Urban Regeneration, London
Harris JS (1995) Ground Freezing in Practice. Thomas Telford Publishing Limited, London
FURTHER READING
BTS/ICE (British Tunnelling Society and Institution of Civil Engineers) (2000) Specifica-
tion for Tunnelling. Section 4 Ground stabilisation processes. Thomas Telford Ltd,
London: 116–118.
115
BGFS (British Ground Freezing Society) (1995) Technical Memoranda on the Ground
Freezing Process. BGFS, Nottingham. (TM1 Harris JS, Wills AJ: Introduction to artifi-
cial ground freezing; TM2 Harris JS: AGF processes; TM3 Harris JS: Site investigation
for AGF works; TM4 Jones RH: Control of ground movements in AGF works; TM5
Harvey SJ, Wills AJ: Value, risk and cost effectiveness in AGF works; TM6 Harris JS,
Bell MJ: Shaft freezing; TM7 Harris JS: Tunnel freezing; TM8 Auld FA: Casting
concrete against frozen ground.)
116
Chapter 10
Slope stability in temporary
excavations
Neil Smith Director, Applied Geotechnical Engineering
Most construction sites require excavation and the stability of the sides must be
considered in relation to safety and economy. Where possible, it is usually most
economic to form excavations in open cut. The essence is to determine the steepest
practicable slope. Several factors influence slope behaviour: (1) the nature and variability
of the soil; (2) groundwater conditions; (3) loading close to the crest; (4) the time the
excavation will remain open; and (5) the consequences of failure. Factors (1) and (2)
require adequate information from site investigation. Factors (3–5) need good liaison
with the construction planning team. Construction is generally straightforward,
although it must be coordinated with other processes (especially groundwater control).
Slopes in rock require similar considerations, but the behaviour of rock slopes tends
to be dominated by the discontinuities since the blocks of intact rock are comparatively
strong. The important characteristics of rock slope discontinuities are: inclination
relative to the cut slope (noting that there may be more than one set of discontinuities);
frequency; persistence; openness; and smoothness. Safety is a paramount consideration:
unsupported trenches are examples of steep cut slopes and collapses are a major cause of
accidents on construction sites.
10.1. Introduction
This chapter gives guidance on the design and construction of temporary slopes for
excavations on construction sites. Construction of below-ground works within an
unsupported excavation has generally been viewed as the cheapest option. However,
the use of battered (as opposed to vertical) slopes requires excavation of significantly
larger amounts of material and the extra excavation can bring substantial additional
cost if the material has to be taken off site, especially if any of it is contaminated. The
use of sloping sides may also affect the choice of plant, for example, if cranes have to
be sited outside the excavation and so need a longer reach. Figure 10.1 is a decision
diagram showing the steps involved in determining whether excavating to a batter is a
viable option for the project.
Provided that the safety and financial risks associated with slope movement are properly
controlled, relatively simple approaches to the design of temporary slopes on construc-
tion sites can be acceptable. For large temporary slopes, sophisticated investigation,
analysis and monitoring may be appropriate (Kovacevic et al., 2007).
117
No
Design supported
excavation within Is a cut-off around
No Yes Install cut-off. Do not excavate
retaining walls, with the excavation
below groundwater level
dewatering or excavation practicable and
before the cut-off is complete.
under water if necessary. economic?
Slope stability in temporary excavations
In addition to the nature of the ground and the groundwater regime, time is an important
parameter for temporary slope design; the shorter the design life of the slope, the better.
The design must also take account of the consequences of movement occurring. In this
regard, it must be remembered that soil can tolerate much larger pre-failure strains than
most buildings or hard-paved surfaces. Failure of adjacent structures may therefore take
place before the slope itself fails.
Eurocode 7 (EC7; BSI, 1997) gives general guidance on slope design in Section 11.
Clause 2.2(1)P requires that both short-term and long-term design situations shall be
considered. In this context, the distinction between short and long term depends on
the nature of the ground as well as the design life of the slope. In coarse soil, long-
term conditions are established quickly and there is generally no requirement to consider
short-term conditions. In fine soil, initial stability is governed by undrained parameters;
an indefinitely long period is required for fully drained (long-term) conditions to be
established. Tomlinson (2001) presents information from a number of excavations in
London Clay; the period prior to slippage ranged from 1 day to no failure after four
months. The design life of temporary slopes in clay lies somewhere between fully
undrained and fully drained conditions.
1 Try to keep excavations away from the site property boundary to avoid a risk to
neighbours or the public (and party wall problems).
2 Ensure there are no critical services close to the crest or toe of the slope.
3 Plan drainage or surface-covering measures to minimise rainwater infiltration to
the ground beneath the slope, including the elimination of ponded water close
behind the crest.
4 Avoid routing haul roads or locating heavy plant just above the crest of a slope.
5 Avoid stockpiling spoil near the crest.
6 Where relatively steep slopes are needed, make an initial cut to a flat angle and
then cut to the steepest slope in short lengths (as short as practicable and less than
the slope height). Do the work then reinstate the slope to a flat angle before
excavating the next bay. Consider constructing in ‘hit and miss’ bays.
7 Keep the period for which the slope is open down to a minimum, for example, by
using a two-stage approach as in (6) above.
119
When an excavation is cut into clay, the surrounding ground experiences lateral and
vertical stress relief. In simple terms, this produces a tendency for the soil to expand
and hence generates a corresponding suction in the pore water while the effective
(intergranular) stress between the soil particles, which controls strength, remains
largely unaffected. An example of the extent of suction within a clay slope is given in
Figure 10.2, which is a prediction by Kovacevic et al. (2007) for a major excavation at
Heathrow Terminal 5. Over time, groundwater flows towards the zones of high pore
suction and an equilibration process takes place which reduces the suction and decreases
the stability of the clay slope. In a homogeneous clay of low permeability, this process is
slow, so that a relatively steep slope may remain stable for a substantial time (months or
years).
The suction pressure which can be maintained in a clean gravel is negligible and
equilibration is rapid so that, on excavation, the effective stress (and thereby the shear
strength) within the soil reduces; a steeply cut slope will quickly degrade to the angle
of repose of the soil.
Figure 10.2 Predicted contours of pore pressure within the London Clay at Heathrow Terminal 5.
Note the extent of the suction zone and the high suction pressures
Reproduced from Kovacevic et al. (2007)
Terrace gravel
Suction zone
120
In real soils, gravels and sands will generally contain some fine grains which may be
sufficient to allow small suctions to be maintained for short periods. This leads to
slopes which may stand longer than expected at steeper angles than the angle of
repose, but this additional stability may only persist for a short time. Another possible
factor which may maintain a steep slope in coarse soil is the deposition of salts at
grain contacts when groundwater evaporates. Reliance on these factors is risky unless
supported by good local experience.
On the other hand, real clays (especially glacially deposited clays) contain discontinuities
such as fissures or thin bands, pockets or ribbons of coarser soils. The discontinuities or
occurrences of coarser soil allow suction pressures to fall more rapidly and the result is
that the slope is likely to be less safe than predicted (or safe for a shorter time) and there
may be a higher risk of small failures within a larger slope (see Kovacevic et al., 2007).
In silts, sands and more permeable soils, flows can be established quickly such that
excavation side slopes can be rapidly eroded by groundwater issuing into the excavation.
This leads to severe problems on site. The answer is (a) to ensure that the groundwater
regime is properly known from investigation and monitoring prior to the design of
temporary works, and (b) to install suitable groundwater control measures before
excavation starts (see Chapter 6 on control of groundwater).
In rocks, adverse groundwater conditions reduce the stress across discontinuities and
hence the friction available to resist movement. They also impose lateral forces on
steeply inclined discontinuity planes.
EC7 Clause 2.1(14–21) defines three geotechnical categories (1, 2 and 3) as described in
the following sections.
10.4.3.1 Category 1
Clause 2.1(8)P of EC7 states that ‘small earthworks [shall be identified] for which it is
possible to ensure that the minimum requirements [see National Annex (BSI, 2007)]
121
10.4.3.2 Category 2
Clause 2.1(17) of EC7 defines Category 2 works as having ‘no exceptional risk or difficult
soil or loading conditions’. Most temporary slopes on construction sites would fall into
this category, but if the consequences of failure can be reduced sufficiently they could be
treated as Category 1.
10.4.3.3 Category 3
Clause 2.1(20) states that this category includes slopes which fall outside the other two
categories. Clause 2.1(21) states that this category ‘should normally include alternative
provisions and rules to those in this [EC7] standard’. The notes to this clause give the
examples of ‘unusual or exceptionally difficult ground or loading conditions’.
Temporary slopes classified as Category 3 are outside the scope of this guidance, and
specialist advice should be sought at an early stage in the planning of the project.
Figure 10.3 Small slip caused by temporary removal of the toe of a 100 year old cutting in stiff clay.
Some fill has been placed in front of the slip to stabilise the ground and to re-establish the access
road on which the men are standing
122
into a stiff clay to make an access road on which the two men are standing. At the time of
the photograph, fill had been placed in front of the slipped mass to restrain further
movement and to re-establish the access road. The slipped clay was subsequently
replaced with coarse granular fill. The slip that occurred could be considered a
Category 1 event, since the consequences of the failure were trivial and easily accommo-
dated by the contractor within a contingency.
However, the boundary of the site was very close to the slope crest (at the fence in
Figure 10.3) and the designer’s main concern was to avoid any ground movement
which would affect the neighbouring property. The risk of such a failure classifies as a
Category 2 event. It would therefore be possible in principle for the designer to accept a
lower factor of safety against a small failure such as the one which did occur, but to
require a higher factor against any failure which would reach the crest of the slope. This
example indicates that it is possible for the same slope to have more than one category.
As an example, 0crit of a well-graded sub-angular sand and gravel may be taken as 368.
The factor of safety to be used on slopes cut into coarse-grained soil will depend on the
circumstances. Where the consequences of slope movement are small (Category 1),
F ¼ 1.0 may be acceptable provided that the value of 0crit is reasonably conservative.
For longer-term slopes or where failure is to be avoided (Category 2), then F ¼ 1.25
may be prudent. The factor of safety is applied to tan(0crit); if the slope angle is , then
tan ð0crit Þ
F¼
tan
Hence, for 0crit ¼ 368, F ¼ 1.25 would give a slope () of 308 (1 : 1.73).
123
Figure 10.4 Suggested temporary cut slope gradients for different materials
2:1
1.5:1
1:1
Cl
ay
(te
mp
1:1.5
or
ary
)
1:2
San
W
Mas
ell ock
siv
rock e
Gr
(perm
d
joi
r
ave
nt
Clay
1:3
l
ed
anent
)
1:6
It may of course be possible to cut slopes to angles steeper than 0crit, provided that some
failures can be accepted. Factors such as the presence of some fine soil combined with
slight cementing of the soil grains can give a considerable increase in the stable slope
angle over that which theory would predict.
For the case with the water table at the ground surface, stable angles are approximately
half that of dry slopes.
124
A – Angularity1 A: 8
Rounded 0
Subangular 2
Angular 4
B – Grading of soil2 B: 8
Uniform 0
Moderate grading 2
Well graded 4
1
Angularity is estimated from a visual description of the soil.
2
Grading can be determined from grading curve by use of: Uniformity
coefficient ¼ D60/D10 where D10 and D60 are particle sizes such that in the
sample, 10 % of the material is finer than D10 and 60 % is finer than D60.
Grading Uniformity coefficient
Uniform <2
Moderate grading 2 to 6
Well graded >6
Conventional limit equilibrium analyses may be used to analyse slopes. A problem with
these analyses is that the failure mechanism must be assumed (or guessed) in order to
perform the analysis. If the mechanism is wrong, the analysis is likely to give misleading
results. A recent development which can help with the determination of the critical
ultimate limit state mechanism is discontinuity layout optimisation (Smith and
Gilbert, 2007). However, the early approaches of Bishop (1954) (circular surfaces) and
Janbu (1954) (general surfaces) are still in wide use. An advantage of the assumption
125
of a circular plane of sliding is that a family of potential surfaces can be defined in terms
of a grid of centres and a range of radii. The most critical of these can then be found.
In uniform clay soil, the critical slip surface is likely to be moderately deep and to emerge
somewhere close to or a little beyond the toe of the slope. Where a weak layer is present,
the most critical surface found in a circular analysis will be tangential to the weak layer,
but a distinctly non-circular surface will probably be the most critical. In some cases,
non-circular surfaces can be simplified to an active wedge at the upslope side, a
passive wedge at the downslope side and a polygon sliding on the weak layer between
the two wedges.
In view of the uncertainties involved, the analysis is best viewed as a tool to guide
the judgement of the designer in the selection of the most appropriate slope to adopt.
Whatever method is used, sensitivity analyses are essential to give an appreciation of
the possible effects of differences in shear strength. It is normal to use undrained soil
parameters in the analysis of the short-term stability of clay slopes, but to allow for
some loss of strength dependent on the design life of the slope and the nature of the
material. A range of strength values should be used to examine the effects of softening
(loss of suction) on stability.
10.5. Monitoring
There are two principal principles in the construction of temporary slopes. As mentioned
in Sections 10.2 and 10.3, one is to minimise the angle, length and open period of the
slope and the other is to monitor the slope. Monitoring systems must be properly
designed, bearing in mind that it is no use gathering data which cannot be analysed
and interpreted. It is possible to acquire so much information that important trends
cannot easily be discerned.
All monitoring systems should include regular visual inspection of the slopes, especially
the areas just behind the slope crest and at the toe. Substantial slope movements can be
preceded by smaller movements which are manifested in cracking behind the crest
126
(tension cracks) or bulging of the ground near the toe. If such movement is seen, there is
often time to carry out some stabilisation work such as placing backfill at the toe before a
major failure takes place. The ground should also be checked for indications of ground-
water emerging from the slope; this may simply be dampness or softening, rather than
flowing water.
A wide range of monitoring systems is available, from very simple to very sophisticated. At
the simple end of the spectrum, a row or rows of ranging poles arranged in straight lines
along and above the slope can be visually monitored on a daily basis. It should be noted,
however, that the predominant component of movement of the ground at the top of a
slipping mass will be vertical, so level measurement is important. Targets fixed to pegs
can be surveyed precisely, either manually or by automated instruments. In-ground
instrumentation can take the form of slip indicators, inclinometers or piezometers.
Slip indicators are essentially simple tubes installed in boreholes drilled through the zone
of likely movement into stable ground below. Slippage is evidenced by a closure of the
tube at the surface of shearing; however, these instruments do not provide information
on the amount of movement.
There are two types of inclinometer. The early instruments consisted of tubes grouted
into boreholes (taking care to try to match the stiffness of the installation to that of
the ground). Like slip indicators, these must extend into fully stable ground at depth
in order to provide a datum for lateral movement. A ‘torpedo’ is lowered down the
tube, using locating grooves to orient readings of the tilt of the torpedo at (normally)
0.5 m intervals down the tube. In that way, a continuous profile of the tube is obtained
from each monitoring visit. More recently, inclinometers have been developed to
measure the tilt of devices which are permanently installed in the ground. This allows
remote reading and the distribution of data through the internet.
Note that it is generally less practicable to install instruments at or close to the toe of the
slope since this is the main area of construction activity.
Finally, the contractor and designer (and any other concerned parties) must agree trigger
levels and contingencies to stabilise a moving slope. The simplest and quickest approach
to movement is often to place new fill in front of the slipped or slipping mass (see
Figure 10.3). This may be only an emergency measure to arrest movement and to
allow work on re-design of the slope. Careful excavation of failed fine-grained soil and
replacement by coarse granular material such as hardcore can be a sufficient remedy.
A further alternative may be the installation of soil nails, which is effectively a way of
reinforcing the soil mass. Nails are usually drilled and grouted in place although
127
driven nails can be used. Unlike ground anchors, they are not pre-tensioned and require
further movement of the ground to develop tensile resistance. While costlier than repla-
cement, they may allow the reinstatement of a steep slope required for construction.
REFERENCES
Bishop AW (1954) The use of the slip circle in the stability analysis of slopes. European
Conference on the Stability of Earth Slopes, Vol. I, Stockholm.
BSI (1994) BS 8002: Earth Retaining Structures. BSI, London.
BSI (1997) BS EN 1997-1: Eurocode 7 Geotechnical design. BSI, London
BSI (2007) UK National Annex to Eurocode 7 Part 1. NA to BS EN 1997-1. BSI, London.
Janbu N (1954) Application of composite slip surfaces for stability analyses. European
Conference on the Stability of Earth Slopes, Discussion Vol III, Stockholm.
Kovacevic N, Hight DW and Potts DM (2004) Temporary slope stability in London Clay
– back analyses of two case histories. Advances in Geotechnical Engineering.
Proceedings of the Skempton Conference, London, 3: 1–14.
Kovacevic N, Hight DW and Potts DM (2007) Predicting the stand-up time of temporary
London Clay slopes at Terminal 5, Heathrow Airport. Ge´otechnique 57(1): 63–74.
Pettifer GS and Fookes PG (1994) A revision of the graphical method for assessing the
excavatability of rock. Quarterly Journal of Engineering Geology 27(2): 145–164.
Smith CC and Gilbert M (2007) Application of discontinuity layout optimization to plane
plasticity problems. Proceedings of the Royal Society A: Mathematical, Physical and
Engineering Sciences 463(2086): 2461–2484.
Tomlinson MJ (2001) Foundation Design and Construction. Seventh edition. Pearson
Education, London.
FURTHER READING
Bromhead E (1992) The Stability of Slopes, 2nd edition. Taylor & Francis, London.
Hoek E and Bray JW (1981) Rock Slope Engineering. E & F N Spon, London.
128
Chapter 11
Sheet piling
Ray Filip Temporary works consultant and training provider – RKF Consult Ltd
Interlocking steel sheet piles are profiled steel sections which are installed vertically into
the ground by a variety of methods to support the ground and exclude water when
excavating or to carry applied loads. They are used in temporary works to support the
ground and to exclude water while a structure or service is installed below ground
level. They are generally extracted to be re-used; however, they may be incorporated
into the permanent works such as sealed basement walls where they may also be used
to carry vertical loads from the building or structure. Sheet piles can also be used as
anchorages and in conjunction with other types of retaining walls. They may be used
with tubular or I-section piles to form a high modulus combination wall. ‘Cofferdam’
is a term given to a closed sheet pile ‘box’ used for deep retention schemes and which
may include provisions for excluding water. Isolated steel I or H sections or steel
tubes can be driven into the ground to form bearing piles. The design concept of sheet
piling is unusual in that it utilises the shear strength of the ground to support itself;
the element being supported is therefore also part of the supporting mechanism.
129
Figure 11.1 Examples of U profile, Z profile, combination wall, straight web and box pile
used in straight lines or can also be installed to a gentle radius. Combination walls are for
high modulus work (e.g. for marine work) and comprise steel I beams or large diameter
steel tubes with U- or Z-profile sheets between. Where a line of piles is required to turn a
corner or form a junction, special corner sections are available (see Figure 11.2).
Sheet piles are available as hot-rolled sections with standard steel grades varying from
grades S240GP–S460GP to EN 10248 (BSI, 1996a) and cold formed sections with stan-
dard steel grades from S235JRC to S355JRC to EN 10249 (BSI, 1996b). From grade 270
to 355 the increase in stress is 30% but the increase in procurement cost is approximately
2% (also improved driveability). Specialist products such as marine grade steels are also
available. For further information on the available range of products, rolling tolerances,
corrosion protection and achievable radii, see the Arcelor Mittal Piling Handbook (2008)
or similar information from other manufacturers.
130
~45
~45
~50
90° ÷ 135°
~50
C 14 OMEGA 18
Mass ~14.4 kg/m Mass ~18.0 kg/m
~45
~5
~50
60
~30
°÷
12
0°
C9 DELTA 13
Mass ~9.3 kg/m Mass ~13.0 kg/m
11.3.3 Vibrators
Vibrators are suitable for use in predominantly granular soils and soft to firm silts or
clays. Small versions known as excavator-mounted vibrators are attached to a standard
site excavator, whereas larger (more powerful) versions are attached to a piling rig or
freely suspended from a crane. Spinning eccentric weights generate the vibration (spin
in opposite direction to cancel out the horizontal components leaving the vertical
component) which is used to liquefy the soil and reduce resistance rather than drive
the pile. With rig-mounted vibrators, a crowding force is often provided to concurrently
push the piles and the self-weight of the vibrator is utilised for crane-suspended versions.
Vibrators are not very suitable in stiff clay.
131
Hammers can be single-acting (where the power is used to lift the drop weight so it can
fall under gravity) or double-acting (using the impulse to accelerate the drop weight
downwards as well as lift it back up for the next blow). Double-acting hammers
provide a much faster impact rate (120 blows per minute), but single-acting have their
place in highly cohesive or tough ground conditions. Hammers generate significant
noise (even when acoustically shrouded) and develop vibration in hard driving condi-
tions.
Noise and vibration are significant issues and local authorities can set limits under the
1974 Control of Pollution Act (see Figure 11.3).
11.3.6.1 Pre-augering
A small auger is screwed into the ground at the sheet pile clutch interlock positions. This
loosens the dense ground locally and probes for buried obstructions. For retention
schemes, pre-augering below formation level should be avoided as it reduces passive
resistance.
132
Piles are often installed by tracked machines such as leader rigs or with piling
equipment suspended from crawler cranes. Consideration must be given to the
working platform beneath the machinery (see Chapter 5 on site roadways and
working platforms).
133
g bituminous sealants: applied pre-driving and mainly used for temporary sheet
piling
g hydrophilic or resin-based sealants: applied pre-driving, more effective than
bituminous sealants and mostly suitable for permanent works but can be damaged
during installation particularly by heat generated by hard vibratory installation
g seam welding clutches: applied after installation and provides the highest degree of
water tightness; predominantly used for permanent sealed basements.
11.3.9 Cofferdams
Cofferdams are essentially a sheet pile ‘box’ formed to support the ground and possibly
to exclude water from the excavation (Figure 11.4). Cofferdams are conventionally single
skin but, for very deep excavations, double-skin walls are used which are formed from
two parallel lines of piles filled with rock and tied together with tie rods. These are
generally designed as a gravity-retaining structure. Corner sections (see Figure 11.2)
are used to form a continuous wall in square or rectangular cofferdams. Circular
cofferdams are generally installed using circular piling gates by the panel-driving
method though leader rigs can also be used.
For all cofferdams (and retaining walls in general), internal framing or external anchorage
can be used. The design and detailing of a cofferdam and its framing or anchorage requires
careful consideration by a competent designer with experience of the construction process
that will occur within the cofferdam. The sequence of installation and removal must be
considered in the design to identify the critical loads. It is not unusual for the critical
load case to be during the removal process rather than installation.
It is difficult and generally not economic to exclude 100% of the water, hence sump
pumps are used to pump water out of the cofferdam. Dewatering can be used to
134
Figure 11.4 Sheet pile excavation with heavy duty proprietary framing
Courtesy of Groundforce Shorco Ltd
control water ingress but can be expensive and has significant risks associated with it (see
Chapter 6 on control of groundwater). Water may also enter the cofferdam over the top
of sheet piles (when positioned in a river or sea) or from beneath. Care should be taken to
avoid base failure by piping or heave. The sheet piles can be driven to a greater depth to
penetrate an impervious soil layer (cohesive material) as a cut-off typically around 2 m. If
it is not possible to utilise a cut-off, an underwater concrete plug can be formed in the
base of the cofferdam and the water pumped out once the concrete has hardened.
The supports to cofferdams can be provided by ground ties and anchorage, fabricated
steel internal framing or heavy-duty proprietary hydraulic framing (hydraulic struts
with capacities of 2500 kN and more are available). Further information is available
in CIRIA SP95 (2003a).
11.4. Eurocode 7
A variety of design methodology and guidance documents have been developed such as:
Arcelor Mittal (2008), CIRIA 104 (1995), CIRIA C580 (2003b) and CIRIA SP95
(2003a). The most recent code is EC7 (BSI, 2004a) and EC3 (BSI, 1990) and the UK
National Annexes. EC7 is a comprehensive limit state design code based on the use of
135
g use of calculation
g adoption of prescriptive measures such as the use of conservative rules, attention
to specification and control of materials, workmanship, protection and
maintenance procedures
g experimental models and load tests
g observational methods.
Section 2.2 of EC7 requires that both short- and long-term design situations are
considered. Typical ultimate limit states from section 2.4.7.1 of EC7 are
Partial factors are defined in Annex A of EC7 with the factors being increased for
abnormal risk situations or reduced for less severe cases, temporary structures or
transient design situations. A partial factor of 1.0 should be used for accidental
situations.
Section 2.4.7.3.4 of EC7 provides details of three design approaches for the various limit
states. Section 4 of EC7 states that: ‘to ensure the safety and quality of a structure, the
following shall be undertaken, as appropriate’.
136
For flexible retaining walls supported by anchors or struts, the magnitude and
distribution of earth pressures, internal structural forces and bending moments depend
to a great extent on the relative stiffness of the structure and the stiffness, stress and
strength of the ground. For problems of ground–structure interaction, analyses should
use stress–strain relationships for ground and structural materials and stress states in
the ground that are sufficiently representative, for the limit state considered, to give a
safe result.
Sections 7–9 of EC7 provide guidance on the design of piled foundations, anchorages
and retaining structures (where diagrams of modes of failure are shown). EC7 (BSI,
2004a) does not provide specific guidance to the supervision, monitoring and main-
tenance of retaining structures. This specific guidance can be found in EN 12063 (BSI,
1999) for sheet pile walls and EN 12699 (BSI, 2001) for displacement piles. Specifiers
may use ICE Specification for Embedded Retaining Walls (2007) for guidance.
Issues in EC7 (BSI, 2004a) that remain to be resolved include tie bar and ‘dead-man’
anchorage systems and factoring water pressures.
11.5. Design
The first and perhaps the most important stage in design process is to interpret the site
investigation information and establish a geotechnical model for analysis. Most site
investigations are commissioned for the permanent structure and it is not uncommon
for additional investigation to be required to aid the temporary works design. EC7
Part 2 (BSI, 2004a) now gives guidance and rules for the extent of the site investigation
required for embedded retaining walls and bearing piles.
There are two philosophies of design so that we can utilise the inherent shear strength of
the soil to help the sheet piling support it, namely
Although SSI is a very precise analysis, it uses the elastic properties of soil which can only
be determined to a very imprecise level. Historically, LE has therefore been the common
method of calculation. It permits very simple pile designs to be carried out manually;
whether the use of SSI or LE provides the best solution is best determined by use of
specialist software.
A design should aim to be both safe and economical and the structure classification
should determine the expertise employed in the design. Codes and standards that
should be followed have recently changed; British Standards were effectively withdrawn
and ceased to be maintained on 31 March 2010, to be replaced by Eurocodes. There are
associated European Standards effectively in place to be used by the designer in conjunc-
tion with Eurocodes, but the designer should be aware that, although limit state methods
137
and analysis are common to present and former Codes of Practice, application and
values of partial factors and resulting Factors of Safety are different.
The design of sheet pile walls, cofferdams, etc. and their supports need to consider the
elements described in the following sections.
11.5.4 Deflection
Deflection should be limited to prevent damage to surrounding, roads, buildings or
services caused by settlement or ground movement. Deflection limits should be
established and a monitoring regime established if necessary. Readings should be
taken on a regular basis during excavation and also once excavation is completed.
Trigger levels should be established, such as
138
Deflections as well as the required penetration of cantilevered sheet pile walls can be
significant for retained heights over c. 4.0 m (unless very heavy piles are used).
Torsion on the waling beam should be considered due to the vertical reaction imposed
from raking props, unless the vertical reaction can be carried by corbels. Self-weight
bending and impact loading (in the least favourable position) should be allowed for.
Prevention of progressive collapse from impact is usually achieved by back-analysing
assuming any one strut is removed and limit state conditions. The design of steel
framing was developed from BS 449 (BSI, 1989), BS 5950 (BSI, 1990) and EC3 Part 5
(BSI, 2003), with all three methodologies currently in use.
139
11.5.14 Driveability
Drivability refers to the ability of sheets to be installed into the ground without buckling,
damage or end-resistance or wall-adhesion preventing penetration. It should be noted
that, in all soils other than those of low strength, the calculated structural requirements
of the sheet piles will almost invariably be less onerous than the strength needed for
driveability. The Piling Handbook (Arcelor Mittal, 2008) provides the calculation
method for suitable sections relative to items such as Standard Penetration Test (SPT)
N values and soil descriptions, length and steel grade of pile and installation methods,
but tends to give overly optimistic findings for drivability. Generous factors are
advised in the calculations.
140
g Use coefficients and the Rankine formula to develop active and passive
pressure diagrams. Over-dig, tension cracks filling with water on the active side
and soil softening on the passive side should be considered. The active and
passive pressure diagrams should then be combined to give a design analysis
model.
g Supports may be introduced at appropriate levels using a framing/anchorage
arrangement to allow the work to be carried out as easily as possible and limit
deflections.
g Forces and moments are resolved to test equilibrium of the total pressure diagram
in the ultimate limit condition. This is the maximum load situation and will give
the required penetration (adding a length to permit reverse passive reaction).
Partial factors are applied to variable and applied loads in accordance with EC7
(BSI, 2004a); the worst case for Design Approach 1 (UK National Annexe)
combinations 1 or 2 determines the required pile length with the design effect
moment and forces.
g The depth of penetration determines the length of pile required.
g The maximum moment multiplied by a factor plus the yield strength of the pile
steel will give the minimum required pile section modulus and therefore the pile
section designation for structural adequacy. The pile section modulus may need to
be increased to account for ‘driveability’ (see Section 11.5.14). After checking
driveability, the chosen section is verified for structural adequacy in accordance
with EC5 and the National Annexe after taking into account effects of corrosion
on the section properties including the check on section classification and the
effects of that. Piles can be verified for bending strength using plastic section
properties if remaining in Class 2 but, if the section drops to Class 4, then
buckling and shear may determine the section. It is best to adopt sections
remaining in Class 3 or higher for structural walls.
g Multiply the forces put into the supports by a factor FT if using the limit
equilibrium (LEQ) method to allow for the interaction of the soil and piles. This
gives the design forces for anchors or frames.
g Back-analyse the frames to test for any one strut or tie accidentally removed for
the serviceability limit state, to check that progressive collapse will not occur.
g Maximum potential deflections and ground settlements may be estimated using
SSI or Finite Element methods of analysis. Alternatively, the appendix to CIRIA
C580 (2003b) shows deflections and settlements from extensive site studies for use
with LEQ.
141
REFERENCES
Arcelor Mittal (2008) Piling Handbook, 8th edition. Arcelor Mittal.
BSI (1989) BS 449: Specification for the use of structural steel in building. BSI, London.
BSI (1990) BS 5950: Structural use of steelwork in building. BSI, London.
BSI (1996a) BS EN 10248: Hot rolled sheet piling of non-alloy steels. Technical delivery
conditions. BSI, London.
BSI (1996b) BS EN 10249: Cold formed sheet piling of non alloy steels. Technical delivery
conditions. BSI, London.
BSI (1999) BS EN 12063: 1999, Execution of special geotechnical work, sheet pile walls.
BSI, London.
BSI (2001) BS EN 12699: Execution of special geotechnical work, displacement piles. BSI,
London.
BSI (2003) BS EN 1993: Eurocode 3, Design of steel structures. BSI, London.
BSI (2004a) BS EN 1997: Eurocode 7, Geotechnical design. BSI, London.
142
FURTHER READING
BSI (British Standards Institution) (1984) BS 8002: Code of practice for earth retaining
structures. BSI, London.
Byfield M and Mawer R (2001) The development of section modulus in Larssen U-shaped
sheet piles. University of Cranfield, Cranfield.
CIRIA (Construction Industry Research & Information Association) (1984) Design of
retaining walls embedded in stiff clay. Report 104. CIRIA, London.
CIRIA (2008) EC7 implications for UK practice. Report C641. CIRIA, London.
Dawson R (2001) Steel to replace concrete. Proceeding of the Institution of Civil
Engineers, Geotechnical Engineering 149(4): 205–207.
Day RA and Potts DH (1989) Comparison of design methods for propped sheet pile walls.
SCI Publication 077.
Filip RK (2004) Recent advances in quiet and vibration-less steel pile installation.
Proceedings of Conference of Deep Foundations Institute, Chicago: 1–11.
ICE (Institution of Civil Engineers) (1996) The Observational Method in Geotechnical
Engineering. Thomas Telford Publishing, London.
Packshaw S (1962) Cofferdams. Proceedings of Institution of Civil Engineers 21: paper
Number 6588.
Rowe PW (1955) A theoretical and experimental analysis of sheet pile walls. Proceedings of
ICE 4(1): 32–69.
Rowe PW (1957) Sheet pile walls in clay. Proceedings of ICE 7(3): 629–654.
Symons IF, Little JA, McNulty TA, Carder DR and Williams SGO (1987) Behaviour of a
temporary anchored sheet pile wall on A1(M) at Hatfield. TRRL Research Report 99,
TRL, London.
Yau JHW and McNicholl DP (1990) Failure of a temporary sheet pile wall: case study. In:
Proceedings of the Seminar on Failures in Geotechnical Engineering, University of Hong
Kong.
Useful Websites
ArcelorMittal Ltd, www.arcelormittal.com
143
144
Chapter 12
Trenching
Ray Filip Temporary works consultant and training provider – RKF Consult Ltd
A trench is defined as an excavation whose length greatly exceeds its width and depth. It
is generally considered that a shallow excavation is less than 6 m deep (most trenches)
and a deep excavation is greater than 6 m deep (specialist advice should be sought).
Trenches can be excavated by hand or mechanical digger to allow a service, pipeline
or foundation to be installed or, when backfilled with granular fill, used to improve
drainage. A trench may have battered sides when site circumstances allow, or temporary
shoring may be installed (traditional timbering, sheets and frames, proprietary trench
boxes) to protect workers entering the trench. Controlling water ingress into a trench
is often a major issue (see Chapter 6 on control of groundwater). A safe system of
work and adequate planning is necessary, especially when workers are required to
enter a trench, and the trenching operation can also have a detrimental effect on the
surroundings when appropriate management control measures are not provided or
inappropriate solutions used. Good health and safety practice is critical to trenching
operations.
12.1. Introduction
12.1.1 Major alternatives
Trenchless techniques (to minimise surface disruption) include methods such as
micro-tunnelling, auger/thrust boring, pipe ramming, impact moling, directional
drilling, mole ploughing and rock boring. Trenchless techniques might not be possible
or practical because soils are unsuitable or very large diameter or very long lengths of
pipework are to be installed, or due to cost limitations, site-specific restrictions and
the non-availability of specialist machinery and labour (CIRIA, 1987).
12.1.2 Soils
Soils can be classified by particle size (broadly cohesion-less or cohesive), compactness
and structure with a separate classification for organic soils (soil classifications are
provided in various documents such as BS 5975 (BSI, 2011; table 18) and BS 8002
(BSI, 1994). Accurate soil descriptions are essential to allow soil design parameters to
be determined. Designers will summarise and approximate actual soil conditions to
145
form an idealised model for design and these approximations have to be confirmed by
monitoring the soils being excavated.
Granular soils are sands and gravels which have an angle of repose. This angle is
dependent on particle shape, particle roughness, compaction and grading. A well-
graded material has a variety of sized particles and these tend to pack together, giving
higher angles of repose. Poorly graded materials have a predominance of a single-size
soil and tend to have lower angles of repose. Granular soils have high permeability
and close sheeting will generally be required for deeper trenches or when a high water
table is encountered to prevent collapse, significant ground movement and fine materials
being washed out. If the groundwater pressure is not relieved (de-watering) or released
through any gaps in the support scheme, there will be a large increase in the pressure
on the support scheme.
Cohesive soils are bound together due to the effects of water between the fine particles in
clays and silts (this is known as cohesion). When excavated, clay soils can stand near
vertical without support for a period of time. The cohesion in silts will vary with moisture
content and they are less stable than clays. The permeability is low (particularly in clays)
and drainage occurs over a significant time period, hence extensive dewatering is often
not required in cohesive soils (sump pumping is often adequate) unless bands of sands
or coarse silts (known as partings) are present or the soil is heavily fissured. Even
though these soils can be relatively stable in the short term, a risk assessment should
determine whether support is necessary (considering fissures, surcharges, adjacent
works, timescale, etc.).
Rock has cemented particles and even though the rock mass itself may be very strong,
inclined bedding planes and fractures can lead to instability of blocks. Heavily weathered
rock may need to be considered as a very dense cohesionless material. Excavating rock
can require heavy machinery or specialist techniques.
When groundwater is encountered, the following items should be observed: the level at
which the water was encountered, the rate of inflow and rise of water and the final level
which the water reaches.
146
and ditches dug to cut off surface water. In layered soils, different batter angles may be
appropriate for each of the strata.
Regular checks should be carried out looking for potential signs of instability or
degradation. Warning signs of rotational slip failures are: bulging at the toe and
tension cracks appearing at the top of the batter. If these signs are noted then the
batter can be re-graded, weight added to the toe of the slope or the drainage of the
batter improved.
In dry conditions over a short duration, granular soils have an internal angle of friction
(’) which will typically range c. 25–508 and batters may be safely cut at an angle slightly
less than ’. When groundwater is present, the safe angle will be reduced. In the longer
term (months), the angle will need to be less than the effective internal angle of friction
(’0 ). Cohesive soils can be cut near vertical in the very short term (up to a week), but
drainage occurs and collapse will inevitably follow. Where men are required to work
in a confined excavation, then the sides will need to be safely battered or sheeting and
propping or boxes used. Safe batter angles are listed in many publications (e.g.
CIRIA, 1983).
147
Some contractors and utility companies have standard solutions for relatively
shallow low-risk trenching operations. These pre-designed solutions indicate sizes for
trench sheets, walers, struts, traffic clearances and surcharge limitations. Appropriate
risk assessments and method statements are provided by a competent person and
operatives must have been appropriately trained. Such solutions are very useful for
short duration or repetitive ‘reactive maintenance’ contracts (repairing a damaged
service).
148
12.2. Techniques
Over a very short period of time and for a relatively shallow depth most soils will stand
near vertical, allowing sufficient time for certain shoring systems to be installed. If the
sides do need to be supported (as a result of a risk assessment being carried out) due
to the nature of the soil, longer-term use, sensitive surroundings or for deeper trenches,
then a trench support scheme will be required. A competent person should make the
decision (by risk assessment and consideration of site-specific conditions) if any of
the below types of equipment are suitable and which type to use. Work should not be
permitted outside the protected area. All equipment should be assembled and used as
per the manufacturers’ instructions and consideration should be given to a safe
method of installation and a safe method of removal.
Drag boxes are suitable for relatively shallow trenches (up to about 4 m depth). Trench
boxes can be assembled in lifts to achieve significant depths (over 6 m is not uncommon).
Trench sheets can be placed at the ends of the boxes to provide additional protection
to the open ends. With both versions the pre-excavated trench is wider than the box
to allow the box to be installed and moved; they do not support the sides of the
excavation, but act as a shield to protect the workforce (even when backfill is placed
149
between the box and the surrounding ground there is a period of time when the ground
has to be self-supporting). There is a risk of the ground collapsing onto the box affecting
operatives, plant and services on the surface around the box, whereas those inside the
box are shielded.
Excavation should not be permitted beneath the box (known as ‘flying’ a box) as friction
between the box and the surrounding ground cannot be relied upon to support the box.
An alternative installation method for a trench box known as ‘dig and push’ exists (see
Figure 12.3). With this method the soil is supported throughout installation; although
this reduces the risk of significant ground movement, it will be slower.
150
Figure 12.2 Trench box on site (edge protection not installed). Note backfilling between the box
and ground, dewatering and excavated spoil kept away from edge of trench
Courtesy of Groundforce Shorco Ltd
151
Figure 12.3 ‘Dig and push’ sequence for installing a trench box
Courtesy of Groundforce Shorco Ltd
A simple guide system (baulk timbers or steel beams) should be used to ensure the sheets
are driven near vertical and to maintain a straight line. When site conditions allow, the
sheets can be partially installed into a pre-excavated ‘lead trench’ which is then back-
filled. Multiple levels of horizontal waling beams with hydraulic, adjustable struts (or
timber walings and struts) are used with the spacing of the struts suiting the pipe
lengths to be installed. The trench depth is limited by the length of sheets that can be
installed by light machinery, especially with interlocking (rather than overlapping)
sheets which require additional height for ‘clutching’ the sheets together. The slenderness
of the sheet can also preclude driving into dense or stiff soils. These systems can also be
installed using the ‘dig and push’ method.
152
Small gaps can be left in the sheeted wall to allow for services crossing the trench and, if
relatively stable ground is encountered, then intermittent sheeting can be used. This is the
principal reason why the horizontal waling system is so adaptable and popular. Simi-
larly, manholes can be constructed in a square- or rectangular-sheeted excavation
using proprietary hydraulic supports known as manhole braces, whereby all four
walers in a frame are hydraulically driven off a common manifold (timbers can also
be used). Gaps can be left in the sheets or some sheets left above formation level to
allow pipework to be installed.
153
Figure 12.5 (a) Post and plank excavation and (b) side rail system trench
Courtesy of Groundforce Shorco Ltd
(timber, precast concrete, in situ concrete or steel sheets) which span between the flanges
of the vertical H-section. When placed into augured holes, the bottom of the H-section is
often held in concrete. The planks are installed and wedged in position, one beneath the
other, as the excavation progresses downwards. When using timber, steel or precast
154
planks, the soil behind the planks has to be excavated slightly larger to accommodate the
planks, with sand backfill placed behind the planks to fill the gap, but some ground
movement is inevitable. In situ concrete planks accommodate construction tolerances
and are cast against the excavated face; backfilling is, therefore, not required. This
method has the advantages that the main structural supports are installed prior to
excavation, can be easily altered and gaps can be left to accommodate services. However,
it can be time-consuming, labour-intensive and some ground movement is possible.
A variant uses light steel sections spanning horizontally between the flanges of the posts
with trench sheets pushed down progressively, and this dramatically reduces the labour
content. A proprietary version is known as a ‘side rail system’ which comprises posts,
sliding panels (each c. 1.5–2.5 m deep) and struts, which form a continuous ground
support and are installed by excavator. They can be used for widths up to 7 m and
depths in excess of 7 m. The equipment is relatively large and heavy and requires
heavy excavators for installation. For deeper trenches with large diameter pipes, an in
situ concrete blinding is cast at the base and lower levels of struts can be moved
upwards or removed. Additional support such as trench sheets may be required to
provide additional support at each end.
Proprietary equipment is designed for the most robust loadings envisaged, and in all but
obviously excessive conditions the need for detailed calculations is unlikely. Where it is
felt wise to calculate, empirical methods as given in the following section may suffice.
Supports for trenches often have several levels of supports; for such cases traditional soil
mechanics theory such as the triangular Coulomb pressure diagram have been found not
to apply (because these support systems are relatively flexible and there is a degree of soil
arching and displacement which leads to a redistribution of pressures). From full-scale
tests, Terzaghi and Peck (1996) derived empirical trapezoidal pressure diagrams for
calculating maximum strut loads. From further experience, these pressure diagrams
were developed to incorporate rules for the design of walings and sheets. These are
incorporated into simple charts in the CIRIA 97 (1983) guide, which was prepared
assuming the following conditions.
g Dry conditions (assumes water is dealt with by dewatering) as water pressure has
a great effect on soil behaviour.
g Up to 6 m in depth with supports (walings and struts). Experience has shown that
in a wide range of soils the strut loads are similar up to this depth; specialist
advice must be sought and an experienced person must design the trench support
scheme for trenches of depth greater than 6 m.
155
g Short-, medium- and long-term trenches in granular soils and mixed soils but only
short–medium-term in clays. Pressure increases with time in clays, so different
parameters are used in long-term trenches in clay.
g Steel sheets with a minimum section modulus z of 35 cm3/m (or timber poling
boards of 32 mm thickness) which can be driven using non-specialist driving
techniques are assumed. There is no minimum ‘toe in’ of the sheets quoted and
the maximum cantilever of the sheeting should not exceed 500 mm.
g Timbers are a minimum grade of SC4 (C24) and steel walers are a minimum grade
S275 to BSEN 10025 (previously grade 43A). Adjustable props meet regulations
as described by BS 4074 (BSI, 2000).
g The slope across the trench does not exceed 1 : 4.
g Surcharge limited to 10 kN/m2 and deflection is not considered a major
problem.
g Not to be used in soft clays (un-drained cohesion <30 kN/m2) and saturated silt.
g There is no heave or boiling at the base of the trench.
g The supports should be installed tight against the sides of the trench to ensure soil
arching takes place and the load on the waling is relieved.
The effective depth of a trench is the actual depth, but an adjacent batter is considered to
increase the effective depth. Timber waling and trench sheets are designed for loads
equivalent to 50% of the pressure diagram to allow for arching. Steel walings, being
stiffer, are designed for the full pressure diagram.
The chart in Figure 12.6 is for use in granular soils, mixed soils and short-term trenches
in clay. It uses a rectangular pressure diagram and the pressure is calculated as:
A second case is provided for medium-term trenches in clay as the earth pressure
increases with time. Unless the water table is lowered by dewatering, then the full
hydrostatic pressure diagram should be used. The pressure due to soil and due to
surcharge are calculated as follows:
where ¼ 17.5 kN/m3 is the density of soil, ; ¼ 358 is the soil friction angle, H is effective
depth of excavation (m) and S ¼ 10 kN/m2 is surcharge. The coefficient of active earth
pressure is defined Ka ¼ 1 sin ;, where Ka ¼ 0.65 approximates maximum triangular
pressure to rectangular pressure.
156
1
2 3 20
Example 1 10
3
Load on strut: w kN/m run of waling 4
r timber
5 be
5 30 tim
150 5
6 5×
20
22 40
250 × 200
Example 2 timber
6 250 × 250
timber
30 350 × 350 50
timber
300 × 150 250 × 250
timber timber 60
40
1.0 1.5 2.0 2.5 3.0 3.5 1.0 1.5 2.0 2.5 3.0 3.5
Maximum Maximum horizontal Maximum Maximum horizontal
vertical spacing spacing of struts: m vertical spacing spacing of struts: m
of walings: m of walings: m
g lowering the water table may lead to fine soils being drawn out of the soil, leading
to settlement and possible damage to surrounding buildings, services and roads
157
Site investigation data should be studied to determine the groundwater level and if
artesian water pressure may be encountered. The permeability of the soils should also
be estimated to determine the rate at which groundwater will flow into the trench.
Trial holes can be a simple way of crudely estimating inflow rates and the stability of
soils. It should also be noted that soils tend to be laid down in layers; vertical and
horizontal permeability will therefore differ.
REFERENCES
BSI (1994) BS 8002: Code of practice for earth retaining structures. BSI, London (replaced
by BS EN 1997-1, 2004).
BSI (2000) BS 4074: Specification for steel trench struts. BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
CIRIA (Construction Industry Research & Information Association) (1983) Trenching
practice, 2nd edition. Report 97. CIRIA, London.
CIRIA (1987) Trenchless construction for underground services. Technical Note 127.
CIRIA, London.
HSE (Health & Safety Executive) (2007) The Construction (Design and Management)
Regulations. HSE, London.
Terzaghi K and Peck R (1996) Soil Mechanics in Engineering Practice. John Wiley & Sons,
London.
TRADA (Timber Research and Development Association) (1990) Timber in Excavations,
3rd edition. Thomas Telford Publishing, London.
FURTHER READING
Arcelor Mittal (2008) Piling Handbook, 8th edition. Arcelor Mittal, London.
BSI (2002) BS EN 13331-1: Trench lining systems. BSI, London.
CIRIA (Construction Industry Research & Information Association) (1986) Proprietary
trench support systems. Technical Note 95. CIRIA, London.
CIRIA (1986) Control of groundwater for temporary works. Report 113. CIRIA, London.
CIRIA (2001) Control of groundwater. Report R515. CIRIA, London.
158
159
Chapter 13
Diaphragm walls
Chris Robinson Technical Manager, Cementation Skanska Limited
Andrew Bell Chief Engineer, Cementation Skanska Limited
Diaphragm walls (also known as slurry walls) can be used to form cofferdams, quay
walls, foundation elements and embedded retaining walls (e.g. to construct tanks,
shafts, basement excavations or cut and cover tunnels). Their use is becoming more
prevalent in congested urban environments where construction below ground is
becoming ever more common; this is due to the availability of development land and
associated high land prices and to the development of deep urban infrastructure. The
economics of diaphragm wall use become increasingly attractive when they can be
used to form part of the permanent works. This chapter will describe common uses
of diaphragm walls, their method of construction (including aspects of additional
temporary works required to facilitate construction), a summary of typical construction
details which can aid construction of a high-quality finished product and a description of
the most significant design considerations.
13.1. Introduction
Diaphragm walling refers to the in situ construction of vertical walls by means of deep
trenches. Stability of the trench excavation is maintained by the use of a support fluid;
this is most commonly a bentonite suspension, although polymer support fluids can
also be used. Diaphragm walls are constructed in discrete panels typically ranging in
length from 2.8 m to 7.0 m using purpose-built grabs or milling machines. The panels
are constructed in the required sequence (discussed in Section 13.3.1) abutting each
other to form a continuous structural wall. T-shaped panels or counterforts can be
constructed where very high bending capacity elements are required, typically to
minimise propping requirements. A variety of details are available to form the ends
(stop ends) of each panel which will generally incorporate a water bar, and can
include details to facilitate the transfer of shear forces between panels.
13.2. Applications
Although an expensive form of temporary works, generally used where other forms of
temporary ground support are not practical, diaphragm walls become very economical
when they can be used to form part of the permanent works.
In addition to providing ground support and water exclusion, diaphragm walls can also
be utilised to carry vertical loads. Where the magnitude of vertical loading exceeds the
available wall capacity (based on an embedment length required to maintain wall
161
stability), the panels can be taken deeper to carry the vertical load. Discrete isolated
panels known as barrettes can be constructed to carry vertical loads alone. These
might be adopted where diaphragm walls are being constructed elsewhere on a scheme
and it would be uneconomic or impractical to mobilise additional plant to construct
conventional bearing piles, or where the founding level is so deep as to preclude these
from being adopted.
Figure 13.1 depicts the typical sequence of operations for diaphragm wall construction.
Upon construction of the first diaphragm wall panel (opening panel), further panels are
constructed to suit the specific site conditions and constraints. Panel sequencing is
typically ordered such that a minimal number of stop-ends (peel off/re-usable type
stop ends) are in the ground at any one time. This is not a constraint where non-reusable
stop ends (e.g. pre-cast concrete stop ends) are used.
As well as diaphragm walls being a temporary works solution in their own right, their
construction can require some significant additional temporary works to ensure
successful execution. These include guide walls, temporary anchorages and/or temporary
props, jet-grouted props at depth, working platforms for tracked plant (see Chapter 5),
slope stability assessment (see Chapter 10), temporary screening to protect public and
personnel, protection of water courses and adjacent services, and foundations for wet
and dry silos (for support fluid).
As noted by Fernie et al. (2001), ancillary works (broadly speaking all operations prior to
bulk basement excavation) can cause a significant proportion (up to around 80%) of
recorded settlements of adjacent structures. Caution should therefore be taken to
ensure that any precautionary measures do not cause more problems than they are
intended to mitigate.
162
163
Diaphragm walls
avoid high overbreak occurring. This would adversely affect the finished surface of the
diaphragm wall, which may require trimming back if the overbreak is in excess of the
specification limits. Suitable backfill materials often used are stiff clays or clay-bound
hogging. Properly placed and compacted, these materials will give a relatively smooth
finish to the wall over the backfill depth and potentially aid wall verticality.
Care should be taken to ensure that the guide wall itself remains stable with the
diaphragm walling plant operating in close proximity to it. This may be achieved
through provision of suitable backfill or strutting. Once the panel has been constructed
and cured for a day or so, it will provide the required support to the guide wall in lieu of
the original backfill.
It may be necessary or prudent to design the guide wall to span laterally while supporting
vertical loads, for example, trapping off the panel reinforcement, in the event that the
ground beneath it collapses during panel construction.
In principle, there is no reason why polymers cannot be used alone without bentonite
minerals. There is quite a large body of experience of polymer support for diaphragm
164
wall construction in North America and Asia, but very limited experience within the
UK.
The support fluid will require cleaning or de-sanding throughout the construction
process to ensure that the required properties are maintained.
The size of the bentonite farm (the area required for mixing, cleaning and storing the
support fluid) will depend on the scope of the works, number of rigs used, the
maximum panel volume and the nature of the ground supported.
Deep panels in fine granular materials will require far more fluid cleaning than for stiff
cohesive ground.
It should be noted that one of the major costs of diaphragm walls is the establishment
and removal of the bentonite farm. Diaphragm walling, therefore, tends to be less
economic for small areas of wall compared to alternative solutions. There are,
however, certain circumstances which will preclude the adoption of alternatives, for
example, secant piled walls. These might be factors such as the depth of excavation
required and watertightness.
165
A suitable area of the site should be identified for location of the bentonite farm (or other
support fluid mixing and storage area). The bentonite farm will occupy a relatively large
proportion of a typical site; Figure 13.2 depicts a typical set-up.
A check should be undertaken to ensure that the bearing capacity of the materials
forming the foundations for the dry and wet silos is adequate. It is common for a
reinforced concrete slab to be provided to ensure sufficient stability of the bentonite
farm components, which may require to be bolted down to withstand wind loading
when empty.
Lifting of diaphragm wall reinforcement cages will almost certainly require a tandem lift
(see Figure 13.3) to ensure that the cage is not damaged when lifting from the horizontal
to the vertical plane. This is classified as a complex lift and requires greater consideration
than more conventional lifting operations. Refer to Chapter 20 for further guidance on
lifting operations.
Where cage sections are to be spliced, the splice detail should ensure that the
reinforcement configuration is as free from bar congestion as possible. This will assist
construction of a high-quality finished product. Reinforcement bar congestion is often
the primary cause of poor-quality concrete, particularly at the near face of the panel.
The splice should ensure that operatives are not required to place their hands within
the cage where serious injury could be incurred should the cage sections move relative
to each other.
Careful checking of diaphragm wall cages is required, particularly when there are a large
number of relatively complex details, for example, box-outs for pull-out bars (such as
Kwikastip1 or Startabox1 ), starter couplers, etc. BS EN 1538 (BSI, 2010) specifies
tolerances for reinforcing elements (including couplers) both in the plane of and perpen-
dicular to the plane of wall. Failure to address these areas sufficiently can lead to support
fluid inclusions remaining within the completed panel.
13.3.7 Concreting
Diaphragm wall panels are concreted in an identical manner to bored piles constructed
under support fluid or which have groundwater in the bore. Sectional tremmie pipes
are employed which typically have a diameter of between 200 mm and 300 mm. The
diameter of tremmie pipe will depend on individual contractors’ equipment and the
depth of tremmie pipe required (i.e. the depth of panel or barrette to be concreted).
166
167
Diaphragm walls
The reinforcement cage detailing should fit the dimensions of pipe to be used and the
number of tremmie positions required to ensure a high-quality finished product. A
short panel (around 2.8–4.5 m length) would typically require a single tremmie location,
whereas longer panels (around 6.5–7.5 m length) would require two positions. Corner
panels will require particular consideration.
Concreting of diaphragm wall panels can be significantly more complex than for bearing
piles. Single diaphragm wall panels may have multiple tremmie locations for concrete
placement and, perhaps more significantly, the concreting operation can take many
hours due to the relatively large volumes of concrete required to be placed. When very
large concrete pours are required, a prudent measure is to ensure that the concrete
supplier has a standby batching plant available. To comply with the ICE specifications
(2007), tremmie pipes must be maintained at levels to ensure that they have a maximum
168
To attain the highest possible quality finished product, it is essential that panel construc-
tion is completed in the shortest practicable time without unnecessary delays.
Rope grabs are more basic than the more modern hydraulic grabs. They consist of a
short body, at the base of which is a clamshell bucket. The grab has one set of ropes
169
to carry the weight of the grab (and spoil when the clamshell bucket is full) and a second
set to close the clamshell bucket and cut the panel. The grab is generally supported by a
crawler crane. While rope grabs are old technology, they still have a place in modern
construction. For example, since a conventional crawler crane is used in conjunction
with the rope grab, the system lends itself to low-headroom scenarios where a short
jib crawler crane is required. Alternative low-headroom hydraulic systems are commer-
cially available, but these may be significantly more costly. However, as with considera-
tion of all alternative systems, the system to be adopted also needs to be compatible with
the site-specific ground and project conditions such as programme requirements, etc.
Hydraulic grabs have a much higher closing force than rope grabs which can improve
production rates or enable them to be used to excavate panels in stronger ground. The
grab can be supported on a kelly bar or, more recently, a rotator itself supported by a
modified crawler crane unit. Rotators are available for 1808 and 3608 plan movement.
The rotator allows a high degree of flexibility of panel orientation, which is particularly
170
important on tight sites and at corner panel locations. The use of a rotator in conjunction
with steering teeth on the grab clamshell bucket allows the grab to be steered according to
ground conditions to maintain the panel verticality within tight specification tolerances.
Hydraulic grabs are also fitted with electronic inclinometers which enable the grab operator
to see a graphical representation of panel geometry and verticality in real time. The data
enables the grab operator to react quickly to maintain panel verticality.
13.3.9 Hydrofraise/hydromill/trenchcutter
For diaphragm walls constructed within relatively strong strata, or for deep diaphragm
walls, the use of hydromills becomes necessary (see Figure 13.6).
A hydromill is essentially a reverse circulation trench cutter. This cuts the panel rather
than digging it (as is done with grabs). Hydromills can have counter-rotating cutting
wheels or cutting chains, depending upon the manufacturer and the ground conditions.
Deeper diaphragm wall panels require a longer cycle time to lower a grab into the panel,
take a bite from the base of the panel, be lifted from the panel and deposit the spoil.
Reverse circulation hydromills, on the other hand, continuously cut the ground which
171
then goes into suspension within the support slurry. This is circulated from the panel
through de-sanding equipment and returned into the panel. The hydromill thereby
negates the requirement for the cutting tool to be continuously lowered and lifted
from the panel. The use of hydromills is costlier than the use of grabs; however, they
will usually be employed where grabs cannot be used to progress panel excavation due
to the strength of the ground or where the greater production rates outweigh the
greater plant costs.
As stated, the adoption of particular plant or techniques needs to be compatible with the
ground conditions. The use of hydromills within cohesive materials can be problematic in
clays with a relatively high plasticity, such as London Clay. The mills can become blocked
by the clay and their efficiency dramatically reduced. One solution would be to employ a
hydraulic grab to excavate the upper section of a panel (e.g. to the base of London Clay)
and then employ a hydromill to continue panel excavation to final depth.
As with hydraulic grabs, hydromills are fitted with electronic instrumentation to enable
the panel excavation to be undertaken to achieve the required tolerances.
Stop ends most often used in modern practice take the form of either re-usable ‘peel off ’
profiled steel plate (e.g. CWS (Continuous Water Stop) stop ends; see Figure 13.7) or
precast concrete profiles.
When using a hydromill it may be possible to grind a slotted profile into the end(s) of the
adjacent panel(s) to provide both a shear key and a significantly more tortuous path for
groundwater.
If water ingress is not critical, then stop ends may be entirely omitted.
13.4. Design
13.4.1 Scope
This section on design is not intended to provide full guidance on how to design
embedded retaining walls. This subject is covered in detail within the various codes of
practice and best practice guides referenced listed at the end of this chapter and the
next. In particular, the reader is referred to CIRIA Report C580 (2003), BS 8002 (BSI,
1994) and BS EN 1997-1 (2004).
The intention for this section is to provide commentary on those aspects specific to the
design of diaphragm walls.
172
The site investigation should also provide the designer with other relevant information
which may have an impact upon the design, including
173
During construction of the embedded retaining wall, any significant variation in ground/
groundwater conditions should be communicated to the design team without delay. It is
imperative that the contracting organisation has a sufficient understanding of the design
to decide what constitutes a significant variation. The wall designer should provide a wall
manual in accordance with the ICE specifications (2007). Depending upon the design
route adopted for the contract, the intention of the ICE specifications (2007) is to
provide sufficient information to facilitate communication of pertinent design and
construction information between parties.
The design of embedded retaining walls will often involve both undrained (total stress)
and drained (effective stress) design. This of course will depend upon the geological
materials present at the particular site under consideration. Where the retaining wall is
required to provide ground support/groundwater exclusion for a longer time than
undrained behaviour can be assumed, effective stress analysis will be necessary.
Most practitioners adopt a limit state design approach when designing embedded retaining
walls. The limit states considered are ultimate limit state (ULS) and serviceability limit
state (SLS). ULS involves consideration of the safety of people and safety of the structure
itself, for example, instability of the structure, failure by rupture of the structure or any
part of it or excessive deformation of the structure such that adjacent structures reach
their ultimate limit state. SLS describes the consideration of conditions relating to the
performance of the structure under normal operating conditions, for example, deforma-
tion of the structure or deformation of the ground supported by the structure.
Two sets of analysis calculations are generally performed when considering the ULS
design of embedded retaining walls. The first ULS analysis assesses the required wall
toe depth to maintain stability for the relevant earth pressures, applied loads and
specified factors of safety (partial factors on resistances, actions, soil mobilisation,
etc.). The second set of calculations determines the reinforcement requirements based
upon the ULS bending moments and shear forces derived from the first analysis.
Project specifications (as well as design codes/guidance documents) usually require the
design of the retaining wall to allow for unplanned excavation, usually 10% of the retained
174
height up to a maximum of 0.5 m. While this can be considered to be best practice in most
circumstances, it may be possible to make some economies to the design by gaining the
agreement of all relevant parties that this design requirement be removed; sufficient
control measures must be implemented during the bulk excavation operations and this
design constraint must be clearly communicated to all parties. Most design analyses are
generally undertaken using some form of soil–structure interaction (SSI) analysis. SSI
analyses can be further divided into subgrade reaction/pseudo finite element methods and
finite element/finite difference methods. When considering SLS calculations, these preclude
the use of limit equilibrium methods and will require some form of SSI to be undertaken.
The selection of analysis method should be compatible with the problem and the avail-
able design input information. There is clearly little point opting for the more complex
SSI analysis method where there is little accurate information on design soil parameters.
Modern complex analyses can appear deceptively easy to run and can lull the uninitiated
into a false sense of security regarding the appropriateness of their analyses.
The relative size of the embedded retaining wall element is generally governed by the
magnitude of the ULS bending moment which it is required to resist. SLS considerations
are often involved in determining the temporary support requirements (e.g. propping/
waling beam stiffnesses and locations) rather than in consideration of the wall element size.
Fernie and Suckling (1996) report that embedded retaining walls in stiff UK soils will
generally exhibit excavation-induced lateral movements of around 0.15% of the retained
height. The relationship with wall stiffness is reported as being more relevant to canti-
lever and single-propped systems than for multiple-propped embedded retaining walls.
Bolton et al. (2010) further postulated that wall displacements are unavoidable regardless
of the embedded retaining wall element constructed (within the bounds of conventional
element sizes and stiffnesses). The wall displacements are controlled to a much greater
degree by the wall support system adopted (temporary propping and bottom-up
construction versus top-down construction using permanent floor slabs, and even the
construction of propping/lateral support elements by pre-excavation techniques).
175
the wall from the anchorages should be checked. In most cases, the increased vertical
load will not require any increase in wall embedment, but the design must demonstrate
sufficient vertical capacity of the wall.
Depending on the nature of the diaphragm wall, there may be an opportunity to refine
the reinforcement design according to the distribution of bending moment and shear
force magnitude with depth. Bars can potentially be curtailed or bar sizes changed
across splice zones to provide the most economic reinforcement configuration.
The detailed design of diaphragm wall reinforcement cages must consider the potential
impact of cage congestion. BS EN 1538 (2010) requires that the horizontal clear space
between single vertical bars or groups of vertical bars parallel to the face of the panel
should be at least 100 mm. For horizontal shear reinforcement, BS EN 1538 recommends
a minimum vertical spacing of 200 mm.
The detailed design of the reinforcement cages should also consider how the cages
are to be fabricated. For example, such detailed design should consider if closed or
open shear links will facilitate easier (and potentially safer) fabrication. While it may
seem that such considerations are going to the extremes of detail, they can have a
significant impact on the overall quality of the wall, particularly where the reinforce-
ment cages have a large number of complex details (coupled box outs, pull-out starter
bars, etc.).
Where water-bearing strata are retained by the embedded retaining wall, it is often a
requirement to take the toe of the retaining wall to a depth where it will provide a vertical
cut-off by effectively sealing in to a low permeability strata. This relies on the presence of
such a suitable stratum within a sensible depth of the toe level, which is designed for wall
stability.
176
Often savings through adoption of the observational approach arise from a reduced con-
struction programme through revised propping requirements and excavation sequencing.
Reference should be made to CIRIA Report 185 (1999) for comprehensive guidance on
the subject of the observational approach.
REFERENCES
Bolton M, Lam SY and Vardanega PJ (2010) Predicting and controlling ground move-
ments around deep excavations. In Mair RJ and Taylor RN (eds) Geotechnical Aspects
of Underground Construction in Soft Ground. Balkema, Rotterdam.
BSI (1994) BS 8002: Code of practice for earth-retaining structures. BSI, London.
BSI (2004) BS EN 1997-1: Eurocode 7: Geotechnical Design: Part 1 General Rules. BSI,
London.
BSI (2010) BS EN 1538: Execution of special geotechnical works – Diaphragm walls. BSI,
London.
CIRIA (Construction Industry Research & Information Association) (1999) The observa-
tional method in ground engineering: principles and applications. Report 185. CIRIA,
London.
CIRIA (2003) Embedded retaining walls – guidance for economic design. Report C580.
CIRIA, London.
Fernie R and Suckling T (1996) Simplified approach for estimating lateral wall movement
of embedded walls in UK ground. In Mair RJ and Taylor RN (eds) Geotechnical Aspects
of Underground Construction in Soft Ground. Balkema, Rotterdam.
Fernie R, Shaw SM, Dickson RA et al. (2001) Movement and deep basement provision at
Knightsbridge Crown Court, Harrods, London. Conference on response of buildings to
excavation induced ground movements, CIRIA, SP199, July 2001.
Huder H (1972) Stability of bentonite slurry trenches with some experience in Swiss prac-
tice. Proceedings of 5th European Conference on Soil Mechanics and Foundation Engineer-
ing, Madrid, 517–522.
ICE (Institution of Civil Engineers) (2007) ICE Specification for Piling and Embedded
Retaining Walls, 2nd edition. Thomas Telford, London.
177
ICE R&D (Institution of Civil Engineers Research & Development Enabling Fund) (2009)
Reducing the Risk of Leaking Substructure: A Clients’ Guide. ICE, London.
Potts DM and Burland JB (1983) A parametric study of the stability of embedded earth
retaining structures, Transport and Road Research Laboratory, Supplementary Report
813.
FURTHER READING
BSI (1986) BS 8004: Code of practice for foundations. BSI, London.
BSI (1997) BS 8110-1: Structural use of concrete – Part 1: Code of practice for design and
construction. BSI, London.
BSI (2004) BS EN 1992-1-1: Eurocode 2: Design of concrete structures. Part 1-1 General
rules and rules for buildings. BSI, London.
BSI (2009) BS 8102: Code of practice for protection of below ground structures against
water from the ground. BSI, London.
CIRIA (Construction Industry Research & Information Association) (1999) Temporary
propping of deep excavations – guidance on design. Report C517. CIRIA, London.
CIRIA (2000) Prop loads in large braced excavations. Project Report 77. CIRIA, London.
Puller MJ (1994) The watertightness of structural diaphragm walls. Proceedings of Institu-
tion of Civil Engineers 107: 47–57.
Puller MJ (2003) Deep Excavations: A Practical Manual, 2nd edition. Thomas Telford
Publishing, London.
178
Chapter 14
Contiguous and secant piled walls
Chris Robinson Technical Manager, Cementation Skanska Limited
Andrew Bell Chief Engineer, Cementation Skanska Limited
Contiguous and secant piled walls can be used to form embedded retaining walls, for
example, to construct tanks, shafts, tunnel portals, road/rail cuttings, basement excavations,
cut and cover tunnels and other forms of underground structures. They are commonly
adopted where alternative methods cannot be successfully implemented because of the
nature of the ground or performance requirements. In common with diaphragm walls,
the economics of contiguous and secant piled walls become increasingly attractive when
they form part of the permanent works. This chapter includes a description of common
uses of contiguous and secant walls and their method of construction (including aspects
of additional temporary works required to facilitate construction), a summary of typical
construction details which can aid construction of a high-quality finished product and a
description of common design approaches and design considerations.
14.1. Introduction
Contiguous and secant piled walls are two common forms of embedded retaining walls.
The most significant difference between these two forms of retaining wall is that
contiguous piles are designed to be totally independent reinforced elements with, typically,
150 mm between adjacent pile elements at commencing level, whereas secant piles are
designed to form an interlocking wall. Secant piles can therefore retain both ground and
groundwater (subject to construction tolerances as discussed in Section 14.3).
179
180
14.2. Applications
Contiguous and secant piled walls can be adopted to provide solutions to a variety of
constructed situations. They tend to be adopted where the general site conditions,
depth of excavation, ground conditions or watertightnesss requirements (note: secant
piled walls only) preclude the adoption of alternative (potentially cheaper) methods,
for example, construction in open-cut excavations, king post walls or sheet piling (see
Chapter 11).
Where particularly onerous ground conditions are present, for example, very loose fine
sands, a secant piled retaining wall may represent the preferred solution even if ground-
water is not present above excavation level. The nature of such ground may pose an
unacceptable risk of ground loss through insufficient arching capability of the ground,
which may in turn pose a risk to the stability and serviceability of adjacent structures,
services or ground to be supported.
The drawing sequence presented in Figures 14.4 and 14.5 shows the typical sequence of
site operations for contiguous pile wall and secant pile wall construction, respectively.
Pile construction needs to be sequenced to ensure that construction of any pile does
not adversely affect any previously constructed piles. For secant walls, the sequence
needs also to take account of the strength gain of the female pile concrete (female pile
concrete can potentially crack when male piles are cut into them if the concrete is too
strong, which may have a significant impact on the wall watertightness).
181
182
Courtesy of Cementation Skanska Limited
Contiguous and secant piled walls
183
Contiguous and secant piled walls
As well as being a temporary works solution in their own right, both contiguous and
secant pile wall construction can require some considerable elements of additional
temporary works for their successful execution. As described in Chapter 13 on
diaphragm walls, these include aspects such as guide walls, temporary anchorages
and/or temporary props, working platforms for tracked plant (Chapters 5 and 7), consid-
eration of slope stability (Chapter 10), temporary screening to protect public and
personnel, protection of water courses and protection of adjacent services.
Precautionary measures may be adopted such as using long temporary casings, limiting
the number of piles constructed along sensitive elevations within one shift or under-
pinning of particularly sensitive adjacent structures.
When constructing contiguous piled walls, guide walls tend to be used less often since
the verticality tolerance is usually less critical (contiguous piled walls are not designed
for groundwater exclusion and therefore do not have any watertightness performance
criteria to satisfy). Without a guide wall, a tolerance limit of 75 mm is typically
specified.
Unlike contiguous piled wall construction, secant pile wall construction will almost
invariably require a scalloped guide wall to be constructed in advance of piling
184
operations. The very nature of secant piled walls requires the piles to interlock, usually
for watertightness.
A large range of piling techniques can be adopted for construction of contiguous and
secant walls, although practical considerations may limit the use of some of these for
constructing interlocking secant walls.
Some of the above piling techniques (e.g. rotary bore piling/case and auger piling, etc.)
can be employed adopting a short length of temporary casing (although care needs to be
taken when assessing pile spacing for cased/uncased diameters) where the ground is self-
supporting. Others may require full-length casing in unstable ground or the use of a bore
support fluid, while CFA/SFA provide full bore support without the need for temporary
casings or support fluid.
The number of longitudinal reinforcing bars forming the pile cage needs to be compatible
with the pile diameter and minimum reinforcement requirements. In most instances, the
reinforcement requirements of wall piles will satisfy minimum reinforcement require-
ments without explicit consideration.
Female piles to secant piled walls are not normally reinforced. However, when designing
the wall as a hard/hard wall, the female pile will be constructed from a structural concrete
mix (with a specified characteristic strength) and will be suitably detailed to facilitate
subsequent male pile construction. The requirement for the male piles to cut into the
female pile clearly imposes constraints on the amount of reinforcement that can be
accommodated within the female pile. Notwithstanding this limitation, reinforcement
of the female pile can provide the additional wall capacity (in terms of bending
resistance), which may otherwise require larger-diameter male piles or alternative
construction methods to be adopted.
185
It is possible to vary the pile reinforcement within fabricated cages to the structural
requirements at different levels within the pile. Bars can potentially be curtailed or
changes to bar sizes made across cage splice zones. It should be noted that it is not
generally possible to economise pile cages to the same degree as within diaphragm
wall cages; this is partly due to the requirements to consider unfavourable cage orienta-
tion within the pile bore and the design of the pile as a column element.
Where cage lengths are to be spliced, the splice detail should be carefully detailed to limit
bar congestion. This will assist in maintaining a high-quality finished product. Reinfor-
cement bar congestion is often the primary cause of poor-quality concrete at the exposed
face of the pile. As with detailing of reinforcement cages for diaphragm wall panels, the
cage splices should be detailed to ensure that operatives are not required to place their
hands within the cage where serious injury could be incurred should the cage sections
move relative to each other.
It may be possible to introduce couplers and box-outs within pile reinforcement cages.
However, it is the authors’ experience that a high degree of redundancy (in terms of the
number of couplers fabricated into the reinforcement cage) is usually required to ensure
that sufficient couplers can be located when the box-out is exposed. Pile reinforcement
cages have a tendency to twist within the pile bore. This is particularly the case where
pile cages are plunged within wet concrete, but can also occur where cages are hung or
suspended within the empty pile bore with concrete subsequently being placed via
tremmie pipe. Another consideration of pile box-out design is to ensure that there is
sufficient space for concrete to flow between all the reinforcing elements and to ensure
that the box-out (typically constructed with polystyrene or Styrofoam to facilitate
exposure of the couplers) does not cause buoyancy instability of the pile cage.
14.3.7 Concreting
As described in Section 14.3.5, a variety of piling techniques can be adopted. Where piles
are constructed adopting CFA or SFA techniques, concrete will be introduced from the
toe of the pile and brought up to piling platform level. Reinforcement cages are installed
following the concreting operation.
For open bore piling techniques (e.g. large-diameter rotary bored piling, cased and auger
piling, etc.), concreting of the pile (or grouting for small diameter piles) will usually be
undertaken following installation of the reinforcement cage. For open bore techniques,
sectional tremmie pipes are generally employed which typically have a diameter of
between 200 mm and 300 mm. The diameter of tremmie pipe will depend on individual
contractors’ equipment and the depth of tremmie pipe. Consideration of the diameter
of reinforcing cage is also important in order to ensure the tremmie pipe can easily be
inserted and removed without causing damage to, or snagging on, the cage.
14.4. Design
14.4.1 Scope
This section on design is not intended to provide full guidance on how to design
embedded retaining walls. This subject is covered in detail within the various codes of
186
practice and best practice guidance listed at the end of this chapter. In particular, the
reader is referred to CIRIA Report C580 (2003), BS 8002 (1994) and BS EN 1997-1
(2004).
The intention for this section is to provide commentary on specific design aspects of
contiguous and secant piled walls. High-level guidance on the most relevant aspects to
the design of embedded retaining wall is provided in Chapter 13, Section 13.4.3.
Where the various factors are unfavourable, some form of temporary support will be
required. This may take the form of temporary berms (generally to facilitate installation
of another temporary support measure), temporary props or temporary ground
anchorages. Each of these measures will require suitable design. Should temporary
ground anchorages be adopted, this may require a wayleave if the anchorages are to
be installed beyond the site boundary. The wall will also require a design check to be
undertaken to ensure that vertical stability is maintained, since the anchorages will
impart a downward vertical load onto the embedded retaining wall.
Temporary ground anchorages and temporary props will generally require a waling
beam to span along the wall to distribute the load between wall and anchors/props.
Often an in situ capping beam will be constructed which can be used in lieu of a
waling beam. Waling beams for ground anchors can be more complex, particularly if
they are fabricated from steel. A twin parallel flange channel arrangement is often
adopted which allows the anchorage to pass between the two channel sections at the
required angle of declination (often 30–458 below horizontal). An alternative solution
for ground anchorages is to construct an in situ head block detail (see Figure 14.6)
which spans adjacent reinforced piles.
14.4.3 Instrumentation
Where wall deflections are a critical element of the design (most particularly where the
observation method is being employed), inclinometer reservation tubes can be installed
within the reinforced piles. A suitably sized thin-walled steel tube (with a sealed end) is
attached to the reinforcement cage (in sections with threaded couplings if spliced cages
are required). The reservation tube facilitates subsequent installation (grouting in) of
inclinometer tubes, which generally take the form of plastic ducts with orthogonal
guides for the inclinometer torpedo. It is usual for the inclinometer duct installation to
be undertaken outside the piling contract.
187
Female piles
Male piles
Pile concrete
broken/scabbled
back to sound concrete
Steel bearing
plate
2 No. dowels
Other forms of instrumentation can be installed within wall piles, generally fixed to
reinforcement cages (e.g. vibrating wire strain gauges or fibre-optic strain gauges to
monitor the state of stress within wall piles). Again this is most commonly associated
with the observational method, for example, when assessing whether additional
temporary support measures may be omitted. For full details on the observational
method, refer to CIRIA Report 185 (1999).
Care needs to be taken to ensure that the reinforcement design is compatible with the
method of pile construction proposed. For example, plunging cages within CFA/SFA
piles has practical limitations on the depth attainable for a given cage weight and rigidity.
A maximum practical reinforcement cage plunge depth of around 15 m is often adopted.
Although it is possible to adequately reinforce these pile types to deeper levels, specific
guidance from specialist piling contractors should be sought.
There may be opportunity to refine the reinforcement design according to the distribu-
tion of bending moment and shear force magnitude with depth. Bars can potentially
188
be curtailed or bar sizes changed across splice zones to provide the most economic
configuration.
The reinforcement cage design should take into account any requirements for handling/
lifting the cages to ensure that they are sufficiently robust. This may require the inclusion
of suitable lifting bands, etc. Where cages are required to be spliced, sufficient attention
to detail should be given to the method used to form the splice such that operatives are
not required to put their hands within the cage during the splicing operation.
While it is usual to maintain male/female pile interlock to below dredge level, there may
be opportunity in certain circumstances to found the female piles above dredge level if
a suitable low-permeability stratum is present. It is important to clearly communicate
any risks that may be present in such an arrangement, and have suitable contractual
arrangement in place covering risks associated with watertightness and the costs of
any potential remedial works.
Where water-bearing strata are retained by the secant wall, it is often a requirement to
take the toe of the retaining wall to a depth where it will provide a vertical cut-off by
effectively sealing in to a low-permeability stratum. This relies on the presence of such
a suitable stratum within a sensible depth of the toe level which is designed for wall
stability. Additionally, a check should be undertaken to ascertain the level to which
male/female pile interlock can be maintained.
189
unreinforced, subject to the net vertical loading being compressive in nature. Clearly the
additional depth would need to be suitably reinforced to carry any tensile loads (e.g.
hydrostatic uplift forces). When designing embedded retaining walls to withstand
vertical loads, it is usual to consider only that part of the wall below the deepest
excavation level (i.e. the fully embedded length of the wall).
Where temporary lateral support to the wall is provided through ground anchorages, the
vertical stability of the wall should be checked due to the increased vertical loading on the
wall from the anchorages.
Where cohesive materials are present at dredge level, consideration should be given to
the potential for these materials to degrade and soften, usually just within the top 1 m,
thus potentially adversely affecting the passive resistance afforded to the retaining
wall.
As well as considering the required wall toe level to maintain lateral and vertical stability,
certain ground conditions may require the wall toe to be taken deeper by consideration
of more global factors. One such scenario is where an embedded retaining wall is to be
constructed within deep soft cohesive deposits. The potential for basal failure of the
excavation, whereby the soft cohesive materials affecting flow around the toe of the
wall should be assessed in such situation. Piping failure may need to be avoided and
potential slip planes may need to be intercepted (see Chapter 10).
REFERENCES
BSI (1994) BS 8002: Code of practice for earth-retaining structures. BSI, London.
BSI (2004) BS EN 1997-1: Eurocode 7: Geotechnical Design. Part 1 General Rules. BSI,
London.
BSI (2009) BS 8102: 2009 Code of practice for protection of below ground structures
against water from the ground. BSI, London.
BSI (2010) BS EN 1536: Execution of special geotechnical works – Bored piles. BSI,
London.
CIRIA (Construction Industry Research & Information Association) (1999) The observa-
tional method in ground engineering: principles and applications. CIRIA Report 185.
CIRIA, London.
CIRIA (2003) Embedded retaining walls – guidance for economic design. CIRIA Report
C580. CIRIA. London.
190
FURTHER READING
BSI (1986) BS 8004: Code of practice for foundations. BSI, London.
BSI (1997) 8110-1: Structural use of concrete. Part 1: Code of practice for design and
construction. BSI, London.
BSI (2004) BS EN 1992-1-1: Eurocode 2: Design of concrete structures. Part 1-1 General
rules and rules for buildings. BSI, London.
CIRIA (Construction Industry Research & Information Association) (1999) Temporary
propping of deep excavations – guidance on design. CIRIA C517. CIRIA, London.
CIRIA (2000) Prop loads in large braced excavations. Project Report 77. CIRIA, London.
ICE (Institution of Civil Engineers) (2007) ICE Specification for Piling and Embedded
Retaining Walls, 2nd edition. Thomas Telford, London.
ICE R&D (Institution of Civil Engineers Research & Development Enabling Fund) (2009)
Reducing the Risk of Leaking Substructure: A Clients’ Guide. ICE, London.
Puller MJ (2003) Deep Excavations: A Practical Manual, 2nd edition. Thomas Telford
Publishing, London.
191
Chapter 15
Caissons and shafts
Andrew Smith Contracts Director, Joseph Gallagher Ltd
This chapter deals with the construction of shafts and caissons using pre-cast
segmental linings together with the use of sprayed concrete lining (SCL) in shaft
sinking operations.
Caissons in the marine environment are not covered in this chapter. Construction
methods used including types of shaft linings generally available, other materials and
construction plant commonly used together with a brief description of specialist
processes that can be used to provide ground stability when required are all discussed.
Comments on design considerations for the construction of shafts and caissons using
pre-cast segments are also included.
15.1. Introduction
Shaft sinking in the UK is generally carried out using circular pre-cast segmental linings,
originally developed after the war as an alternative to the original and more expensive
cast-iron linings. They are typically used to provide access for tunnelling operations
where they can then be converted to permanent access chambers. More recently, they
are increasingly being used to form storage chambers and pumping stations, etc.
where they can offer a cost-effective solution to more expensive alternatives such as in
situ construction within a piled cofferdam. The advantages are that the permanent
works materials are, in effect, used as the temporary works during construction and
also that the construction ‘footprint’ is kept to a minimum (an important consideration
in urban areas).
Specification clauses for shaft construction and break-outs from shafts can be found in
the British Tunnelling Society (BTS) Specification for Tunnelling (2010).
For small-diameter shafts in the range up to 4 m, full-circle segmental rings are available
but high unit weights need to be considered when specifying the construction plant
required.
193
exposed ground is supported with the minimum of delay and there are no large
temporary working spaces to be backfilled on completion.
Pre-cast shaft linings were originally designed to mirror the earlier cast-iron linings so
that they had a rib and recessed panel appearance. As such, they had a limited use in
the context of depths exceeding around 20 m and are no longer readily available.
These linings have now been replaced with solid units, normally 1000 mm wide, in
standard diameter ranges from 4 m to 25 m with different manufacturers having their
own bolting systems. The individual segments are cast to very exacting tolerances to
ensure accurate alignment between components. The design of these units normally
incorporates a sealing system utilising hydrophilic strips or rubber gaskets, the latter
being factory fitted. Increasingly, these segments are being manufactured using fibre
reinforcement which has a beneficial effect in terms of fire resistance and also makes it
easier to break-out openings and attach fixings. Although the majority of manufacturers
produce a standard range of products to a set design suitable for the vast majority of
projects, they will also provide a design service and special manufacture of linings to suit
more demanding or specific design conditions. Additionally, as part of their service, manu-
facturers are able to provide specialist items such as corbel units to accommodate landing
slabs and also complete roof slabs designed to the engineer’s loading requirements.
A more recent development, often used in conjunction with segmental shaft construc-
tion, is the use of sprayed concrete lining (SCL). SCL is used to replace the segmental
lining and can be used in conjunction with reinforcement and/or lattice arches or, alter-
natively (and now more commonly), with fibre reinforcement. Typically, segmental ring
construction might be used for the upper levels of the shaft where ground conditions and
surface features make this appropriate with a change to SCL in the underlying more
stable strata. For permanent works, use of a secondary in situ lining may be required
to create a smooth finish. SCL can be used in conjunction with either sprayed-on or
sheet-membrane waterproofing. One considerable advantage with SCL is that openings,
etc. required in the shaft lining can be formed by the use of additional framing reinforce-
ment but without the need to introduce expensive temporary works support while the
portal structures are constructed. On the other hand, openings in shafts built with
segmental rings generally require considerable temporary support during construction
(particularly at deep levels) as the inherent compressive strength of a shaft ring is lost
once an opening is formed in it, until the permanent works are completed.
194
15.3.1 Underpinning
Underpinning involves the excavation and erection of each ring of the segmental lining
beneath the previously constructed ring. As each ring is completed cementitious material
is injected behind the lining to fill any voids and secure it in the ground, ready to support
the next ring which is bolted up from beneath. Different manufacturers have their own
bolting systems to do this. This method is normally used in firm self-supporting ground or
where ground treatment processes have created stable ground conditions. However, it can
also be used to recover a situation where a shaft being sunk as a caisson has become stuck
although additional ground stabilisation processes will probably be required in this situation.
The initial segmental ring is placed in a pre-dug excavation and keyed in to a concrete
collar cast around it, normally by inserting dowels through the grout holes. It is vital
during this process that this initial ring is built within the correct tolerance and fully
supported, as any settlement during the casting of the collar could have serious repercus-
sions in keeping the rest of the shaft vertically aligned. Once the first ring of the shaft is
fixed in the ground, it is common to fix plumbing brackets around the top of the ring to
check verticality as the shaft is sunk.
The grouting process requires the base of each ring to be sealed. There are geotextile
hoses available that can be fixed behind the ring before it is built; after building, the
hose is inflated with grout to seal the annulus before void grouting begins. It is
however more common to push excavated material in under the ring once built to
achieve the same result: so-called ‘fluffing up’. When using this method, care must be
taken not to damage the seals. If an excavator is being used it is possible to fit a
purpose-made blade for this purpose. It is also good practice to form small voids in
the previously grouted annulus up to the grout holes of the ring above to release any
trapped air as grouting takes place.
Traditionally carried out by hand, adherence to current HAVS (hand arm vibration
syndrome) regulations can make this a time-consuming process. One method of over-
coming this is to reverse the bucket on the shaft excavator so that it can dig upwards to
assist the trimming process (Figure 15.1). Segments are usually placed by crane using
specially manufactured underpinning frames, supplied by the segment manufacturers.
Grouting normally uses bagged cementitious material supplied shrink-wrapped and on
pallets for weather protection and ease of handling by forklift. It can be mixed and
pumped in special composite units driven by compressed air. The segment manufacturers
generally provide threaded grout sockets in their segments, and it is important to check
that the grout gun nozzle is compatible with the fittings supplied.
195
As well as underpinning using segments, the same basic process can be used with SCL
methods (Figure 15.2). Once the shaft has been excavated for the pre-determined
length, the SCL is applied using either a hand-held nozzle or a robot sprayer. For
most operations, this material is supplied ready-mixed and either discharged directly
into the pump or held in a re-mixer on site. As with most operations involving SCL,
the material is supplied retarded and an accelerator is added at the nozzle. Reinforcement
can be provided in the form of mesh or pre-fabricated arches, but it is becoming increas-
ingly common to use fibre-reinforced concrete which speeds up the process considerably.
Openings typically use steel reinforcement locally, and can be formed incrementally as
the excavation proceeds without the need for temporary support. If required, sprayed
waterproof membranes can be incorporated into the lining, normally by sandwiching
them between two separate layers of SCL.
Caisson sinking typically involves constructing the first one or two rings of the shaft at
ground level within a substantial reinforced concrete collar using a special cutting ring at
the leading edge. As with underpinning, it is vital that the initial rings are built accurately
196
and held in position while the collar is concreted. These rings are surrounded by poly-
styrene sheets before concreting the collar to create a sleeve through which the shaft
can slide. Sacrificial jacking bases are also positioned before concreting within the
collar onto which the shaft jacks are then fixed. For shaft sizes up to around 10 m
diameter, most segment suppliers manufacture their own pre-cast cutting edges. Over
this size, it is necessary to use a fabricated steel unit which must be designed to suit
the sizes and fixing patterns of the rings to be used. For larger diameters and demanding
ground conditions, it is essential to have a steel unit which can be welded on site to
increase rigidity and prevent shaft distortion during sinking.
The cutting edge (Figure 15.3) must provide an overcut to the rings to be used so that an
annulus is formed as the shaft sinks, enabling a lubricant to be introduced. This annulus
is typically of the order 50 mm. There are a number of products on the market suitable
for this operation. The caisson is sunk by excavating from within and then letting the
shaft sink in a controlled manner, almost always by the use of vertical hydraulic jacks
positioned around the collar. The size and hence weight of this collar must be sufficient
to counteract the anticipated jacking loads required. As the shaft sinks further, rings are
added at the surface with specially designed working cages needed for this operation
(Figure 15.4).
The annulus created by the cutting edge is kept filled with a thixotropic material such
as Bentonite or one of a range of synthetic products currently available to support the
197
Figure 15.3 Steel cutting edge (during trail erection) ready to receive pre-cast units
Courtesy of PL Manufacturing Ltd
198
Once a caisson becomes badly out of alignment the consequences can be severe, including
getting it stuck and/or segment damage. In this regard, careful attention should be paid
to the lubrication process, particularly where there is a risk of ground coming onto the
caisson. In addition, a careful analysis of the ground conditions should include a deter-
mination of the likelihood of large obstructions such as boulders blocking the cutting
edge. Caissons have far less danger of becoming stuck in fine-grained homogeneous
soils than in, say, gravels or boulder clay, where alternative methods might be more
appropriate.
Caisson excavation can be carried out ‘dry’ or ‘wet’ depending on ground conditions. If
the ground is naturally stable or has been stabilised by, for example, dewatering (see
Chapter 6), excavation can be carried out from the surface or from within the shaft
using excavation plant described for underpinning above. If the conditions are unstable
and/or waterlogged, or where the hydrostatic conditions could cause the base of the
excavation to ‘blow’, excavation must be carried out with the shaft flooded to the
prevailing hydrostatic level. In these circumstances the excavation plant normally used
is either an excavator-mounted pole grab, where special telescopic models can reach
depths of around 20 m (see Figure 15.5), or a rope-operated digging grab mounted on
a crawler crane. With the latter, the digging ability is governed by the hardness of the
material and the submerged weight of the grab. In hard material it is possible to add
weights to the grab; if this is not successful, measures such as pre-auguring or the use
of chisels suspended from the shaft crane must be considered.
It is becoming more common to use caisson sinking, even in stable ground, because the
method eliminates the need for the trimming process required when underpinning. The
method also minimises the need for personnel to be in the shaft, as the ring building takes
place at the surface.
The same plant is used for mixing and pumping the lubricant as for the final grouting,
and is described in Section 15.3.1 on underpinning.
With wet caissons it is normally necessary to seal the base with the shaft submerged. This
is because dewatering, once the shaft has reached its depth, might cause the base to heave
or ‘blow’ under hydrostatic pressure. Even if dewatering is a possibility, sealing the base
in wet ground conditions can be extremely difficult. The depth of the so-called concrete
‘plug’ must be sufficient to provide enough resistance to the hydrostatic uplift in conjunc-
tion with the weight of the shaft rings and the weight of the collar. The latter is normally
attached to the shaft, once sunk to its final position, by fixing dowel bars through the top
rings into the collar designed to provide the shear resistance required. The concrete plug
is placed by tremmie methods, almost always using concrete pumps. To provide a key it
199
is usual to install recessed panel rings in the plug location or, alternatively, corbel rings.
Segment manufacturers usually supply these as part of their shaft segment range. The
plug must be left in place for a minimum of 5 days to cure before dewatering begins.
Preparation of the surface can then commence, usually by placing a regulating blinding,
to allow construction of the structural base above.
Where the base of the shaft is founded in stable ground or it has been rendered tem-
porarily stable by dewatering/pumping or another ground stabilisation process, as an
alternative to a deep plug it is possible to provide uplift resistance by under-reaming.
The shaft is first stabilised by normal annulus grouting and the cutting edge is usually
removed, which in the case of a steel fabricated unit can be re-used. The base excavation
is then under-reamed, using temporary supports if required, to extend it beyond the shaft
footprint. Once the reinforced concrete base is cast, this mobilises passive resistance of
the undisturbed ground above the toe to counteract uplift.
There are a number of ground stabilisation processes that can be used to aid shaft sinking
and to reduce construction risks; these are discussed in detail in Chapters 6, 8 and 9. It is
200
worth bearing in mind that if the shaft construction involves excavating, moving and
disposing of large amounts of saturated material, particularly in urban areas, it may
be prudent to consider ground stabilisation on environmental grounds to lessen the
impact (see Chapter 7). Likewise, if the shaft is to be sunk using sump pumping to
control groundwater, the issues of silt separation and discharge facilities should be
seriously considered; very exacting standards are normally demanded from licensing
authorities before such discharges can be accepted into surface water disposal systems.
If deep well dewatering is being considered, there are issues to be addressed with
regard to abstraction and discharge licenses.
The use of such processes needs to considered and decided upon at the construction
planning stage as installation is more difficult to achieve once construction has
started, likely to be less effective and can be very disruptive and costly.
The most important factor which influences shaft design is the type of ground and
whether it is unstable or competent and self-standing when excavated. The presence of
groundwater exacerbates the unstable ground conditions and imposes hydrostatic
pressures which increase linearly with depth in both unstable and competent strata.
Once ground conditions are known from site investigations, the basic parameters for
excavation and muck removal, groundwater control, ground stability control and
lining installation can be evaluated and the shaft lining type chosen. Calculation of
active pressures from the ground on to the walls is covered in Chapter 11. These
figures are also used to give upward pressures to check the stability of the base in the
temporary as well as the permanent condition. Note that it is the situation before the
base is installed that is the critical condition, and that particular dangers are created
by the hydrostatic forces.
201
It is always prudent to design out any plan shape with intrusions or a non-uniformcross-
section, not only for the high stresses at angles but because of the danger of such shapes
becoming stuck when being sunk as a caisson.
REFERENCES
BTS (British Tunnelling Society) (2004) Tunnel Lining Design Guide. Thomas Telford
Publishing, London.
BTS (2010) Specification for Tunnelling, 3rd edition. Thomas Telford Publishing, London.
202
Chapter 16
Bearing piles
John Hislam Director, Applied Geotechnical Engineering Ltd
A review of differing piling methodology and design principles is presented for short-
term temporary works bearing piles. Depending upon the nature of the works, it is
quite normal to adopt lower load factors compared to permanent works since the
duration of use is short and activity adjacent to the installations decreases the risks
associated with non-observation of possible problems. It is accepted practice to adopt
the observational method as a means of realising economic solutions; an intrinsic
requirement of such methodology is to ensure contingency plans are assessed and
prepared.
16.1. Introduction
In temporary works, bearing piles are used for a variety of purposes such as: test pile
loading reaction systems, jetties, plant working platforms, crane bases, façade retention,
supports for site access/accommodation/existing structures, etc. (see Chapters 4, 17 and
26). In extreme cases, they can be utilised for the provision of general site working
conditions in very weak or soft ground conditions. As well as environmental considera-
tions, very often the type of pile adopted is an adjunct to main works piling as lead-in
times and mobilisation/demobilisation costs and time are saved. Both driven and
bored piles are used; driven piling provides the advantage of instant load-carrying
capability outweighing the usually more acceptable environmental advantages of
bored piles. While timber piles can be included here, it is more usual that steel or concrete
piles are adopted (dependent upon material availability).
Crane base piles are more usually of the bored pile type as environmental considerations
associated with urban development tend to rule out noise and vibration issues associated
with driven piles. In-service and out-of-service crane loadings have to be considered for
any pile layout group in order to determine the worst case design. Pile cap self-weights
203
have to be included within such considerations, although it is quite normal to neglect the
pile self-weight.
In many cases temporary piles are required to provide support or access for permanent
works. The ‘heavy’ engineering cases are usually associated with marine works, where
either a temporary jetty or pile guidance system is required. In these cases, it is more
normal to adopt driven piles, either of H- or circular-section. Modern pile-driving
equipment is readily able to pitch and drive a variety of steel sections to reasonably
precise tolerances, usually with the ability to predict load-carrying capacity via wave
equation analysis.
Other installation methods such as vibratory are more usually used for sheet piling
which, if being used elsewhere on the site, could be also used for bearing loads.
Ground-borne vibration, from either driven or vibratory methods, is generally less of
a problem than it is conceived to be. Relatively light hand-held equipment including
adapted breakers and impact moles are used to install scaffold pole-size steel piles in
difficult access conditions, particularly useful for the provision of access/working
platforms and safety fences on embankments and cuttings. Screw piles, installed by
rotation, are very useful for lighter loads for similar duties as above, together with
falsework support. As driven piles, their other advantages are negligible spoil removal
and the ability to be instantly loaded.
In many cases, temporary works for façade restraint or the support of existing structures
while manifest changes are made require relatively small equipment as either working
area and/or headroom can be restricted. This has possibly been one of the largest
growth areas in piling technology over recent years, only matched by similar strides in
continuous flight auger piling. There is now a wide variety of specialist piling equipment
suitable for pile diameters ranging from around 150 mm to over 500 mm that can work in
such restricted areas. The health and safety requirements associated with such equipment
must be observed, as these require the use of guarded rotary machinery.
In all cases, any piling methods which are specified for permanent works should be
considered and the benefits of avoiding duplicating establishment charges, etc. compared
against the greater efficiency of using the technically best method for the temporary
works. An example of this could be on a project where diaphragm walling has been
specified, where a choice might be made to use barrettes for load-bearing (a method
not normally considered unless large numbers could justify the high set-up costs; see
Chapter 13).
204
The design of driven piles is usually based upon the use of the empirical formulae (e.g. the
Hiley formula), but there are many alternatives and derivatives. It is suggested that the
reader reviews explanations of this type of predictive methodology (see Further reading
list at the end of the chapter); note however that the technology is based upon large-
diameter piling and could therefore be misleading for smaller-diameter piles. It is best
to seek advice from specific practitioners who maintain databases of case histories
upon which they can judge more accurate predictions. As a fundamental point, it is
not unreasonable to require statically installed pile design theory to be used as a basis
for driven piles.
The design of bored piles follows fairly well-understood principles that rely on the calcu-
lation, or estimation, of the combination of shaft friction and end bearing. In clays the
shaft friction can be derived from the product of undrained shear strength and pile shaft
surface area, the undrained shear strength being empirically factored to realise a match
with many years of practical feedback from pile testing. The base resistance is derived
from the product of a bearing capacity factor (for isolated piles, usually 9), the pile
base area and the undrained shear strength at the pile toe.
In granular materials the shaft friction is derived from a factored product of the average
effective overburden pressure over the shaft length and the material’s characteristic or
average angle of friction between the pile shaft and soil. The factors that are used are
essentially based upon the method of pile installation; driven piles have larger values
than bored piles. In rocks (particularly weak rocks), much empirical work has been
carried out as reviewed, for example, by Tomlinson and Woodward (2008) and specia-
lists should be consulted in order to assist with such design. Chalk is somewhat of a
special case in that pile design has become accepted as being dependent upon the factored
effective overburden pressure rather than standard penetration test empirical data
(CIRIA, 2002 and 2003).
Where end-bearing can be located on rock, the Arcelor Piling Handbook (2010;
Chapter 10) provides some useful capacities.
205
Figure 16.1 Cross-bracing of piles for lateral loading and effective length
The adoption of lower design load factors of less than 2 can lead to the requirement for
verification of such designs, especially in cases where such piling works may not be in
areas where adequate ground condition information is fully available. Static and
dynamic load testing can be used, although it is preferable to have correlation data
available for site conditions in order to add confidence to the dynamic results.
16.3.3 Loadings
Lateral loading conditions can arise from impact loading from either berthing vessels or
operational impacts in the case of marine works or land-based construction plant. In
addition, wind and/or water flow (possibly with floating debris impact or damming load-
ings) must be included in possible load cases. These loadings are described in greater
detail for jetties and load platforms in Chapter 17 and for crane bases in Chapter 4.
Guidance can also be found in BS 5975 (2011), pertinent to all temporary works.
Where vertical loading from deck beams or walers on piling frames are taken on to the
sides of the piles (e.g. by corbels), eccentric loading will be induced. The eccentric effect
should be treated as another form of horizontal loading for its effect on the pile.
206
REFERENCES
ArcelorMittal (2008) Piling Handbook, 8th edition reprint. ArcelorMittal, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
CIRIA (Construction Industry Research & Information Association) (2002) Engineering in
chalk. Report C574. CIRIA, London.
CIRIA (2003) Shaft friction of CFA piles in chalk. Project Report 86. CIRIA, London.
SCI (Steel Construction Institute) (1989) Steel Bearing Piles. P156. SCI, Ascot.
Tomlinson MJ and Woodward J (2008) Pile Design and Construction Practice, 5th Edition.
Spon, London.
FURTHER READING
AGS (Association of Geotechnical and Geoenvironmental Engineers) (2006) Guidelines for
Good Practice in Site Investigation. AGS, Kent.
BSI (2000) BS EN206-1: Part 1 Specification, performance, production and conformity.
BSI, London.
BSI (2000) BS EN 1536: Execution of special geotechnical work. Bored piling. BSI,
London.
BSI (2001) BS EN 12699: Execution of special geotechnical work – Displacement piles.
BSI, London.
BSI (2004) BS EN1992-1-1: Eurocode 2: Design of concrete structures. Part1-1: General
rules and rules for buildings. BSI, London.
BSI (2007) BS EN 12794: Precast concrete products – Foundation piles. BSI, London.
BSI (2007) EN1993-5: Eurocode 3 Design of steel structures. Part 5: Piling. BSI, London.
BSI (2009) BS5228-1: Code of practice for noise and vibration control on construction and
open sites. Part 1: Noise. BSI, London.
BSI (2009) BS5228-2: Code of practice for noise and vibration control on construction and
open sites. Part 2: Vibration. BSI, London.
Fleming WGK, Randolph M, Weltman A and Elson K (2009) Piling Engineering, 3rd
edition. Taylor & Francis, London.
FPS (Federation of Piling Specialists) (2006) Handbook on Pile Testing. FPS, Kent.
Healy PR and Weltman AJ (1980) Survey of problems associated with the installation of
displacement piles. Report PG8 1980. CIRIA, London.
ICE (Institute of Civil Engineers) (2007) Specification for Piling and Embedded Retaining
Walls. Thomas Telford Publishing, London.
Turner MJ (1997) Integrity Testing in Piling Practice. Report 144. CIRIA, London.
207
Chapter 17
Jetties and plant platforms
Paul Boddy Head of Temporary Works, Interserve Construction Limited
This chapter explores the options available to a site and designer when a jetty or platform
is required. It highlights the materials and design approaches possible and the
importance of obtaining good information about the expected loadings. The chapter
also highlights the importance of good communication between the site team and the
designer.
17.1. Introduction
In marine and river temporary works terms, a plant platform is a structure constructed at
the edge of a watercourse to allow the loading and off-loading of materials, plant and
personnel from a vessel. A plant platform would be necessary where the existing bank
is not considered strong enough to withstand the plant loadings.
If the river bed profile is such that a barge cannot approach the bank, then the plant
platform may need to extend out further to where the river depths are greater. If this
is the case, the structure becomes a jetty (see Figure 17.1).
Plant platforms are also used to gain high-level access in dry sites, for example, at street
level for craneage and trucks to stand above a building basement under construction. In
structural terms, these can equate to jetties but without being subjected to the mooring
and environmental loadings as described in Sections 17.5.2.3 and 17.5.3.
Three types of structure could be considered when specifying a plant platform or jetty:
solid structure, open jetty structure and floating jetties, as described in the following
sections.
209
Figure 17.1 Situations for jetty and plant platform: (a) barge can approach the bank, but the bank
is weak; (b) plant platform installed; (c) barge cannot approach river bank; and (d) jetty installed
(a)
(b)
(c)
(d)
shallow. Ensuring that the entire structure is removed when the works are completed can
also be difficult.
The stone could be installed by floating plant such as a pontoon and excavator working
with a hopper. To ensure the stone structure meshes together, the machine driver would
need to choose the stone sizes to ensure they interlocked as far as practically possible.
Issues for access would have to be considered, as described in Chapter 18.
210
Existing riverbank
(a)
(b)
In both cases, ties and walers should be provided to prevent lateral movement. For long
sheet-piled jetties, it is prudent to form compartments with regular sheet-pile cross-walls.
The decking design will depend on the use of the jetty and therefore what loading it will
need to resist. Steel plate is a durable option but can become slippery when wet; a timber
decking (timber sleepers or ekki mats) may be more appropriate. As the structure is
temporary, the main structure will almost certainly be in steel as concrete will require
time to cure and be difficult to remove afterwards. Proprietary systems such as steel
211
soldiers may also be considered, as they can be hired and returned when the job is
completed.
The span sections will almost certainly be steel and can either be a beam/column section
or a truss section. Beam/column sections are more economical for short spans, but when
they increase then a truss section may be more feasible (Figure 17.4).
As spans increase then the beam/column-section size will increase considerably to resist
not only static moments but also deflections due to dynamic loading. Deflections under
loading need to be controlled to ensure that the platform does not become unstable.
Lateral stability also becomes an issue at long spans and bracing will therefore be neces-
sary. For shorter spans, column sections provide better resistance against torsion and
impact (although structurally less efficient than beam sections).
The pier sections will be chosen for simplicity of installation and so will be circular H or
circular sections which can be installed by plant and removed afterwards. To resist lateral
forces, raking piles should also be considered as they require smaller sections than if the
vertical piles have to resist all lateral loads (Figure 17.5). For information on installation
and design of temporary bearing piles, refer to Chapter 16.
212
Guide pile
Water range
The bollard and connection should be designed to prevent it breaking off under the force
from a moving boat. The designer may also wish to consider the design case where the
vessel has set sail, forgetting to untether.
17.4.2 Connections
All sections for a working platform or jetty will typically be large, heavy and generally
difficult to connect, particularly if being installed by a crane on a pontoon. All
213
connections should therefore be large, simple and with plenty of redundancy should it be
found that not all the designed connections can be installed. If welding is to be consid-
ered, then ensure that the specified weld is larger in length and size to that required to
allow for difficulties in maintaining weld consistency. This is particularly the case if
welding is required underwater by divers, who will be relying on touch and feel to under-
take their works (assume a throat strength of 110 N/mm2). It may also be prudent to
provide lugs to hold two sections together while the connection is being undertaken
(Figure 17.8).
17.4.3 Consents
Any works being undertaken in or around a watercourse in the UK require Temporary
Works Consent from the Environment Agency (EA), who may issue conditions on the
use and design of a loading platform or jetty. Such conditions will be issued to ensure
that any pollution risk is kept to a minimum and that the structure does not impede
river flows, particularly in flood conditions.
A prudent temporary works designer will therefore ask for a sketch from the site team
with initial sizes based on what is available and where connections are required. The
designer can then back-calculate to confirm with the site team about whether the sections
work and, if not, enquire as to what else is available.
17.5. Loadings
When designing a loading platform or jetty, the loads to be imposed on it must be
considered. At the beginning of the design process the designer and the site supervisor
214
should agree what plant and materials are to be transported over the structure in order to
determine the possible loadings.
g self weight
g imposed load from plant
g environmental loads (wind and water)
17.5.2 Plant
17.5.2.1 Tracked plant
Tracked plant typically includes excavators, crawler cranes or piling rigs. The loading
will be from the tracks. The loading from excavators and crawler cranes clearly
varies depending on working radius and load being lifted. This will typically induce a
trapezoidal load distribution under the tracks. The designer may then wish to convert
this load using a rectangular or Meyerhof distribution (Figure 17.9).
For a crawler crane, the pressure diagrams should be available from the crane supplier
and will be based on crane type and counterweight, boom length and maximum load
at working radius.
For the excavator this information may not be as readily available; some judgement
therefore may be necessary to determine the track loads. The designer should consider
the three load cases of working over front, over side and diagonally. Plant may
need to be controlled in its operation to ensure that it does not overload the platform
or jetty.
215
A piling rig will also induce track loads onto the jetty. A Federation of Piling Specialists
(FPS) contractor will provide a pressure diagram similar to that of a crawler crane
(e.g. Chapter 5). The designer must review the criteria on which the pressure diagram
is based to ensure it is as required, in particular that dynamic load is factored up.
17.5.2.3 Craft
The jetty should be designed for a nominal impact loading from craft coming alongside
the structure and should allow for some accidental loadings; refer to BS 6349-1
(BSI, 2000) for assistance. Fendering may need to be fitted to the structure or, if the
loadings are too high, it will need to be free-standing.
q ¼ 500 v2.
An allowance should be made for the shape of the structure and a cw value given
accordingly to derive the force on the structure. Further reference can be sought from
BS 6349-1 (BSI, 2000), which also provides guidelines on wave actions.
Where flood conditions are possible, consideration should be given to the loading from
the damming effect of trapped debris against the structure. If flow is restricted, then this
loading will be considerable.
17.6. Analysis
The analysis of platforms and jetties can be conducted using traditional methods of
analysis; only specific aspects are covered here. Once the forces are known and the
216
materials chosen, then the design is generally straightforward. The designer must
however consider the factors of safety being applied to the loadings and be satisfied
that an allowance for dynamic loading has been taken into account, particularly with
excavators or wheel plant that may need to brake when on the structure.
REFERENCES
BSI (1997) BS 6399: Loadings for buildings. Part 2: Code of practice for wind loads (incor-
porating Amendment No. 1 and Corrigendum No. 1 2002). BSI, London.
BSI (2000) BS 6349: Code of practice for general criteria (including amendment 1 2005).
Part 1-Maritime structures. BSI, London.
BSI (2010) BS EN 1991-1-4: Eurocode 1. Actions on structures. Part 1–4: General Actions
Wind Actions (incorporating corrigendum A1, January 2010). BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures and
the permissible stress design of falsework (incorporating corrigendum No. 1). BSI, London.
Gaba AR, Simpson B, Powrie W and Beadman DR (2003) Embedded retaining walls –
guidance for economic design. CIRIA Guide C580. CIRIA, London.
HA (Highways Agency) (2001) The Assessment of highway bridges and structures (includes
correction dated August 2001). Report BD21/01. HMSO, London.
FURTHER READING
Blake LS (2004) Civil Engineers Reference Book, 4th edition. Elsevier Butterworth-
Heinemann, Oxford.
Ehrlich LA (1982) Breakwaters, Jetties and Groynes: A Design Guide (Coastal structures
handbook series). Sea Grant Institute, New York.
Elson WK (1984) Design of Laterally Loaded Piles. CIRIA Report 103. CIRIA, London.
Williams BP and Waite D (1993) The Design and Construction of Sheet-piled Cofferdams.
CIRIA Special Publication 95. CIRIA, London.
217
Chapter 18
Floating plant
Paul Boddy Head of Temporary Works, Interserve Construction Limited
This chapter explores the range of equipment types available when it is necessary to go
afloat. It highlights the importance of good planning, both in terms of using the plant
and also ensuring the vessel can actually be used at the location. This chapter gives
advice on the necessary calculations and checks that need to be undertaken for a given
scenario, and stresses the importance of these checks being carried out by an engineer
or naval architect who has both technical competence and practical experience.
18.1. Introduction
There are many situations in the field of construction where conventional plant cannot be
used, and to undertake works, the contractor may need to go afloat. This would typically
be for works being undertaken either at the edge of a water course or actually in it.
g The works involved are too remote from the edge of a watercourse for plant to
physically reach. Examples of this may be the driving of piles for bridge piers in a
wide river or undertaking works to an inlet structure in a reservoir.
g The works are located at the edge of a watercourse, for example, sheet piling but
there is no easy access for conventional plant, for example, a piling rig. This
situation may arise if the topography is too steep to allow plant to approach the
bank, or the ground simply cannot support the weight of the proposed plant. The
distance travelled to reach the site may also be great, making it cheaper to procure
floating plant rather than to provide a haul road.
Other non-engineering reasons why access cannot be gained to a river bank include where
there are issues with third parties not permitting access across their land. This can be
common place with Inland Waterways, where there is history of conflict between
landowners and the waterway. There may also be environmental factors which may
prohibit the use of plant in sensitive areas (e.g. Sites of Special Scientific Interest).
As with all engineering projects, and in particular marine engineering projects, the access
to the site is key and the planning engineer should ensure that all options are considered
both in terms of practicality and cost before the decision is made. In many circumstances,
gaining access to a site can be more expensive than actually carrying out the works
themselves.
219
Once the decision has been made to go afloat, further logistical issues may arise as marine
plant is generally heavy and will have to be lifted into the water course at a convenient
location. This may itself require further temporary works, such as constructing jetties or
strengthening river banks to accommodate a heavy crane. Alternatively, it may need to
be floated from a remote location. If the latter is viable, then the planner should consider
the time required for the plant to be moved from the ingress location to the work site.
In either situation, the planning engineer must also consider possible obstructions such
as locks or bridges which would impede the movement of plant and also whether there is
adequate water depth for the plant being considered. Tidal and river flows should also be
considered and what contingencies are necessary should be river be in spate. The time of
year should also be considered.
If the plant needs to be supported from a canal/river bed then other environmental
considerations need to be taken into account, such as whether disturbing the riverbed
will damage habitats. Canal beds are historically lined with puddle clay to prevent
water loss, which could be punctured by plant.
The use of marine plant provides challenges for both the operator and design engineer
alike to ensure it is used effectively and safely. The planning engineer should not
underestimate the amount of work required in not just the use of floating plant itself,
but also in its transportation to and from and set-up at the work site.
A barge is generally a single structure which is large in dimensions, meaning it will remain
in the water and be towed from location to location by a tug boat or similar. In most
circumstances, the plant on the barge will remain permanently on board.
A pontoon is typically modular and is designed to be dismantled and moved from one
location to another, typically by road, and then craned into the water along with the
necessary plant.
Clearly the primary concern for the use of barges or pontoons with plant is to ensure the
vessel is large enough to be stable while the given plant is undertaking its specified duties.
Each load case should be considered by an appropriately trained design engineer or naval
architect to ensure the barge or pontoon remains stable.
It is also essential that the site team understand the limitations of the plant use on the
barge or pontoon, and do not stray outside of the boundaries set. Any required
changes to site operations should be referred back to the designer to confirm it is safe
to proceed.
220
As well as cranes, excavators can also be used on barges for such activities as dredging
or grading of river banks or for positioning of bank protection such as rip rap (large
stonework). An excavator working is more dynamic than a crane and the pontoon
can experience more roll, which can make working difficult. In these circumstances,
the barge may be fitted with ‘spud legs’ which are generally circular tubes running
through the deck of the pontoon. These can be lowered so they embed themselves into
the riverbed to provide extra stability (see Figure 18.2).
Care does need to be taken when using spud legs purely for stability if the watercourse is
subject to wave action or the pontoon is operating during an ebb tide. The reason for this
is that in either scenario an air gap could appear below the pontoon, meaning its weight is
transferred onto the spud legs rather than being supported by the water. If the pontoon
has not been designed for this scenario, then the pontoon could either become unstable
or individual elements could fail due to the change in load path.
If any of the above scenarios could occur, then a jack-up barge may be more appropriate
(described in Section 18.2.3).
Where access is required but a barge cannot reach the location (e.g. the river is not wide
enough or there is an obstruction such as a canal or river lock, which can be as narrow as
7 feet) then a modular pontoon may be a more appropriate solution. The Uniflote is the
classic example (Hathrell, 1968). It is a steel box approximately 5 m long by 2.5 m wide
221
by 1.2 m or 1.8 m deep, and can be connected to others to provide a range of different-
sized working platforms depending on the requirements. It was developed for the army
and had three requirements
There are now numerous copies of the original concept; an example is depicted in
Figure 18.2. These are used very frequently on inland waters either for dredging
works, lifting operations and other typical construction activities. Whatever the activity,
the ensemble still needs to be checked for stability and pontoon size adapted if necessary.
In addition, the connection details should be considered to ensure load can be safely
transferred from one unit to another.
The decision of whether to use a barge or a pontoon will depend greatly on the plant size
being proposed and the access restrictions at the location where the floating craft is
required. The barge shown in Figure 18.1 could only be transported by water because
of its size and would also be restricted by any constrictions along the route; it could
be used anywhere on the River Thames within London and could be transported to
other estuarial locations in the UK and abroad. However, it could not be used upstream
222
of Teddington Lock as the barge width is considerably wider than that of the locks at
that location. On the other hand, a modular pontoon could be transported via road to
an ingress point and constructed on the water.
Clearly, if considering a pontoon, the planning engineer must ensure that a suitable loca-
tion is available to launch the pontoon, which may require vehicular and crane access. As
already stated, this will require additional considerable temporary works (Chapter 17).
Pontoon sections can allow flexibility in the shape of the working platform allowing
for non-standard shapes to be developed, thus optimising the size of the pontoon for
the required works. This is particularly useful in activities such as inland dredging
where the width of the watercourse can be limited.
The perceived weak spot of these pontoons units is the connection lugs. Various
independent studies have been undertaken to check the long-term durability of the
connections and Finite Element analysis has been performed on the stability of the
pontoons due to the inherent flexibility of the product.
A jack-up barge is in effect a barge or pontoon which can be lifted clear of the water.
Typically, it will have four spud legs (one in each corner) which embed themselves in
the watercourse bed, allowing the pontoon to be raised using jacks (see Figure 18.4).
Jack-up barges have been used extensively in offshore works such as wind farms and in
the petrochemical industry, but are now also used in tidal rivers and estuaries. They can
either consist of a single section or be built up of individual units connected by lugs. This
gives the user some flexibility in transporting the barge and the shape.
223
Spud leg
Bed
Embedment
224
18.2.4 Other
18.2.4.1 Crane barges
These are bespoke vessels designed with an integral crane unit. These vary from the small
boats working on inland waterways to the large vessels constructing oilrig platforms.
In basic terms, the stability of any floating plant is a function of its combined centre of
gravity and a theoretical height known as its ‘metacentric’ height. It is easiest to relate
distances from the keel, K. Figure 18.5 depicts the basic principals; see also Webber
(1990).
For the pontoon and its plant to be stable, the theoretical distance KM should be greater
than KG, i.e. the metacentric height is higher than the centre of gravity. If not, the craft
will rotate until the above is true, which may mean that it capsizes. KB is computed by
calculating the centre of buoyancy, which is a function of the water displaced by the
M – Meta-centre
G – Centre of gravity
B – Centre of buoyancy
K – Keel
B
K
225
bd 3
I=
12
d V=a×b×d
a
weight of the pontoon and the plant thereon. KG is computed by determining the
combined centre of gravity for the pontoon and the plant.
The distance BM is defined as I/V where I is the second moment of area for the plan of
the pontoon and V is the volume displaced by the pontoon. We therefore have
KM ¼ KB þ BM
Increasing the size of the pontoon or reducing the volume of water displaced will clearly
increase the metacentre height. However, increasing the volume of water displaced will
increase KB.
The weight of the pontoon, crane, load and ancillary items depicted in Figure 18.1 are all
known as well as their centres of gravity about the pontoon keel; the stability of pontoon
can therefore be confirmed. The crane would have been given particular examination and
split up into its tracks, body and jib, and the weight and centre of gravity of each item
considered separately.
The location of the plant on the deck will also tend to make the pontoon rotate, reducing
freeboard on one side and increasing it on the other (Figure 18.7). This should be checked
because if the roll experienced exceeds 78, the pontoon can be difficult to work on. In
addition, if the freeboard is reduced excessively, there is the possibility water could
flood onto the deck.
The crane would have also been closely examined because when it slews around, the
amount of roll of the pontoon will vary. It may be necessary to down-rate the capacity
of the crane while working on the barge. The reason for this is that as the crane lowers its
jib to reach out further, the pontoon will roll accordingly. This rolling effect will increase
the working radius of the crane and could mean it is working outside of its load envelope.
To assist with stability, the barge may have been constructed with a series of internal or
baffle walls. These compartments can be individually filled with water to assist with
stability by rebalancing the barge should a heavy item of plant be required at one end.
The barge itself would need to be designed or checked to ensure it is structurally adequate
to support the weight of the plant on the pontoon.
226
Environmental effects such as wind, water current and tides also need to be considered
in the stability of a barge and the plant thereon. The wind load on a lattice jib can be
considerable.
The actual types of plant may also give rise to different load patterns. A crawler crane
generally operates slowly, both when lifting and slewing, and the loads that are being
dealt with are generally fully understood. On the other hand, an excavator undertaking
dredging duties will be operating faster as it first excavates the river bed and then drops
the arisings into a hopper. It will therefore exert a larger dynamic load onto the barge,
and this must be taken into account by the engineer checking the pontoon and barge.
A situation where the machine needs to dig into a stiff ground, so that the machine is
pulling harder than expected, may occur. The engineer needs to consider the likely
conditions the driver will be operating within and that additional loading criteria are
checked. This can increase the expected loads on the pontoon between 25% and 100%
of those calculated in theory. Good judgement and experience is just as important as
being able to ‘crunch’ numbers.
The barge will generally be towed to the works location and may for some of its duties act
as a floating platform. It therefore needs to be checked for stability in the same manner as
a floating pontoon or barge.
When it is in its elevated position, the barge must also be analysed to ensure that it is not
overloaded due to the loaded plant. A structural analysis is therefore needed to ensure
227
that the structure can transfer the plant loads onto the spud legs. In a barge, this will
ensure the structure is adequate in bending and in shear. In a modular system, the
connecting lugs between units will also need to be reviewed for shear and tensile capacity.
The lugs are a critical item and should be visually inspected on a routine basis; there have
been several examples of lugs failing under load.
The loads in the spud-leg jacks should be compared against the jack capacity. The
capacity of the leg itself will depend on the length of leg below the barge. The analysis
in its simplest term can consider the leg as a pinned column and the capacity determined
by the leg’s radius of gyration and the effective length.
However, the critical item with a jack-up barge is the foundation for the spud legs. The
legs would typically need to embed into the river bed far enough so that when the barge
is lifted, no further settlement occurs. When planning a jack-up barge operation,
topographic and geotechnical investigations should ideally be undertaken. This is not
always possible, however, particularly if the jack-up barge is being used to obtain the
geotechnical information in question. A successful jack-up relies as much on the skill
of the barge master as it does on design and analysis. The barge master will monitor
for settlement during its deployment and will undertake load-tests to be satisfied the
set-up is stable. Topography is important to ensure the barge is not being deployed on
a slope.
Once jacked-up, the spud legs will be subjected to forces from the current and wave
action. Care has to be taken to ensure the barge is positioned higher than the top of
the waves, so it is not lifted and dropped by the resonating wave action. The designer
needs to identify the water flows in the location and the expected high-water level. In
an estuary or further out at sea, the barge legs will be further subjected to wave
action. In this situation, the designer must indentify the critical wave height about the
maximum high-water level.
These heights should be related to chart datum, allowing the designer to specify the
height of the barge above the bed level so that, in normal working conditions, the
water level does not approach the underside of the barge. In an extreme weather
event, the barge will be jacked-up as high as possible so that it is clear of high wave
action.
The other main concern for a jack-up barge is when lifting the spud legs so the barge
can be moved. Considerable suction may need to be overcome by the embedded
length of tube, and the skill of the barge master is critical for the safe extraction of the
spud legs.
The use of the jack-up barge has considerable risk involved and the International
Jack-up Barge Owners Association has been formed to share best practice. There are
also many HSE documents on the subject. A best-practice guide, Guidelines for Site-
specific Assessment of Mobile Jack-up Units has also been drafted by the Society of
Naval Architects and Marine Engineers (SNAME) (Bennett et al., 1994).
228
This section on design has only covered aspects pertinent to stability. Structural
adequacy checks follow standard design procedure.
REFERENCES
Bennett WT, Hoyle MJR and Jones DE (1994) Guidelines for Site Specific Assessment of
Mobile Jack-Up Units. SNAME, New Jersey.
Hathrell JAE (1968) The Bailey and Uniflote Handbook, 3rd edition. Acrow Press, London.
Tupper EC (2004) Introduction to Naval Architecture, 4th edition. Elsevier Butterworth-
Heinemann, Oxford.
Webber NB (1990) Fluid Mechanics for Civil Engineers, SI edition. Chapman and Hall,
London.
FURTHER READING
Blake LS (1994) Civil Engineering Reference Book, 4th edition. Elsevier Butterworth-
Heinemann, Oxford.
229
Chapter 19
Temporary bridging
Bernard Ingham Engineering Manager, Mabey Hire Services Ltd
Temporary bridges combine the performance of a permanent structure with the ability
to re-use most or all of the constituent parts. This has set a challenge for military and
bridging suppliers to develop solutions which are easily transported, installed and
removed, but which offer features associated with permanent structures for relatively
short periods. A great deal of knowledge has been built up in bridging systems and
working methods over many years. The key to success with any temporary bridge
scheme is to find the specialist skilled and experienced personnel who are able to make
the best use of knowledge developed over many years.
19.1. Introduction
Temporary bridges may be required to carry pedestrians, services, public highways,
site access roads, site haul roads, special loads, railways or military vehicles. The
obstacles to be crossed can include rivers, canals, footpaths, public roads, railways,
site roads, services and construction areas. Bridge spans may also be required to form
link spans to jetties or ramps, thus eliminating the use of temporary fill. They can
also be used to provide a low-level removable canal crossing, for example, to move
construction plant.
The bridge may be required for just a few hours, for example, to carry a special load
over a weak bridge or canal, or for several years. Lead times from initial concept to
installation can also vary greatly from several months for larger schemes to just 1–2
days for emergency applications such as following flood damage.
Whatever bridge is required, its foundations will also be temporary works and
require design. The interface responsibilities between the procurer, temporary bridge
supplier and the constructor of the foundations need establishing at an early stage; see
Chapter 2 on management. The Temporary Works Coordinator (TWC) has an
important role in temporary bridging.
Many different temporary bridge systems and installation solutions are available from
launching from one side to using heavy craneage to lift into position. The military
solutions have led the development of many of the modern bridge systems.
Detailed design of temporary bridging systems and individual scheme designs can be
influenced by the following factors
231
g the need for easy and rapid transportation to site, installation and removal
g the availability and versatility of components to maximise utilisation
g the need to take economic advantage of the relatively short-term design life
g the use of suppliers’ technical knowledge due to lack of EuroCode or UK design
codes or standards specifically written for the design of temporary bridges
g the ground conditions for the construction of temporary foundations
g the suppliers’ data available from testing of full-scale bridges and their
components.
Although the following sections include information about procedures adopted in the
UK, many of the principles apply to other sites throughout the world. Information on
the temporary works involved in the installation of permanent bridges, for example,
‘push-launch’ structures, is given in Chapter 27 on bridge installation techniques.
Named after the civil servant Donald Bailey who designed the radical new steel bridge
system, the Bailey Bridge was adopted as the standard Military Bridge in 1941. It was
revolutionary in that its light but strong and versatile system could be manually
erected without craneage, and proved to be one of the greatest inventions of the war.
It played a significant part in the allied victory; by 1947 it had been used to build
more than 1500 bridges in northwest Europe.
Other temporary bridge systems have subsequently been developed for both military and
civilian use. Figure 19.1 illustrates a modern military bridge, which incorporates many of
the principles of Donald Bailey’s 1940s design.
232
A feature of the panel bridge is the use of high-strength pins to connect panels together.
The pins are in shear, and design of the pin housing has always been a critical area for
stress concentrations/welding.
Current panel bridges such as the Mabey and Leada Acrow products remain the most
widely used (Figure 19.2 depicts a typical example). They have the advantages of
233
Figure 19.2 Panel bridge with reinforced concrete pads on temporary reinforced fill abutments
Courtesy of Mabey Hire Services
are the Mabey Delta and Unit Construction Bridges that are designed for spans in the
range 50–130 m. Certain types use bolted connections in line with the trusses. This can
allow pre-cambering of the structure to cater for deflections during use, a significant
benefit when temporary bridge equipment is used as falsework (see Chapter 22 on false-
work). Other lightweight truss bridge systems exist for footbridge applications.
234
Figure 19.3 Deck-type bridge with trestle piers and 3 m high concrete lower piers for impact
resistance
Courtesy of Mabey Hire Services
g railway bridges (although temporary systems have been adapted for rail use)
g sites where dimensional constraints eliminate the use of any standard system
g simple short-span footbridges where the use of scaffold systems, unit beams, etc.,
is acceptable.
Key factors in the choice of type of foundations are economy, speed of installation and
method of removal or making good following dismantling of the bridge. Cost and time
savings can be made by not adopting specifications of permanent structures
235
Figure 19.4 Typical bank seat detail for temporary panel bridge
Courtesy of Mabey Hire Services
g minimising concrete reinforcement (e.g. anti-crack steel for durability may not be
required)
g allowing greater settlement than for a permanent design
g considering reductions in vertical and horizontal specified loading to the bridge
deck (e.g. some load values such as highway longitudinal loads that may be
deemed inappropriate for temporary site construction applications)
g the use of pre-cast concrete and timber foundation elements
g the bridge supplier can provide intermediate piers using a proprietary braced
trestle system.
19.3.1 Programme
Establishing an outline programme from the outset is a vital part of the design process.
The requirement for bridging will have been identified in the temporary works (TW)
register. The likely start date and period of use of the bridge will influence the choice
of system due to considerations of economy, performance and availability. Sufficient
time should be given to checking the design.
236
19.3.2 Loading
19.3.2.1 Public use
For public highways and footpaths, the relevant highway authority will specify the
imposed loading, including any guardrail and/or barrier loading. They are most likely
to adopt a load rating from a standard specification such as the Highways Agency
HA, HB and pedestrian loading. Traffic loads on bridges is provided in BS EN 1991-2
(BSI, 2003). Where side protection is specified, the solidity of any such barriers can
impart significant wind forces onto the bridge.
g Single vehicle loading: adopting one vehicle per span is accepted practice as long
as measures are established to control vehicle movements and allow for vehicle
breakdown (e.g. towing off by second vehicle). Emergency vehicles should be
considered where lighter vehicles are specified.
g Bridge Assessment Standards: Highways Agency’s bridge assessment standard
BD21/01 (HA, 2001) is a useful document for specifying loading for multiple
vehicles where the maximum gross vehicle weight is under 44 tonnes.
g Vehicle numbers for fatigue: the approximate total number of vehicle passes
should be established.
g Edge protection requirements depending on use of the bridge, guardrails and/or
vehicle edge protection barriers may be required; these can impart high lateral forces.
g Pedestrian loading for non-public applications can be specified in a number of
ways providing measures are established to control the loading on site. It is
recommended that Service Class 1 (0.75 kN/m2) be the minimum, although each
member of any platform should be designed for a minimum Service Class 2
loading (1.50 kN/m2).
With regard to spans, the minimum clearance envelopes for vehicles, pedestrians, rail-
ways, water-borne traffic, flooding and any other obstacles should be established with
the relevant authorities and included in the TW design brief. Consider areas that may
be required for construction, as this can affect the choice of span arrangement. Estab-
lishing the dismantling method is important, as the site layout will often change signifi-
cantly while the bridge is in use and the method of dismantling may not be the reverse of
the assembly procedure.
The TWC should ensure that the deck and foundation design are compatible to find
the optimum arrangement. Choice of articulation will be a part of this process. An
237
Figure 19.5 Severe site constraints determined the bridge type and erection method on this site
in Scotland
Courtesy of Mabey Hire Services
g Confirmation of the design brief and submission of any client technical approval
documentation (e.g. Form C on railway works).
g General arrangement drawings, including foundations.
g Calculations from (a) suppliers and (b) foundation designers.
g Drawings and calculations for any special scheme-specific components.
g Design risk assessments. Particular considerations for temporary bridges are:
control of loading on the deck; clearances below for vehicle impact or flooding,
etc.; temporary loads and stability during installation and dismantling; ground
conditions for installation and dismantling; site security (bridges are often outside
scope of regular site control measures).
g Design check certificates.
Note that although the bridge supplier may issue a design check certificate to its own
check category, there should be a separate design check certificate for the temporary
works as a whole, including the foundations (see Chapter 2 on management).
238
g Deck surface: the use of anti-skid surfacing with high levels of skid resistance and
durability. These can be factory applied to temporary bridge steel deck.
g Deck drainage: most temporary bridge decks are formed of a series of relatively
lightweight deck units which flex independently under load. Such units are
designed for water to pass through the joints, and sealing is not a practical option.
g Parapet: various proprietary parapet systems are available from the bridge
suppliers.
g Impact protection: consideration must be given to the risk of vehicle, train or
vessel impact to temporary bridge soffits and supports. Common solutions include
increasing clearances outside the zone of risk, barrier systems, earth bunds,
sacrificial goal posts or beams and concrete piers (see Figures 19.2 and 19.3).
239
g Assembly in place: this simply involves assembling the span(s) in place, for
example, where the bridge is being built on top of an existing bridge or on the
ground where excavation beneath will follow installation. Site craneage would be
required.
g Lift into place: the development and availability of large mobile cranes has
permitted the lifting of ever larger and heavier spans. Safety and speed are the
main beneficiaries of lifting. Particular attention is required for the design and
installation of the temporary works of the crane pad foundations to cater for the
outrigger loads.
g Traditional advancing launch: this is often the only practical method for medium-
and long-span applications. Standard bridging is used to form a temporary
lightweight cantilevered nose connected to the front of the structure. The
completed deck is then easily moved forward and controlled using site plant,
winches or tirfors. As the bridge is advanced on ground-mounted rollers, the
position of the centre of gravity of the bridge must be carefully monitored
(particularly where the bridge is advanced and sections of completed bridge
added to the rear end). As the launching nose advances, the tip will deflect and
suitable allowances made. The physical act of rolling a bridge will cause top and
bottom booms to change from tension to compression and vice versa; in such
cases it is even more important that the bridge supplier’s instructions are
followed.
g Other launch methods such as the tail launch, which uses additional bridging at
the rear of the structure, can be used. Other installation methods include floating
on pontoons and cantilever build, as illustrated in Figure 19.6.
Figure 19.6 Side spans of a panel bridge being installed using the cantilever method
Courtesy of Mabey Hire Services
240
19.4.3.5 Inspections
Inspection requirements should be established before construction commences, and will
include short- and long-term inspection regimes. Many factors will influence the
frequency and extent of inspections such as the bridge type and intensity of loading.
Conventional means of access methods are usually employed, such as the use of
mobile elevated working platforms (MEWP). Although temporary bridging would be
inspected on a construction site as part of the statutory inspection, the specialised
nature of the equipment often requires specific regular checks to be carried out by
241
persons familiar with the equipment who also have the training to know which part of
the structure can be susceptible to fatigue or wear. The advice of the specialist supplier
should be sought.
g Panel bridge truss elements used for spanning falsework girders. As the elements
are joined by pins in shear, there is no control over the deflection of the assembly
and allowance has to be made either by tapered firring pieces and/or a falsework
skeletal system seated on the girders with facilities for height adjustment (see
BS 5975; clauses 19.5 and 19.7.1).
g Truss elements in the vertical plane to form supports for bridges and other
structures.
g Bridge decking elements used with panel or trestle systems to form site working
platforms.
g Deck elements placed on compacted fill and used to spread the applied load and
reduce the applied bearing capacity required on the compacted fill.
REFERENCES
BSI (2003) BS EN 1991-2: EuroCode 1: Actions on structures. Part 2: Traffic loads on
bridges. BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
HA (Highways Agency) (2001) BD21/01: The assessment of highway bridges and struc-
tures. HMSO, London.
FURTHER READING
HA (2005) BD2/05: Technical Approval of Highway Structures. HMSO, London.
Harpur J (1991) A Bridge to Victory. HMSO, London.
Joiner JH (2001) One more River to Cross. Leo Cooper (imprint of Pen and Sword,
Barnsley).
Network Rail (2004) Technical Approval of Design, Construction and Maintenance of
Civil Engineering Infrastructure. Network Rail, London.
242
Chapter 20
Heavy moves
Martin Haynes Sales & Marketing Director, Fagioli Ltd
Andrea Massera Engineering Director, Fagioli SpA
This chapter deals with the horizontal and vertical movement of heavy loads. It outlines
the reasons for choosing this method of construction and discusses the various methods
that are available and used within the industry. Although conventional cranes are
mentioned, priority is given to alternative techniques including trailers, hydraulic jacks
and various skidding systems.
20.1. Introduction
Firstly, what constitutes a ‘heavy move’? For the purposes of this chapter, a heavy move is
something that cannot be installed using conventional site plant. This could be as low as
five tonnes (e.g. retrofitting a vault into an existing bank basement) or as heavy as
30 000 tonnes (lifting of a complete oil platform). The idea of constructing a component
somewhere other than its final location must offer a significant advantage over in situ
construction for it to be considered. These are usually seen as one or more of the following.
Occasionally, the programme will dictate that a heavy move technique is necessary. This
is commonly seen during replacement of rail and motorway bridges. The new bridge is
constructed close to the existing bridge, and the old bridge is moved out (by skidding
or trailers) and the new bridge installed in a similar manner. The on-site programme
and therefore disruption to the public is kept to a minimum (refer to Chapter 27 on
bridge installation techniques). In addition, off-site prefabrication permits the erection
of industrial plants, for example, in locations where climatic conditions are adverse
and the available site construction period restricted to a few months a year.
243
20.2. Techniques
20.2.1 Cranes
Two types of crane are widely available for heavy lifts: these are truck-mounted
telescopic boom cranes and crawler-mounted lattice boom cranes. Many other types
of crane are available (e.g. tower cranes) but are not considered because of lack of
capacity and/or availability.
244
boom attached; they are ready to lift after placing their outriggers and extending their
boom. They can lift, slew (rotate) and boom up or down, but cannot travel under
load. Note that this type of crane can only lift its quoted capacity with the boom at
minimum extension and radius; capacities reduce quickly as the boom length and/or
radius is increased.
g the weight and centre of gravity location of the lifted item (the hook of the crane
must be over the centre of gravity or the load will tip)
245
Once these points have been established, the best practice is to consult the crane hire
companies who will select the most suitable crane and prepare the necessary rigging
studies. This will allow the selection of the most suitable crane that is available.
An alternative is to carry out the rigging studies oneself. Starting with the information
above, the crane location should be set on plan together with the build/delivery and
final locations of the lifted item. The distance between the centre point of the crane
slew ring and the centre of gravity of the load is the lift radius. The length of the
boom is decided next and will depend upon the lift height required. Once these physical
dimensions have been established, crane duties (a series of charts/tables showing lift
capacities with various combinations of boom length and lift radii) must be consulted
to see which crane and configuration is most suitable. Cross-referencing the lift radius
with boom length will give the safe lifting capacity of the crane.
Operationally, crane lifts are relatively quick when compared to alternative methods
with most lifts being completed within one hour. When in operation, however, all
cranes have a limiting wind speed; in general the longer the boom, the lower the opera-
tional wind speed. The crane duties will specify the limiting wind speed for the crane
which can be as low as 9 m/s.
20.2.2 Trailers
Self-propelled modular transporters (SPMTs) have the ability to move huge loads over
long distances with minimal ground preparation (see Figure 20.3). The item to be moved
will generally be built on temporary supports leaving a clear space 1500 mm high
underneath to allow the SPMTs to be inserted. The SPMTs have an integral jacking
system to lift the item from the temporary supports. Once the load is on the trailer
bed it can be moved to location and, if required, the trailer hydraulics can be used to
transfer the load into the permanent supports.
SPMTs come in two basic module sizes of four axles and six axles with capacities around
32 tonnes per axle line, which is in excess of even major road capacities. These modules
can be connected together to provide the correct capacity and physical configuration
necessary for a particular move. Once inter-connected, they act as a single unit and
are operated from a single point. Linking SPMTs together allows massive structures
to be moved; loads in excess of 10 000 tonnes are not unusual.
If considering the use of SPMTs, then consult the specialist transport companies offering
such equipment. They will need to know the following information
246
Operationally, SPMTs are quicker than alternative systems (e.g. skidding) with opera-
tional speeds of up to 5 km/h; speeds in excess of 500 m/h will however result in a
reduced load-carrying capacity. There is no wind speed rating for using SPMTs as
there is with cranes; however, the wind load on the item needs to be considered and
may result in a limiting wind speed for any particular operation.
In cases where a relatively short lift (e.g. less than 2 m) is required, then cylinder jacks can
be used to jack and pack the load (see Figure 20.4). This technique can be used in
247
conjunction with trailers or hydraulic skid shoes to allow the item to be built closer to
ground level.
Most cylinder jacks will operate with a full load at around 700 bars of pressure, so expect
bearing loads around 70 N/mm2 above and below the jack. Flat jacks operate at 150 bars
and air bags even less, so there are generally no bearing pressure problems.
Strand jacks are only one part of the overall lifting system; in addition, they will require
pumping systems and control systems. It is important that the systems are correctly
matched to ensure a properly controlled and synchronised operation. The control
systems available are very precise, which is useful when setting weld gaps or fitting bolts.
248
Figure 20.5 Strand jacks mounted on a purpose-built structure (2000 tonne lift)
Courtesy of Fagioli Group
Strand jacking arrangements are very flexible and can easily fit with the shape of the item
being lifted. They can be used in situations where cranes would be impractical or even
impossible (e.g. an aircraft hanger roof where there are multiple lift points with differing
loads).
Strand jacks need a support structure (see Figure 20.6). Wherever possible, the supports
should utilise the permanent works with only minor modifications. There are no specific
design requirements for the strand jack supports; most are steel which is designed in
accordance with normal codes. The speed of operation of strand jacks (m/h rather
than m/min) means that dynamic loads are not really applicable.
If it is not possible to utilise the permanent works then it may be necessary to consider a
complete temporary support system. This can be purpose-built or configured from a
modular support system available to hire. Such support systems are available with
very high capacities (500 tonnes on a 100 m unguyed tower and much more for guyed
towers) and the combination of a modular support system and strand jacks can be a
viable alternative to heavy cranes.
249
Figure 20.6 Strand jacks mounted on a modular support structure (1100 tonne lift)
Courtesy of Fagioli Group
Limiting wind speeds will depend on the support structure and the wind area of the lifted
item. The typical limiting wind speed for strand jack operations is 16 m/s.
20.2.4.1 Rollers
Mechanical rollers have been in use for many years. They work by using an endless chain
around a central bearing plate to provide a low-friction interface. Typical friction values
are quoted at around 5% but could go as low as 2% so beware, especially if moving
downhill. Rollers do not normally require a breakout force (i.e. a greater force to start
the movement) and are smooth in operation.
250
Rollers are available in capacities from very low values up to 1000 tonnes per unit, but
this size would be unusual. Multiple units are typically used to give the desired load
capacity. The rollers should be suitably sized to suit the load at every support point,
and should also have sufficient additional capacity to cope with possible unknown
affects (e.g. differential settlement). For loads over 1000 tonnes, the recommendation
is that rollers should be used at 50% of capacity.
Some of the rollers will need to be provided with guides to ensure that the item to be
moved follows the correct path. Note, however, that providing too many guides can
lead to binding during the move. Care is also required at joints in skid tracks to
ensure continuity of line and level.
The main drawback with rollers is that they have high localised bearing loads which need
to be spread through the skid tracks, possibly using mats or cribbing as an additional
layer. This will affect the build height and may also require extensive jacking down on
completion of the operation.
251
space of at least 1500 mm underneath to allow the shoes to be inserted. The 1500 mm
height may need to be increased to allow for the depth of the skid tracks and any
load spreading. The low-friction interface is normally stainless steel–PTFE (polytetra-
fluoroethylene) which needs to be lubricated.
The control systems available with skid shoes allow a high degree of control over the
loads and displacements; high tolerances of accuracy can be achieved both during
the movement and final placement.
The limitation of skid shoes is that they work in one direction only. Changes of direction
will require a set down of the load onto temporary supports followed by re-alignment of
the track and shoes.
All sources for coefficients of friction advise the use of care in using the values. For the
purposes of a heavy move, this means that the movement system should consider a
worst-case scenario and be conservative in specifying the forces required. For
example, it would be normal to specify a pulling force of 10% if using PTFE on steel.
With low-friction interfaces there may be a requirement for breakout (a higher pulling
force to start movement), particularly if the item has been static for an extended
period of time.
20.3. Design
20.3.1 General
The heavy move technique requires accurate preparation. Engineering design is the key
element for the success of the operation from the point of view of safety, technical,
budget and schedule. Engineering targets are to identify the most safe, robust, economic
method and equipment and to produce complete and clear technical documentation for
252
the safe execution of the operations, avoiding any miscommunication from engineering
to site.
Input for the engineering activities, which is carefully reviewed and assessed by a
competent and experienced engineer, includes the following
g transport and lifting drawings as well as installation phases drawings for the item
g installation procedures, operation manuals
g calculation reports for installation hardware
g design, shop drawings for structures as well as fabrication/inspection plans
g 3D simulations where appropriate.
Engineering documents are the base for analysis of all safety aspects of the operations.
HAZID (hazard identification study), HAZOP (hazard and operability study) and
Safety Job Analysis are structured review techniques for identification and assessment
of operation hazards, techniques which need sound and detailed engineering documents.
The load distribution in a lift is normally calculated as a static load case by applying the
dynamic hook load (DHL) at the hook position and distributing weight and any special
load to each element. The DHL is the product of lifted load (including rigging) and the
dynamic amplification factor (DAF). The skew load factor (SKL) allows for the extra
loading on slings caused by the effect of sling-length and lift-point manufacturing
tolerances as well as rigging arrangements, which will affect a statically indeterminate
253
lift. For statically indeterminate lifts, such as 4 slings in a pyramid arrangement, the SKL
could vary over a wide range of 1.0–2.0, dependent upon sling-length tolerance and load
shape. A figure of 1.25 is often taken, where sling length tolerance is within 0.25% of its
nominal length.
The maximum dynamic forces calculated as explained above are the design forces for
lifted item structure and its pad-eyes for slings, grommets and lifting hardware (e.g.
lifting beams).
Shackles are selected on the base of the static hook load (SHL) since their safety factor
also covers the dynamic effect. Cranes are selected on the basis of SHL applied at the
hook position and by determining the reactions for each crane involved in the lifting
operation, considering the CoG envelope as appropriate.
Recommended DAFs for onshore lifting, according to DNV (2000; Part 2, Chapter 5)
Rules, are as shown in Table 20.2. As a general rule, lift operation shall be studied in
such a manner that the maximum tilt of the lifted item is less than 2%, unless there is
a different requirement of the client. The maximum operational wind speed shall be
defined according to the crane manufacturer instructions, taking into consideration
the weight and shape (drag factor) of the lifted item. A normal value for the
maximum operational wind speed is v ¼ 9.0 m/s (taken at top of crane boom). The
maximum ratio of exposed surface/weight for the lifted item is 1 m2/tonne (otherwise
the operational wind speed shall be reduced).
Lifting a load with two or more cranes requires greater planning since many factors
affect the operation: accuracy of weight/CoG, capacities of lifting accessories,
synchronisation of crane motions, crane instrumentation, site conditions and super-
vision. When such factors cannot be accurately evaluated, good engineering and industry
practice dictates that an appropriate down-rating should be applied to all cranes
involved (BSI, 2006) as follows.
Higher values for the de-rating factor can be used, but should be based on individual
cases. The load transferred by the crane to the ground should be carefully evaluated.
The pressure under a crawler crane varies over the range 40–60 tonne/m2 (crane
working at more than 70% of its capacity). Adequate spreader mats should be placed
254
under the crawler to reduce the soil bearing pressure to less than 20 tonne/m2. Good
practice is to require a ground bearing capacity for crane operational area of
25 tonne/m2 (refer to Chapter 5 on site roads and working platforms).
The crane lifting area should be adequately prepared and tested for the maximum ground
pressure to ensure that (1) the maximum inclination of the ground is within 0.38
(0.5%) in any direction and (2) the maximum ground settlement under pressure of
25 tonne/m2 for crane lifting is within 30 mm (this will vary with crane model).
Where required, adequate installation guides should be provided for placing the lifted
items to ensure that the load lands at the correct location. Design load for the guides
is a horizontal force of not less than 5% of the weight of the lifted item.
As a general rule, the lifting operation should be studied in such a manner that load does
not get any closer than 1 m to the crane boom or any other structure. This limit can be
changed depending on the operational conditions (height, visibility, wind, etc.).
Clearance from energised power lines should be as described in the American Society of
Mechanical Engineers document B 30.5-2000 (ASME, 2000), see Table 20.3. Typical
output documents of lifting engineer include
g analysis of the item structure to be transported (inputs for this analysis are the
transport drawings and the SPMTs reactions against the item structure, issued by
the Heavy Transport Contractor)
<50 10 (3.05)
>50–200 15 (4.60)
>200–350 20 (6.10)
>350–500 25 (7.62)
255
For the design transport weight (DTW), the vertical DAF is 1.0. The transported item
is subject to horizontal inertia loads in the longitudinal direction (direction of motion)
and in the transverse direction due to acceleration/deceleration. It is also subjected to
horizontal loads due to the road slope, SPMT type, number of driven axles, convoy
arrangement and operational criteria.
For SPMTs (maximum design speed 5 km/h), normal ranges for horizontal accelerations
are as follows
When on slope, the acceleration a3 is added to a1 or a2. Between the underside of the
transported item and top steel of SPMTs frame/stools, plywood is used for a higher
friction coefficient. An adequate lashing/stopper system should be provided, designed
to withstand (with the contribution of the friction, when appropriate) the total
horizontal load.
The loads/reactions distribution in the transport system is calculated as a static load case
by distributing the DTW to each temporary transport element and to SPMTs. As a
general rule, the SPMT hydraulic circuits are arranged in such a way that a 3-point
system is achieved. With this system (isostatic), the reactions on the trailers are
practically constant and there is no overstress in the transported item or in the SPMT
structure/hydraulic system. If tall items are transported with only one SPMT line,
then the 4-point system is considered to have more stability. SPMTs are also suitable
equipment for the execution of load-in/load-out operations, since they can spread
heavy loads within the capacity limits of a barge or ship deck.
The design operation conditions which are generally considered are as follows
256
g ship maximum trim 1%, maximum heel 1% (for load-in load-out operations)
g ship in level with the quay: 100 mm (for load-in load-out operations).
In the operational condition, considering the most unfavourable loads combination, the
SPMTs axle loads should not exceed the allowable loads stated by the manufacturer
(which depend on the travelling speed). When several axles are overhanging, the
trailer spine beam bending capacity should not be exceeded.
The SPMTs tractive power (as per the Manufacturer’s data sheet) should be at least 20%
greater than the theoretical power required to overcome rolling friction and slope.
As general rule, the transport path should be checked in order to verify there is a
minimum lateral and top clearance of 1 m from existing structures. Less clearance can
be accepted after checking on a case-by-case basis.
When SPMTs are loaded to full capacity (36 tonnes per axle), the maximum ground
gross pressure is 10 tonne/m2 (with reference to the SPMT frame projected area). The
local contact pressure under the wheels is c. 9.5 kg/cm2 (tyre inflation pressure of
10 bar). Maximum ground pressure for each specific transport shall be transmitted to
the client to check the existing bridges, culverts, underground services, retaining walls,
quays, etc.
REFERENCES
ASME (American Society of Mechanical Engineers) (2000) B 30.5-2000: Mobile and loco-
motive cranes. ASME, New York.
BSI (2006) BS 7121: Safe use of cranes. BSI, London.
DNV (Det Norske Veritas) (2000) Rules for planning and execution of marine operations.
DNV, Høvik.
FURTHER READING
Bates GE (1998) Exxon Crane Guide: Lifting Safety Management System. Specialized
Carriers and Rigging Association.
GL Noble Denton (2010) Guidelines for Marine Lifting Operations. 0027/ND. Noble
Denton.
MacDonald JA, Rossnagel WA and Higgins LA (2009) Handbook of Rigging, 5th edition.
McGraw-Hill Professional, New York.
Shapiro H, Shapiro JP and Shapiro LK (1999) Cranes and Derricks, 3rd edition. McGraw-
Hill Professional, New York.
257
258
Chapter 21
Access and proprietary scaffolds
Peter F. Pallett Pallett TemporaryWorks Ltd
Ian Nicoll Chief Engineer, Interserve Industrial Ltd
The provision of temporary safe working platforms for the erection, maintenance,
construction, repair, access or inspection, etc. of structures is known as scaffolding. It
can be formed from individual tubes with fittings or proprietary components. The design
philosophy for stability of scaffolds and loading limits, together with the information
necessary to source the safe height of most UK scaffolds without the need for further
calculations, is discussed. The designation and simple rules for inspection are included.
21.1. Introduction
Wherever it is required to provide a safe place of work for the erection, maintenance,
repair or demolition of buildings and other structures and to provide the necessary
access, a temporary structure known as a scaffold is erected.
In the middle ages, a ‘skaffaut’ was a mobile tower equipped with battering rams used for
assaulting castles. Shakespeare referred to the gallery structure at the Globe Theatre in
London as the ‘scaffoldage’. The first tubular steel scaffolding was seen in the UK
around 1920, using standard 2 inch diameter water pipes that were available in 21 feet
lengths. This type of scaffolding with 2 inch tubes connected together remains a
common method, but more recent scaffolding equipment comprises proprietary
components connected together. By its nature, such structures are usually temporary
and, unlike permanent structures, require to be dismantled after use. This introduces a
re-use aspect to scaffold structures which are almost always erected using material
previously used. This generates an industry of supply (generally on hire), erection, use,
inspection and maintenance, followed by dismantling, removal and inspection prior to
reuse. Managing the scaffold is important in terms of maintaining its capability to
provide adequate access.
Falls from height account for over 50% of fatal accidents in construction (see Chapter 1
on safety) and scaffolders are particularly at risk. The recommended procedures for
controlling scaffolds and all temporary works are given in BS 5975: 2008 þ A1: 2011
Code of practice for temporary works procedures and the permissible stress design of
falsework (BSI, 2011). Industry guidance on scaffolding recognised by the Health and
Safety Executive (HSE) is provided by the National Access and Scaffolding Con-
federation (NASC). TG20 Guide to good practice for scaffolding with tube and fittings
(NASC, 2008a) is the technical guidance on the European Standard BS EN 12811-1
259
(BSI, 2003a). NASC has also published a very useful ‘toolbox talk’ booklet on TG20.
Guidance on proprietary scaffolding should be sought from the supplier/importer of
the particular scaffold. Training courses are run through Construction Skills and
NASC, and there is also a National Certification scheme for all scaffolders.
All users of scaffolding on site should be aware of the statutory requirements under the
Work at Height Regulations and the Construction (Design and Management) Regula-
tions (CDM). They should also be aware that it is their duty to provide a ‘safe place
of work’. The authoritative guidance is given in SG4:10 Preventing Falls in Scaffolding
(NASC, 2010a) and its useful site booklet SG4:You (NASC, 2010b).
A typical arrangement of an independent tied scaffold using tube and fittings is shown in
Figure 21.1, which also includes the terms regularly used in scaffolding.
260
Toe-board
Right angle
coupler Guard-rail
Transoms
Joint pin
Window tie
Ledger
to ledger
brace
Swivel Sleeve
coupler coupler
Right angle
coupler Standards
Ledgers
Façade brace
Foot tie
Base plates
Sole board
With the exception of mast climbing scaffolds, each type of scaffold discussed above is
available either as traditional tube and fitting scaffolds or as proprietary prefabricated
system scaffolds. The big advantage of traditional scaffolding is its adaptability on
complex-shaped structures. Adaptability is often hard to achieve within the dimensional
constraints of a proprietary system. Further decisions would include whether the scaffold
is required to be left unclad, fitted with debris-netting or fully sheeted. Obviously there
261
Once the type of scaffold has been selected, the procurer needs to consider the activity for
which the scaffold is required, the number of platforms, its width, any projections and,
most importantly, the imposed loading per platform which will change depending on the
activity to be performed. BS EN 12811-1 (BSI, 2003a) has introduced six designated
imposed load categories for scaffolds, discussed later in this chapter.
21.2.2 Designation
NASC introduced a three-number designation for scaffolds which describes the type of
scaffold used and assists procurers in selection of the majority of scaffolds, namely N1–
N2–N3 where
A typical designation would be 3–5–2, meaning load class 3 with 5 scaffold boards
between the uprights and with 2 scaffold boards fitted on the inside adjacent to the
structure being scaffolded. The cantilevered inside board is assumed lightly loaded.
Where the inside boards are required to have the same load as the platform, a suffix F
is added. Certain scaffolds are limited to only 1.8 m bay lengths, and these are designated
with the suffix S.
The importance of the designation is that it defines the size and loading of the scaffold
platforms, and is used by the designer to establish the safe height. The safe height of
basic tube and fitting scaffolds for a range of wind conditions in the UK and Ireland,
as a function of the scaffold designation, is listed in the tables in TG20 (NASC, 2008a).
The safe load of any strut (member in compression) is limited by its buckling, so its
effective length LE and slenderness ratio LE/r should be considered. For scaffold tubes
262
Notes: Axial loads derived using a quasi-permissible stress approach according to BS EN 1993-1-1 (BSI, 2005b)
with partial safety factors of f ¼ 1.5 and m ¼ 1.1.
The shaded values are for comparison purposes, and should not generally be used in scaffolds unless the scaffold
has been designed
in compression, the slenderness ratio should meet the condition LE/r < 271 for struts and
braces intended to carry wind and lateral loads, i.e. the lacing and diagonal bracing
should be less than 4.25 m long. Safe working loads of scaffold tubes related to the
effective length LE are listed in Table 21.1. Scaffold tube struts which have a free canti-
lever at the end, such as at the bottom of a freestanding scaffold, are regarded as having
an effective length of the adjacent strut length plus twice the length of the free cantilever;
see TG20 (NASC, 2008a; appendix D).
The safe axial tensile load for Type 4 scaffold tube is 79.1 kN but is usually limited by the
safe slip load capacity of one coupler, i.e. 6.1 kN.
263
Table 21.2 Safe working loads for individual couplers and fittings
264
When considering the design of all scaffolds, the fact it can buckle or fall over in any
direction should be taken into consideration. The accepted rule is: ‘Think vertical,
think horizontal, then think horizontal again’. The procurer will have previously
decided whether the scaffold will be unclad, sheeted or debris-netted.
g Service class 1: 0.75 kN/m2; inspection and very light duty access.
g Service class 2: 1.50 kN/m2; light duty such as painting and cleaning.
g Service class 3: 2.00 kN/m2; general building work, brickwork, etc.
g Service class 4: 3.00 kN/m2; heavy duty such as masonry work and heavy
cladding.
265
BS EN 12811-1 (BSI, 2003a) states that a scaffold in use shall be loaded with one
platform (generally the top) with the full service class load and an adjacent platform
with 50% of that service class load. If more platforms are in use or the scaffold is
taller than permitted, then the scaffold needs to be designed. TG20 Volume 2 (NASC,
2008a) provides a method for calculating the safe height of an unclad independent tied
scaffold which has more than the two platforms loaded, by considering individual
loads in the standards.
BS EN 12811-1 (BSI, 2003a) also requires a notional horizontal load of 2.5% of the
vertical imposed load (minimum of 0.3 kN per bay) applied to each bay of the scaffold,
and also states specific concentrated and partial area loads on the platforms.
To cater for the wind on all temporary works a simplified method of calculating the wind
forces was developed for BS 5975: 2008 þ A1: 2011 (BSI, 2011) and TG20 Supplement 1
(NASC, 2011) by introducing a wind factor Swind for any particular site (see also
Chapter 22 on falsework). It should be noted that the determination of the wind
factor Swind is only required once for each site.
To check the safe height of a scaffold at a site location using TG20 Supplement 1 (NASC,
2011), the wind factor Swind needs to be evaluated. This is defined:
A
Swind ¼ Twind vb;map 1 þ
1000
where vb.map is the fundamental wind velocity at the location (m/s), Twind is the
topography factor considering the ground conditions and A is the altitude of the site
(m). Values for vb,map and Twind are listed in both BS 5975: 2008 þ A1: 2011 (BSI, 2011)
and TG20 Supplement 1 (NASC, 2011).
266
The Work at Height Regulations requires calculations to be completed for all scaffolds
unless there is either a pre-existing design, i.e. already completed in the office for a similar
job, or the scaffold is erected in accordance with a recognised standard configuration. The
latter refer to the use of basic scaffolds detailed in TG20 (NASC, 2008a), erected correctly
and conforming to the bracing, loading and tying patterns established. Additionally, the
latter also refers to technical data contained in the user guide for each system scaffold.
As with all structures when carrying out a full design, the three design requirements to
consider having established the loads are the strength and stability of the members,
followed by the lateral stability of the structure and finally the overall stability of the
structure.
Designers will be aware of the effect of effective lengths on the load capacity of the
scaffold members. Buckling of struts is prevented by adequate sideways restraint in
two directions at right angles, known as ‘creating effective node points’. The value of
the restraint force is usually only 2.5% of the load in the strut. The amount of fixity
of the coupler to the tube is generally neglected; Table 21.1 lists the safe load for
various effective lengths.
The effective length of the vertical standards in an independent tied scaffold is not always
obvious. Lift heights were previously considered relevant, but engineering analysis
supported by full-scale testing of scaffolds by NASC (2008a) has shown conclusively
that two parameters control the effective lengths in a tube and fitting scaffold: the vertical
spacing between the levels of the tie positions and the bay length. A fuller treatise is given
in TG20 (NASC, 2008a; volume 2). The effective lengths of Type 4 scaffold tube related
to the vertical tie arrangement and the bay length for normal 2.0 m lift scaffolds are listed
in Table 21.3.
267
Table 21.3 Effective lengths (LE) for fully ledger-braced independent tied scaffolds with 2.0 m lifts
Data taken from TG20 (2008)
A scaffold tied at alternative lifts (4 m) with any bay length therefore has an effective
length of 3.2 m, giving a safe axial load of 13.6 kN using Type 4 ‘as new’ tube; see
Table 21.1.
This spacing applies to unclad, debris netted and sheeted scaffolds. TG20 (NASC, 2008a)
classifies three capacities of tie: light duty (capacity in tension 3.5 kN), standard duty
(capacity in tension 6.1 kN) and heavy duty (capacity in tension 12.2 kN).
Four main types of tie are used, namely the through tie, box tie, reveal tie and anchor tie.
Ties should preferably be fitted to the standard or at least within 300 mm of the standard
on the ledger.
The selection of tie positions, whether the tie fixing into the building should be tested
before use and even the suitability of the fabric of the building to carry ties are all
part of the discussions at the early stage of procurement of scaffolds. Structures with
minimal provisions for ties will obviously require more detailed design.
Although BS EN12811-1 (BSI, 2003a) states that ledger bracing should be omitted at
working lifts to give ‘a completely unimpeded area’, the best-practice opinion of the
HSE and the NASC is that, for the majority of scaffolds in the UK, all scaffolds
268
should have full-height ledger bracing on alternate bays. Where a client requests the
omission of up to two lifts of ledger bracing to a scaffold, referred to as a ‘part ledger
braced scaffold’, information on the significantly reduced safe height is given in TG20
(NASC, 2008a; volume 2).
The bracing is fitted as diagonals at an angle between 358 and 508, creating stiff triangles.
Loads are transmitted down through the scaffold to ties and/or the ground, etc. and the
brace direction is unimportant (see Figure 21.3). If the bays are less than 1.5 m, then
ledger bracing is fitted onto every third pair of standards. Façade bracing is fitted to
the outside standards at least every 6 bays, either across 2 bays as shown on the right
side of Figure 21.3 or continuously from bottom to top. Where façade bracing is
across a single bay or not fitted between ledger braced frames, plan bracing is required.
269
As soon as any of the parameters alter, such as the decision to have the full imposed load
on more than one platform at a time, the structure is invalidated as a basic scaffold and
calculations are required. A method of evaluating the safe heights for such scaffolds with
more levels loaded, for unclad scaffolds only, is also given in TG20 (NASC, 2008a;
volume 2).
270
Suppliers/importers have duties under the Sale of Goods Act to ensure the items are
supplied as described. This imposes duties on the users of such scaffolds to ensure
that they are supplied, used and subsequently dismantled following the instructions of
the suppliers or importers. This is usually in the form of a user guide which provides
the loading capability and safe heights of scaffolds, including tying patterns. Proprietary
equipment suppliers will often provide erection manuals, safety DVDs and, in certain
cases, run training courses and/or workshops to familiarise users with the system. It is
important that such scaffolds are used as intended.
The use of fixed-length transom units with patented fittings to each end are regarded as
proprietary scaffold and not tube and fittings. Information on their use should be sought
from the supplier of the system.
One significant design consideration is the lack of continuity along the length of the
scaffold. The ledgers are not continuous and stop at each standard; each pair of
271
standards generally requires to be tied to the building and this can increase the amount of
tying needed. It is always recommended to refer to the latest supplier’s technical data.
The imposed load on proprietary systems is often higher than the equivalent tube and
fitting scaffolds. This can have advantages due to the speed of erection, particularly
on large and straightforward façades. The platforms have to fit at the modular connec-
tion points of the scaffold, so platforms cannot always be fitted to suit the ideal work
location. This is in contrast to tube and fitting scaffolds, where they can be fitted to
suit the work required.
Work at Height Regulation 12 requires that all working platforms be inspected prior to
use after exposure to conditions likely to have caused deterioration (i.e. after high winds
or local flooding) and at suitable intervals. The latter refers to scaffolds erected for some
considerable time, say more than 3 months, where the fittings may become loose. It
should also be considered that the lifespan of a scaffold should not be greater than
2 years. After this time, more permanent loads (especially wind loading) should be
considered and the scaffold may require additional design.
In addition, for all platforms used in construction where a person could fall 2.0 m, the
Work at Height Regulations state that they should be inspected every 7 days. All
contractors and subcontractors should be aware of this and be maintaining a register
of such inspections.
REFERENCES
BSI (1982) BS 1139: Part 1 Metal scaffolding. Specification for tubes for use in scaffolding.
(Withdrawn October 1990 and replaced by BS EN 39.) BSI, London.
BSI (2001) BS EN 39: Loose steel tubes for tube and coupler scaffolds. Technical delivery
conditions. BSI, London.
BSI (2003a) BS EN 12811-1: Temporary works equipment. Part 1: Scaffolds – Performance
requirements and general design. BSI, London.
BSI (2003b) BS EN 12810-1: Façade scaffolds made of prefabricated components. Part 1:
Products specifications. BSI, London.
BSI (2003c) BS EN 12810-2: Façade scaffolds made of prefabricated components. Part 2:
Particular methods of structural design. BSI, London.
272
BSI (2005a) BS EN 74-1: Couplers, spigot pins and baseplates for use in falsework and
scaffolds. Couplers for tubes. Requirements and test procedures. BSI, London.
BSI (2005b) BS EN 1993-1-1: EuroCode 3. Design of steel structures. General rules and
rules for buildings. BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
NASC (National Access and Scaffolding Confederation) (2008a) TG20 Guide to good
practice for scaffolding with tube and fittings. NASC, London.
NASC (2008b) TG20: Overview Toolbox Talk. NASC, London.
NASC (2010a) SG4:10 Preventing Falls in Scaffolding. NASC, London.
NASC (2010b) SG4: You. User Guide to SG4:10 Preventing Falls in Scaffoldings. NASC,
London.
NASC (2010c) SG25: Access and Egress from Scaffolds. NASC, London.
NASC (2011) TG20 Supplement 1, The effect of the introduction of the European Wind
Code BS EN 1991-1-4: 2005 on Basic Scaffolds and TG20; Appendix H. NASC,
London.
273
Chapter 22
Falsework
Peter F. Pallett Pallett TemporaryWorks Ltd
Godfrey Bowring Consultant
Falsework comprises the temporary structure used to support other structures, usually
permanent, until they can support themselves. Falsework is loaded for a short term,
often highly stressed, and after loading has to be destressed under load. Constructed
generally of pre-used equipment, the skeletal nature of such structures makes their
stability during erection, use and dismantling a crucial design consideration. In
certain cases they may rely on parts of the completed permanent works for stability.
The design philosophy of falsework differs to that for permanent works, and good
procedures and workmanship can contribute significantly to its safety.
22.1. Introduction
Any temporary structure used to support a permanent structure while it is not self-
supporting is known as falsework. Although usually considered as the support of in
situ concrete for slabs, bridges, etc., it is also used to support precast units/segments,
structural steelwork and timber structures and as backpropping. Falsework has a
different design philosophy to that used in the permanent works: it is loaded for a
short time; rarely is it an item in a Bill of Quantities; it often uses reusable components;
it has the unique structural requirement to be de-stressed under load (so that it can
be removed); it is rarely tied down and therefore relies on its own weight for
stability; and it is usually stressed to 90% of its safe capacity. Often assembled from
proprietary components, the falsework is then re-used. The design process has to
make allowances for erection tolerances and take into account that the components
are re-used many times.
The first document in the world of falsework TR4 (CS/ISE, 1971) was a Concrete Society
Report. It is interesting to note that it proposed a number of classes of falsework. A BSI
code committee on falsework was formed shortly afterwards in 1972. Coincidentally, this
was only two weeks before the collapse of heavy falsework over the River Loddon near
Reading, which in turn generated a Government Advisory Report known as the Bragg
Report (Bragg, 1975). BS 5975 was eventually published in 1982, taking account of the
recommendations in that report. It contained procedures as well as sufficient technical
content to enable a permissible stress design of falsework to be completed without
reference to other sources of information. The current version of BS 5975: 2008 þ A1:
2011 (BSI, 2011) has made clear that the procedures apply to all temporary works,
including falsework (see Chapter 2).
275
A limit state code on falsework, BS EN 12812 (BSI, 2008) was first published in
December 2004 and classifies falsework design into three classes as A, B1 and B2, but
makes no provision for the safe management of either falsework or temporary works.
Note that the existing code BS 5975 exists in parallel with BS EN 12812.
This chapter provides information on the design and use of falsework in the UK.
The different philosophy of the two design codes means that you cannot combine them.
Permissible stress uses the elastic properties of the material, whereas limit state uses the
plastic properties.
Permissible stress (BS 5975 (BSI, 2011)) assumes that the applied load is less than the safe
working load of the material, defined
failure load
safe working load ¼
factor of safety
The usual recommendation for the factor of safety for temporary works is 2.0.
Limit state (BS EN 12812; BSI, 2008) assumes that the design value of the loads is less
than the design resistance at ultimate load. Allowing for the partial safety factors, we
have
characteristic strength
ðactual loadsÞ ðpartial safety factorÞ 4
ðpartial material factorÞ ðclass factor)
where the partial safety factor is 1.35 for self weight and 1.50 for all other loads and the
partial material factor for both steel and aluminium falsework is 1.1. BS EN 12812
introduces different class factors for Class B1 (1.0) and B2 (1.15).
Note that the partial safety factors used in falsework are not the same as those used in
permanent design, and that the class factor for B2 is a 15% penalty for designing
276
Note that using a factor of safety of 2.0 against failure in permissible stress is not directly
comparable with using a combined 1.50 1.1 ¼ 1.65 factor in limit state design, as the
former is based on the elastic properties of the material and limit state uses the plastic
properties.
What happens if the checking organisation uses BS 5975, but the falsework has been
designed to BS EN 12812? The philosophies and factors discussed in Section 22.1.1
are different; you cannot ‘pick and choose’. The design method and the falsework
class to be considered therefore needs to be established at a very early stage of procure-
ment (see Chapter 2 on management).
Proprietary aluminium support systems and bearers from suppliers, which are being
increasingly used, have high strength to weight ratios. Since aluminium is more ductile
than steel, it results in greater elastic deflection and the stiffness is more critical in
designs. To utilise the higher load capacities, the vertical legs are often spaced further
apart. This may not always be economic when you consider other factors such as
providing access for operatives to strike out the soffit forms and the permanent work
deflection limits as set in the specification. Figure 22.2 depicts a typical example of an
aluminium support system.
277
Figure 22.2 Typical partially braced proprietary aluminium falsework with ledger frames
Courtesy of Jim Murray, PERI (UK) Ltd
278
The majority of aluminium systems rely upon the soffit formwork being restrained by
elements of the permanent works to provide lateral restraint for the falsework. This is
known as ‘top restrained’ falsework and is discussed later in this chapter. Refer to
BS EN 12813 (BSI, 2004) and BS 5975 (BSI, 2011) for the methods of structural
design for load on bearing towers when using prefabricated components.
Adjustable aluminium props are now used extensively in falsework, mostly for building
construction. Their low weight and high capacity make them easily transportable and
economic to use on site as falsework and as individual back props. Many of the
systems have ‘gates’ or ‘ledger frames’ to connect the legs together. BS 5975 refers to
these as ‘bracing frames’. The structural capacity of such systems should be obtained
from the suppliers. ‘Ledger frames are often fitted as aids to erection and, while not
designed as such, may contribute to the stability of the falsework. Designers and users
should be clear as to their purpose. When joined together, the systems often make
tables of soffit formwork. This is discussed in more detail in Chapter 24 on soffit
formwork.
As already inferred, components of steel and aluminium should never be mixed because
of their different elastic properties. The steel will not compress elastically or shorten
sufficiently under load (being elastically stiffer than the aluminium) to allow the
aluminium to carry its ‘share’, implying that the majority of the load would be carried
by the steel.
Loads on falsework described in BS 5975 (BSI, 2011) are compatible with BS EN 12812
(BSI, 2008). Loads can either be permanent actions, i.e. self weight, or variable actions,
for example, the supported structure and weight of operatives.
279
For design purposes, BS EN 12812 (BSI, 2008) recommends the density of concrete to be
taken as 2500 kg/m3 (i.e. assume 25 kN/m3). Where smaller building slabs are under
construction, the flat slab guide (CS140; CSG, 2003) recommends that the concrete
density for backpropping be 24 kN/m3 but that the falsework supports be designed for
25 kN/m3.
g Working area load: To allow for the imposed load from construction operations, a
distributed load is applied over the whole working area of 0.75 kN/m2 (i.e. service
class 1).
g Variable transient in situ loading allowance: applied over an area 3 m 3 m
between 0.75 kN/m2 for slabs up to 300 mm, increasing to a maximum of
1.75 kN/m2 for solid slabs greater than 700 mm, when placing in situ concrete.
Intermediate thicknesses are designed on 10% of the slab self weight (BSI, 2011;
clause 17.4.3.1). See Figure 22.3 for a depiction of working area and transient in
situ concrete loading. The implication of this transient load when using permanent
3m
3m
(0.75 kN/m2 to 1.75 kN/m2)
Class 1 minimum
(0.75 kN/m2)
280
metal decking formwork, often used on up to 3 m spans, is that all such metal
decking should be designed for a minimum imposed load of 1.5 kN/m2.
g Striking formwork load: to assist striking and handling of individual items from
under a soffit, BS 5975 (BSI, 2011; clause 19.1.1) recommends a working platform
fitted about 2 m underneath the soffit forms, designed for an imposed service
class 1 load of 0.75 kN/m2.
The basic equation for the peak wind velocity pressure qp at a location for a falsework is
established from the following relationship:
where qp is the peak velocity pressure (N/m2); cprob is the probability factor (a minimum
value of 0.83 is recommended); ce(z) ce,T is the combined exposure factor (BSI, 2011;
clause 17.5.16) and Swind is the wind factor at the location (m/s) (see Section 21.4.3.3
in Chapter 21). The wind factor Swind incorporates the fundamental wind velocity
(Vb,map) for the site, a topographical factor to allow for the terrain and the altitude
(m) of the site (BSI, 2011; clause 17.5.1.33).
Generally, placing concrete on falsework will not occur when the wind speed exceeds the
operating limits set for the construction plant. This is usually at Beaufort Scale Force 6
and corresponds to a wind speed of 18 m/s. This is known as the working wind, which
generates a working wind pressure of 0.20 kN/m2.
The procurer will have previously decided whether the falsework will be unclad, sheeted
or debris-netted. This will have a significant effect on the magnitude of the wind forces
which have to be considered.
g for steel falsework, c. 0.5 mm/metre of height (þ0.5 mm per joint þ 1.0 mm per
timber joint)
g for aluminium falsework, c. 0.9 mm/metre of height.
281
BS EN 12812 (BSI, 2008; clause 9.3.4.2) for Class B2 falsework considers the sway imper-
fection ’ for structures taller than h ¼ 10 m is calculated using
rffiffiffiffiffi
10
tan ’ ¼ 0:01
h
where h is the height (m) and ’ is the angular deviation from the theoretical line. Where
the structure is <10 m, a lower limit of tan ’ ¼ 0.01 is considered (i.e. use 1% of applied
loads for falsework up to 10 m in height).
Note that the foreword to BS EN 12812 (BSI, 2008) states (referring to BS 5975) that ‘the
application of this force has made a significant contribution to the safe use of falsework
since its introduction’.
282
1 the structural strength of the members and connections, i.e. are the node points
restrained and are web stiffeners fitted to beams?
2 the lateral stability of the falsework structure
3 the overall stability
4 the positional stability, i.e. will it slide.
The method adopted for the design, whether limit state or permissible stress, will
consider these four checks which are written into BS 5975 (BSI, 2011; clause 19.4.1.1)
and are recommended as the starting point of any falsework design. The importance
of an accurate and relevant Falsework Design Brief, prepared usually by the TWC
(see Chapter 2 on management) for use by both the Temporary Works Designer and
the design checker cannot be overstressed. A good brief will provide safe, economic
and effective falsework structures.
Many of the proprietary falsework systems have patented joints which provide some
moment capacity. These can reduce the effective length factor below unity, thus
increasing their safe load capacity. Follow the recommendations of the suppliers; in
particular, look out for unrestrained cantilever extensions on falsework at the head
and bases. The effective length of such members in cantilever depends on whether or
not the structure can sway relative to the cantilever tip; see BS 5975 (BSI, 2011; clause
19.4.2.4.4).
For structural steel beams carrying concentrated loads, web stiffeners should be
provided at all loading transfer points including supports, unless calculations are
provided to show that such stiffeners are not required (BSI, 2011; annex J). The omission
of web stiffeners has been the cause of fatal accidents, so their importance must not be
underestimated.
283
284
The three load cases depicted in Figure 22.4a–c all illustrate free-standing falsework
structures with diagonal bracing providing the restraint. There are falsework systems
or arrangements that rely on the top soffit formwork to provide the lateral stability.
These are known as top-restrained falsework, depicted in Figures 22.2 and 22.4d. The
assumption made by the designer in this case is that the head of the falsework is
restrained from movement in both lateral directions. This requires that the permanent
works are able to resist these notional stability forces, which is why Table 2.2 in
Chapter 2 excludes top-restrained falsework from design check Category 1 (simple
design); top-restrained falsework requires the approval of other parties, generally
including the designer of the permanent works. More information on top-restrained
falsework is given in BS 5975 (BSI, 2011).
All parties should be absolutely clear as to whether falsework has been designed on the
assumption that it is top restrained and, if so, whether the permanent works has been
designed (i.e. it is stiff enough in the temporary condition?) to provide the required
restraint. If not, the falsework should be designed as free standing.
Lateral stability also applies to the webs of steel beams at reaction points and at positions
of concentrated loads. Web stiffeners are required at all loading transfer points,
including supports, unless calculations are provided to show that such stiffeners are
not required (BSI, 2011; annex J). Note that the default condition for steel beams with
concentrated loads is to fit them with web stiffeners unless proven otherwise. The
comment at the end of Section 22.4.2 applies equally here.
The compression flanges of all beams should also be considered for lateral torsional
buckling. Permanent Works Designers (PWD) are aware of the potential instability of
steel beams on composite bridge designs (where relatively small top flanges are designed
for shear connection to the concrete), but not necessarily for the compression loads
during erection which can require additional lateral bracing. See BS 5975 (BSI, 2011;
annex K) for the effective length of steel members in axial compression.
When using systems that have cantilever working platforms, there is always a risk of
overturning. This is particularly the case when narrow aluminium tables are used,
where it may be necessary to tie these down to the slab.
285
where Ff is the frictional restraint force (kN), W is the vertically applied force (kN), R
is the reaction normal to the surface, is the minimum angle (degrees) to the horizontal
at which sliding will commence and P (kN) is the force applied to overcome the static
friction (Figure 22.5). The friction restraint does not depend on area of contact. The
applied load multiplied by the coefficient of static friction gives the value of frictional
restraint (the value at which it slides) with no factor of safety.
BS 5975 (BSI, 2011) recommends that minimum factor of safety against sliding is 2.0,
whereas BS EN 12812 (BSI, 2008; Table 2) gives different partial safety factors for stabi-
lising (
i ¼ 0.9) compared to destabilising loads (
i ¼ 1.35–1.5).
g to ensure stability during erection and dismantling (minimum lateral force 0.5%
of W )
g to improve the load-carrying capacity of the elements by stabilising node points
and controlling their effective lengths (normally a notional force of 2.5% of W )
286
The important aspect to remember when designing falsework is that the above four
criteria are not cumulative; checking for one criterion may actually be sufficient for all
four! For example, bracing a free-standing falsework sheltered from the wind for the
minimum horizontal disturbing force of 2.5% of W would also satisfy the structural
requirement for creating node points (also 2.5% of W ).
All of the above failures of falsework could have been avoided had correct procedures for
checking/inspection by competent persons and trained TWCs been adopted. The use of
more proprietary items with patented connections means that the integrity of the whole
structure may be impaired by the omission of a bolt or wedge, or even just failing to
tighten them up correctly. The supplier’s advice should always be sought.
Care must be taken when erecting and dismantling the falsework. The NASC booklet
SG4:10 (NASC, 2010) is a useful guide on preventing falls in both scaffolding and false-
work. The size of falsework structures means that there should be some clearly defined
stages in the construction. For example, formal checks might need to be established
after completing the foundations, when it reaches its support level and, obviously,
prior to loading. In addition to the provision of safe working areas for erecting, striking,
etc., are the conditions of use as envisaged in the design brief ?
287
A common error on falsework is failing to take account of the elastic settlement under
load (Section 22.3.6). For this reason, experienced erectors will set the top level of
the falsework higher than the final required level to allow for shortening under load.
For example, a reservoir roof slab support 5.4 m high would be expected to elastically
shorten by c. 5–6 mm if steel falsework was used, increasing to nearer 10 mm if
aluminium falsework was used.
Safe falsework requires competent experienced people, with an eye for detail. Failures
usually occur because small items are missed or forgotten or through lack of proper
process and control.
REFERENCES
Bragg SL (1975) Final report of the Advisory Committee on Falsework. HMSO, London.
BSI (1999) BS EN 1065: Adjustable Telescopic steel props: Product specifications, design
and assessment by calculation and test. BSI, London.
BSI (2002) BS EN 1991-1-1: EuroCode 1. Actions on structures. General actions. Densities,
self-weight, imposed loads for buildings. BSI, London.
BSI (2003) BS EN 1991-1-3: EuroCode 1. Actions on structures. General actions. Snow
loads. BSI, London.
BSI (2004) BS EN 12813: Temporary works equipment. Load bearing towers of prefabri-
cated components. Particular methods of structural design. BSI, London.
BSI (2008) BS EN 12812: Falsework: Performance requirements and general design. BSI,
London.
BSI (2010) BS EN 1991-1-4: 2005 þ A1: 2010 EuroCode 1. Actions on structures. General
actions. Wind actions. BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
Burrows M, Clark L, Pallett P, Ward R and Thomas D (2005) Falsework verticality:
leaning towards danger? Proceedings of ICE, Civil Engineering, Feb 2005: 41–48,
Paper 13624.
CS (Concrete Society) (1999) Checklist for Erecting and Dismantling Falsework. Concrete
Society Construction Group Leaflet CS123. Concrete Society, Camberley.
CS/ISE (Concrete Society and the Institution of Structural Engineers) (1971) Falsework:
Report of the joint committee. Technical Report TRCS 4. The Concrete Society, London.
CSG (Concrete Structures Group) (2003) Guide to Flat Slab Formwork and Falsework.
Concrete Society Ref No. CS 140. Concrete Society, Camberley.
NASC (National Access and Scaffolding Confederation) (2010) SG4:10 Preventing Falls in
Scaffolding. NASC, London.
288
Chapter 23
Formwork
Laurie York Consultant
Peter F. Pallett Pallett TemporaryWorks Ltd
‘Formwork’ is the term used to describe the fabrications and constructions used to form
the shape of concrete structures, acting as the mould. It is normally removed once the
concrete has achieved sufficient strength, although it is sometimes left in place as
permanent formwork which may or may not contribute to the structural capacity of
the formed concrete. Formwork is either vertical for walls, columns, beam sides, etc.
or horizontal for supporting slabs, cantilevers, underside of beams, etc. Horizontal
formwork is described in Chapter 24 on soffit formwork.
23.1. Introduction
Traditional formwork is fabricated using timber-based materials but steel, glass-fibre-
reinforced plastics (GRP), glass-fibre-reinforced cement (GRC) and other materials
are also used. Proprietary formwork includes the walings, bearers, soldiers, supports
and various panel systems available from specialist formwork supply companies.
Choice of material is often based on the required number of uses of the formwork and
the finish specified for the concrete by the Permanent Works Designer, with more
expensive materials often justified by more onerous surface finish and durability require-
ments for the concrete.
Designing formwork and ensuring its subsequent control on site requires a thorough
understanding of the pressures generated during the placing, compacting and setting
of concrete. Knowledge of formwork materials and practical experience are both essen-
tial to ensure that formwork is economic, easy to fabricate and use. The procedures for
and management of formwork are described in Chapter 2.
289
Working platform
(min 600 mm) with
toeboards, guardrails Vertical soldiers
Tie rod centres
and end guardrails
The span of the
soldiers will be
the centres of
the tie rods
Waler plate
Bearers/
walings
The span of the
face contact
material is the Through tie
centres of the assembly
bearers/walings
Push–pull
Starter props
bars
Span of the
bearers/walings
Kicker
Vertical
soldiers Stop ends not shown for clarity
walings and vertical stiff steel soldiers. The tie rods balance the forces and the inclined
props provide stability and allow alignment of the forms.
Where there is only one face of the form, such as against an existing wall or the ground,
then the formwork is known as single-faced formwork. The forces can be significant and
are discussed in Section 23.8.3.
23.3. Economy
Many people in the process can affect the economy of the construction. These include the
Permanent Works Designer, the contractor, the subcontractor and specialist suppliers,
as well as the site operatives. The cost of formwork will vary depending on the contract
specification and also on the number of uses required of the formwork. On a concrete
frame building, the cost of formwork and falsework can be 39% of the cost of the
structure, increasing on a ‘civils’ contract to as much as 55% of the cost of the concrete
structure. A better-quality face contact material may be more expensive but may be
necessary either to produce the required finish on the concrete or to enable re-use of
the formwork without costly refurbishment. Similarly, a more robust form may be
more expensive but allow more re-use. A considerable percentage of the costs is
associated with the labour involved; formwork design that reduces the labour
290
requirement will generally be more economic. Examples include the reduction in the
number of tie rods, the use of proprietary panels to reduce initial make-up costs and
use of aluminium bearers to reduce the weight of the formwork panels (permitting
larger form areas for the same size of crane).
British Standard BS EN 13670 Execution of concrete structures (BSI, 2009) defines the
generic surface finishes. The two main UK specifications are: National Structural
Concrete Specification for Building Construction (NSCS) (ConStruct, 2010; Section
8.6) and the Highways Agency specification (HA, 2004; volume 1, clause 1708).
Table 23.1 compares the two specifications. Other specifications used include the UK
Water Industry Research Civil Engineering Specification for the Water Industry
(CESWI, 2011; clause 4.28) and the National Building Specification (NBS) Formwork
for in situ concrete (Section E20) which is published annually online.
The NSCS (ConStruct, 2010) refers to full-size reference panels at seven locations in the
UK which may be viewed by visiting the various locations. Other sources of information
include the Concrete Society Technical Report 52, Plain Formed Concrete Finishes
(CS, 1999). As well as generic surface finishes, special finishes required under the contract
may be defined.
291
23.5. Tolerances/deviations
Tolerances specified for the formwork are not necessarily the same as those specified for
the permanent works. The three sources of deviations in the surface of the finished
structure are: (1) inherent deviations, such as the elastic movement of the formwork
under load; (2) induced deviations, such as the lipping between two sheets of plywood
(possibly within plywood manufacturing tolerances); and (3) errors, such as inaccurate
setting-out. Inherent deviations include deflection.
The acceptable magnitude of the deviations will depend on the specified quality of the work
and also the distance from which the concrete surface will be viewed. Typical deviations in
verticality for normal work will be 20 mm from a grid line for a 3 m high wall. More
information is available in Section 2.6 of the Formwork Guide (CS, 2011; Section 2.6).
The use of expanded metal steel products (Hy-rib) in construction joints, in place of
plywood, almost eliminates the need for scabbling (i.e. removes the risk problems
associated with vibration known as ‘white finger’) since the metal is left in place.
Trials by the British Cement Association have shown that there is a significant reduction
in concrete pressure when expanded metal products are used. See details in the
Formwork Guide (CS, 2011).
The use of fabric such as Zemdrain on the face contact material introduces control over
the permeability. This is known as controlled permeability formwork (CPF); see CIRIA
Report C511 (CIRIA, 2000). By allowing the pressure of the concrete to force the surplus
pore water and any air into the fabric and to be drained away, there is an almost
complete elimination of blowholes on the concrete surface. The face strength of the
concrete is increased by 30% in the critical cover zone of the concrete. A recent
292
development is the use of composite sheets having a foamed polystyrene core with a
synthetic thermoplastic material face to both sides or, for wall formwork, with a thin
reinforcing sheet of aluminium to both sides. Repair techniques on such sheets are
specialised.
23.6.2 Bearers
In wall formwork, bearers are typically aligned horizontally supporting the face contact
material; in such cases they are often referred to as ‘walings’. They are often softwood
constructional timber using strength classes C16, C24 or C27. Site conditions, such as
exposure to wetting, affect the strength of timber. The duration of loading and facility
for load sharing will also affect the apparent strength of the timber, as will the depth
of the timber section used. The Formwork Guide (CS, 2011) provides details of safe
working properties for various commonly used timber section sizes.
Alternatively, proprietary beams fabricated from timber, aluminium or steel can be used.
Many of the European systems of formwork incorporate proprietary timber bearers
fitted vertically. They should be used in accordance with the manufacturers’ instructions
at all times.
23.6.3 Soldiers
Soldiers can be timber or proprietary beams; in Figure 23.1, the soldiers have central
slots for the positioning of tie rods. The soldiers support the bearers which span hori-
zontally between them, either using design properties given in the Formwork Guide
(CS, 2011) or using information from manufacturers.
After use, the bar is removed and re-used. The safe load taken by the bar can be large (of
the order 110 kN) and waler plates of sufficient size will be required to spread the load
into the soldiers or bearers. The Formwork Guide (CS, 2011) recommends that, when
designing with these recoverable tie assemblies, a minimum factor of safety of 2 is
adopted based on the minimum guaranteed ultimate strength of the bars used. Alterna-
tively, safe working loads may be supplied by the tie manufacturer.
Other types of tie rods include taper ties, lost ties, wire ties, coil ties, mild steel all-thread
ties, etc. The lost-tie system utilises tapered-ended ‘she-bolts’ screwed to each end of
threaded tie rods. This allows the entire assembly to be placed through the wall form
from one side during erection and permits the she-bolts to be subsequently removed,
293
Waler plate
Knock-on nut
Lost spacer tube
Recoverable tie
leaving the tie rod embedded in the concrete. The length of the lost tie rod is specified to
suit the wall thickness and the cover required to the reinforcement.
It is important to note that suppliers of tie-rod systems will state recommended safe
working loads for the ties used in tension in wall formwork. Whenever used in other
applications, the safe loads will alter as the factors of safety increase, and reference
must be made to the tie-rod supplier.
Formwork comprising proprietary panels gives a greater potential for re-use for the
plywood by reducing damage during handling. Panels are either small enough for
manual handling (say 1200 mm 600 mm), or they may be large crane-handled panels
(up to 2400 mm 2700 mm). The main advantage is the speed of initial use, with little
‘make-up’ time. It should be noted that such panel systems, when on hire, will generally
give ‘ordinary’ finish (Class F2, suitable for rear, unseen faces of retaining walls), rather
than ‘plain’ finish (Class F3, F4, F5, suitable for quality surfaces to walls). They will often
have far fewer components and they have the significant benefit of being available on hire.
294
295
P P
P P
the site team so that those placing the concrete will be able to limit the rate of placing the
concrete such that the design pressure is not exceeded. Drawings of formwork should
state the maximum pressure in units of kN/m2, and users of proprietary panels should
be aware of the limiting design pressure for their particular arrangement. Changing
the tie-rod diameter and/or locations can affect the limiting pressure on the proprietary
panel system. Whenever a fluid such as wet concrete is placed in a vessel, the pressure
exerted by the fluid at any depth P (Figure 23.4) is defined:
P¼hD
where P is the pressure (kN/m2); h is the depth (m) and D is the density (kN/m3). For
concrete, the density is taken as 25 kN/m3. Under 3 m head of fluid concrete, the pressure
is therefore 75 kN/m2.
The pressure always acts at right angles to the face. The force on the formwork can there-
fore be calculated as pressure area acted upon.
For the pressure imposed by concrete on formwork, basic research was published in
CIRIA Report R108 (Harrison and Clear, 1985). Since then, there have been changes
296
Pmax Pmax
P P
to the types of cement and to concrete specifications (to suit BS EN documents, etc.).
University of Dundee research (Dhir et al., 2004) concluded that the methods adopted
by CIRIA R108 were safe for use, subject to understanding of the type of concrete
used. Concrete mixes using different cement combinations were first classified into
three groups for purposes of concrete pressure determination by Pallett (2009). The
groups are
For the specified concrete, temperature, form height, required rate of rise of the concrete
within the form and the maximum concrete pressure Pmax can be determined. On site, the
formwork arrangement is known, the limiting concrete pressure Pmax is known (from
drawings or from suppliers’ data sheets) and the concrete group is known, but the
concrete temperature can vary from day to day. To limit the maximum design pressure,
the rate of rise of the concrete within the form must be determined so that the required
rate of delivery of the concrete can be arranged. There are two ways to determine the rate
of rise where the value of Pmax is known: either use the pressure tables or, more
297
298
Data courtesy of the Concrete Society
Formwork
299
Formwork
Formwork
accurately, use the Rate of Rise Tables RWA, RCA, RWB, RCB, RWC and RCC
published in the Formwork Guide (CS, 2011; section 4.4.3).
Note that where pour heights are small (such as in building work), if the formwork is
designed for full fluid head (Figure 23.4) then there is no limit to the speed of concreting
as it is not possible to generate a pressure larger than fluid head. Designing a 3 m column
form for maximum 75 kN/m2 pressure would therefore not require control of the rate of
rise at any concrete temperature or type of concrete mix.
1 Check the safe span of the face contact material or plywood. This is usually
derived from tables plotting pressure against span. This obviously gives the
maximum spacing for the bearers fixed horizontally (Figure 23.6a) or vertically
(Figure 23.6b). The span of the face contact material is taken as being the spacing
of the centres of the bearers/walings.
2 Check the safe span of the bearers. This can also be derived from tables but more
economic designs will be produced by analysing the bearers as continuous beams
(where appropriate). It is good practice to design the bearers to project past the
last support by a maximum of 1/3rd of the adjacent span, since this improves the
economy of the design. The span of the bearers when used horizontally as walings
is taken as being the spacing of the centres of the soldiers (or of the stiff
horizontal members when the bearers are fixed vertically).
300
3 Check the design of the soldiers. The tie rod spacing determines the span
requirements for the soldiers but additional checks may be required if, for
example, there are joints in the soldiers. As the design pressure reduces near the
top of the wall, fewer ties are required (see Figure 23.7). The layout of ties (and
therefore the soldiers) may have been specified by the Permanent Works Designer
to suit, for example, a particular surface finish with a regular pattern. The span of
the soldiers is taken as being the spacing of the tie rods.
4 Check the capacity of the tie rods for the chosen arrangement of soldiers.
5 Check the deflection of the face of the forms. With such large pressures being
imposed on the formwork system, the design should take careful account of
deflections with reference to the acceptable tolerances (see Section 23.5). It should
be remembered, however, that the deflection limit (normally 1/270th of the span)
is applied individually to the face contact material, to the bearers and then to the
soldiers; it is not cumulative.
Other considerations to be observed during the design process will include safety issues
(working platforms for placing the concrete, CDM Regulations, Work at Height Regula-
tions) and handling of the formwork (weight limit for manual handling and crane or traveller
capacity for mechanical handling). See Chapter 1 on safety and the section on mechanical
handling of formwork in the Formwork Guide (CS, 2011; section 5.9).
301
A simplified method to establish the wind forces (both ‘maximum’ and ‘working’)
on freestanding wall formwork can be found in the Formwork Guide (CS, 2011;
section 4.5.1).
The most common method of support for small single-faced formwork is to use inclined
props as shown in Figure 23.8. Because of the inclined props, there will always be a
Walings
Kicker
302
Proprietary frames
with adjustable feet
and twin angled ties
into fixing in base
Stop end and
its support
not shown Joint
Walings
corresponding and significant uplift force at the base of the form. This introduces the
need for fixings into the existing slab to provide restraint against the uplift, as well as
providing the necessary horizontal restraint.
Control of movement of the top of the single-faced formwork is always difficult. Care
is required to ensure that the specified tolerances are not exceeded. Loads imposed on
the existing slab by the anchors, fixings and props are significant; approval from the
Permanent Works Designer should always be sought.
Columns may have circular, square, rectangular or other irregular cross-sections. These
irregular cross-sections can often be formed by placing inserts or box-outs within
standard square or rectangular column moulds. Circular column forms are available
made from steel, GRP or single-use cardboard forms; the latter is particularly useful if
303
only a few columns are required. Steel and GRP column forms can incorporate splayed
column heads.
Variations in column shapes and sizes, particularly on the same project, should be kept to
a minimum. Ideally, column sizes should be selected in modular increments of 75 mm
(e.g. 300, 375, 450, etc.) enabling both the designer and the builder to standardise (and
hence economise).
A range of column forms using proprietary panels is available, with some designed to
cater for varying rectangular shapes by adjustment on site using ‘universal panels’
arranged in a ‘windmill’ pattern. As panel formwork is designed for a specific limiting
concrete pressure, it may be beneficial to sequence the construction of the columns
with other site concreting work and/or to pour multiple columns simultaneously to
reduce the risk of overloading the column formwork. Restricting the pour rate so
that the allowable pressure on the forms is not exceeded may not always be a workable
solution, and specifically designed formwork may be more appropriate.
A lower criterion is generally possible for striking forms with basic or ordinary concrete
finish (Classes F1 and F2; BSI, 2009), usually related to the maturity of the concrete. In
practice, provided the mean concrete temperature is above 108C overnight, the vertical
formwork can be struck next morning. If the mean air temperature is above 108C, it
can safely be assumed that the concrete temperature was above 108C. Where the
newly struck concrete may be exposed to frost, a minimum in situ concrete equivalent
cube strength of 2 N/mm2 is recommended.
Note that certain project specifications, for example, the Specification for the Water
Industry (CESWI, 2011), may require higher strengths prior to removal of formwork.
23.11. Workmanship/checking
To produce concrete of a high standard of appearance and precision, a high standard of
site workmanship is required; this will apply even if formwork is specially constructed for
304
the job in question. Where the amount of work does not justify the use of specially
designed and fabricated formwork, the degree of success in achieving the required
result will depend solely on the skill and expertise of the site operatives. In a few
cases, the economic solution may be to fabricate a high-quality form for only one or
two uses.
There are cases, however, when good-quality work can be achieved with less skilled
labour. These will normally be where work is simple, of a highly repetitive nature and
where the size of the job justifies the use of sophisticated special-purpose formwork
which is designed for simplicity. A considerable amount of the work carried out on
site does not demand good appearance or close tolerances, although appreciable skill
and experience is still necessary to ensure that the formwork is stable, safe and to the
required standard.
All blemishes in the formwork will appear on the surface of the finished concrete.
Blemishes will obviously not become apparent until the formwork is struck, by which
time the concrete has hardened making the repairs difficult.
When using proprietary systems, it is important that site management ensure that
the operatives construct the systems correctly using the manufacturer’s approved
components. Once the forms have been constructed the site must ensure that the
concrete placement methods are appropriate for the system being used, and are being
used within the pressure limits set by the manufacturer and/or formwork designer (see
also Chapter 2).
REFERENCES
BRE (Building Research Establishment) (2007) Formwork for Modern, Efficient Concrete
Construction. BR 495. HIS BRE Press, Bracknell.
BSI (2009) BS EN 13670: Execution of concrete structures. BSI, London.
CESWI (Civil Engineering Specification for the Water Industry) (2011) Civil Engineering
Specification for the Water Industry, 7th edition. WRc plc, Swindon.
CIRIA (Construction Industry Research and Information Association) (2000) Controlled
Permeability Formwork. Joint Report CIRIA/Concrete Society, Report C511, London.
ConStruct (Concrete Structures Group) (2008) A guide to the safe use of formwork and
falsework. Ref: CSG/005. Concrete Society, Camberley.
ConStruct (2010) National Structural Concrete Specification for Building Construction, 4th
edition. Ref. CCIP-050. Concrete Society, Camberley.
CS (Concrete Society) (1999) Plain Formed Concrete Finishes: Illustrated Examples.
Concrete Society Technical Report No. 52, Crowthorne.
CS (2003) Checklist for the assembly, use & striking of formwork. Ref. CS144. Concrete
Society, Crowthorne.
305
CS (2011) Formwork: A Guide to Good Practice, 3rd edition. Special Publication CS 030.
Concrete Society, Camberley.
Dhir RK, McCarthy MJ, Caliskan S and Ashraf MK (2004) Design formwork pressures
for the range of new cement, superplasticised and self-compacting concretes. DTI
research contract No. 39/3/739 (CCC2399) University of Dundee Report CTU/3004.
HA (Highways Agency) (2004) Manual of Contract Documents for Highways Works.
Volume 1 Specification for Highways Works. HMSO, London.
Harrison TA and Clear C (1985) Concrete pressure on formwork. Construction Industry
Research and Information Association, Report 108. CIRIA, London.
NBS (National Building Specification) (annually) Formed finishes, Section E20. National
Building Specification Ltd, Newcastle-upon-Tyne.
Pallett PF (2009) Concrete groups for formwork pressure determination. Concrete 43(2):
44–46.
306
Chapter 24
Soffit formwork
Peter F. Pallett Pallett TemporaryWorks Ltd
The support to the underside of in situ concrete, flat or inclined, requires formwork to
contain and mould the concrete to the desired shape. Often supported on foundations
by means of falsework, the soffit formwork is an important part of temporary works.
The supports can only be removed when the concrete has gained sufficient strength, so
the procedure and sequence for striking needs to be understood and approved before
concreting.
24.1. Introduction
The forming of the underside of concrete structures while the concrete gains strength
utilises soffit formwork, either inclined or nominally level. Unlike vertical formwork
(discussed in Chapter 23), the concrete has to have gained sufficient strength before
the structure can support itself thus allowing the formwork to be removed. Generally,
in situ concreting on soffit formwork will have some supporting falsework (see
Chapter 22 on falsework) to transfer the load to suitable foundations. In certain cases,
such as permanent formwork or in the case of cantilever soffit formwork, supporting
falsework may not be required and the loads are transferred directly to the permanent
works.
Soffit formwork is found in civil structures and to the undersides of bridge decks, beams
and parapets. In building it is used to form the underside of slabs: ideally flat slabs or,
with moulds, slabs with downstands. It is also used to form the underside of beams
and in a wide variety of situations. A typical arrangement of soffit formwork to a
structure with supporting falsework is depicted in Figure 24.1.
This chapter provides information about the different philosophies of design and use of
soffit formwork for both civil and building applications. It includes bridge and slab
formwork, cantilever soffits and the latest considerations for striking of soffit formwork
in the UK.
307
Primary bearer
Span of the
primaries Span of the
secondaries
Consider the components of soffit formwork. The face contact material in contact with
the concrete could be plywood, wood-based panels (particleboard or oriented strand
board), plastic composites, steel or a proprietary panel system. These are usually
supported on bearers known as secondary bearers. In turn, the secondary bearers are
supported on more substantial bearers, known as primary bearers. The primary
bearers fit on to the falsework uprights, usually centralised in adjustable forkheads.
The primary bearers may be considered as part of the formwork and designed using
the Formwork Guide (FGTGP) (CS, 2011) or as part of the falsework (BSI, 2011); for
example, if the primary bearers are constructional softwood the safe load tables are
identical in FGTGP and BS 5975.
British Standard BS EN 13670 Execution of concrete structures (BSI, 2009) defines the
generic surface finishes. The two main UK specifications are the National Structural
Concrete Specification for Building Construction (NSCS) (ConStruct, 2010; section
8.6) and the Highways Agency specification (HA, 2006; volume 1, clause 1708). Table
23.1 in Chapter 23 provides a comparison of the two specifications. Other specifications
used include the UK Water Industry Research document Civil Engineering Specification
for the Water Industry (CESWI, 2011; clause 4.28) and the National Building Specifica-
tion Formwork for in situ concrete (NBS, 2011; section E20).
Particular care is necessary in specifying the surface finish for bridges, especially for their
parapets. The CIRIA report R155 Bridges: Design for improved buildability (CIRIA,
1996) recommends that the HA Class F3 finish be limited to small vertical areas of the
parapet, such as those visible from the highway. Unfortunately, if specified for the
entire parapet edge and soffit, the ‘no-tie’ requirement of Class F3 makes restraint of
such soffit formwork extremely complex and unnecessarily expensive, and often leads
to grout loss at connections with unsightly marks. The solution is to specify Class F4
for most of the parapet, leaving the small vertical upstand (visible) face as Class F3
finish. Realistic specifications will produce better results.
309
but also the two construction operations loads (see Chapter 22). These are (1) the
working area load distributed over the whole working area of 0.75 kN/m2 (i.e. Service
Class 1) and (2) the variable transient in situ loading allowance applied over an area
3 m 3 m, which varies from 0.75 kN/m2 for slabs up to 300 mm up to a maximum of
1.75 kN/m2 for solid slabs greater than 700 mm. Intermediate thicknesses are designed
on 10% of the slab self weight (BSI, 2011; clause 17.4.3.1).
24.3.2 Horizontal
In addition to vertical forces, soffit formwork can have applied horizontal forces. These
could be from surges in concrete pump lines, impact forces and possibly from
arrangements of the stop-ends. All slabs will have stop-ends and/or construction
joints. Particular care is necessary on all slab/deck stop-ends greater than 400 mm
depth.
When casting a slab against an existing structure or against a previous cast section, there
will be a lateral force generated from the reaction to the pressure of concrete (i.e. the
fluid head of concrete on the connection) acting on the existing face. For example, a
900 mm slab cast against an existing slab generates a lateral force of about 10 kN per
metre run of joint.
A common site error is to ignore discontinuities in the soffit formwork. Where there is a
discontinuity in the formwork, a lateral force (equivalent to the fluid head of concrete at
that point) exists laterally and attempts to move the forms apart. Generally the
arrangement of staggered bearers supporting the face material will provide the restraint,
but where a section of falsework and formwork is not connected (e.g. between tables of
formwork) or where the falsework is staggered to allow for changes in levels, discontinu-
ities can be formed. The solution is usually quite simple: join the sections of falsework
together below the soffit formwork. This is discussed in more detail in both the
Formwork Guide (FGTGP) (CS, 2011) and BS 5975 (BSI, 2011).
24.4. Design
The design of soffit formwork is different from that for walls as discussed in Chapter 23;
the loads are usually less. For example, a 900 mm thick in situ concrete bridge
deck imparts an imposed load of 22.5 kN/m2 on the soffit and a 250 mm slab only
6.25 kN/m2, compared to the pressures on the face of wall and column forms in the
order of 60–130 kN/m2. As a result, the supporting members will often safely span
greater distances although this will lead to larger deflections in the bearers. Often it is
the deflection criteria that will govern the span.
310
Figure 24.2 Beam reactions for one, two or three spans with distributed load
10 kN 10 kN 10 kN 10 kN 10 kN 10 kN
The basic design principle is that the loads are transferred from the face to the supporting
falsework. Refer to Figure 23.5 (Chapter 23) and consider the similar arrangement
turned through 908 to become a soffit.
The distribution of load into the vertical members from the soffit formwork bearers will
often be very random. The face contact material, the secondary bearers and often the
primary members in the forkheads will be continuous over several supports, giving
rise to increased reactions at internal supports from their elastic reactions. The
support reactions of beams change when they are continuous over more than two
supports, and this applies to all types of members used as beams.
Figure 24.2 shows the support reactions caused by a distributed load on each equal span
of 10 kN per span for one, two or three spans. The designer’s worst case is a single beam
continuous over two spans, i.e. with three supports (Figure 24.2b), giving a central
reaction of the static load times 1.25 for continuity, i.e. a staggering 25% increase in
load. The ‘10% continuity rule’ in BS 5975 (BSI, 2011; clause 19.3.3.2) accepts that, in
the case of falsework comprising random bearers with the formwork and falsework all
in various lengths, the vertical load is calculated on the area supported by the standard
plus 10% to allow for continuity. (Note that the 10% is added once, not for each level of
bearers.) In certain cases, for example, over two spans, a more precise calculation may be
justified. Certain proprietary systems incorporate simply supported beams and this ‘rule’
may not apply.
In the UK, most contractors will use table systems for flat slab construction; generally,
with aluminium beams in both directions at the top. A typical example is shown in
Figure 24.3. Table systems comprise large diameter aluminium props (typical diameters
are 100–150 mm) with long threaded sections to allow for adjustment, connected
together with ledger frames and/or cross-bracing.
The stiffness of the assembly is derived from long lengths of aluminium primary beams.
These tend to limit the lengths of table handled to 12 m with standard components, but
longer lengths are possible. Tables have the benefit that once made up they enable rapid
construction, and the benefits increase with repetitive use. They can be used with cross-
wall construction or columns, but are most economic when there is access to opposite
faces of the building for direct removal of the tables. Table systems ideally suit flat
slab and repetitive construction and become economic at over eight uses. Cycle times
as short as 4 days have been achieved with careful planning; see the flat slab guide
(ConStruct, 2003).
311
Tables need space to be ‘flown’ out of the building (either to one side or to both sides),
with a minimum end allowance of 500 mm for clearance from the building to adjacent
structures/objects and to allow for cantilever access platforms. A minimum clearance
to columns/walls of 40 mm per side should be allowed to the sides of each table; some
infill support is therefore necessary at arises when used with crosswall construction.
The various methods of handling tables are outside the scope of this book; see the
Formwork Guide (FGTGP) (CS, 2011) for detailed information. Consideration must
be given to the handling and operation of the system. For example, if handled as
individual tables, the lengths of aluminium bearers become critical in striking out after
pouring. Physically removing long lengths is extremely hard and places unnecessary
risk on the operatives. Note that BS 5975 (BSI, 2011; clause 19.1.13) recommends
that, when handling individual units, a working platform of about 2 m below the
underside of the soffit should be fitted.
312
method adopted, such as deflections caused by the elastic extension of the tie rods, etc.
The effective length of cantilevered beams is given in BS 5975 (BSI, 2011; table K.3)
where, for example, when the tip of the cantilever is free to rotate with the applied
load, the effective length is 7.5 times the actual length. Design is not easy, and considera-
tion of differential deflections and subsequent movements can become complicated.
Further guidance is given in the Formwork Guide (FGTGP) (CS, 2011).
Unlike wall formwork, in order to remove soffit formwork the structure has to be
capable of carrying its own weight plus any imposed load at the time of striking. The
Basic, ordinary, plain or Use specifications, codes of practice or tables (e.g. CIRIA R136, 1996)
special finish
HA all classes Assess the concrete strength at time of striking knowing the maturity,
and concrete mix, etc. using either a method based on pro rata the
strengths or, for flat slabs, an assessment based on crack width
313
total service load on the member at the time of striking will depend on the sequence of
construction and, in multi-storey construction, whether there is any backpropping.
The load can include the following.
Particular consideration should be given to the striking of suspended slabs which are
designed for light imposed loads; examples are roof slabs and reservoir roofs which
will not have earth cover. In these cases, the self weight may be the predominant load
and the construction operations load will represent a load very similar to the Perma-
nent Works Designer’s imposed load. For such slabs with very low design imposed
loads, it may not be possible to strike until the concrete approaches its characteristic
strength. Instances of early cracking and excessive deflections on reservoir roofs have
been attributed to incorrect early striking. The permanent works slab or beam will
generally be allowed to take up its instantaneous deflected shape before other loads
are applied.
Some proprietary systems actually have quick strip arrangements, allowing the expensive
formwork face components to be struck early (normally at a minimum concrete strength
of 5 N/mm2 to avoid damage) while leaving the main slab supported until approval to
strike is received. The management and control of such soffit schemes is discussed in
Chapter 2.
314
the design service load is similar to the ratio of concrete strength at the time of striking to
the characteristic concrete strength.
The assumption for loading a concrete slab is that the crack width is proportional to the
stress in the steel reinforcement which, in turn, is proportional to the load. Hence if load
is removed or added, there will be a proportional reduction/increase in crack width.
Although the slab is designed for the ultimate limit state, the actual maximum load on
the slab at the time considered will be the summation of the unfactored loads because
the consideration of crack width is at serviceability limit state, not ultimate. The load
applied to a slab during any stage of construction should not be greater than the
designer’s unfactored design service load. Obviously if the concrete slab is struck
earlier than intended, then the structure may be permanently damaged.
This method was first published by the BRE in BR 394 (Beeby, 2000) and is fully detailed
in the Formwork Guide (CS, 2011; section 5.3.7.2). Using this method of evaluating
crack width criteria will give a faster method for safely striking flat slabs than that
derived from the ratio of loads.
The method relies on accurately evaluating the actual concrete strength of the new
concrete. This subject is covered in more detail in both the Formwork Guide (CS,
2011) and the Flat Slab Guide (ConStruct, 2003). Cube strengths at an early age (say
19 hours) can be unrealistic and alternative methods of strength assessment at early
age (e.g. LOK test, maturity measurements or the Capo test) are preferred.
Where there are several other levels of construction, there may also be a requirement to
support additional loads. There will be occasions in multi-storey construction when the
slab immediately below the level to be supported has not achieved full maturity, and
construction loads will require to be supported through several levels. This is known
as backpropping; see BS 5975 (BSI, 2011).
1 slabs spanning between walls: commence striking at the middle, working towards
the walls
2 slabs supported on beams: strike the slabs commencing at the middle, working
towards the beams, then after slabs are fully struck strike the beams commencing
at the middle, working towards the columns
3 cantilevers: commence at the tip and strike towards the support.
315
REFERENCES
Beeby AW (2000) A radical redesign of the in-situ concrete frame process, Task 4: Early
striking of formwork and forces in backprops. The University of Leeds, Building
Research Establishment Ltd. Report BR 394, London.
BSI (2008) BS EN 12812: Falsework: Performance requirements and general design. BSI,
London.
BSI (2009) BS EN 13670: Execution of concrete structures. BSI, London.
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
CESWI (Civil Engineering Specification for the Water Industry) (2011) Civil Engineering
Specification for the Water Industry, 7th edition. WRc plc, Swindon.
CIRIA (Construction Industry Research and Information Association) (1995) Formwork
striking times: Criteria, prediction and methods of assessment. Report R136. CIRIA,
London.
CIRIA (1996) Bridges: Design for improved buildability. Report R155. CIRIA, London
ConStruct (Concrete Structures Group) (2003) Guide to Flat Slab Formwork and False-
work. Ref No. CS 140. Concrete Society, Camberley.
ConStruct (2008) A guide to the safe use of formwork and falsework. Ref: CSG/005.
Concrete Society, Camberley.
ConStruct (2010) National Structural Concrete Specification for Building Construction,
4th edition. Ref. CCIP-050. Concrete Society, Camberley.
CS (Concrete Society) (2003) Checklist for the assembly, use & striking of formwork.
Ref. CS144. Concrete Society, Crowthorne.
CS (2011) Formwork: A Guide to Good Practice, 3rd edition. Special Publication CS 030.
Concrete Society, Camberley.
HA (Highways Agency) (2006) Specification for Highway Works, Manual of Contract
Documents for Highway Works. HMSO, London.
NBS (National Building Specification) (2011) National Building Specification. NBS
Building, Newcastle-upon-Tyne.
316
Chapter 25
Climbing and slip forms
Jim Murray Engineering Director, PERI Ltd
Climbing and slip forms are systems for the construction of in situ reinforced concrete
vertical wall elements, several lifts high, that do not rely on support or access from
other parts of the permanent works. A combined formwork assembly and access
platform is either supported on anchors/tracks bolted to the previous section of wall
or, for continuous pouring, supported on climbing rods cast into the concrete. Protection
screen systems, either for use with climbing formwork or as separate entities, have similar
design considerations. The type of structure and suitability on repetitive wall elements
affects the economy of these systems.
25.1. Introduction
The operation of climbing formwork, often referred to as ‘jump form’, involves
individual pours with the formwork moved between the pours; alternatively, the
operation of slip form requires the concrete to be poured continuously and the forms
are raised to suit the speed of placing concrete.
Climbing or jump-form formwork (see Section 25.3) typically comprises the formwork
and safe working platforms for cleaning/fixing of the formwork and access for fixing
reinforcement and concreting works. Unlike conventional formwork discussed in
Chapter 23, it supports itself on the previously cast concrete and does not rely on
support or access from other parts of the structure or permanent works.
Climbing formwork is suitable for vertical elements in high-rise structures such as bridge
piers/columns or stair/lift shafts, core walls, shear walls, etc. in buildings. These are
constructed in a staged process. It is a highly productive system designed to increase
speed and efficiency while minimising labour and crane time. Systems are normally
modular and can be joined together to form long lengths to suit varying construction
geometries. Further detailed guidance on use of climbing formwork is given in
Formwork: A guide to good practice (CS, 2011; section 6.2).
Slip forming (see Section 25.4) is a system whereby the entire formwork system with
its safe access platforms are incrementally jacked upwards as the reinforcement and
concrete is placed. Slip forming is used for constructing chimneys, silos, water tanks
and shaft linings as well as towers, lift shafts and bridge piers. There are no construc-
tion joints, which is particularly important for leak-free structures. Fast rates of
production can be achieved, but these systems require more initial site preparation
317
and are generally carried out by specialist subcontractors. See also the Good Concrete
Guide 6 (CS, 2008).
Slip-form formwork has a longer initial set-up time than climbing or jump forms and will
be more expensive per square metre of formwork, but the resulting equipment cost per
square metre of slip-formed surface will be much lower. As it is often a 24 hour process
in shifts, the labour costs can be higher but production rates will be faster. Slip-forming
will usually suit tall structures that are not less than 20–25 m tall and which do not have
frequent plan changes. Slip-forming may be economic for structures as low as 12 m tall,
or even lower if several identical structures can be constructed in sequence.
The construction in advance of such core/lift shaft walls can provide stability to the main
structure during its construction and can have the beneficial effect of taking the core off
the project critical path. This then permits different phases of work to be carried out
concurrently without interference.
318
The benefits of reduced demand on the use of cranes and the efficient construction
method achieved due to the repetitive nature of the construction method leads to fast
cycling times. Where sites have sufficient ground level space and a well-considered
construction program linked with good site management and experienced operatives,
using more traditional techniques can be just as effective.
The Permanent Works Designer (PWD) is the only person who has sufficient knowledge
of the structure to assess whether it can withstand the loading induced, particularly by
the climbing formwork system. The Construction (Design and Management) (CDM)
coordinator has a duty to ensure that the permanent works and temporary works
designers (TWD) cooperate, and that the permanent structure can withstand the loads
applied to it by the temporary works (see Chapter 2 on management). The temporary
works coordinator (TWC) has an important role at this stage, as it is highly probable
that the use of climbing formwork as a method of construction was not considered
during the structural design phase. The PWD should verify whether the vertical elements
can support climbing formwork, albeit within defined constraints (e.g. under extreme
319
wind conditions). However, it is the responsibility of the PWD to review their design and
liaise with the contractor’s TWD team.
The PWD also has a role to play when the contractor is considering the climbing cycle or
rate of slip. The concrete mix design and the early age strength of the concreted section
both have a significant impact on whether a short cycle time/climbing rate is achievable.
The cycle time can be a critical factor on the effectiveness of the system. For example, if
the climbing system cannot be climbed after 4 days it may not be a viable solution in
relation to the costs involved, particularly when large systems utilising hydraulics are
proposed. When slip-forming, materials such as concrete, reinforcement, hoist exten-
sions, etc. must suit the rate of construction envisaged. Changes to the design, such as
increased reinforcement or improved concrete mix for early higher strength, may be
viable to produce the output required for a given site.
The proprietary supplier will provide details of the loads being imposed onto the
structure by the system proposed and how the brackets then transmit the loads into
the structure. The ability of the structure to resist the applied forces is equally as impor-
tant as the concrete strength local to the anchor. To be able to check these conditions
requires a detailed knowledge of the concrete structure and associated site activities.
The TWC has a duty to ensure that the PWD is satisfied in the ability of the structure
to withstand the loads being imposed and remain in accordance with the design
specification. Working together with the contractor’s TWD team, they can then consider
an appropriate lifting cycle.
25.2.4 Design
The design of climbing formwork and slip-form operations is generally carried out by the
design offices of the proprietary supplier or specialist subcontractor, as they have the
product knowledge, competency and expertise. With the reduction in contractor’s
temporary works departments, contracts increasingly rely on proprietary suppliers to
provide relevant designs with equipment. Those contractors who still have a design
capability should be able to understand and produce designs for climbing jump-form
formwork systems (see Section 25.3.2). However, it is recommended that design
responsibility for the specialised design of rail climbing formwork systems and slip
form should remain with the proprietary suppliers and/or specialist subcontractors,
due to their in-depth knowledge of and familiarity with the systems.
320
within the site compound is limited, the platforms can be pre-assembled off site with
integral guard rails, toe-boards and lifting points. They are generally used to allow
steel fixers access to the reinforcement, access to the finished concrete face of a structure
and to fix soffit edge formwork. They can be designed to provide a continuous deck
around outside corners and can be easily adapted to suit buildings with geometrically
complicated shapes.
g Climbing jump-form formwork: units are individually lifted off the structure and
relocated at the next construction level using a crane (see Figure 25.1).
g Rail-guided climbing formwork: units are connected to a rail/anchor configuration
attached to the structure which offers greater safety and control during lifting
operations. The proprietary systems available differ in that some have a fixed
sectional rail attached onto the structure which the units climb up, recycling the
rail as they climb, while others have an ‘anchor and shoe’ assembly attached to
the structure which the rail climbs by recycling the ‘shoe’ as it climbs.
g Rail-guided self-climbing formwork: hydraulic jacks are introduced to climb the
units up the structure, avoiding the requirement of using a crane. This can be
done either with portable hydraulics resulting in platforms being raised
sequentially or, alternatively, with hydraulics positioned at every rail and the
platforms lifted in unison. The latter option is more expensive; it only becomes
viable on structures of significant height or of a complex plan layout.
The operation must be carefully planned and coordinated. Access to the working area
during lifting should be restricted to essential personnel since unprotected edges are
created between the individual platforms when they are being moved; this presents a
significant safety risk and must be cordoned off.
321
Concreting
platform
Working
platform
Wind
brace
Finishing
platform
(a) First Phase – Clean (b) Second Phase – (c) Third Phase – (d) Fourth Phase – Once
and prepare formwork, Once the casting cycle With the formwork sufficient concrete
erect reinforcement and is complete and retracted the system strength system is
attach anchors to be sufficient concrete anchor assembly can craned (’jumped’) up
cast in with concrete. strength has been be fitted for the next onto the next anchor
gained the formwork cycle. The wind brace assembly and wind
can be retracted. can be disconnected. brace connected.
The process for rail-mounted systems is similar, although the system is fixed to a rail and
therefore does not leave the wall once attached until construction is complete. Often such
systems incorporate devices to strike the formwork laterally on the working platform to
facilitate cleaning, fixing anchors, etc.
Different systems provide the option of varying platform widths. Wider platforms provide
the space to allow the associated wall formwork to be fully retracted when struck, thus
allowing clear access for cleaning and fixing reinforcement. Whereas, narrower platforms
generally incorporate a tilting mechanism for access to the formwork face. Self-climbing
systems can be designed to incorporate additional platform levels above the main
construction level. This enables multiple operations to be performed simultaneously,
for example, fixing reinforcement for a wall prior to casting concrete at a lower
section, which offers the potential for realising shorter construction cycle times.
322
The achievable platform length is governed by the spacing of anchors and the associated
wind parameters to be considered. The longer the platform, the higher the load on the
anchors; this then has an impact upon the height and width of the associated wall form-
work. All of these factors will dictate the allowable operational wind speed at which the
climbing system can be safely worked on and moved. If this is misjudged, then part of the
reason for selecting a climbing system from the outset may be compromised.
A typical slip-form example is shown in Figure 25.3. Vertical slip-forming can involve
large volumes of concrete, generally of high-strength low-water to cement ratio with
significant admixtures to suit the consistency necessary for placing, but supplied in
323
small amounts continuously. The upward form movement speed can be adjusted to a
significant degree by varying the admixture dosage. The concrete surface finish is peculiar
to the process.
The layout is normally simple. It is possible to vary the wall thickness and layout over the
height, although each complication will increase the overall cost. The ideal structure
should have significant dimensions in both major axes to ensure stability. Variants of
the process can operate horizontally and on sloping structures.
Slip-forming is a highly specialised process and advice should be sought from people with
the relevant experience. A useful guide is the Concrete Society Good Concrete Guide 6
(CS, 2008).
324
jacking rods are either left in place or subsequently withdrawn from the concrete by
reversing the jacks. The various walls are braced together, and working areas and
access scaffolds are added. The PWD has an important role in ensuring that reinforce-
ment is detailed to suit the method.
The forms are generally only 1.2 m high and are set up with a slight taper to prevent the
form gripping the concrete and lifting it at the trailing edge. It is set up with the correct
wall thickness at approximately mid-height. The face contact material will most likely be
steel panels although, where vertical striations are needed, film-faced plywoods with
hardwood features have been used. Trailing platforms can be used to carry out any
remedial or finishing work on the concrete.
Openings can be formed by using inserts narrower than the wall, thus avoiding displace-
ment of the former vertically as sliding proceeds. An allowance of 18 mm less than the
nominal wall thickness will clear any build-up of hardened concrete on the upper edge
of the form.
Screens can be clad in either solid panels (e.g. plywood) or with perforated mesh or
netting and are supported on rails attached to the already cast slabs of the permanent
structure. Commonly perforated mesh is chosen, as it provides a balance between
325
protecting those below who may be struck by falling debris while still permitting natural
light to enter each working level. Some screens can incorporate working platforms; this
can be of benefit where post-tensioning work is required or to provide space for following
trades undertaking work on the façade.
Wind effect is often the governing factor to be considered when designing protection
screens, as they often travel up to two levels in advance of the construction works
thus generating significant cantilever forces (especially if solid cladding panels are
used). Anchor spacing is directly affected by wind load which in turn determines the
spacing of slab rail supports. Additional propping may even be required, in particular
at corner sections which induce even higher loads as described in Formwork: A Guide
to Good Practice (CS, 2011; section 4) and the British Wind Standard BS EN 1991-1-4
(BSI, 2009).
326
Slip-forming requires a continuously controlled inspection regime and needs a small but
highly skilled workforce on site; this requires team effort and there is no place for restric-
tive working practices on slip-form operations. For this reason, specialist subcontractors
are often employed to carry out the slip and the various responsibilities are established
before slip-forming commences.
Each stage of a climbing sequence, such as a climbing sequence for jump forms as
described in Section 25.3.2, will result in load being transferred into the supporting perma-
nent structure. The TWD must pass on all relevant information to ensure that stability is
considered at every stage and that permissible imposed loads are not being exceeded.
REFERENCES
BSI (2009) PD 6688-1-4. Background Information to the National Annex to BS EN 1991-1-4.
BSI, London.
CS (Concrete Society) (2011) Formwork: A Guide to Good Practice, 3rd edition. Special
Publication CS 030. Concrete Society, Camberley.
CS (2008) Good Concrete Guide 6: Slipforming of vertical structures. Ref. CS162.
Concrete Society, Camberley.
FURTHER READING
BRE (Building Research Establishment) (2007) Formwork for Modern, Efficient Concrete
Construction. BR 495. HIS BRE Press, Bracknell.
ConStruct (Concrete Structures Group) (2010) National Structural Concrete Specification
for Building Construction, 4th edition. Ref. CCIP-050. Concrete Society, Camberley.
CS (2003) Checklist for the assembly, use and striking of formwork. Ref. CS144. Concrete
Society, Crowthorne.
327
Chapter 26
Temporary façade retention
Ray Filip Consultant
Stuart Marchand Director, Wentworth House Partnership
The work involved in supporting existing façades or party walls for renovation, during
rebuilding or after damage is a specific type of temporary works with different
risks, durations and types of loads than the temporary structures of falsework and
scaffolding. Although similar equipment is used, the philosophy, procedures and
relationship to the permanent works are different. Unlike formwork and falsework,
the structure to be supported already exists; it may be old, in poor condition or not
vertical. Procedures will need to reflect the status stressing the importance of initial
surveys and possible extensive monitoring during use to reduce the risks. Permanent
works designers have an important role in the management and control of façade
retention schemes.
26.1. Introduction
The elevations of a building are generally known as façades and the front façade is often
highly decorative. Building façades are retained because they are considered to be of
architectural, historical or visual importance. Behind the façade, the interiors may
have deteriorated or simply may not meet the requirements of modern usage. To meet
with planning restrictions and conservation orders (listed buildings), the interiors are
rebuilt behind the retained façade using modern methods and materials such as steel
frame or reinforced concrete. A shoring (façade retention) scheme is generally required
to support the façade and protect workers (and the general public) while the work is
carried out. The shoring scheme must be designed and installed with care. The façade
will eventually be connected and supported by the new internal structure. Occasionally
other external walls (not just the front elevation), internal walls, floors and roof structure
will be retained and restored. Emergency unplanned retention may also be required after
fire or explosion to make the remaining structure safe for rebuilding or safe demolition, if
deemed beyond repair.
The authoritative guidance document is the CIRIA Report C579 Retention of Masonry
Façades: Best Practice Guide (CIRIA, 2003a), which gives information for all the parties
involved in the planning, design and construction of façade retention schemes. For
specific site guidance, CIRIA C589 Retention of Masonry Façades: Best Practice Site
Handbook (CIRIA, 2003b) is recommended. Both documents highlight the importance
of appointing a competent temporary works coordinator as recommended in BS 5975
(BSI, 2011); see also Chapter 2 on management.
329
This chapter gives information about the types of façades and the important differences
in design principles relative to other temporary works.
When considering façade retention schemes, the following philosophy and sequence
should be adopted.
g Planning: desk study to establish the age of the building, neighbouring properties
and their owners, history and previous usage of the building and listing issues.
Visual inspection to identify potential hazards, existing services, working
restrictions, other site constraints, etc. CDM coordinators (appointed by the
client) have a duty to be involved at this stage so that conceptual design for the
temporary works and new permanent works can demonstrate that the works can
be built safely. Relevant permissions may be needed and other statutory
obligations may have to be carried out.
g Secure the site: public protection, protection of neighbouring properties and safe
means of access, etc. should be provided.
g Understand how the building works: never underestimate the importance of this
and a full structural investigation and dimensional survey (which should include
ground conditions, groundwater levels and foundations) should be carried out. How
stability is achieved and the position of load-bearing elements should be established.
The CIRIA Report R111 Structural Renovation of Traditional Buildings (CIRIA,
1986) is a useful guide on how old buildings may have been constructed. Particular
attention should be paid to areas of obvious distress such as cracks, bowing or
ingress of water (signs of poor maintenance) and to any area where repairs or
alterations have previously been carried out. The façade may not necessarily be of
solid construction; it may be stone faced on brickwork backing or rubble in-filled
stone walls. The surroundings should be investigated, including pavement vaults,
cellars, services and the condition of nearby pavements and roads. Preliminary
shoring and propping may be necessary. Surface finishes may need to be removed
to expose brickwork and embedded steel or timber and their condition assessed.
Floorboards may also have to be removed to assess floor capacity.
g Unstable areas should be made safe: temporary supports should be installed to
allow repairs to be carried out or to remove hazardous items (such as asbestos).
Chimney stacks are prone to weathering and can be removed. If laid with lime
mortar, the bricks can be re-used for rebuilding (or the chimneys will require
330
It is important to realise that the building has achieved a state of equilibrium over many
years and, by changing the support conditions, this equilibrium can be jeopardised; the
designer must consider this. Deflection is one of the main criteria for the design, and the
support system should be sufficiently stiff to limit the deflection.
Some of the above points may be given lower priority or even be bypassed in the case of
emergency retention schemes. Emergency schemes will be selected on availability of
materials, ease and speed of assembly.
331
Are there any hazards (e.g. asbestos)? How is vertical and lateral stability achieved? Are
neighbouring properties affected by the new works? What is the form of construction?
Particular attention should be paid to cracks and signs of damage. Chimneys should
be thoroughly investigated (e.g. position of flues). Verticality of the façade (out of
plumb of more than 10% of the wall thickness is a matter of concern).
Are there any basements or pavement vaults? What is the position and nature of existing
services? Is the building listed or part listed or is it in a conservation area? Are there
any tree preservation orders (TPOs)? Soil conditions, groundwater level and existing
foundations must all be considered. As part of the survey, some materials testing may
be necessary to establish strength, etc.
26.3.2 Scaffolding
Scaffolding in tube and fitting (see Chapter 21 on scaffolding) is suitable for relatively
low-level façades (up to 3–4 storeys with sufficient space around the base of the façade
for installation, approximate width required will be 50–100% of the height; CIRIA,
2003a). Above this height the quantity of scaffolding involved and labour costs may
make installation, inspection, rebuilding and removal challenging. A typical example
is depicted by Figure 26.1a.
332
333
To minimise disruption to the internal rebuilding, vertical towers are installed to the
external face of the façade (where practicable). The frame acts as a vertical cantilever
with wind loads transferred to the foundations through bracing. Deflection of the
frame is usually the critical design consideration to prevent damage to the façade
(through cracking). Push-pull props or tie rods are used for bracing. Horizontally
spanning waling beams, which are tied together through window openings (to minimise
drilling though the façade), are placed internally and externally, ‘sandwiching’ the
façade. The towers are connected to the waling beams.
It is important to consider the combined vertical and horizontal loads on the foundations
as the uplift (due to the overturning moment) will reduce the horizontal capacity. The
greater the width of the frame (i.e. greater lever arm), the smaller the uplift will be. An
allowance may need to be made for pedestrian access. Portal frames are utilised for
this purpose, as shown in Figure 26.2b. Particular attention is required in the design
of such schemes, as the eccentric loading from the self-weight onto the portal structure
induces lateral sway deflection in the portal. Vertical towers can also be utilised where
party walls are to be retained.
334
With all support systems, timber or proprietary frames may be used to brace up major
openings in the façade, to maintain the shear capacity and prevent lateral displacement.
335
Figure 26.3 Typical horizontal frame arrangements with diagonal bracing: façade supported by
(a) external framing and (b) internal framing
(a) Façade supported by external framing (b) Façade supported by internal framing
‘characteristic values’ for equipment; note, however, that a characteristic value is not a
safe working load. The following loads and practical considerations typically need to
be addressed in the design process from concept to installation, monitoring and use.
336
(CIRIA, 2033a) recommends that working platforms be normally designed for service
class 2 loading (1.5k N/m2) minimum for minor repair work, or higher if significant
rebuilding is necessary. Material storage areas, loading bays or site cabins need to be sepa-
rately considered. Further guidance can be found in C579 (CIRIA, 2003a; section 8.4).
Load case Weight Self weight Wind load Impact Out of Other
of actual of retention load plumb vertical
façade structure Maximum Working loads
337
factor cseason (formerly referred to as Ss) and the probability factor cprob (formerly Sp) are
taken as 1.0. Unlike falsework, scaffolding and formwork, there is no reduction for the
short-term loading (less than 2 years) for façade temporary works. The basic equation
for the peak wind velocity pressure (qp) is given in Section 22.33.5 of Chapter 22 on false-
work. The façade is generally considered as a solid impermeable face (due to sheeting and
boarding-up of openings). If the wind can be ‘trapped’ in a corner then the wind force
may be increased. Positive wind pressures as well as wind suction are considered; see
CIRIA (2003a; section 8.6) for further guidance.
g 1.5% of the vertical load at that point acting horizontally which includes the self
weight of façade, retention structure, other dead loads and imposed loads affecting
the retention structure (such as storage and site cabins), but the actual lateral load
should be calculated (the lateral load could be as much as 10% of the vertical load).
g The actual lateral load from eccentricity of the façade (due to out of plumb,
corbels, balconies, etc.) plus 1.5% of the total vertical load on the retention
structure which will include its self weight plus other dead and imposed loads (self
weight of the façade is not included).
This lateral load is considered as acting as a uniformly distributed load over the façade
surface and is applied to the retention scheme at the connection points. It is added to the
wind loading as shown in Figure 26.4. The wind loading will generally be significantly
Figure 26.4 Wind loading plus lateral load applied to support points
Courtesy of RFK Consult Ltd
338
higher than the self weight component of lateral load. A clear load path to the ground
should be identified; see CIRIA (2003a; section 8.8) for further guidance.
CIRIA C579 (2003a; section 8.9) provides advice on other loads to be considered.
Dynamic loads (such as traffic) generating vibration are generally ignored and any
loosening of fixings can be addressed by using lock nuts or regular inspection. Fatigue
is generally not taken into account in temporary works. However, there have been
cases whereby a retention scheme has had to support a façade for many years due to
planning or financial issues, and fatigue and corrosion can become issues. Seismic
activity is not taken into account in the UK. Thermal movement, due to expansion
and contraction of the materials of the façade and support system, may occasionally
need to be considered when specifying ‘acceptable’ levels of movement in the façade.
Older façades tend to be of substantial stone or masonry thickness and have greater
thermal mass, hence respond slowly to daily temperature variations. Excessive thermal
movement can lead to cracking of the façade. CIRIA Technical Note 107 Design for
Movement in Buildings (CIRIA, 1981) and BRE Digest 251 (BRE, 1995) give further
advice. If a retained façade is within a potential flood risk area or next to water then
these effects should be considered; ice formation is not deemed to be a major issue.
External schemes need to consider interfaces with the general public and appropriate
safety measures should be taken. Security is a significant consideration with adequate
protection necessary.
Provision may be required to allow for the scheme to be altered due to variations in wall
locations discovered during demolition and/or any unexpected items discovered on site.
Large openings in the façade may require bracing with timbers or steel frames to
maintain overall shear capability. The effects of fire are occasionally considered and
an appropriate risk assessment should be carried out. Lightning protection with
adequate earthing should be provided to steel support frames.
339
balconies, not plumb, etc.) and imposed loads to provide overall stability. The structure
may also tilt during demolition or reconstruction work as support conditions change due
to re-supporting the vertical weight of the façade, settlement or heave. The new support
positions should be provided as close as possible to the original condition in the existing
structure. It must, however, be remembered that wind loading on the façade may change
after the existing building has been demolished. The points of support to the façade
should be at relatively close centres to prevent excessive deflection.
Section 8.14 of CIRIA C579 (CIRIA, 2003a) recommends a factor of safety of 1.5
against overturning; This is larger than that used for falsework in BS 5975 (BSI,
2011). Any kentledge required is then calculated as:
H
lateral deflection H 4
750
where H is the height to the restraint level under consideration above the point at which
the façade is considered fixed and H is the lateral deflection. The limit was first quoted
by Goodchild and Kaminski (1989).
CIRIA C579 (2003a) further states that: ‘. . . there appears to be no evidence that
working with this limit has resulted in distress to façades. Equally there is no evidence
to warrant recommending a more liberal limit.’ In addition, a maximum floor to floor
deflection of 5 mm over 3 m is often quoted. Designers of temporary works façades
should be aware that the movement of the ground in response to partial demolition
and subsequent reconstruction are inherent in the scheme and should be considered by
the client’s consultant. The lateral deflection limit of the system defined above will be
used by the temporary façade designer, unless tighter limits are set. This limit could
be unacceptable for a party wall, creating large gaps between it and the return walls/
340
floors within the adjacent building. BRE Digest 251 (BRE, 1995) discusses crack widths
and repairs. Damage category 2 is generally considered appropriate with cracks of
<5 mm requiring only decorative repair, but agreement should be reached with the
party wall surveyors. Owners of supporting party walls of occupied buildings may be
less tolerant of movement causing cracking, and deflection limits may need to be
reviewed. Excessive vibration during demolition should be avoided.
Local restraint is provided by the strength and stiffness of the support scheme prior to
demolition of the existing supporting structure. When the existing structure is
removed, the effective length of the façade dramatically increases and the supporting
connections must provide sufficient restraint to the façade. CIRIA C579 (2003a;
section 8.11) recommends that connection should be designed for the greater of the
following loads (to be effective in resisting lateral buckling of the wall).
g 2.5% of the total gravity load on the façade at the level of the connection being
considered, plus the wind force on the area of the façade restrained by the
connection.
g The lateral load arising from offsets and out of plumb of the façade at the level of
the connection, as a uniformly distributed load along the length of a linear
element (waling).
Where kentledge blocks are used to cater for the lateral loads in sliding, then sufficient
friction must be generated. If the blocks are partially or fully buried, then passive resis-
tance can be considered. The kentledge blocks can also be designed to spread vertical
load and limit settlement. If the kentledge is to be cast in situ, then polythene must
not be used to protect the pavement as it will act as a slip membrane. CIRIA C579
(2003a; section 8.14) recommends a factor of safety of 2.0 against sliding. When consid-
ering sliding, only permanent loads should be used to provide beneficial restoring
341
REFERENCES
BRE (Building Research Establishment) (1995) Assessment of damage in low rise build-
ings. Digest 251. BRE, London.
BSI (2000) BS 6187: Code of practice for demolition. BSI, London.
BSI (2005) BS EN 1991-1-4: Eurocode 1: Actions on structures. Part 1–4: General Actions:
Wind Loads. BSI, London (including Amendment No. 1, January 2011).
BSI (2011) BS 5975: 2008 þ A1: 2011 Code of practice for temporary works procedures
and the permissible stress design of falsework. BSI, London.
CIRIA (Construction Industry Research & Information Association) (1981) Design for
Movement in Buildings. Technical Note 107. CIRIA, London.
CIRIA (2003a) Masonry Façade Retention: Best Practice Guide. Report C579. CIRIA,
London.
CIRIA (2003b) Masonry Façade Retention: Best Practice Site Handbook. A5 size Report
C589. CIRIA London.
342
FURTHER READING
BRE (1992) Good Building Guides. GBG1, GBG10, GBG15. HSE, London.
BSI (1981) BS 5977-1: Lintels. Part One Method for assessment of load. BSI, London.
BSI (1993) BS 5080 Part 1. Structural Fixings into Concrete and Masonry. BSI, London.
CIRIA (1994) A Guide to the Management of Building Refurbishment. Report 133.
CIRIA, London.
Doran D, Douglas J and Pratley R (2009) Refurbishment and Repair in Construction. CIOB
(Chartered Institute of Building), Ascot.
Highfield D (1991) The Construction of New Buildings Behind Historic Façades. Spon Press,
London.
HSE (Health and Safety Executive) (1984) Health and Safety in Demolition Work (four
parts). Guidance GS29. HSE, London.
HSE (1985) Safe Erection of Structures (four parts). Guidance GS28. HSE, London.
HSE (current version) Roof-work: prevention of falls. Guidance GS10. HSE, London.
HSE (current version) Façade Retention. Guidance GS51. HSE, London.
Knight LR (1984) The façade can be a nightmare. Civil Engineering Magazine, March: 29–31.
Lamsden BS (1988) Remedying defects in older buildings. Technical Information Service
89. CIOB (Chartered Institute of Building), Ascot.
NSWC (New South Wales Construction) (1992) Façade Retention. Code of Practice. New
South Wales, Australia.
Thorburn S and Littlejohn GS (1992) Underpinning and Retention, 2nd edition. Taylor &
Francis, London.
343
Chapter 27
Bridge installation techniques
Keith Broughton HOCHTIEF UK Construction
John Gill HOCHTIEF UK Construction
27.1. Introduction
Installing bridges is one of the more high-profile aspects of Civil Engineering, and
generates significant media and public interest. It takes a brave project team to invite
public scrutiny, however, as the risks in these projects can be high. The rewards for
engineers working in this field are great.
Most people will have heard of the Millau Viaduct, watched a local bridge construction
project or seen one of the ‘Mega Structures’ films. This chapter will help readers
appreciate the effort needed to make the techniques look simple, and enthuse them to
investigate further.
Bridge installation techniques discussed use the structure itself to provide all or most of
its support during erection. The required design is covered in this chapter, but only in
concept.
Some of the concepts are subject to patents and intellectual property rights. Many
companies have particular specialist knowledge which goes much deeper than described
within this text. Current best practice is generally project specific and many details do not
necessarily translate to another project.
345
The rate of development is rapid and techniques quickly evolve such that some of the
details here will become obsolete within a few years. As with permanent design, future
developments in material technology and mechanical plant will change our view of the
possible. However, engineers will always return to the basic principles and concepts
such as those in the erection schemes described here.
Selecting a bridge installation technique is a major decision, which requires careful and
thorough preparation in order to have all the required information to hand. The range of
information includes
A useful debate at an early stage of a project is whether the client should prepare a fully
developed scheme, or leave the design and installation techniques entirely to the
tendering contractors. Experience suggests that the optimum is part way between
these extremes. Early contractor involvement is successful when developing a scheme
prior to a full tender. A negotiated contract can also be successful, where tenderers
develop and submit their technical solutions and obtain client feedback during a
tender process leading to a final pricing stage, but this is expensive for both parties.
For such schemes, Clients should carefully consider and review the allocation of
responsibility of the major risk events.
A large number of solutions are available for bridge installation, and most can be
grouped into the headings used for the following sections: partial deck erection
schemes, deck erection as a single unit, erection by tunnelling and mining and segmental
bridge construction.
346
concrete deck, bonded with welded shear studs to an upper steel flange. Bolted or
air welded joints would be located between 1/3 and 1/4 points of span. Pre-cast beam
solutions also require an in situ stitch arrangement. Large-piece erection uses either
tracked or telescopic cranes.
Stability and robustness. Considering the stability of the beams during the lift, it is prefer-
able to erect the beams as braced pairs; otherwise, temporary bracing is required. The
main beam design also needs to consider the construction load cases including placing
of deck concrete. Bracing should be designed to ensure that any permanent formwork
cannot slip off its bearing on the steel flanges, particularly in an area over road, rail or
occupied sites. Pay particular attention to the lateral stability of edge beams as lifted
and landed.
Foundations and support. Temporary trestles will require foundations designed to spread
short-term loads into the ground while supporting beams until splices are completed (and
sometimes until the deck is complete).
Plant and equipment. Erection cranes will typically be of the range 250–800 tonne mobiles,
but the largest 1000 tonne mobile cranes or 1200 tonne gantry cranes may be required for
high load and large radius lifts. Cranes utilising super-lift require a clear working area of
up to 30 m diameter. A limitation on crane set-up near railways will be a significant
constraint, and the congestion of city centre areas often prevents super-lift work.
Temporary stresses. A critical load case during concreting of an in situ composite deck,
where additional temporary bracing will be required, occurs when concrete cast on the
beams creates locked-in stresses as they deflect under wet concrete load.
347
Independent review and validation. For any installation over railways, a full design and
independent check of the temporary supports, stability and method statements will be
required.
The large pre-cast concrete sections are lifted onto a pre-constructed edge foundation
that acts as a springing point for the arch. The most important aspect of the erection
concerns the designer’s assumptions of the shear capacity of the foundation and place-
ment of the structural backfill. The temporary loads on the foundation generally
govern design. The structural fill must be placed equally on both sides of the arch,
taking care not to create an out-of-balance loading.
348
Figure 27.2 Lifting in a pre-cast arch at the Greater Bargoed Community Regeneration Scheme
Courtesy of Hochtief UK
349
Stability of the initial arch is critical and may require temporary propping, kentledge or
anchors to prevent toppling or spreading. The subsequent arch units are generally stable
once the first units are fixed in position. Units are generally robust during lifting and
handling provided attention is paid to positioning lifting points and the layout of the
working areas. Two-piece arches are not robust when placed until all joint details are
completed; for example, a stitch joint may be required at the crown and grouting may
be required at the springing points. They are best used only where the load of a one-
piece arch would be beyond the limit of the crane.
Contingency measures. A temporary trestle may be required to support the first section if
two cranes cannot be used. Rapid-strengthening concrete could be used to complete
joints or stitch details to prevent arch spread or to connect units (e.g. Ductal produced
by Lafarge).
Control and monitoring. Control measures should cover final position, alignment and
inspection of bearing areas to ensure no point loads are introduced. Monitoring of
arch units during and after installation is important to ensure that deflections are
within expected limits.
Independent review and validation. It is crucial that the interface between specialists’
design of the arch system and the global design into which it fits is considered and
that all temporary load cases are developed. A number of failures during construction
have occurred and it is recommended that the designer review literature on these while
developing the design.
350
A variation less frequently used is to launch the bridge sideways from a temporary
construction position alongside the final alignment. This requires temporary abutments
and piers with a slide track to the permanent abutments and piers. A benefit is the shorter
timescale for the final installation, which usually requires an existing road deck to be
closed for a very short time.
Such methods were employed during the construction of the A38 Marsh Mills Viaducts
project (see Figure 27.3) near Plymouth, where traffic disruption was minimised using a
side launch technique (permitting the two viaducts to be replaced with only two weekend
road closures). The picture shows the viaduct deck on temporary supports, carrying live
traffic, before the demolition of the existing pre-cast concrete deck and the eventual slide
onto the newly constructed piers.
351
The minimisation of friction at bearing positions will reduce jacking loads. However,
consider restraints for sideward sliding and the requirement for a braking system.
Foundations and support. Normally the permanent substructures and foundations are
used to provide support during the deck launch. This is often supplemented by
temporary supports to reduce bending moments, shear forces and deflections. Careful
attention is important to the interface between the sliding deck and the support to
ensure adequate guidance is provided and to limit the resistance loads on the support.
352
The pathway on the support can be provided by sliding or rolling bearings. The bearing
design accommodates static friction (when the launch commences) and lesser dynamic
friction (as launch proceeds) as well as lateral forces (imposed in restraining any
lateral movement). The choice of rolling or sliding bearing will depend on requirements
for tolerances, friction resistance, intensity of load at bearings and robustness. Final
fixity of the deck is completed after the launch; permanent bearings will sometimes be
carried in during the launch, but isolated out of use to prevent abnormal stresses.
Plant and equipment. The design of a deck launch should only be undertaken with the
knowledge and support of specialist plant and equipment suppliers. Indeed, the design
erection scheme will be partly defined by the size, capacity and applicability of the
suppliers’ hydraulic jacks, strand jacks and the associated power-pack. The motive
power required to move the structure can be calculated from the frictional values esti-
mated in the slide tracks, the roller bearings or at the structure/soil interface. Guidance
to estimated frictional coefficients can be found in Chapter 20 (Section 20.2.4.3) on heavy
moves and in the online Engineer’s Handbook (http://www.engineershandbook.com).
353
Control and monitoring. It is essential to monitor both movement and stress to the struc-
ture during erection. The structure should be modelled through all of the temporary
design load conditions and allowance made for the effects of temperature. There may
be a requirement for a braking restraint. The onsite teams should be highly integrated
with well-understood lines of communication and clearly identified decision makers.
This can involve specific specialists taking control of the site operations for the critical
jacking operations, who then hand over to the overall site supervision team upon
completion.
Independent review and validation. An independent design check must consider all
aspects of the method of erection on the design. The checking engineer should be
selected from consultants with the experience and ability to undertake similar complex
schemes.
354
Figure 27.5 Installation of a pre-cast portal box at CTRL 342 using SPMT units
Courtesy of Hochtief UK
Foundations and support. While the transporter will spread the load well, the competence
of the surface layers must be assured such that the transporter does not settle under load.
The risk of transporter failure due to settlement of the wheels is significant in determining
whether the work can be completed within a road or rail possession.
Plant and equipment. Appropriate transporters are readily available from a number
of companies including Mammoet UK Ltd, Fagioli Ltd and ALE (Abnormal Load
Engineering) Ltd.
Temporary stresses. The arrangement of the support system located on top of the
transport bogeys will determine the stresses during the lifting process. If possible, it is
preferable to incorporate the temporary load case in the permanent design. Beware of
the danger of stresses created by distortions generated by the transport path.
Independent review and validation. An independent review is needed to ensure that all
load cases have been considered within the final design and method, and incorporated
either in the permanent design or within temporary works.
355
The calculation of the frictional resistance to jacking is critical and should only be under-
taken by an experienced designer using reputable data sources. The final solution will be
356
obtained by numerous design reiterations in which the estimated costs and risks are
repeatedly measured and balanced. On the A23 Coulsdon Town Improvement Scheme
(Figure 27.6), a horizontal line of 600 mm diameter steel tubes were installed at high
level beneath the rail ballast to act as the ADS. The tubes were filled with reinforced
concrete, and locked into a concrete slab over the jacking pit to provide horizontal
restraint to the rail track.
The safety of the miners, the loads from their machinery and the requirement to support
the open face of the excavation determines the robustness of the design of the front
mining shield. The shield is typically made either from heavy steel sections or from
reinforced concrete with steel cutting edge. Guidance is available in the preparation of
357
the specification for jacked box tunnelling including the requirements for the working
environment, predicting and monitoring ground conditions and grouting (PCE, 2010).
Various means are available to the designer to reduce the jacking loads. These might
involve low friction materials such as PTFE on the slide paths, tunnelling muds such
as bentonite injected into the over-excavated void around the structure and drag
shields to the roof incorporating steel sheets or steel ropes.
An excellent example of a segmental erection scheme is shown in Figure 27.7 for the
construction of the STAR Light Railway in Kuala Lumpur, where the deck is being
358
Figure 27.7 The erection gantry of a glued segmental bridge used for the STAR Light Railway in
Kuala Lumpur
Courtesy of Vinci PLC and Mark Raiss, Benaim
REFERENCES
NCE (New Civil Engineer) (2005) Backfilling thought to be culprit in Gerrards Cross
Tunnel collapse. New Civil Engineer (EMAP), August 2005.
ICE (Institution of Civil Engineers) (2010) Specification for Tunnelling, 3rd edition. British
Tunnelling Society and Institution of Civil Engineers. Institution of Civil Engineers,
London.
Parag CD, Frangopol DM, Nowak AS (1999) Current and Future Trends in Bridge Design,
Construction and Maintenance. Thomas Telford Publishing, London.
FURTHER READING
Rosignoli M (2002) Bridge Launching. Thomas Telford Publishing, London.
Troyano LF (2003) Bridge Engineering: A Global Perspective. Thomas Telford Publishing,
London.
359
360
Index
access, 1, 6, 7, 11, 38, 44, 103, 122, 123, 142, bearers, 277, 286, 287, 289, 290, 293, 300, 301,
148, 162, 181, 188, 193, 204, 219, 221, 307, 308, 309, 310, 311, 312
222, 223, 224, 231, 237, 239, 311, 314, bearing piles, 203–207
317, 321, 325, 326, 330, 331, 334, 347, 352 analytical process, 207
planning, design, and cost considerations, cross-bracing of piles for lateral loading
table, 28–29 and effective length, illustration,
site compounds and set-up, 34–37 206
see also platforms; scaffolding; site roads ground parameters, design, 205
access mats, 75 installation methods, 204
ALE Engineering Ltd, 355 load factors, 205–206
aluminium, 75, 150, 263, 276, 277, 278, 279, loadings, 206
281, 285, 287, 288, 291, 293, 294, 307, use of, 203–204
311, 312 Beaufort Scale Force, 281
American Society of Mechanical Engineers Beeby, AW, 315
(ASME), 253, 255 Bell, FG, 93
anchorage, 137, 139, 140, 175–176, 187, 190, Bennett et al., 228
321, 322–323 bentonite, 164–166, 197
in-situ anchor head block, illustration, table, Bentonite support fluid compliance
188 testing, 165
anti-drag system (ADS), 356, 357 typical diaphragm wall Bentonite farm
Approval in Principle (AIP), 23 set-up, illustration, 167
Berry et al., 106
Bailey, DC, 221, 225, 232 Bill of Quantities, 275
Bailey Bridge, 222, 232 birdcage scaffolds, 260
modern day military logistical support Bishop, AW, 125
bridge, illustration, 233 Black and Lister, 65
barges, 209, 210, 213, 214, 220–223, 225–227, Bolton et al., 175
256 Bond, A, 87
crane barges, 225 boreholes, 21, 29, 37, 39, 83, 84, 85, 106, 110,
hopper barges, 225 126, 127
jack-up barges, 223–224, 227–229 Boulanger, R, 103
typical barge with crane and ancillary plant, boundary of site, 11
illustration, 221 box-out, 186
barriers, 8, 11, 237, 337 BR394, 315
basic soakage tests (BRE), 65 BR470, 72
beams, bridge installation techniques, 346, Bragg Report (1974, 1975), 3, 4
347, 349 1975, 6, 16, 17, 19, 275
beams, soffit formwork, 311, 313, 314, 315 bridges, 318
361
362
363
climbing formwork (continued ) ConStruct, 291, 305, 309, 311, 313, 315
typical sequence of a jump-form system, Construction (Design and Management)
illustration, 322 Regulations (CDM) 2007, 5–6, 16, 17, 21,
protection screens, 325–326 25, 27, 142, 148, 241, 260, 272, 301
design, 326 coordinators (CDMC), 7, 9, 17–18, 319,
types of, 325–326 326, 330
typical protection screen, illustration, 326 Construction (Health, Safety and Welfare)
system selection, 320–321 Regulations (CHSWR), 5
viability assessment, 318–320 Construction Industry Training Board
design, 320 (CITB), 6, 16
economy of construction, 318–319 contiguous piled walls, 179–191
suitability of structure, 319–320 application, 181
typical climbing formwork to a lift core, construction and plant, 181–182, 184–190
illustration, 319 construction sequence, illustration, 182
see also formwork; formwork, soffit; slip pile construction techniques, 185
form design, 187–190
cofferdams, 134–135, 139 suitability, 179
illustrations, 131, 135 typical example, illustration, 180
communications, 37–38 continuous flight auger (CFA), 185, 186, 188
compound contractor, 27–28, 31, 34
electricity, 38 see also Principal Contractors (PCs);
layout, illustrations, 35, 36 subcontractors
locating, 34–37 contractual obligations, 8–11
requirements, 32–33 Control of Pollution Act, 132
for more detail see site, compounds and controlled permeability formwork (CPF), 292
set-up computers, 24, 31, 68, 127, 286 corrosion, 84, 130, 139, 141, 148, 339
see also IT Coulsdon Town Improvement Scheme, 357
Con-Form Contracting, 97 Coupland, J, 102
concrete, 9, 204, 235, 239, 280, 329 couplers, 166, 186, 264, 265, 266, 277
bridge installation, 346–347, 348–350, 351, crane barges, 225
357, 358 crane erection, large, single and multi-span
concrete pressure calculation, 295–300 bridges, 346–348
contiguous and secant piled walls, 186 cranes, 61, 70, 71,73, 203, 213, 221, 226, 240,
diaphragm walls, 166, 168–169 244–246, 247, 254–255, 318, 319, 337, 347
formwork, 289, 290, 291, 292, 293, 295–300, see also crane barges; crane erection;
301, 302, 303, 304–305 crawler cranes; tower cranes; truck
formwork, climbing, 317, 321–322 mounted telescopic boom cranes
formwork, slip form, 317, 323, 324, 325 crawler cranes, 133, 170, 199, 215, 227, 245, 257
formwork, soffit, 307, 309, 310, 312, 313, cylinder jacks, 247–248
314, 315 jack and pack, illustration, 248
foundations for tower cranes see tower
cranes Darcy’s Law, 86
pre-cast concrete arch, erection of, 348–350 Dawson Construction Plant Ltd, 133
see also sprayed concrete lining de Wit et al., 102, 105
Concrete Society, 3, 275, 298–299, 316, 324, deep wells see wells
327 deflection, 138–139
FGTGP, 289, 292, 293, 294, 297, 300, 301, Designated Individual (DI), 18, 20
304, 305, 309, 310, 312, 313, 314, 316, design brief, 20, 21–22
325, 326 design check, 20, 22–24
Conditions of Contract (ICE) 1999, 8, 15 list of categories, 23
364
365
366
367
368
369
370
371
372
373
374
375
truss bridges, 232–234, 242 water, 29, 39–40, 134, 138, 140, 146, 162, 165,
Tunnel Lining Design Guide (BTS), 194 199
tunnelling, 356–358 see also cofferdams; drainage; groundwater,
Tupper, EC, 225 control of; plant, floating; jetties;
wastewater; water industry
underpinning, 162, 184, 194, 195–196, 199 water industry, 9, 291, 304, 309
underpinning showing trimming, wave action, 216, 220, 227
illustration, 196 Webber, NB, 225
Unifloat, 221–222, 225 welfare, 5, 27, 28, 31, 35, 38, 39, 41–43
Unit Construction Bridges, 234 see also safety
University of Dundee, 297 wellpoints, 80, 81–83
unlimited limit state (ULS), 174, 175 wells,
deep wells, 80, 82, 118
vegetable oil release agents, 295 development of, 83
vertical capacity of walls ejectors, 80, 83
contiguous and secant piled walls, 189–190 main features of a wellpoint, illustration,
diaphragm walls, 175–176 81
vertical H-sections, post and plank, 153–155 pumped well dewartering systems, 87–88
post and plank excavation and side rail relations between pumped well systems,
system, illustration, 154 permeability and drawdown,
vertical shores, 150 illustration, 80
vertical towers, 334, 335 three stages of wellpoint system, illustration,
vibration, 204 82
vibrators, 131, 134 wellpoint header pipe around excavation,
Vinci PLC, 359 illustration, 81, 82
voice and data equipment, 30 wellpoints, 81–83
voids, 83, 92, 101, 115, 121, 147, 195, 280, 307 white finger, 292
wind loading, 166, 216, 227, 247, 265,
walls see contiguous walls; diaphragm walls; 266–267, 272, 281, 287, 326, 327, 334,
formwork, wall formwork design; guide 337–338, 341
walls; party walls; retaining walls; secant Work at Height Regulations (WAHR) 2005,
piled walls; sheet piling 5, 7–8, 260, 267, 272, 301
Waste and Resources Action Programme working platforms see platforms
(WRAP), 74 workshop buildings, 33, 38
waste management, 33
wastewater, 31, 37, 40, 106–107 Zemdrain, 292
376