0% found this document useful (0 votes)
94 views343 pages

SC070056 Report

scour

Uploaded by

murtaza05
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
94 views343 pages

SC070056 Report

scour

Uploaded by

murtaza05
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 343

Toe structures management manual

Project: SC070056/R
The Environment Agency is the leading public body
protecting and improving the environment in England and
Wales.

It’s our job to make sure that air, land and water are looked
after by everyone in today’s society, so that tomorrow’s
generations inherit a cleaner, healthier world.

Our work includes tackling flooding and pollution incidents,


reducing industry’s impacts on the environment, cleaning up
rivers, coastal waters and contaminated land, and
improving wildlife habitats.

This report is the result of research commissioned and


funded by the Environment Agency’s Evidence Directorate.

Published by: Author(s):


Environment Agency, Horison House, Deanery Road, Andy Bradbury
Bristol, BS1 9AH Jonathan Rogers
www.environment-agency.gov.uk Dick Thomas

ISBN: 978-1-84911-290-1 Dissemination Status:


Publicly available
© Environment Agency – December 2012
Keywords:
All rights reserved. This document may be reproduced Coastal structures, structure toe, asset management,
with prior permission of the Environment Agency. condition assessment, beach, scour

The views and statements expressed in this report are Research Contractor:
those of the author alone. The views or statements Halcrow Group Ltd
expressed in this publication do not necessarily Ash House, Falcon Road, Sowton, Exeter, Devon
represent the views of the Environment Agency and the EX2 7LB
Environment Agency cannot accept any responsibility for Tel: 01392 444252
such views or statements.
Environment Agency’s Project Manager:
Further copies of this report are available from our Dave Hart, Manley House, Exeter
publications catalogue: http://publications.environment-
agency.gov.uk or our National Customer Contact Collaborator(s):
Centre: T: 08708 506506 None
E: [email protected].
Project Number:
SC070056

Product Code:
LIT 7651

ii Toe structures management manual


Evidence at the
Environment Agency
Evidence underpins the work of the Environment Agency. It provides an up-to-date
understanding of the world about us and helps us to develop monitoring tools and
techniques to manage our environment as efficiently and effectively as possible.
The work of the Environment Agency’s Evidence Directorate is a key ingredient in the
partnership between research, guidance and operations that enables the Environment
Agency to protect and restore our environment.
This report was produced by the Research, Monitoring and Innovation team within
Evidence. The team focuses on four main areas of activity:

• Setting the agenda, by providing the evidence for decisions;


• Maintaining scientific credibility, by ensuring that our programmes and
projects are fit for purpose and executed according to international standards;
• Carrying out research, either by contracting it out to research organisations
and consultancies or by doing it ourselves;
• Delivering information, advice, tools and techniques, by making
appropriate products available.

Miranda Kavanagh
Director of Evidence

Toe structures management manual iii


Executive summary
This manual is intended to provide those responsible for flood defences, coastal
erosion protection structures and other coastal structures with practical guidance on
how to determine, monitor, assess and mitigate for the effects of toe scour. It describes
the different types of toe protection structures and provides illustrations of typical
designs that are often used as solutions. The case studies on beach lowering and
scour management provide real examples of both good and bad practice, and discuss
lessons learnt from past schemes.
Scour, in specific relation to coastal engineering projects, can be defined as ‘the
removal, by hydrodynamic forces, of erodible bed material in the vicinity of coastal
structures’. This definition distinguishes scour from the more general erosion and notes
that the presence of a reflecting structure contributes to the cause of scour. Scour that
affects coastal structures can lead to partial damage, or in extreme cases, complete
failure of the structure.
A comprehensive survey published by CIRIA in 1986 concluded that scour at the toe of
structures represented the most prevalent and serious form of damage to seawalls in
the UK. Toe scour is a serious and costly problem – moreover, it is one that is not
limited to any particular environment or generally to any particular type of seawall.
Toe protection provides insurance against scouring and the undermining of a structure.
It provides additional armouring of the beach or base of a defence in front of the
structure which prevents waves and currents from scouring and undercutting it.
This manual presents, in a logical way, how relevant issues should be covered in the
assessment, management and design of the toe of coastal defence structures. These
issues include:
• monitoring of beach levels and structure condition (including links to coastal
monitoring programmes) – Chapter 3;
• trigger points for action during the life of the asset (including links to
‘performance features’) – Section 3.4;
• maintenance and replacement of elements of the asset or of the whole
structure – Chapter 4;
• environmental and sustainability issues surrounding scheme/structure
planning, design and operation – Chapters 2, 3, 4 and 5;
• inputs to the assessment and design processes (including relationship to
defence fragility, both for asset management and new design) – Chapter 5;
• selection of mitigation options to enhance or prolong performance of the
asset – Chapter 2 and Section 5.3.
The manual is supported by appendices containing information on scour processes,
methods of predicting scour and a number of case studies from around the UK to
illustrate particular management approaches and draw on the experience of the
techniques that have been employed.

iv Toe structures management manual


Acknowledgements
This manual has been developed from the outputs of the Environment Agency’s Toe
Structures for Coastal Defences project (SC070056) undertaken by HR Wallingford,
which provided much of the research base for the content of this manual.
The development of the guidance was directed by a steering group consisting of:
Nick Lyness (chairman) Environment Agency
Geoff Baxter Environment Agency
David Hart Environment Agency
Assistance and review was also received from:
Brian Farrow North Norfolk District Council
Chris Hayes Environment Agency
Steve McFarland SEPA
Adrian Phillpot Environment Agency
Alan Williams Coastal Engineering UK Ltd
The project team is grateful to the members of the steering group and those who
contributed time, information and images to the development of the manual.

Toe structures management manual v


Contents
1 Introduction 1
1.1 Background and context 1
1.2 Concept of toe protection 3
1.3 Structure of the manual 5
1.4 The asset management cycle 8
1.5 Target audience 8
1.6 Acknowledgements 9

2 Toe structure types and materials 10


2.1 Introduction 10
2.2 Types of toe structure 11
2.3 Materials 21
2.4 Choice of structure type and materials 33

3 Asset management 36
3.1 Introduction 36
3.2 Management overview 37
3.3 Life-cycle failure modes 39
3.4 Planning a management programme for structure toes 44
3.5 Defining ‘critical’ and ‘trigger/alert beach levels’ 46
3.6 Beach monitoring 48
3.7 Beach depletion and foreshore down-cutting 68
3.8 Structure condition monitoring 69

4 Maintenance 81
4.1 Introduction 81
4.2 Issues associated with maintenance of existing toes 83
4.3 Maintenance of toe structures 92

5 Toe structure design 109


5.1 Introduction 109
5.2 Identification of the problem 110
5.3 Project appraisal 111
5.4 Design principles 114
5.5 Undermining 118
5.6 Liquefaction at the toe 123
5.7 Geotechnical stability 124
5.8 Resistance to wave and current loading 134
5.9 Hydraulic performance 143
5.10 Effects on coastal processes 143

vi Toe structures management manual


5.11 Public safety 144
5.12 Natural environment 145
5.13 Heritage and visual impact 146
5.14 Amenity 147
5.15 Construction issues 148

List of abbreviations 152

References 154

Appendix A: Scour processes 164

Appendix B: Predictive methods 195

Appendix C: Case studies 225

Table 2.2 Loss of thickness (mm) due to corrosion for piles and sheet piles in sea water 1 23
Table 3.1 Examples of detailed intrusive and non-intrusive assessment techniques for defence structures and/or
beaches 77
Table 4.1 Structure repair/restoration options and their applicability to the maritime environment (based on CIRIA
2010b) 94
Table 4.2 Options for repair works related to defects in concrete and concrete reinforcement and their applicability
to the maritime environment (adapted from BSI 2006) 94
Table 5.1 Toe related failure modes of main structure types 115
Table 5.2 Relevance of criteria to toe structure yypes 118

Figure 1.1 Definition of the structural toe of a defence as opposed to the ‘visible’ toe 3
Figure 1.2 Layout of the guide 7
Figure 1.3 Document information in relation to the asset management cycle 8
Figure 2.1 Underpinning of seawall toe with steel sheet piles (courtesy HR Wallingford) 15
Figure 2.2 Sheet pile underpinning 16
Figure 2.3 Cribwork and concrete block fill, Norfolk (courtesy of North Norfolk District Council) 16
Figure 2.4 Extensive crack in masonry toe (courtesy HR Wallingford) 17
Figure 2.5 Gabion baskets of rock (courtesy HR Wallingford) 18
Figure 2.6 Rock infill of scour trough, Le Dicq, Jersey (courtesy HR Wallingford) 19
Figure 2.7 Timber bulkhead with rock toe protection at Lepe, Hampshire 19
Figure 2.8 Extended scour apron, masonry steps and armour, St Ouens Bay, Jersey (courtesy HR Wallingford) 20
Figure 2.9 Concrete stepped revetment, Crosby, UK (courtesy Sefton Council) 21
Figure 2.11 The weathering zones of a seawall (from Thomas and Hall 1992) 22
Figure 2.12 Grouted stone/masonry (courtesy HR Wallingford) 24
Figure 2.13 Grouted stone toe apron, Thames Estuary (courtesy HR Wallingford) 25
Figure 2.14 Concrete toe beams (courtesy HR Wallingford) 26
Figure 2.15 Reinforcement rebar corrosion and spalling of concrete (courtesy of HR Wallingford) 27
Figure 2.16 Surface salt deposits and associated cracking from alkali–silica reaction in concrete (courtesy of HR
Wallingford) 28
Figure 2.17 Failure of a concrete toe beam at St Ouens Bay, Jersey (courtesy HR Wallingford) 28
Figure 2.18 Asphalt grouted stone revetment in the Thames Estuary (courtesy HR Wallingford) 29
Figure 2.19 Gribble attack on Timber pile at the sediment line (courtesy of HR Wallingford) 31
Figure 2.20 Close up of gribble damage to timber (courtesy of HR Wallingford) 31
Figure 2.21 Rounding of groyne timbers by beach shingle (courtesy HR Wallingford) 32
Figure 2.22 Necking in derelict timber piles (courtesy ENBE) 32
Figure 3.1 The PAS 55 Plan – Do – Check – Act framework (from BSI 2008) 38
Figure 3.2 Key phases for asset management of existing coastal defence structures 39
Figure 3.3 Overturning and settlement of a gravity wall due to removal of passive resistance 40
Figure 3.4 A seawall in the process of overturning (courtesy of Black & Veatch) 41
Figure 3.5 Exposure of interlocking steel piled toe arising from beach lowering (courtesy A P Bradbury) 42
Figure 3.6 Frequent overtopping causes build-up of pressure on landward side of seawall (courtesy A P Bradbury)42
Figure 3.7 Repaired wall showing failed realignment. Rock armour has been added at the toe (courtesy A P
Bradbury) 43
Figure 3.8 Promenade wall under construction (courtesy Poole Borough Council) 44
Figure 3.9 Information required to assess toe scour 45
Figure 3.10 Assessment of toe scour decision process 46
Figure 3.11 Relationship between long-term statistics on beach levels at the structure, uncertainty, and ‘trigger’ and
critical levels 47
Figure 3.12 Beach survey using global positioning equipment (courtesy Channel Coastal Observatory) 49
Figure 3.13 All terrain vehicle fitted with GPS equipment for coastal topographic surveying (courtesy of Worthing
Borough Council) 50

Toe structures management manual vii


Figure 3.14 Digital terrain models showing beach evolution adjacent to seawall with changing structure alignment
(courtesy Channel Coastal Observatory) 51
Figure 3.15 Migration of MHWS contour towards seawall (courtesy Channel Coastal Observatory) 52
Figure 3.16 Migration of MHLS contour towards seawall (courtesy Channel Coastal Observatory) 53
Figure 3.17 Beach profile and structure descriptors used in structure toe analysis (courtesy AP Bradbury) 57
Figure 3.18 Beach profile cross-sectional trend analysis 58
Figure 3.19 Beach profile evolution in the presence of seawall (courtesy Channel Coastal Observatory) 59
Figure 3.20 Typical hydrodynamic conditions used in the analysis of a storm event (courtesy Channel Coastal
Observatory) 60
Figure 3.21 Temporal and spatial distribution of beach and structure toe intersection elevations (courtesy Channel
Coastal Observatory) 62
Figure 3.22 Temporal distribution of beach and structure toe intersection elevations (courtesy HR Wallingford) 62
Figure 3.23 Scour monitors in operation on a shingle beach to obtain data on within tide changes in bed levels
(courtesy HR Wallingford) 64
Figure 3.24 Fluctuations in sand bed levels over a tidal cycle (from Pearce et al. 2006) 65
Figure 3.25 Comparison of traditional profile with Argus profile envelope (from Green and Illic 2009) 67
Figure 3.26 An extreme example of beach depletion, foreshore erosion and down-cutting – note the level of the
base of the access steps in relation to the level of the beach (courtesy HR Wallingford) 68
Figure 3.27 A traversing erosion beam instrument 69
Figure 3.28 Abraded concrete toe and onset of undermining (courtesy of HR Wallingford) 70
Figure 3.29 Exposed reinforcing mesh due to steel oxidation and resultant cracking and flaking of concrete cover
layer (courtesy HR Wallingford) 70
Figure 3.30 Beach lowering at access steps (courtesy HR Wallingford) 74
Figure 3.31 Schematic of backfill washout 75
Figure 3.32 Undermining and fill washout of stepped revetment and wave return wall 75
Figure 3.33 An example of a page from the ‘Seawalls’ section of the Environment Agency’s Condition Assessment
Manual (2006) 76
Figure 4.1 Application of monitoring data to identify beach zones beneath trigger levels (courtesy Worthing
Borough Council) 85
Figure 4.2 Beach that has reached trigger level prior to repair and recharge (courtesy Poole Borough Council) 86
Figure 4.3 Application of monitoring data to identify long term beach performance in context with structure toe
triggers (courtesy Channel Coastal Observatory) 87
Figure 4.4 Typical damage to tops of steel interlocking piles (courtesy AP Bradbury) 88
Figure 4.5 Temporary access road constructed from local beach material (courtesy New Forest District Council) 90
Figure 4.6 Improvement of access by construction of rock toe with wide crest berm suitable for plant access
(courtesy New Forest District Council) 90
Figure 4.7 Improvement of access by construction of concrete roadway suitable for plant access (courtesy Alan
Williams) 91
Figure 4.8 Temporary rock bund to provide safe working area for plant during toe repairs (courtesy New Forest
District Council) 92
Figure 4.9 Unsupported section of reinforced mass concrete seawall following major undermining (Chesil Cove
1962) © Stuart Morris 96
Figure 4.10 Preparation of foundation excavation beneath wall for underpinning, using bespoke plant to ensure safe
working beneath seawall (courtesy New Forest District Council) 97
Figure 4.11 Underpinning detail for seawall foundation repair (courtesy New Forest District Council) 98
Figure 4.12 Typical collapse of promenade decking following undermining of backfilled mass concrete wall and
washout of fine core (Chesil Cove 1962) © Stuart Morris 98
Figure 4.13 Contrast between concrete toe and modified rock armour toe Folkestone Warren (courtesy Bryan
Curtis) 99
Figure 4.14 Manoeuvring armourstone blocks in a revetment (courtesy Dean and Dyball Ltd) 99
Figure 4.15 Armour displaced from a rock toe, Llanfairfechan, North Wales (courtesy Alan Williams) 100
Figure 4.16 Weathered gabions at East Head, Chichester Harbour, West Sussex (courtesy HR Wallingford) 101
Figure 4.17 Vegetation growth on a revetment (courtesy ENBE) 102
Figure 4.17 Vegetation colonising joints in a masonry coastal defence (courtesy HR Wallingford) 103
Figure 4.19 Prior to and following maintenance of grouted masonry toe (courtesy Alan Williams) 104
Figure 4.20 Dislodged masonry blocks from revetment toe (courtesy HR Wallingford) 104
Figure 4.21 Washout and collapse of block revetment following block removal (courtesy Canterbury City Council)105
Figure 4.22 Deteriorated sheet piles 105
Figure 4.23 Complete loss of section of sheet piles in an area exposed to sediment movement 106
Figure 4.24 Damaged interlocking steel piles requiring maintenance to make site safe (courtesy A P Bradbury) 106
Figure 4.25 Repairs to open stone toe at Prestatyn, North Wales (courtesy Alan Williams) 108
Figure 5.1 Toe failure example 110
Figure 5.2 Goodrington Sands Seawall, Paignton (courtesy of ENBE) 115
Figure 5.3 Undermining of wartime coastal structures at Kilnsey, Holderness (courtesy of ENBE) 119
Figure 5.4 Undermining failure of a revetment 120
Figure 5.5 Typical toe details (after McConnell, 1998) 122
Figure 5.6 Sinking of armour units into liquefied seabed 123
Figure 5.7 Sliding failure of gravity wall 125
Figure 5.8 Overturning failure of gravity wall 125
Figure 5.9 Rotation failure of sheet piled wall 125
Figure 5.10 Simplified forces on a gravity wall 126
Figure 5.11 Example of thrust block and piled toe 131
Figure 5.12 Geotechnical support for the concrete seawall at Lowestoft (courtesy Waveney District Council) 132
Figure 5.13 Example of rock used to provide passive support 132
Figure 5.14 Slip circle failures 133
Figure 5.15 Relative depth of toe 135
Figure 5.16 Sliding of toe armour on hard substrate 137

viii Toe structures management manual


Figure 5.17 Geotextile being incorporated as part of a toe structure 138
Figure 5.18 Example of rock apron (courtesy Jersey States Government) 140
Figure 5.19 Example of a flexible armoured revetment 141
Figure 5.20 Caister on Sea, UK 1979: cross-section phase I with open stone asphalt revetment and a sand mastic
asphalt toe slab (from Schönian 1999) 143
Figure 5.21 Heritage interest – the flint-faced seawall at Lowestoft South Beach 147
Figure 5.22 Toe apron proves walking platform – Paignton, Devon (courtesy of ENBE) 147

Toe structures management manual ix


1 Introduction

Chapter 1 introduces the manual. It gives the background to the


document, outlines its objectives, its scope, intended readership and
use.

Key links to other chapters


o Chapter 2 – Toe structure types and materials
o Chapter 3 – Asset management
o Chapter 4 – Maintenance
o Chapter 5 – Toe design

Who will be interested in this chapter?


o All users of the manual

1.1 Background and context

1.1.1 A brief history


Coastal engineering dates back to the ancient world bordering the Mediterranean
Sea, Red Sea and Persian Gulf where the ancient civilisations began developing
harbours to aid maritime traffic, around 3,500 BC (USACE 2012).
The ancient Egyptians and Minoans developed many sophisticated port structures in
the pre-Roman era, including rubble-mound breakwaters on the River Nile at Djoser
(c.2,500 BC) and Alexandria, Egypt (c.1,800 BC). However, these ancient ports had a
common problem that needed to be addressed; they had to be kept clear of silt in a
time when dredging was unknown. This was often achieved through flushing by
means of channels, tanks, sluices or diverting rivers through canals such that silt
would be driven away from the harbour.
The Romans continued the evolution of coastal structures, introducing many
innovations in coastal engineering including the construction of walls underwater and
solid breakwaters. In the Mediterranean, they replaced many of the ancient rubble-
mound breakwaters with vertical concrete walls. These colossal coastal structures
could be built rapidly and required little maintenance. In some cases, wave reflection
was actively used to prevent silting. In most cases, rubble or large stone slabs were
placed in front of the walls to protect against undermining of the structures. The
Romans were apparently the first to recognise the problem of toe scour.
In the UK, coastal defences have been constructed since at least Roman times, with
flood embankments protecting areas on the Medway and the Severn Levels. The
sole objective of coastal defence up until the 19th century was to protection low-lying
agricultural land and the reclamation of fertile inter-tidal areas. However, by this time,
the popularity of leisure pursuits at the seafront, such as bathing and promenade
walking, had intensified and this led to a concerted effort to capture land and protect

Toe structures management manual


1
the coast from erosion by the construction of seawalls. This period also saw
defences, predominantly vertical seawalls, built to protect critical infrastructure such
as roads and railways, for example Brunel’s South Devon Railway along the
coastline between Dawlish and Teignmouth, while seawalls and particularly
breakwaters were built for military use and to provide fishing harbours.
In the early 20th century, defences were also built to stimulate and regenerate local
economies, protecting reclaimed areas. After the Second World War, many seawalls
were constructed to enable the public to have better access to the coast and to
protect valuable agricultural areas during a time of food rationing. The North Sea
surge of 1953, which flooded many low-lying coastal settlements and agricultural
land in eastern England, led to a major upgrade of flood defences around the coast.
Renewal, upgrade and construction of new coastal defences has continued to the
present day, albeit with a far more enhanced system of economic and benefits
justification than in the past.
Around the coastline and tidal estuaries of England and Wales there are now an
estimated 2,935 km of coastal defences, which amounts to 36 per cent of the total
length of the coast (Halcrow 2011).
These defences are subject to many types of failure mechanisms with crest failure,
settlement and outflanking among many other forms of failure. In 1986, CIRIA
undertook a comprehensive survey of coastal authorities in England, Scotland and
Wales with regard to the performance of seawalls (CIRIA 1986). This included an
examination of the main types of damage experienced by coastal defences. The
survey concluded that erosion at the toe of structures represented, by far, the most
prevalent and serious form of damage to seawalls in the UK; over 12 per cent of all
seawalls reported erosion at the toe, which represented over a third of all damage
reported.
Just like the Romans millennia before, toe scour, at that challenging area where the
land meets the erosive force of the sea, is still recognised as a major problem for
coastal defence structures.

1.1.2 What is toe scour?


Scour, in a hydrodynamic context, can be defined as ‘the erosive force of moving
water’. This broad definition includes erosion of sediment under any circumstances,
such as beach profile change and inlet channel migration.
Scour, in specific relation to coastal engineering projects can be defined as ‘the
removal, by hydrodynamic forces, of erodible bed material in the vicinity of coastal
structures’. This definition distinguishes scour from the more general erosion, while
the presence of a reflecting structure contributes to the cause of scour.
Scour that affects coastal structures can lead to partial damage, or in extreme cases,
complete failure of the structure.
The toe of a coastal defence structure is shown in Figure 1.1.

Toe structures management manual


2
Defence structure
Visible toe (or sediment / structure interface)

Beach
The toe of the
structure

Figure 1.1 Definition of the structural toe of a defence as opposed to the


‘visible’ toe
Toe scour is particularly prevalent at vertical seawalls where waves, enhanced by
reflections, combine with localised currents to mobilise and move sediments. This
can undermine the structure leading to tilting, slumping and other failure
mechanisms. Scour-induced damage also occurs at sloping-front structures when
scour undermines the toe so it can no longer support the armour layer, which then
slides downslope. Scour at vertical sheet pile walls can result in seaward rotation of
the sheet pile toe due to pressure of the retained soil.
Structural damage or failure brought about by scour affects coastal projects in a
number of ways including:
• decreased functionality;
• increased costs to repair or replace the structure (scour-related damage
is often difficult and expensive to repair);
• assets protected by the structure may become at risk of flooding or
erosion;
• clients and other stakeholders will lose confidence in the project's
capability to perform as required.
Toe scour is a common problem that needs to be understood and considered fully in
the design and maintenance of coastal structures.

1.2 Concept of toe protection

1.2.1 The need to protect the toe


Toe protection provides insurance against scouring and undermining of a structure. It
provides additional armouring of the beach or base of a defence in front of the

Toe structures management manual


3
structure, which prevents waves and currents from scouring and undercutting it.
Factors that affect the severity of toe scour include:
• wave breaking (when near the toe);
• wave run-up and backwash;
• wave reflection;
• the grain-size distribution of the beach or bottom materials.
The toe of a defence often requires special consideration because it is subjected to
both hydraulic forces and changing beach profiles fronting the structure. Seasonal
variations in beach profile will be a contributing factor in determining the type and
extent of toe protection.
Toe protection is also often needed along structures that cause concentration of
currents, such as training walls and breakwaters extending from the shoreline.
Furthermore, highly reflective structures such as impermeable vertical walls are
much more susceptible to near-structure scour than sloping rubble-mound structures.
Toe stability is essential because failure of the toe will often lead to failure of the
entire structure.

1.2.2 Management criteria


Management of the toe of coastal structures should meet various criteria (for
example, policy, sustainability and cost) including the need to:
• provide and maintain a level of performance;
• monitor changes in condition and performance and to gather data on
these changes;
• consider the impact of extreme forcing conditions (storms, surges and so
on);
• consider changes in associated shoreline or beach management
practices.
The practitioner also has to work within various constraints be they political (policy),
financial, environmental or simply practical.
This manual presents, in a logical way, how relevant issues should be covered in the
assessment, management and design of the toe of coastal defence structures. These
issues include:
• monitoring of beach levels and structure condition (including links to
coastal monitoring programmes) – Chapter 3;
• trigger points for action during the life of the asset (including links to
‘performance features’) – Section 3.4;
• maintenance and replacement of elements of the asset or of the whole
structure – Chapter 4;
• inputs to the assessment and design processes (including relationship to
defence fragility for asset management and new design) – Chapter 5;

Toe structures management manual


4
• selection of mitigation options to enhance or prolong performance of the
asset – Chapter 2 and Section 5.3.
Environmental and sustainability issues are considered in detail throughout the
manual where these relate directly to toe structures, their construction materials and
related remedial interventions. Generally, as the toe is only one element of a larger
structure, such issues would normally be considered at scheme scale.

1.2.3 Overview of toe structure types


The area where the land meets the sea is complex, often changing its
geomorphology in relatively short timescales. This can lead to confusion over the
identification of the toe – and also its type if not previously known. For the purpose of
this manual, the toe of a coastal defence is defined as the physical toe of the defence
structure as illustrated in Figure 1.1. This should not be confused with that which
might be described as the visible toe – the interface between the beach and the
structure. The structural toe of defences can of course become visible due to the
removal of sediments during storms, or over long periods of drawdown, even if only
temporarily.
Furthermore, the toe of structures that are located permanently below the level of
lowest astronomical tide (LAT) will not be easily observed. For the most part, this
manual refers to structures with emergent toe structures, that is, those that are
located above LAT levels. This is partly because, historically, LAT was about the
lowest level that could practically be reached for most land based methods of
construction.
Because of the role that sediment and shore platforms play in the performance of toe
structures, this manual also discusses the management of beaches and foreshores
to a limited degree including sediment control structures such as groynes and
interventions such as replenishment.
The main focus, however, is on toe structures. These structures are described further
in Chapter 2, where they are considered as:
• vertical, including sheet piling, cribwork, blockwork, solid infill and toe beams,
in materials such as concrete, masonry, steel and timber alone or in
combination as appropriate;
• sloping, including proprietary flexible revetment systems, irregularly placed
rock slopes, gabion mattresses and grouted stone (which are flexible to a
degree), concrete and stone stepwork, and placed blockwork which is more
rigid.
Toe structures may form an integral part of the original defence structure, being a
specific element included at the design stage, or they may be supplemental (that is,
added to the original structure after its construction).

1.3 Structure of the manual


Existing guidance for coastal managers includes:
• Coastal Engineering Manual (USACE 2012);
• The Rock Manual (CIRIA et al. 2007);

Toe structures management manual


5
• The European Overtopping Manual (Pullen et al. 2007);
• Beach Management Manual (CIRIA 2010a);
• The Use of Concrete in Maritime Engineering: A Good Practice Guide
(CIRIA 2010b);
• Condition Assessment Manual (Environment Agency 2006).
This manual aims to complement these documents rather than supersede them. To
reduce duplication, where these other documents contain more detailed information
pertinent to the design, assessment and management of toe structures, the reader is
referred to these where appropriate.

1.3.1 Route map


The manual provides information on the key aspects of the management of toe
structures as depicted in Figure 1.2, which summarises the contents of the manual’s
chapters. Throughout the manual, reference is made to case studies from around the
UK. These are detailed in Appendix C.

Toe structures management manual


6
Background and context
Chapter 1 – Introduction Concept of toe protection
A general introduction to the manual Structure of manual
The asset management cycle
Target audience

Toe structure types


Chapter 2 – Toe structure Materials
Choice of structure type and
types and materials
materials
Description of the types of materials
and structures used to mitigate against
the effects of scour

Life-cycle failure modes


Chapter 3 – Asset Planning programmes
management Defining trigger/alert beach levels
Beach monitoring
Guidance on post-scheme asset
Structure condition monitoring
management activities such as asset
performance and monitoring

Issues associated with


Chapter 4 – Maintenance maintenance
Options for maintenance of different Maintenance of different types of
types of toe protection structures and toe structure
materials and trigger levels for action

Identification of problem
Chapter 5 – Toe design Design principles
Illustration and description of typical Failure mechanisms
designs for toe protection structures Public safety and the environment
and particular considerations in the Construction issues
design process

Scour processes
Appendices Predictive methods
Covering physical processes; methods Case studies
for predicting toe scour and detailed
case studies

Figure 1.2 Layout of the guide

Toe structures management manual


7
1.4 The asset management cycle
This manual provides information and guidance on the toe protection of coastal
defences for practitioners, engineers, designers and managers whose responsibilities
may span the life cycle of these structures. Figure 1.3 illustrates how key decision-
making and management actions feature in the asset management cycle and
indicates where the reader can find relevant information on particular topics in the
manual.

Condition
inspection/monitoring
Chapter 3

Assess
performance
Implementation Chapter 3
Chapter 5
No

Intervention?

Maintenance
Chapter 4
Design solutions Yes
Chapters 2 and 5

Figure 1.3 Document information in relation to the asset management cycle


It should be remembered that any intervention may modify the physical processes
themselves. Monitoring of beach response should therefore continue once a scheme
has been implemented.
Where relevant, the manual contains cross-references to other sections and to other
documents where further or more detailed information on particular subjects or
issues can be found.

1.5 Target audience


This manual is designed primarily for UK readership and practice. However, it is
applicable to the management of toe protection structures anywhere. More
specifically, it is aimed at all those with a direct interest in the management of flood
and coastal erosion risk management assets in England and Wales, such as the
Environment Agency, local authorities, drainage authorities, private owners of coastal
frontages and consultants. In the UK, and specifically in England and Wales, a large
proportion of this type of work is undertaken by or for the Environment Agency, and
as a result the standard procedures tend to be dominated by those produced for the
Environment Agency. Nonetheless, the principles set down in Environment Agency

Toe structures management manual


8
guidance are to a large extent valid for all operating authorities and so this manual
draws heavily on current Environment Agency practice.
This manual is one of the suite of guidance documents covering the full range of
flood and coastal erosion risk management assets. It is structured specifically to suit
the practical needs of flood defence and coastal protection asset managers.
The manual is intended to provide those responsible for flood defences, coastal
erosion protection structures and other coastal structures with practical guidance on
how to determine, monitor, assess and mitigate for the effects of toe scour. It
describes the types of toe protection structures and provides illustrations of typical
designs that are often used as solutions. The case studies on beach lowering and
scour management provide real examples of both good and bad practice, and
discuss lessons learnt from past schemes.

1.6 Acknowledgements
This guidance draws on many sources of information. However, it is strongly guided
and influenced by published outputs from the Environment Agency project,
‘Understanding the Lowering of Beaches in Front of Coastal Defence Structures’
Phases 1 and 2 (projects FD1916 and FD1927).
FD 1916, commissioned under the Joint Defra/Environment Agency Flood and
Coastal Erosion Risk Management Research and Development Programme,
recommended the development of a ‘toe scour’ guide to give practical guidance on
the prediction of toe scour and the options for mitigating toe scour by introducing new
knowledge gained from recent research and translating it into good practice.
This manual has been developed from the outputs of the Environment Agency’s Toe
Structures for Coastal Defences (SC070056) project undertaken by HR Wallingford,
which provided much of the research base for the content of this manual.

Toe structures management manual


9
2 Toe structure types and
materials

Chapter 2 introduces the common types of toe used in terms of


structure, materials and function. The manual is intended for coastal
managers and it is appreciated that most of the time they will be
dealing with existing rather than new structures. This chapter
describes the applicability of different toe types in such a way that it
supports the choice of option when new works are planned.
Importantly, it also highlights some of the advantages and
disadvantages that may arise with existing structures.

Key links to other chapters:


o Chapter 3 – Asset management
o Chapter 4 – Maintenance
o Chapter 5 – Toe design

Who will be interested in this chapter?


o Asset managers
o Coastal engineers
o Contractors

2.1 Introduction
The toe is the seaward edge of the foundation of a coastal defence. It can have a
major impact on its ability to:
• withstand beach lowering and scour;
• resist wave action;
• protect against wave overtopping.
The toe is sited in that critical location where the man-made defence meets the
potentially variable beach, and where the sea dissipates its energy.
The toe can often be a sizeable portion of the defence construction and as such its
effect on the human and natural environment should be carefully considered. This
chapter introduces the common types of toe used in terms of structure, materials and
function.
The manual is intended for coastal managers, who will mostly be dealing with
existing structures rather than new ones. This chapter tends to talk about the
applicability of different toe types in such a way that it supports the choice of option

Toe structures management manual


10
when new works are planned. However, it also highlights some of the advantages
and disadvantages that may well arise with existing structures.
Section 2.2 presents the different types of toe construction and their use. The
materials used in toe construction are discussed in Section 2.3. Toe design is
discussed in Chapter 5.

2.2 Types of toe structure


In order for the toe to be secure over its design life, it needs to be founded at such a
level or in such a way that it will not be undermined. It therefore needs to:
• extend to a level below the lowest expected beach level; or
• be of such a form that it can accommodate lowering beach levels; or
• be of such a form that it can be extended if the need arises.
The toe’s functional design normally serves one or both of two primary functions:
• the retention of the foundation stratum, or subgrade (that is, the
prevention of underscour);
• the ability to take foundation loads directly.
Table 2.1 lists the different types of toe together with a summary of their important
characteristics and their applicability to different situations that may arise. In certain
cases, these types can be used in tandem, for example, using concrete infill to
remedy previous undermining, secured by toe piling to resist future beach lowering.
The various types of toe structure are then described in more detail.

Toe structures management manual


11
Table 2.1 Main characteristics of different toe types
Type of toe General Use on existing Key design Hydraulic Environmental Comments
applicability defences considerations characteristics impact

Vertical Sheet piling – Exposed situations Underpinning, Earth pressures, Toe face is If exposed face is Vulnerable to abrasion
steel, Can accommodate when used together hydrostatic loads, reflective to waves high, can prevent and corrosion.
concrete, rise and fall of with suitable minimum beach if exposed above beach access and Requires access for
timber beach levels within backfill, capable of levels. Achievable beach. Introduction form safety hazard. piling equipment.
acceptable limits. taking major design life limited of sheet piled toe New piling can be
Can be used to foundation loads by abrasion/ can increase wave driven in front of
provide cut-off of corrosion losses. overtopping and existing, if necessary,
groundwater flows. exacerbate local to counter further
beach scour. beach lowering.

Cribwork – Mild to moderately Prevention of Earth pressures, Relatively Vulnerable to


timber or exposed locations erosion of cliff toe retention of rock absorptive if of abrasion.
concrete frame Can accommodate fill by cribwork, sufficient width Requires access for
retaining/ rise and fall of retention of earth piling equipment.
backfilled with beach levels within behind, life limited If materials are readily
rock acceptable limits by abrasion available, can be very
Permeable to losses cost-effective and easy
groundwater flows to construct
Most commonly
used as a toe to a
cliff.

Masonry or Moderate to Filling scour holes, Earth pressures, Toe face is More appropriate to Potentially very
concrete exposed situations subject to good potential for reflective to waves an urban setting from durable
blockwork Relatively rigid foundations settlement if exposed above a landscape
structure that beach. Can viewpoint
requires firm contribute to scour.
foundations.

Rock-filled Areas of mild Unsuitable for Retention of earth Relatively Vulnerable to Can be very cheap
gabions exposure major foundation behind absorptive if rock is vandalism solution, requiring only
Flexible (will loads, but can be Life limited by of adequate size Forms a personal lightweight plant, if
accommodate used to fill scour safety hazard when suitable rockfill readily

Toe structures management manual


12
Type of toe General Use on existing Key design Hydraulic Environmental Comments
applicability defences considerations characteristics impact

settlement) holes, particularly if abrasion and grading. gabions are broken available.
Permeable to only rarely exposed
groundwater flows to waves and
abrasion.

Concrete infill Rigid and Filling scour holes Foundation loads Reflective Care is needed to Can be in situ with a
vulnerable to and taking Achieving achieve good face shutter of
fracture if foundation loads, required strength appearance. permanent in situ
foundations are not subject to good in tidal zone facing, or bagwork can
sound foundations be used. Needs
careful consideration
in tidal conditions.
Concrete toe Rigid Filling scour holes, Earth pressures; Toe face is Care is needed to Choice between
beams subject to good dead weight to reflective to waves achieve good precast or in-situ
foundations retain structure if exposed above appearance. construction,
behind. Requires beach. Can dependant on scale of
suitable contribute to scour. scheme and available
foundation. tidal working window
considerations.
Concrete surfaces
may be vulnerable to
abrasion.
Sloping General note for sloping aprons: of potential use for countering the effects of beach lowering or scour. Requires greater ‘land take’ than vertical solutions.
aprons

Rock Exposed situations Providing a flexible Stability against Depending on Beach access and Potentially very
Flexible layer in front of the wave attack configuration, can safety of beach users durable if using
construction that defence that can Design of reduce wave are important appropriate quality
can accommodate adjust to a limited underlayers to reflections and local considerations. rock
settlement. extent to avoid settlement scour, and also
accommodate the into the beach. change overtopping
effect of beach characteristics.
lowering and limited

Toe structures management manual


13
Type of toe General Use on existing Key design Hydraulic Environmental Comments
applicability defences considerations characteristics impact

scour.

Flexible Protected and Providing a flexible Stability against Depending on Requires careful detail
revetment mildly exposed layer in front of the wave attack configuration, can at edges (edge beam
systems situations defence that can Ability of reduce wave or revetment
depending on adjust to a limited revetment to reflections and local excavated below
weight and type of extent to retain underlying scour, and also lowest design beach
system. Flexible accommodate the material change overtopping level) to preserve
construction that effect of beach characteristics. overall integrity.
can accommodate lowering and limited
settlement. scour.

Concrete Exposed locations Protection to base Stability against Can modify Suitable designs can Requires careful
slopes and depending on of defence against wave attack overtopping. readily allow beach design to achieve
steps strength and weight beach lowering and Requires characteristics. access adequate durability.
of construction scour adequate Precast option may be
foundations. suitable dependant on
scale of scheme and
available tidal working
window
considerations.

Gabion Protected and Protection to base Stability against Can modify Vulnerable to
mattresses mildly exposed of defence against wave attack overtopping vandalism
situations beach lowering and Ability of characteristics. Forms a personal
depending on scour revetment to safety hazard when
weight and type of retain underlying gabions are broken.
system. Flexible material
construction that
can accommodate
settlement.

Toe structures management manual


14
2.2.1 Sheet piling
Steel sheet piling is a common means of securing a defence against the threat of
undermining (Figure 2.1). This technique is commonly used to stabilise foundations
and reduce loss of fill materials, thus prolonging the life of the overall defence.

Figure 2.1 Underpinning of seawall toe with steel sheet piles (courtesy HR
Wallingford)
Steel sheet piling is used both for new works and for restoration to secure a defence
structure against undermining and instability (for example, a seawall and/or higher
ground to landwards). 1 The characteristics that make steel sheet piling particularly
suitable are its tensile strength and its form, which enable it to be driven to
considerable depths (subject to suitable ground) without the need for excavation.
The role of sheet piling in preventing undermining is self-evident. Its role in
restoring/ensuring geotechnical stability of a coastal defence is also often important.
Present day design requirements (factors of safety for stability) can be stricter than
those used in earlier (for example, Victorian) design and construction. When combined
with long-term beach lowering, this can put considerable demands on a newly installed
toe, resulting in considerably longer piles than might be needed on the basis of
undermining alone. Moreover, anchor ties can be difficult and expensive to install
beneath an existing structure and so heavier section cantilever piles are often used to
avoid this complication. The pile section must be chosen to withstand the effects of
future beach lowering and hence geotechnical loading. This, as well as corrosion, may
limit the ‘design life’ of the structure.
Figure 2.2 shows a further example of the use of sheet piling.

1
Examples of the use of sheet piles in toe protection schemes can be seen at the South Beach,
Lowestoft, Overstrand, north Norfolk, and Fort Wall, Canterbury, case studies in Appendix C.

Toe structures management manual


15
Figure 2.2 Sheet pile underpinning

2.2.2 Cribwork
Cribwork is a low-cost form of coast protection comprising rock-filled cages formed by
piling and fabricating timber lattices into a continuous structure (Figure 2.3). In
essence, they are not dissimilar to gabions except that they are made out of timber or
possibly reinforced concrete. Cribwork can protect the toe of defences by absorbing
wave energy and preventing scour of beach material and undermining of the structure.

Figure 2.3 Cribwork and concrete block fill, Norfolk (courtesy of North Norfolk
District Council)

2.2.3 Masonry or concrete blockwork


Many of our urban seawalls, built in Victorian times, are masonry and, in many cases, it
can be an attractive option for new construction. Nowadays, concrete blockwork is an
alternative to stonework in terms of cost and ready availability.

Toe structures management manual


16
The survival of hundreds of miles of Victorian masonry seawalls around the coast of
Britain, many still in excellent condition, is testament to the durability of this form of
construction. However, masonry is essentially rigid, reliant on the structural integrity on
the outer shell of blockwork for its strength. Masonry requires a solid foundation such
as a hard stratum or prepared concrete. Beach lowering, live loads on the wall from
traffic and so on, leaching out of backfill, abrasion and wave impacts can all combine to
damage or remove the blocks at the toe; Figure 2.4 shows an example of an extensive
crack. Failure of the foundation ensues (even if only locally) and this can threaten that
essential structural integrity and lead to rapid failure of the whole wall.

Figure 2.4 Extensive crack in masonry toe (courtesy HR Wallingford)


As soon as deterioration of the toe is identified, steps should be taken to repair the
damage before it spreads. In this case, the use of replacement masonry is clearly an
option as it offers durability and lack of visual intrusion. This is particularly important if
the wall is a listed structure. However, before selecting the option, the cause of the
problem (for example, whether it is the result of long-term beach lowering, a severe
event or progressive attrition of the previous blockwork) should be determined to
identify whether mere reinstatement of the blockwork is sufficient or whether additional
measures (such as toe piling) are warranted. Care must be taken to ensure that any
voids in the backfill are filled and that the replacement blocks are well founded.
Guidance on masonry walls is given in CIRIA publication B13 (CIRIA 1992).

2.2.4 Gabions
Gabions consist of steel mesh forming baskets that are filled with stones (Figure 2.5).
They are available either as gabion boxes (and can be used as a flexible toe), or as
mattresses (in which case they are used as a sloping apron where their flexibility allows
them to accommodate beach lowering with time).
Gabions can be sufficiently flexible to fit an irregular seabed, are relatively cheap to fill
and can be relatively easy to place. They are not suited to exposed coastlines and
beaches with moderately aggressive sea conditions. Although they absorb wave
energy, their resilience to such forces is limited becoming distorted, deformed and
broken with subsequent loss of fill material. Gabion baskets can become a hazard to
beach users and to wildlife when damaged/collapsed.

Toe structures management manual


17
Figure 2.5 Gabion baskets of rock (courtesy HR Wallingford)

2.2.5 Concrete infill


Concrete infill can be used to effect repairs to the toe of a rigid (masonry or concrete)
sea defence. The infill will be rigid and requires good foundation. When placed in situ, a
face shutter may be necessary and the use of special finishes or facings may be
required where visual appearance is an issue, particularly on listed structures. The use
of concrete bagwork can be effective in filling holes where access is difficult.
Concrete infill at the toe, while being an effective means of filling voids, is often not
sufficient in itself and may require other measures as well, such as toe piling.

2.2.6 Rock aprons


Rock can provide a quick and cheap method of filling in a scour hole. Rock dumps
should be monitored to ensure that it continues to provide protection.
Armour rock had principally been used for breakwaters and other industrial
applications, but its introduction as an acceptable material for use in UK coastal
defence led to its increased use in such schemes to extend the life and to improve the
performance of seawalls around the UK.
A modest amount of rock placed at the toe of a defence structure may serve to protect
its toe from undermining, reduce the abrasion of its front face and even modify the
hydraulic performance of the overall defence (Figures 2.6 and 2.7). The latter can
result in changed overtopping or different loads on the defence. However, the
introduction of the rock toe has at times increased overtopping, so it must be carefully
considered in the design process. In addition, there are sometimes concerns about the
impacts on aesthetics, access and public safety especially where such schemes are
installed on beaches of high amenity and recreational usage.

Toe structures management manual


18
Figure 2.6 Rock infill of scour trough, Le Dicq, Jersey (courtesy HR Wallingford)

Figure 2.7 Timber bulkhead with rock toe protection at Lepe, Hampshire

2.2.7 Flexible revetment systems


A revetment is intended to protect a slope, in this case against modest wave and
current action. The protective layer may take the form of rubble, gabions (see
Section 4.2.4), timber, masonry, concrete, concrete bags, concrete block mattress,
concrete armour with rubble underlayers, and other resistant coverings. Using a
revetment as a toe may simply be the continuation of the revetment that forms the main
body of the defence below the beach/ground level or extended as a scour apron. A
flexible and adaptable form of construction is often sought to cater for variations in the
depth of the embankment both along its length and with time. Depending on the
circumstances, a more sophisticated structural toe may be required.
There are a number of proprietary flexible revetment systems available. The systems
consist of small concrete blocks which either interlock or are tied using plastic tendons
to form mattresses that can be placed with a crane. Being proprietary systems their
methods of application are set out in the manufacturers’ literature. Some of this
information is based on physical model tests, and this is to be preferred.

Toe structures management manual


19
These flexible systems can be placed directly on a graded granular slope of a suitable
particle grading, or a geotextile to retain the finer underlayer.

2.2.8 Concrete slopes and steps


The construction of an apron or steps at the base of an existing structure (Figures 2.8
to 2.10) can prolong the life and add amenity value to the defence, albeit at the
expense of losing an area of the beach. This can be achieved at much reduced cost
compared with rebuilding the defence entirely. Such additions to a structure will extend
it seaward. There is a perceived danger that such seaward extensions of a structure
will interfere with longshore sediment transport and hence reduce sediment supply to
downdrift beaches. However, this must be put in perspective given the very small
extent of the structural intrusion compared with the width of the seaward extent of
active sediment transport.
Concrete slopes and steps have been used frequently around the UK and new
techniques have been developed, for example, using a sloping asphalt apron (see
Figure 2.18) to both protect the original structure against undermining and abrasion,
and to reduce wave overtopping.

Figure 2.8 Extended scour apron, masonry steps and armour, St Ouens Bay,
Jersey (courtesy HR Wallingford)

Toe structures management manual


20
Figure 2.9 Concrete stepped revetment, Crosby, UK (courtesy Sefton Council)

Figure 2.10 Supplemental rock armour revetment (courtesy of HR Wallingford)

2.3 Materials
The properties of construction materials for application to seawalls and other marine
structures are described in:
• The Use of Concrete in Maritime Engineering: A Good Practice Guide
(CIRIA 2010b);
• The Rock Manual (CIRIA et al. 2007);
• Potential Use of Alternatives to Aggregates in Coastal and River
Engineering (CIRIA, 2004);
• Chapter 4 of Coastal Engineering Manual (USACE 2012).
A substantial part of the advice given in these references and elsewhere is relevant to
toe design, construction and management.
The position of the toe in a structure and hence the loading on it, together with its
particular exposure to wave and tidal action, means that in relation to the whole
structure, some material properties are highly significant while others are not. This

Toe structures management manual


21
section provides an overview of the relevant properties of materials with particular
reference to those aspects that are especially important to toe protection.
The brief descriptions for each material cover the following key aspects where they are
of particular significance:
• application;
• density;
• strength;
• flexibility;
• resistance to chemical or biological attack;
• durability (strictly speaking this is a relative measure of the ability of a given
material to withstand the pressures of the environment in which it is used).
Figure 2.11 shows a seawall demarked into different zones designated 1 to 4 where
Zone 1 is at the crest of the wall and Zone 4 is at the lowest level. Depending on the
position of the structure in the foreshore, the toe could lie either in Zone 3 (if it is above
low water) or Zone 4. These broad demarcations provide a useful reference in the
discussions which follow.

Figure 2.11 The weathering zones of a seawall (from Thomas and Hall 1992)

2.3.1 Steel
The most serious threats to steel when used in coast protection works are corrosion
and abrasion.
BS 6349 sets out the standards for the design and construction of maritime structures,
including the use of piling (BSI 2010).
The issue of durability is discussed in detail in the Piling Handbook (Arcelor 2008).
The loss of pile thickness due to corrosion only is given in the Piling Handbook, based
on Eurocode 3, Part 5. Relevant values for loss of thickness are given in Table 2.2.

Toe structures management manual


22
Table 2.2 Loss of thickness (mm) due to corrosion for piles and sheet piles in
sea water 1
2 3
Location of pile/sheet pile Required design working life (years)
5 25 50 75 100
Sea water in temperate climate in the zone of
0.55 1.90 3.75 5.60 7.50
high attack (low water and splash zones)
Sea water in temperate climate in the zone of
0.25 0.90 1.75 2.60 3.50
permanent immersion or in the intertidal zone

1
Notes: The values given are for guidance only.
2
The highest corrosion rate is usually found at the splash zone or at the low water
level in tidal waters. However, in most cases, the highest bending stresses occur in
the permanent immersion zone.
3
The values given for 5 and 25 years are based on measurements, whereas the
other values are extrapolated.

The user should refer to the source references when assessing corrosion as it is an
important and complex issue, covering not only general corrosion but also localised
factors.
A particularly aggressive form of corrosion is the microbiological process known (in
Britain) as accelerated low water corrosion (ALWC). In spite of its name, ALWC is not
confined to low water or just above lowest tide level, and it does occur at other levels.
In situations that favour ALWC, surface corrosion rates can be more than 1 mm per
year. CIRIA C634 provides a comprehensive guide to the phenomenon of ALWC and
its management and mitigation (CIRIA 2007).
Depending on the environment and particularly the nature of the sediments, steel loss
will occur to varying degrees through abrasion. This aspect is less well quantified than
corrosion as it depends on a number of variables including exposure of the steel, the
prevalence of abrasive material (for example, shingle), and the level of wave activity –
albeit a coarser sediment such as shingle is more likely to be encountered in a higher
energy environment.
Upper rates of steel loss by corrosion and abrasion are listed by Thomas and Hall
(1992). This reference indicates that, depending on the circumstances, the contribution
to material loss by abrasion can be considerable, for example up to 85 per cent (of
1 mm per year) in Zone 3 (Figure 2.11) where steel is exposed to a high percentage of
gravel and a severe wave climate.
A further important property of steel sheet piling is its relative impermeability. Seepage
losses are generally minor and restricted to seepage through the clutches (the
interconnections), which may be reduced by the presence of fine sediments.
Deep and continuous piling can intercept groundwater flow paths which might be due
to tidal action (flowing both ways) or fresh ground water. The effects of this need to be
assessed on a case by case basis; consequences might include a raised water table
on the landward (geotechnically active) side. Mitigations might include the inclusion of
weep holes or slots, or intermittent shorter piles to relieve the drainage path.

Toe structures management manual


23
2.3.2 Rock and masonry
Rock and masonry both refer to the use of stone. The basic geological characterisation
of different rock types are given by Fookes and Poole (1981). The Rock Manual (CIRIA
et al. 2007) provides a comprehensive introduction to different rock types.
Whereas the term ‘rock’ is usually applied to loose stone used for armouring a sloping
revetment or apron, 2 the term ‘masonry’ always refers to stone placed in a compact
pattern, usually with a relatively smooth outer surface, often grouted (see Figure 2.12).
Masonry is often used both for armouring and to provide a decorative appearance. The
principle mechanical properties of stone (hereafter referred to as rock) that make it an
attractive material for use in toe construction are:
• its high compressive strength;
• its load bearing capacity;
• its high density (for inherent stability);
• its ability when placed irregularly to form a durable flexible toe.

Figure 2.12 Grouted stone/masonry (courtesy HR Wallingford)


Partially grouted armourstones combine the high resistance against currents and
waves of large elements and their flexibility to adapt to ground deformations and the
option of installing comparably thin layers. Stone/rock diameters of 10–40 cm and a
narrow rock size distribution are best for being grouted.
Partial grouting (Figure 2.13) is a reliable and well-established method to meet the
requirements for a long-lasting scour protection including sufficient permeability to
avoid excess water pressure below the armour layer. Partial grouting means filling the
voids in the riprap or stone layer to 35–50 per cent with a special mortar, creating a
bonded layer with high resistance and high permeability and sufficient flexibility.
Cement-bonded grouting materials are recommended for partial grouting (Heibaum
and Trentmann 2010), although the advantages and disadvantages of using asphalt as
a binder is discussed in Section 2.3.4.

2
Examples of the use of armourstone for toe protection can be seen in the Clayton Road,
Selsey, and Corton, Suffolk, case examples in Appendix C.

Toe structures management manual


24
Figure 2.13 Grouted stone toe apron, Thames Estuary (courtesy HR
Wallingford)
The strength of rock differs greatly between the various rock types, but those used for
construction in the marine environment will generally be required to have good
compressive strength.
The degradation of rock is due primarily to mechanical damage by way of fracture,
breakage and abrasion, although ice can become a significant factor in colder climates.
Fracture and breakage are more likely to occur in loose rock used for armour rather
than set in masonry. In this case, the breakage of armourstones yields smaller stones
which, being more susceptible to movement, become increasingly prone to further
breakage and abrasion. Abrasion reduces the mass of stone used in both rock armour
and as masonry. Further spalling can be due to salt attack.
Density is an important parameter if high rock weight is required (for example, for
armouring). High density tends to go hand in hand with high strength and durability.
Other properties of rock relevant to its application in toe construction works are outlined
below:
• Permeability of the armour layer (rather than the rock itself).
Permeability (drainage) is an important issue and introduces a clear
distinction in the application of stone when used as rock armour or as
masonry. Permeability is very important in toe design as it can determine
the drainage of ground water behind a structure. The groundwater level
may be determined both by freshwater and seawater, the latter arising from
retained tidal water or by wave overtopping onto a more porous upper part
of the structure. Clearly, loose rock affords better drainage than a grouted
masonry revetment. Another aspect of permeability, or porosity, relates to
wave energy dissipation. For example, a rough porous rubble slope has
much better energy absorption through internal turbulence than a smooth
wall.
• Compliance/flexibility can be important attributes for toe protection where
it is required to accommodate bed movement, by way of a falling (or
settling) apron for instance. Clearly, this attribute cannot be enjoyed in the
case of consolidated masonry.
• Availability of rock is an important consideration, both logistically and in
terms of cost, as the source and travel distance from source can add
significantly to the price. For large constructions (for example, whole

Toe structures management manual


25
revetments or breakwaters), the cost of importing large rock may be better
justified than for small-scale works involving attention to the toe only. If the
toe is below low water, installation may require the tipping of rock; hence
there is a need to allow for generous tolerances.

2.3.3 Concrete
Concrete consists of three fundamental components: cement, aggregate (sand and
gravel) and water. Different combinations of these ingredients provide for a wide range
of construction materials ranging from mortars used for grouting or rendering, to high
strength structural concrete.
For maritime applications, useful guidance for the design, construction, testing, repair
and maintenance of many different types of such structures can be found in The Use of
Concrete in Maritime Engineering; A Good Practice Guide (CIRIA 2010b). In particular,
it highlights how achieving good quality concrete in the tidal zone is challenging, while
also pointing out how necessary it is to achieve good quality dense concrete in order to
achieve a durable structure in terms of both resisting abrasion and preventing corrosion
of reinforcement.

Figure 2.14 Concrete toe beams (courtesy HR Wallingford)


A significant feature of concrete for use in seawalls is that it can be formed to suitable
shapes which may be both functional and decorative. The latter attribute is not usually
so important in the case of the toe where it is less visible than the upper parts of the
structure. Concrete will more commonly be used in toe protection works to provide
weight and compressive strength. Figure 2.14 shows an example of the use of
concrete toe beams and Appendix C contains a case study describing the use of
concrete toe beams for foundation strengthening along the seawall between Dawlish
and Teignmouth, South Devon, UK.
Although the specific gravity of concrete is usually lower than that of good quality rock
(for example 2.4 versus 2.6), it can be cast into larger volumes of convenient shape to
match the construction details. It follows that concrete is often used where there is a
need to provide ‘fixity’ at the toe without the use of piling (for example, toe beam to a
single layer armour system or as passive resistance to geotechnical pressure).
Good quality mass concrete can be designed for the life of a project, for example, 100
years. Longevity of the concrete is not therefore necessarily a design life determinant,
although service life can be compromised by:

Toe structures management manual


26
• difficulties in obtaining good quality concrete if cast in situ below water;
• chemical resistance – more particularly in respect of reinforcement;
• use of aggregates that are insufficiently durable.
In the case of reinforced concrete, corrosion of steel reinforcement causes the bars to
expand leading to fracture and bursting of the concrete (spalling) (Figure 2.15),
possibly leading to failure. After spalling, the reinforcement is even more vulnerable to
corrosion through direct exposure to air and water. This process is particularly
significant in parts of the structure subject to wetting and drying, as illustrated in Zone 3
of Figure 2.11.

Figure 2.15 Reinforcement rebar corrosion and spalling of concrete (courtesy


of HR Wallingford)
As time goes by and the concrete is subject to repeated cycles of wetting or
submergence, chlorides from the seawater penetrate the concrete and reduce its
alkalinity. The time for this penetration to seriously affect the steel depends on the
thickness of cover and the permeability of the concrete (for example, penetration of 25–
75 mm would typically take about ten years). The quality of the concrete is important as
chlorides can penetrate very small cracks. Figure 2.16 shows an example of salt
deposits with associated cracking, while Figure 2.17 shows the failure of a concrete toe
beam on a Jersey beach.

Toe structures management manual


27
Figure 2.16 Surface salt deposits and associated cracking from alkali–silica
reaction in concrete (courtesy of HR Wallingford)

Figure 2.17 Failure of a concrete toe beam at St Ouens Bay, Jersey (courtesy
HR Wallingford)
There are ways of reducing the ingress of seawater and consequent corrosion of steel
reinforcement including:
• avoiding contact between concrete and seawater during curing;
• inclusion of pulverised fuel ash (PFA), though this increases curing time
which can be problematic for intertidal construction;

Toe structures management manual


28
• reduction in curing time with the use of rapid hardening ordinary Portland
cement (OPC) as a means of reducing exposure to seawater;
• sulphate-resisting cement may be used (especially where waters are near
to warm outfall discharges);
• use of various protective coatings to inhibit the corrosion process.
Clearly it is preferable to avoid the use of reinforcement in concrete wherever possible,
but reinforcement may be required for certain details such as pile capping. Here, the
need for having the capping (at all) could be considered; for example, it may not be
needed if the capping is to remain covered by sand (but the designer needs to be
aware of future risks if the capping becomes exposed).
In simple terms, good quality dense concrete will be durable. In order to achieve good
quality, precast elements can produce a high level of durability and they are often
appropriate in cribwork, toe beams and flexible revetments.

2.3.4 Asphalt and bitumen


Asphalt is the term given to a range of materials which comprise, in various ratios,
bitumen (the binder), fillers and aggregates. Different types include:
• sand mastic (used for grouting blockwork or as a carpet below water);
• lean sand asphalt (uses include as a filter layer to upper armouring, also as
underwater protection against scour);
• open stone asphalt (scour resistant revetment);
• asphaltic concrete (used for load-bearing surfaces).
Figure 2.18 shows the use of asphalt to grout a stone revetment.

Figure 2.18 Asphalt grouted stone revetment in the Thames Estuary (courtesy
HR Wallingford)
In respect of the pressure of the marine environment, asphalt is generally regarded to
be of poor strength, in terms of both tension and compression. Its durability is
dependent on exposure. A key advantage of asphalt in toe construction is that it is

Toe structures management manual


29
relatively flexible, that is, because of its thermo-plastic properties it can accommodate
ground movements better than, for instance, concrete.
The permeability of asphalt is another important consideration. This varies according to
the type of mix; for example, sand mastic is impermeable while open stone asphalt is
porous and relatively permeable. As with sheet piling, the permeability of the asphaltic
material used, and how it is applied, is clearly important in toe protection design for the
avoidance of groundwater build-up and back pressure.
A further consideration is that of abrasion resistance. Asphalt is vulnerable to abrasion
in an environment where shingle is present. This can seriously affect the life of a
structure.
Some further outline notes are given below in respect of the different forms of asphalt.
• Sand mastic is impermeable but is a weak material that can rupture under
pressure.
• Lean sand asphalt has better permeability and is inexpensive, but
depending on circumstances, may only have a short life. It could, however,
be used as an interim or short term measure.
• Open stone asphalt can be used for the body of a wall but should not be
placed underwater as boiling damages the binder; this can limit its
application to toe structures in some cases. While open stone asphalt is
relatively permeable to the extent that it can alleviate tidal differential, in
respect of wave activity it is effectively smooth (that is, it is not a good
energy absorber). It is also vulnerable to erosion. Toe scour protection
using a mastic slab or stone/mastic has been used in the UK for example at
Dovercourt near Harwich in Essex and in several locations on the Wirral.
• Asphaltic concrete is not likely to be used for toe construction.
Further detail on material compositions, properties and general guidance on the use of
asphalt can be found in:
• The Shell Bitumen Hydraulic Engineering Handbook (Schönian 1999);
• The Use of Asphalt in Hydraulic Engineering (van der Velde et al. 1985).

2.3.5 Timber
Timber is widely used in jetties and other marine structures. Good compressive and
tensile strength means that timber can be formed into strong lattice structures, piles,
piers and platforms. Generally, the denser the wood, the stronger and more durable it
is. Greenheart is classed as being of exceptionally high strength and durability.
Though less durable than concrete, the harder woods offer good durability – the life of
a structure being sometimes dictated by that of the bolts and other corrodible fittings
that hold the structure together than the life of the timber itself.
Timber has been used in cribwork (see Figure 2.3). Timber planks were often used as
pilings to seawalls before the introduction of steel piles.
The two main causes of timber decay in the marine environment are biological attack
and abrasion.
• Biological attack. Timber is vulnerable to damage by marine borers. The
resistance to biodegradation depends on the wood species (for a list, see
Thomas and Hall 1992). Two common kinds of infestation are ‘gribble’

Toe structures management manual


30
(Figures 2.19 and 2.20) and ‘Teredo’. Gribble attacks the surface of
exposed timber, mainly below mean tide, penetrating a few millimetres
below the surface, whereas Teredo destroys the internal composition of the
wood. Both gribble and Teredo are encouraged by clean warm water
discharges and are less prevalent in more polluted environments. The more
durable hardwoods used for marine construction tend to be resistant to
biological attack and are thus able to provide a good service life in excess
of 20 years or much longer in certain cases.

Figure 2.19 Gribble attack on Timber pile at the sediment line (courtesy of HR
Wallingford)

Figure 2.20 Close up of gribble damage to timber (courtesy of HR Wallingford)


• Abrasion. Persistent abrasion by both sand and shingle has the effect of
rounding off the corners and details of timber in the sea (Figure 2.21). As
this effect tends to be concentrated close to the bed where sediments move
more, the abrasion tends to be concentrated over a short length – a feature

Toe structures management manual


31
sometimes referred to as necking (Figure 2.22). Evidence of necking high
above the bed can be indicative of a low beach (compared with normal or
former times).

Figure 2.21 Rounding of groyne timbers by beach shingle (courtesy HR


Wallingford)

Figure 2.22 Necking in derelict timber piles (courtesy ENBE)


The specification of hardwoods for use in coastal defence needs to recognise the
availability of source and the use of sustainable forest supplies. Alongside this, there
may be scope to reuse existing timbers from decommissioned earlier works (for
example, estuarial boat mooring piles), possibly not for main structural members, but
for hand-railing or similar ancillary works.
Detailed advice on the use of timber in coastal engineering can be found in Crossman
and Simm (2004).

Toe structures management manual


32
2.3.6 Geotextiles
Geotextiles are used mainly as filters and separators in coastal construction, primarily
to prevent the migration of fine sediments through a coarser over layer of rock. The
design must balance drainage capacity with particle retention (a typical application is
for retained sediment sizes in the range 0.06–2 mm), while satisfying load conditions
for construction and service life survival.
Made from polymer filaments or fibres, geotextiles are formed into textiles by weaving
or other means (for example, welding). Woven textiles tend to have a smaller range of
opening sizes compared with non-woven materials.
Geotextiles can also aid the distribution of armour point loading across soft substrate
conditions.
If the geotextile is to be used, it must:
• be undamaged;
• provide cover in a controlled manner;
• be laid in such a way that it is continuous over the area where it is required
(for example, with sufficient overlap where sheets adjoin).
The use of geotextiles and constructability of toe details was investigated in a prototype
trial at Milford-on-Sea in Hampshire (see Box 2.1).
While geotextiles have a good inherent strength, they must be selected to match the
loading (for example, the placing rocks). The strength of woven materials is determined
by the direction of loading, so care needs to be taken when placing in relation to the
applied load (for example, rock placed on a slope). To avoid the risk of damage, non-
woven geotextiles may be preferred. The loading constraints on a given material
should be obtained from the manufacturer prior to specifying works.
It should also be noted that geotextile has a lower natural angle of friction than rock
such that slopes steeper than 1 in 2 should be avoided.
Geotextiles are susceptible to degradation resulting from ultraviolet (UV) radiation
(sunlight). This can be inhibited by using UV inhibitors and by covering the material
during storage prior to installation. As part of a toe structure, the geotextile is likely to
be covered once installed into the structure.
The placing of geotextile in open seas can be difficult and needs to be carefully
planned. Some form of ballasting, either proprietary or bespoke, is usually devised.
Useful further information on functional design for geotextiles in hydraulic engineering
is described in Section 3.16 of The Rock Manual (CIRIA et al. 2007).

2.4 Choice of structure type and materials


The previous two sections provide an overview of the different types of toe structure
types and materials. The actual choice of a toe structure type(s) for a solution depends
on a number of factors and the selection or short listing of appropriate types for a given
application is the first step towards a successful design.
Factors that a designer should typically include when considering the type and
materials for a toe structure include:
• geotechnical loading;

Toe structures management manual


33
• integration with the structure it is required to support;
• design life;
• level of exposure to wave and/or current attack;
• predicted extent of future scour in front of the toe;
• visual impact, heritage aspects and so on;
• purpose (that is, new toe to new structure or new toe to old structure);
• availability of materials;
• constructability, taking into account aspects such as access, tidal working,
ground conditions, groundwater, integration with any previous toe structure
and environmental impacts (for example, due to excavation in the
foreshore);
• cost;
• level of future maintenance.
As the combination of these factors is likely to be unique for any given site, it follows
that it would be unwise to propose a generic toe structure design specification.
The designs illustrated and described in this manual are included only as case study
examples and information for the benefit of the reader and are not intended as ‘off-the-
shelf’ solutions. Nonetheless, some generic design principles and some tried and
tested toe scour and undermining mitigation options are useful. These are illustrated
throughout this manual and in particular in Chapter 5.

Toe structures management manual


34
Box 2.1 Paddy’s Gap Rock Revetment Works
Maintenance works carried out to coastal defences at Milford-on-Sea, Hampshire,
provided an opportunity to compare a number of different foundation details. The works
comprised the reconstruction of a 60 m rock revetment which protects against the
undermining of a concrete seawall within a groyne bay.
An existing revetment of 1–2 tonne Portland stone armour had been placed as
emergency works in 1995, but performance had deteriorated and it was judged to have
reached the end of its useful life by March 2003.
A new revetment was designed using existing rock as the underlayer and 3–6 tonne
Mendip limestone for armour, with the toe comprising larger stones up to 10 tonnes.
Three different designs, representing varying degrees of complexity, were built with the
time taken and ease of construction monitored. The performance of the different
designs will be assessed during future monitoring.
The most complex design (A) is illustrated below and includes both geotextile and
excavation for the toe rock. The intermediate design (B) omitted the geotextile and the
least complex (C) did not include any geotextile or toe excavation.
Min. 2 stones width
(approx. 2m)
Mendip limestone armour
(3 to 6 tonnes)
Single armour stone
at toe (approx 10 tonnes)

Approx. 2.2m MHWS

Approx. existing beach level


Approx. level of clay stratum
Existing Portland armour Approx. 1.5m
used as underlayer
MLWS
Geotextile laid on clay surface
in option A only Toe armour excavated into clay
in options A and B
Approx. 9m

Potential savings afforded by the designs were discussed with the contractor prior to
the works being undertaken and observed on site. The toe of the structure was found
to be critical since this is most affected by weather and the ‘tidal window’. Construction
of the slope could occur on a rising tide providing the toe had been placed properly.
The time saving resulting from the omission of geotextile was relatively small
(approximately 30 minutes for the 20 m section of revetment) and savings were found
to be minimal since the same plant (one tracked excavator with a bucket and one with
a rock grab) were required on standby for the whole of the work. The omission of an
excavated toe provided a further saving of 10 minutes; however, again both excavators
were still needed as the mobile beach had to be profiled even though the toe was not
excavated into the clay.
The least complex section did, however, provide significantly more flexibility in
construction which enabled work to be carried out during neap tides and in poor
weather conditions. This provided minimal savings in the example, but could have had
a significant benefit, enabling programme to be maintained, during more extensive
works. However, the savings made during construction must be gauged alongside the
longer term performance of the scheme.
Source: Sutherland et al. (2003)

Toe structures management manual


35
3 Asset management

Chapter 3 provides guidance on post-scheme asset management


activities such as asset performance, monitoring, data analysis, risk
assessment, deterioration and trigger levels for action. It establishes
a framework for practical use based on identifying the symptoms,
causes and possible cures for any toe scour problems, based
around a continual data monitoring programme.

Key links to other chapters:


o Chapter 2 – Toe structure types and materials
o Chapter 4 – Maintenance

Who will be interested in this chapter?


o Asset managers
o Coastal engineers

3.1 Introduction
Management of a structure toe can conveniently be considered by reference to two
separate but related elements:
• structure condition and deterioration;
• dynamic changes to the beach or foreshore.
In many instances, the integrity of the structure is controlled by the dynamic changes
arising from changes to the beach in combination with the forcing hydrodynamic
conditions. A structure that may appear to be in excellent structural condition may be
highly vulnerable to dynamic changes on the beach or foreshore, and these can impact
significantly on both performance and condition.
Changes in beach slope and level, relative to any flood defence or coastal erosion
protection asset, affect potential overtopping rates as well as the likelihood of
undermining and failure or breaching of the asset. It is therefore important to employ a
toe management system that is able to cope with dynamic changes to the foreshore.
Coastal defences (including beaches) need to be managed to maintain their level of
performance throughout their life cycle. This management will be required at a policy or
strategic level and at an operational level.
The age of the structure is a significant consideration as some elements will begin to
deteriorate more quickly than others. Elements such as joints in concrete and masonry
may be vulnerable. While other parts of the structure may be continue to be in sound
condition, damage to the weaker elements may lead to structure failure.
Knowledge of the original construction of the structure is crucial and access to ‘as built’
drawings provides a crucial part of the management process.

Toe structures management manual


36
3.2 Management overview
At the strategic level, consideration should be given to issues such as:
• Is the standard of protection afforded by the asset sufficient and in line with
the system standard overall? (That is, is the asset providing the required
level of protection alongside other assets in the system or is it a ‘weak
link’?)
• Planning of future works and maintenance spending (that is, will the asset
be refurbished, replaced or even decommissioned as part of a larger
scheme or a planned change in policy?).
At an operational level, the management considerations are slightly different:
• Is the asset performing as designed and required?
• What needs to be done physically to the asset in order to maintain its level
of performance at the required level?
• How often should the asset be inspected and how? Should it be monitored
on a more regular basis than is currently the case?
• What level of assessment is required?
• What data should be collected, how, and at what frequency and resolution?
Establishing an asset management policy and strategy, together with implementation
plans and operational procedures, should provide a purpose-made set of processes,
tools and performance measures that will enable achievement of an optimum approach
to managing assets. This strategy should identify issues such as those outlined above
and seek to provide answers and solutions. Such a strategy needs to be owned at
executive level, be evidence-based and be auditable in its application. For this a
suitable framework can be adopted from the British Standard Institution’s Publicly
Available Specification for the optimised management of physical assets, PAS 55 (BSI
2008). PAS 55 is applicable to any organisation that depends upon its physical assets
for the performance and continuance of its business operations.
The generic ‘Plan – Do – Check – Act’ framework in PAS 55 (Figure 3.1) provides a
template against which industries can develop or check their own approach to the
management of physical assets. The framework covers:
• Policy and strategy linked to corporate objectives and acceptable risk;
• Information, risk assessment and planning including information
systems, risk identification and assessment, leading to an asset
management plan with its priorities and targets;
• Implementation and operation focused on intervention to maintain,
operate and dispose of assets including such issues as responsibility,
training, awareness, communication and emergency response;
• Checking and corrective action including monitoring of performance and
condition, asset-related failures, corrective and preventive action;
• Management review and audit, completing the cycle and leading on to
continuous improvement.

Toe structures management manual


37
CONTINUAL IMPROVEMENT

gy
M evi

te
an ew

ra
R

St
ag
em

&
y
en

lic
Po
t
ASSET
MANAGEMENT
Asse
king
&
n SYSTEM t
Infor Managem
Chec tive actio plann
matio
n& ent
e c
corr ing

Implementation
& operation

Figure 3.1 The PAS 55 Plan – Do – Check – Act framework (from BSI 2008)
Under the ‘Information, risk management and planning’ phase, there are some
common risk-based issues that need to be considered in both the strategic and
operational management of coastal defences. These issues are:
• Precautionary approach – the need to ensure that proper analysis is
applied to decisions where there are uncertainties and/or a lack of
knowledge, and where there is potential for serious or irreversible
environmental harm (for example, exposing additional lives or property to
flood or erosion hazards);
• Proportionality – the need to ensure that resources are targeted at the
most significant risks and to demonstrate equity of benefits;
• Effectiveness – the need to provide a sound basis upon which to take
consistent flood defence and coastal erosion management decisions;
• Efficiency –the need to take consistent and defendable decisions which
allow the movement from historic and/or reactive unsustainable actions
(subject to legislative constraints) to strategic proactive decisions where
residual risk levels have been reduced to levels deemed acceptable and
sustainable.
This manual provides information on, and analytical tools for, beach levels and
structures at the toe of coastal defences that should aid asset managers or
practitioners in the ‘risk assessment and planning’, ‘implementation and operation’, and
the ‘checking and corrective action’ phases of the asset management system (AMS)
cycle in Figure 3.1. The following sections discuss:
• key failure and damage types;
• risk-based assessment and reliability analysis;
• determining trigger levels for action/intervention;
• beach monitoring and surveys for data collection and the appropriate
analysis of data;
• asset condition assessment;
• when to conduct surveys;

Toe structures management manual


38
• asset deterioration.
In many cases, there is no need for immediate action to mitigate beach lowering or toe
scour. Rather it may be sufficient to monitor the situation and to reduce the
consequences of the erosion and/ or flooding problems that are being experienced.
Equally there are many situations that demand immediate remedial action in order to
avoid brittle structure failure and expensive repair. It should be emphasised that toe
damage or failure can spread rapidly and endanger the integrity of the overall structure.
There is a need to assess the current condition and performance of coastal defences,
and to consider potential future beach lowering. Such an assessment requires
monitoring and analysis of survey results to identify recent patterns of change.
Predictive analysis can be conducted to estimate whether beach levels may lower in
the future and whether the coastal defence structures will be undermined. In addition, it
is necessary to carry out calculations to assess how the defences will be affected by
severe conditions (that is, high tides and large waves) for both present day and future
beach levels.
The majority of toe protection works undertaken are not ‘new build’ or ‘capital’ works,
but are commissioned for other reasons such as remediation, maintenance or repair,
reconstruction and augmentation, or strengthening of existing defence structures.
Within the context of asset management, maintenance, repair and reconstruction can
be seen as stages within the management cycle which occurs throughout the lifespan
of the defence structure. That is, continuous or intermittent monitoring and performance
assessment will/should determine whether interventions are necessary in order to
maintain the level of performance desired and offset deterioration. Figure 3.2 illustrates
this as a cycle of asset management processes and decisions. 3

Figure 3.2 Key phases for asset management of existing coastal defence
structures

3.3 Life-cycle failure modes


Potential failure modes must be considered carefully when developing a monitoring
and maintenance programme. This requires a basic understanding of design principles,
which are discussed in more detail in Chapter 5. The context of the design principles
must be considered in relation to the in-service performance of a system under a

3
This asset management cycle is very similar to the Frame-of-Reference for implementing
coastal erosion management policy in the Netherlands (see Mulder et al. 2011).

Toe structures management manual


39
dynamic loading. These are considered by reference to a simplified representation of
the forces acting on a gravity wall illustrated in Appendix B. The calculations and
approximations discussed in Appendix B require careful review to make fundamental
decisions:
• How much risk is acceptable before an intervention must be enacted?
• How much lead time is required to examine and implement potential
intervention options? That is, how long does it take to consider options,
consult, acquire appropriate permissions, raise funds, complete designs
and secure construction contracts (if required) before the risk of failure
becomes unacceptably high?
Asset managers should consider such time lags before intervention is essential. The
term ‘trigger level’ is usually used to identify the conditions of the beach and structure
at which intervention might be required. The context of beach trigger levels is
discussed further in Section 3.5. Appropriate trigger/alert levels must be set in relation
to the time to ensuing risk of failure at an unacceptable level. The level at which the risk
becomes unacceptable may be set as a matter of coastal policy. Examples of common
practical examples of in-service structure damage that may require remediation are
described and illustrated below.

3.3.1 Overturning and settlement of a gravity wall due to toe scour


Waves can cause scour holes at the toe of the seawall. Scour holes develop
underneath the base of the wall and reduce the surface of the base of the wall, which is
supported by the underlying foundation of the soil. The loading (weight of the structure,
horizontal ground force and horizontal hydraulic force) of the structure thus has to be
distributed over a decreasing foundation area. The developing scour hole results in a
decreasing width of foundation until there is insufficient force to support the structure
(Figure 3.3). If the structure has not been designed to withstand such pressures and
the condition is not prevented or quickly addressed, the structure will move in response
and progressively fail, or even fail completely. The point at which this is calculated to
happen is known as the ‘disturbing moment’ or the ‘ultimate limit state’ (ULS).

Figure 3.3 Overturning and settlement of a gravity wall due to removal of passive
resistance

Toe structures management manual


40
This ‘moment of failure’ can form the basis on which the ‘critical’ level can be set. It is,
of course, not normal to wait until this ‘failure moment’ arrives before a decision is
made to intervene. A factor of safety would usually be inferred from the likely rate of
deterioration towards failure – the critical point. This is problematic for scour processes,
as scour holes can easily form and infill again within a tidal cycle. Trigger or ‘alert’
levels for intervention (in advance of critical levels being reached) therefore need to be
based upon the potential or likely depth of scour in relation to the limit state equation
(Appendix B), given the preceding beach level and probability of dangerous loading
conditions (that is, how likely is it that the next scour event will undermine the structure
sufficiently for it to overturn or collapse).
These points are crucial as they form the basis from which subsequent calculations are
derived. Obviously, the aim is to avoid and prevent instability and failure of the
structure. In order to achieve this, pressures on the structure should not be allowed to
fall below the level at which disturbing moments (causing instability) will occur. If the
ULS is reached it may be too late to prevent failure.
Figure 3.4 shows an example of a seawall in the process of overturning.

Figure 3.4 A seawall in the process of overturning (courtesy of Black & Veatch)

3.3.2 Toe scour resulting in increased overtopping and structure


sliding
Although the symptoms of a seawall failure may appear to result from other causes
such as overtopping, these can often arise as a result of changes at the toe of the
structure. The example below demonstrates changes in conditions at a seawall, which
have arisen as a result of beach lowering:
• Beach levels reached the crest of the wall at the time of construction and the
wall was well protected, with shingle, limiting wave attack on the structure toe.
• Falling beach levels have exposed the structure toe piling (which extends 7 m
below the pile tops) (Figure 3.5). Although the structure continues to resist
overturning and sliding, increased water depths at the structure toe allow larger
waves to attack the structure.

Toe structures management manual


41
Figure 3.5 Exposure of interlocking steel piled toe arising from beach lowering
(courtesy A P Bradbury)
• Regular overtopping has occurred under even relatively benign conditions as a
result of increasing water depths at the toe (Figure 3.6). Water percolation
seawards of the seawall crest has caused a build-up of pressure to landwards
of the seawall.

Figure 3.6 Frequent overtopping causes build-up of pressure on landward side


of seawall (courtesy A P Bradbury)
• Increasing water pressures arising from overtopping have resulted in seaward
displacement of sections of seawall, which have slid seawards. Although the
wall does contain some drainage holes, this does not provide sufficient capacity
to relieve overtopping loads.

Toe structures management manual


42
• Repairs to the structure have been undertaken, which include construction of a
rock toe to resist further lowering (Figure 3.7). Note the kink in the alignment of
the wall which has arisen due to the structure failure and subsequent seaward
displacement of upper wall sections.

Figure 3.7 Repaired wall showing failed realignment. Rock armour has been
added at the toe (courtesy A P Bradbury)
Although structural designers usually incorporate a ‘factor of safety’ into their designs in
order to reduce the likelihood of such failures, these should not be relied upon when
considering reliability and stability years later in the life of the asset. This is because
rates of deterioration may have changed over time. For example, varying degrees of
environmental exposure may result in abrasion or damage to joints. The rate of change
is often a function of the degree of exposure to wave and sediment attack, which in turn
is a function of beach lowering. Occasionally changes in landward loading conditions
(such as superimposed loading from new structures or vehicles, or pore water
pressure) may impact on the structure.
Furthermore, factors of safety used in former times (for example, in the 19th century)
would not necessarily have achieved current day standards. Factors of safety are
discussed further in Section 5.4. Perhaps more significantly, the role of the structure
may have changed since construction with resultant changes in loading conditions. For
example, many structures originally constructed as Victorian promenades were simple
structures with no significant toe formation; these structures were often constructed at
the top of a wide beach and were not expected to withstand wave attack (Figure 3.8).
As beaches have fallen, many of these structures have subsequently assumed the role
of a seawall without further structural modification. These structures are often
increasingly vulnerable to undermining as a result of falling beaches.

Toe structures management manual


43
Figure 3.8 Promenade wall under construction (courtesy Poole Borough
Council)
Specialist engineers should be consulted if critical levels are not already known, or the
loadings on the structure are thought to have changed significantly or are likely to
change due to other factors such as proposed new development which may vary the
loading conditions.

3.4 Planning a management programme for


structure toes
This section introduces the problems associated with managing the toe of coastal
defence structures successfully. An outline approach is provided to assist asset
managers in this task.

3.4.1 Suggested approach for the management of toe scour


For each location being considered, the engineer needs to determine a trigger or ‘alert’
level for intervention (see Section 3.5). This will be based on key parameters related to
structural performance, beach safety and so on. Once this level has been set, a simple
assessment for any section of a structure can be determined (initially) on a seasonal
(for example, summer and winter) six-monthly basis, possibly supplemented following
storm events. The suggested approach is outlined using the eight steps shown in
Figure 3.8.

Toe structures management manual


44
Prescribe critical ‘trigger’ level for beach level at the toe of the structure
(see Figure 3.9 for definition).
The actual line is defined on a case-by-case basis – see Chapter 5 for
design limitations and recommendations).

Determine the maximum water depth at the structure for the next two
seasons based on predicted tide levels with allowance for surge.

Estimate the extreme wave conditions (Hs and Tm) for the next two
seasons.

Determine whether the beach in front of the structure is sand or shingle.

Is the beach slope known?

Estimate the lowest beach level at the structure for the next two
seasons based on a linear trend and the variance of historic bed levels
(unless a more sophisticated approach is warranted) (see Section 3.6.3).

Predict the combined beach level and scour level (see Section 3.6.4).

Carry out condition grade assessment and monitoring as necessary to


confirm the expected position of the beach level (Section 3.8).

Figure 3.9 Information required to assess toe scour


Once this information has been obtained, the decision process shown in Figure 3.10 is
implemented.

Toe structures management manual


45
If the beach is sand, use If the beach is shingle,
the equations in Appendix use a look-up table based
B1.2 to predict scour on Appendix Figure B.3 to
depth. predict scour depth.

Determine the combined beach level and scour depth for the next two
seasons.

Does this cause the beach level to drop below the critical level?

If it does not, then If it does, then plan


reformulate on an appropriate monitoring of
ongoing basis the structure condition and
prediction for the next beach levels at more
two seasons, updating frequent intervals and
the input parameters as implement
appropriate based on site mitigation/intervention
observations. plans.

Figure 3.10 Assessment of toe scour decision process


If a more detailed assessment is required for a particular asset, the predictions can be
made more frequently, given the relevant input data. With some further definition, this
approach can be implemented in a probabilistic assessment (HR Wallingford 2006e).
Certain data and prediction methods are required to conduct this analysis. The
following sections describe the data requirements and present the available methods
for predicting scour.

3.5 Defining ‘critical’ and ‘trigger/alert beach levels’


Establishment of a critical beach level for a structure requires an understanding of the
design limits associated with the structure, that is, knowledge about the physical
conditions it has been built to operate within. If sufficient information is known about an
asset and the forces on it, it may be possible to determine a beach trigger level at
which the probability of failure becomes unacceptably high. In some cases, it may be
possible to have a set of trigger/alert levels, each with its own probability of failure. In
many instances there is currently insufficient information available to determine such
triggers with any degree of confidence. This is the case when as-built drawings are not
available, in which case these need to be regenerated (see Section 3.6.2).

Toe structures management manual


46
A ‘trigger/alert level’ for the beach toe level can be inferred for which the risk of
excessive overtopping, or structural collapse due to loss of toe support is unacceptably
high. If a beach falls below this level, then intervention will be required to reduce the
likelihood of failure given the probable variation in beach level. If the trigger level can
be related to some clearly defined feature (perhaps the top of the first groyne pile or a
ledge on the seawall) or be clearly marked, it makes rapid assessment of the state of
the beach far easier.
The examination of statistics on mean levels at the toe of a structure over a number of
years, together with prediction methods and structure geometry information (level of
the structure toe), provides sufficient information for the asset manager to determine
the physical limits and timescale within which intervention options should be
considered (that is, beach re-nourishment, remedial defence works, decommissioning
and so on). This concept of the critical level is illustrated in Figure 3.11. The trigger for
a more detailed asset inspection could therefore be the exceedance thresholds gained
from results of the extrapolation of beach level data.

Figure 3.11 Relationship between long-term statistics on beach levels at the


structure, uncertainty, and ‘trigger’ and critical levels
Such trigger levels may be based upon visual observations as well as beach profile
measurements.
This approach is also valuable where there is insufficient information available to make
a confident assessment of structural failure conditions. For example, exposure of the
tops of toe piles may serve as a trigger to inspect more regularly when little information
is available about the toe construction. Additionally, structural defects such as damage
to pile tops or loss of jointing materials may also be used as indicators of a
deteriorating structure. If a structure has voids caused by loss of fill due to scour, it will
have an increased risk of failure and further investigation should follow – possibly using
non-destructive testing.
Once the critical level for the beach at a structure has been established, it is also
helpful to define an ‘alert’ level. An alert level is a point at which more detailed or

Toe structures management manual


47
frequent monitoring would be instigated, or intervention options considered and
implemented (for example, beach replenishment), before the critical level is reached.
Having defined these trigger levels, beach condition can then be assessed or graded
according to the volume or height of beach above the trigger level. The combination of
beach level monitoring, time series data, scour depth prediction and the calculation of
failure limit state can provide a representation of the relative reliability of the structure.
The following sections discuss the data required to undertake this analysis, how it may
be collected and how it can be used. Examples are also provided of linear beach level
analyses and trends, along with guidance on the methods of extrapolating data for the
prediction of future beach levels.

3.6 Beach monitoring


The state of a beach at a particular time affects its response to loading and its ability to
provide protection in the short term. Where a beach forms an integral part of the
defence toe, its state will almost certainly require monitoring and maintenance to
ensure that the required level of protection is sustained. Beaches and shore platforms
dissipate wave energy by friction and wave breaking, thereby reducing the wave
energy that reaches a coastal defence. Should beach levels fall, however, higher
waves will reach a structure, which may increase the risk of failure.
Regular monitoring of beach levels allows a picture to be built up of how a beach
responds to the incident wave climate at a scale of storms, seasons and years, which
should inform the management policy. The remainder of this section describes how
beaches and shorelines are commonly measured and monitored. This topic is dealt
with in more detail by Sutherland et al, (2006a) and Section 5 of the Beach
Management Manual (CIRIA 2010a).

3.6.1 Beach level and topography


Baseline beach surveys should enable the beach plan-shape to be described
adequately and be in context with existing control structures (for example, groynes).
This will assist with the understanding of sediment transport patterns and the effect of
sediment control structures on sediment transport rates (and therefore beach volume).
Such surveys will also assist with the determination of the best places for subsequent
monitoring locations, particularly if the orientation of the beach crest(s) varies
significantly along the frontage. Contour and three-dimensional (3D) measurements
can be achieved by a variety of methods and tools including terrain modelling using
results from a real-time kinematic global positioning system (RTK GPS) (Figure 3.12),
light detection and ranging (LiDAR), photogrammetry and bathymetric surveys.

Toe structures management manual


48
Figure 3.12 Beach survey using global positioning equipment (courtesy
Channel Coastal Observatory)
The procedures used to process the data are similar irrespective of the method used to
collect them. Contours can be simply extracted from the digital terrain models,
generated from a grid of elevations.
RTK GPS can now provide a vertical accuracy of ±30 mm and plan accuracy to
±15 mm. For rapid coverage, GPS can be mounted on suitable ‘all terrain vehicles’
(Figure 3.13). For further information the reader is referred to Section 5 of the Beach
Management Manual (CIRIA 2010a).

Toe structures management manual


49
Figure 3.13 All terrain vehicle fitted with GPS equipment for coastal
topographic surveying (courtesy of Worthing Borough Council)
Digital terrain models of beaches are produced routinely for numerous beach
monitoring programmes. These surveys can be used as an excellent indicator that the
structure toe is becoming under increasing loading and that more detailed attention
should be given to assess increasing risks. The primary value is to identify erosion
hotspots.
The sequence of colour-coded beach elevation plots shown in Figure 3.14
demonstrates clearly that the beach is becoming smaller over time. Note that the rate
of change has increased with time and that the survey frequency has been increased
to capture the more rapid changes, as the structure toe becomes more vulnerable.
Surveys identify considerable beach narrowing at the updrift end. Much of the structure
remains well protected and vulnerable areas of the structure toe can be isolated to a
frontage of about 50 m, where detailed monitoring efforts can be focused.
It is recommended that detailed profile analysis should be conducted in conjunction
with structure stability assessments, where the risks appear to be increasing in erosion
hotspots. Changes in beach width can be also indicative of changes in beach
composition and monitoring of foreshore type is important in this respect.

Toe structures management manual


50
Figure 3.14 Digital terrain models showing beach evolution adjacent to seawall
with changing structure alignment (courtesy Channel Coastal Observatory)
The net sediment drift direction at this site is from west to east, and this is reflected by
a declining beach volume at the structure toe at the updrift end of the beach. The
pattern shown demonstrates clearly the influence of an updrift change in structure
alignment, which has an adverse effect on beach supply to the western (updrift) end of
the beach. This change in alignment is a typical cause of falling beach elevations at a
structure toe. Changes in structure alignment are highlighted generally as potential

Toe structures management manual


51
hotspots for beach loss at the structure toe and these should always be examined
carefully for signs of change.
Migration of the mean high water spring (MHWS) beach contour can serve as a useful
indicator that a structure toe is becoming under an increasing frequency of loading.
Figure 3.15 illustrates the migration of the MHWS contour over a period of less than
two years, yet shows a significant trend of migration towards the wall.

Figure 3.15 Migration of MHWS contour towards seawall (courtesy Channel


Coastal Observatory)
The location of the MHWS contour suggests that 50 m of the structure frontage may be
regularly under direct wave attack, in accordance with the tidal cycle. The survey
frequency increased from April 2008, triggered by the rapid movement of the beach
contours towards the wall (Figure 3.16). The pattern indicates that standing water will
have occurred at the seawall at several stages during this period, over the upper part of
the tidal cycle. Any wave activity during these periods will certainly have impacted on
the wall. Parts of the wall will be offered some protection by the beach over most of the
length monitored, but it is clear that the frequency of attack at the western end has

Toe structures management manual


52
increased over time. Such observations should act as a catalyst for more vigilant
observations of structure toe performance. This is generally a useful indicator for
assessment of overtopping risks that may be affected by toe scour.

Figure 3.16 Migration of MHLS contour towards seawall (courtesy Channel


Coastal Observatory)
Tidal range is significant when considering impacts of the beach on toe stability.
Migration of the MHWS contour is a universally applicable indicator, irrespective of tidal
range. Where the beach is narrow, typically when the tidal range is small (<3 m), the
MLWS contour may also be a useful indicator. When this contour approaches the wall,
the frontage will become subject to very regular wave loading and minimal support is
provided to the lower part of the surface emergent element of the structure; in many
cases this may mean that the structure foundations are fully exposed over the entire
tidal cycle. Where available, the MLWS contour is very valuable as an indicator of the
maximum beach width fronting the structure. Such information is not widely available,
simply because of the practical difficulties of conducting surveys over the lowest part of

Toe structures management manual


53
the tidal cycle and is generally less useful in areas of large tidal range where the low
water beach is wide.

3.6.2 Beach profiles


Beach profiles are usually surveyed perpendicular to either the shoreline or a
predetermined baseline. They are used to quantify beach response to storm events,
sediment recovery rates, long-term volume changes and the potential envelope of
cross-shore elevations. When combined with nearshore bathymetric surveys,
morphological changes across the full zone of wave influence can be assessed. Rapid
appraisal of profile data (often in combination with cross-shore empirical parametric
models or trigger levels) is vital to determine whether management works are needed
and for establishing some of the design constraints that will have to be worked with.
The location of profile lines should be considered carefully. The ‘density’ of the lines
should be sufficient to provide an adequate representative coverage of the beach for
the purpose prescribed. Allowance should be made for nearshore features such as
bars, banks and troughs, which induce localised variations in beach response. For the
purposes of structure management, the profiles will generally need to be closely
spaced (30–50 m). Where the effect of groyne systems is to be considered, it may be
necessary to survey lines on either side of the groynes and at the centre of the bays.
However, it may only be necessary to ‘sample’ survey within selected groyne
compartments (due to the repeatability of the bay plan shape in a groyne field). Profiles
taken adjacent to structures are of great value if information is needed on scour, but
the exact line of the survey must be located and repeated exactly to ensure
comparability.
The extent of a particular survey line is also important. The boundaries between the
mobile beach toe and bed rock, and the crest and seawall/cliff should be identified
where appropriate. The landward end of the survey line should extend to the structure.
The seaward end should extend as far seawards as is practical; this may extend at
least to the level of mean low water spring (MLWS) at sites with a small tidal range
(<3 m).Supplementary shore parallel profiles are of great value – particularly at the toe
of defences and the crest of beaches – to provide both a fuller coverage of data and as
a cross check.

Key features to be measured on profiles


A number of key features should be surveyed on or relative to each profile:
Tidal elevation variables (relative to structure details)
These data are used as an indicator of the potential frequency of loading of the seawall
under wave conditions. Structures with toe foundations that lie within the range of the
intertidal zone will be subject to regular wave attack at the structure toe.
Seawall cross-section
This should describe the seawall geometry:
• crest elevation of structure;
• level of structure foundation base;
• level of bedrock base at the foundation position;
• depth of penetration of any toe piling or other toe detail;

Toe structures management manual


54
• structure construction materials:
- concrete grades;
- reinforcing detail;
- pile gauge and type;
- fill type within seawall.
Details of the structure geometry, construction method and materials are crucial for the
determination of structural stability assessment of the structure; this is a basic
requirement for the determination of alarm and intervention levels. Despite this, such
information is not currently available at many sites. This is a major problem for long-
term structure management.
In some cases, adequate as built scheme drawings will provide all the relevant detail,
but in many instances these are not available. Structures may be of any age, perhaps
over 100 years old when as built drawings did not form a standard part of the design
and constructions process. In many other instances, design or as-built drawings have
been lost or destroyed.
Efforts should be made to store historical drawings, since these may be of significant
value to structure management several decades after construction. Whereas built
drawings do not exist, their regeneration is a most valuable investment for long-term
structure management. The basic geometry can be measured or surveyed to provide
the surface emergent structure profile. Details of the structure toe can be more
problematic to determine as these are usually buried in either bedrock or beach
material. Careful excavation of trial holes at the structure toe or other soil investigations
can provide some of the required detail; these should ideally identify the elevation of
the structure foundation base and also the depth to bedrock (if this is at a lower
elevation). It may not be practical to measure such data at many sites where the
bedrock is beneath thick sediments. Toe elevation data should be combined with tidal
elevation data to determine the risks arising from frequency and depth of inundation of
the toe area in combination with wave loading.
An alarming number of promenade type seawalls dating back over the last 100 years
have been constructed on a perched foundation, which has a base well above low
water and which does not close onto bedrock. The stability of such structures is entirely
reliant upon the presence of the beach fronting the wall. Numerous failures have been
observed over many years, usually during or following storm events when the structure
toe is exposed. In the event that beach material is eroded from the structure toe, the
foundation can become rapidly undermined (within a few hours). The zone between
low water and the foundation base is particularly vulnerable, as water can be pumped
by wave activity to scour material from beneath the foundation. Rapid loss of internal fill
material can result, followed by collapse of the concrete deck membrane.
It is unsatisfactory to attempt to manage a structure without scheme drawings that
detail the structure toe construction. A conservative assessment must be made based
on observations and intervention measures planned in accordance with this. The safest
approach is to assume no foundations are present below known levels and to
determine trigger levels on the basis of these.
Where the structure toe is of a piled construction, reconstruction of the drawings is
more difficult since it is neither practical nor desirable to excavate into bedrock
material. A number of specialist non-destructive testing methods for example sonic
testing can be used for determination of the pile length of both concrete and steel piles.
This can be used to establish the depth of penetration of the toe piles. This information
is needed to assess the risk of overturning, undermining or structure sliding.

Toe structures management manual


55
Many structures have been constructed with a concrete membrane that overlies loose
local material – typically local beach material or some other fill. Coring through the
concrete membrane at various locations along the structure will enable identification of
the type of fill material and whether voids are present beneath the surface. Cores holes
should be grouted following testing to ensure that the very method of investigation does
not cause a problem itself.
Bed rock profile from the seawall to seawards
The depth of the bedrock, its profile and its position relative to the structure foundations
has a significant effect on the performance of the structure toe. Bedrock may not be
accessible at many sites due to the thickness of surface sediments.
The bedrock profile is used, where available, to identify the solid surface adjacent to
the seawall that would be exposed in the event of all beach material being removed. In
some instances, changes in bed level beneath this surface arising from erosion can
also be significant; this is typically the case at sites where soft or even medium clays
are exposed. Such information is often difficult to collect and may not be available at
many sites due to the thickness of sediments
Where construction or geotechnical records are not available, small trial holes
excavated adjacent to the seawall should be used where possible to identify the depth
to bedrock at the structure bedrock intersection. The beach may dry onto bedrock at
some sites and this intersection of beach and bedrock can be used to provide the basis
of a linear interpolation between the structure and the toe to provide an indication of the
bedrock surface elevation – though this is not a wholly satisfactory method. Exposure
of bedrock may occur at other locations from time to time and this can be used to build
up a more complete picture of the bedrock profile over time.
Procedures linked with standard specifications for some of the regional coastal
monitoring programmes require surveyors to feature code survey data; this will provide
indications of bedrock elevation where this is seen on each survey. Each point has 3D
coordinates and a surface type attribute. Feature coded survey data can be extracted
from a number of surveys and combined over a period of several surveys to provide
the basis for development of a primitive surface model of the bedrock surface. These
data can be supplemented by surveying elevations of spot heights of bedrock,
determined during excavations for the repair and construction of structures such as
groynes. For example, this has been used effectively at Bournemouth and Milford-on-
Sea to provide a more detailed bedrock surface. Where these data have been
captured, it is possible to measure absolute beach volume changes by using this as a
master profile rather than changes relative to an arbitrary surface (Figure 3.17). Such
an approach may not, however, be feasible at many sites.
Profile surveys
Regular beach profile surveys can be used to develop an indication of changes to
protection of a structure toe. All of the key descriptors identified in Section 3.5.2 are
shown on the simple plot presented in Figure 3.17. The long-term profile envelope
indicates the lowest and highest points that the beach has reached over the course of
the monitoring.
The example shown represents changes over a 20-year monitoring period. These data
can be used in combination with a trend analysis of beach cross-sectional area above
a defined surface (master profile) (Figure 3.18). Where available, the master profile
should be defined by the bedrock profile, thereby enabling absolute changes in beach
volume to be determined; this profile cannot be defined at many sites and only relative
changes can be measured. An at-a-glance assessment of the relative state of the
beach to historical conditions is provided by comparing the most recent survey with the
profile envelope and the mean profile.

Toe structures management manual


56
Figure 3.17 Beach profile and structure descriptors used in structure toe
analysis (courtesy AP Bradbury)

3.6.3 Predicting trigger levels from monitoring data


The variations in beach levels near coastal structures at timescales of the order of tide
to a year are the accumulation of the residual changes that occur during each tide. It is
common to find beach levels lower in winter than in summer due to the increased
occurrence and severity of storms during winter. It also follows that beach levels may
show a greater variation about their seasonal mean during winter.
A variety of types of profile analysis can be conducted as the time series grows.
Analysis of lengthy long-term trends provides the generally most useful indicator of the
beach state. The record length of surveys is significant and several years of data
(ideally at least 10) are required to give confidence to assessments.
Figure 3.18 demonstrates a lengthy time series during which the general trend is of
erosion, but with periods of recovery. A series of storm events of varying intensity
punctuate the long-term profile trend and these are associated with defined
combinations of wave and tidal conditions. Wave data can ideally be derived from a
nearby wave buoy in shallow water (see Figure 3.20). Alternatively, hindcast modelled
offshore data can be transformed to the nearshore zone. Several of these storm
conditions are highlighted in Figure 3.20. Interestingly, the most severe hydrodynamic
conditions are not always associated with the greatest damage or beach loss. A
number of the notable dips in beach cross-section are associated with swell wave or
bimodal wave conditions that would not generally be considered to be damaging
events.

Toe structures management manual


57
Figure 3.18 Beach profile cross-sectional trend analysis
A trend line is fitted through the data set, indicating the mean profile. The profile
envelope is represented by lines parallel with the mean profile. The maximum
departure from the line is associated with the design storm, which denotes the
maximum departure from the envelope line. Application of a storm response line
provides an indication of the anticipated maximum reduction in cross-section on a
defined storm event. This is based on conventional topographic surveys and does not
consider additional lowering that may take place during the storm but which is not
detectable by conventional survey techniques. Additional allowance must be made for
such changes (see Section 3.6.4). Using the critical cross-sections determined for each
site (Section 3.5), a critical acceptable volume trigger can be added to the graph.
Linear projection of the trend, envelope and storm damage lines can be used to
estimate the potential timing for intervention.
Once a trend of position against time has been established, the trend can be
extrapolated beyond the date of the last data point and into the future. The results of an
extrapolation should be interpreted in light of the underlying principles of
geomorphology and sediment transport (that is, tempered with what is actually
realistic). The extrapolation of trends and confidence limits into predictions assumes
that the future hydrodynamic climate will be statistically similar to the climate during the
period the measurements are made.
The shape of individual profiles can provide a useful indicator of beach performance
and its interaction with the structure. A shingle beach will normally form a distinct berm
with a well-defined run-up crest on a healthy beach, where the dynamic equilibrium
profile is able to develop fully (Figure 3.19). Where there is inadequate cross-section
available for the dynamic equilibrium profile to form, the profile may be characterised
by a planed off crest or even a dip at the seawall intersection. The crest invariably
forms at a lower level. The implication here is that the beach and structure intersection
is under wave attack. The planed surface arises from wave reflections from the seawall
increasing the backwash. This observation is generally considered to provide a
preliminary indication that the structure is undergoing increased loading and might
merit more regular inspections.

Toe structures management manual


58
Figure 3.19 Beach profile evolution in the presence of seawall (courtesy
Channel Coastal Observatory)

Post storm surveys (modified from CIRIA 2010a)


This section is concerned with the analysis of beach levels close to the toe of a
structure at seasonal and shorter timescales. The prediction of beach levels on these
timescales is important as they provide the initial conditions for the calculation of toe
scour during a tide or storm.
The beach response to storm conditions may have both long- and short-term
implications and regular post storm surveys should be conducted where possible. In
the short term the surveys may act as a trigger for intervention, when the response is
measured relative to defined critical conditions (see Section 3.5). They may also act to
identify an alarm state, where preparation for intervention may be considered. Under
some circumstances, beach lowering may reach a point where the profile is unlikely to
become restored to a healthy condition naturally, as suggested by Powell and Lowe
(1994). This may result in undermining and destabilisation of structures such as
groynes or seawalls. The failure of the beach to recover following storm events is often
due to the loss of material in longshore transport, which is accelerated as material is
drawn into the subtidal zone. Alternatively, reflections from a vertical structure may
make it difficult for a beach to reform. Where there is a limited longshore feed beach
losses may occur from the updrift zone. It may be desirable to conduct post storm
surveys in any of these circumstances.
It is beneficial to collect hydrodynamic data in conjunction with post storm survey data
and to present these as an event time profile of hydrodynamic conditions (Figure 3.20).
This should normally describe the whole of the event cycle including the build-up, storm
peak and decay, and should where possible include integrated parameters for Hs
(significant wave height nearshore), Tz (zero-crossing period), Tp (peak wave period)
direction and spectral data. These should be supported by tidal profiles and wind data
when possible.

Toe structures management manual


59
Figure 3.20 Typical hydrodynamic conditions used in the analysis of a storm
event (courtesy Channel Coastal Observatory)
In some instances, profile surveys are conducted only following the most damaging of
storm events, primarily to quantify damage to defence systems. However, it is also
beneficial to conduct surveys following more regularly occurring storm events, perhaps
with return periods of about 1:1 year. Although damage might not be expected
following such conditions at most sites, it is advantageous to develop an empirical
framework of measured responses in order to provide a framework of data on either
side of the damage thresholds. This approach may enable the threshold conditions that
cause damage and critical conditions to be defined more accurately, and also develops
confidence in management procedures. This approach has been adopted within the
southern regional coastal monitoring programmes, where post storm surveys are
triggered by exceedance of defined threshold conditions at the network of wave buoys.
The storm threshold used is typically a 1:1 year return period event, at any water level.

Toe structures management manual


60
Events are triggered initially though an automated system of text and email alerts
provided by the real-time wave buoy network. Further consideration is given to the
event including a brief consultation with the relevant operating authority prior to
mobilising a survey team.
Post storm surveys have been used also to calibrate, validate and extend empirical
predictive frameworks used in beach management. For example, validation has been
conducted of empirical models for gravel beaches using the framework proposed by
Powell (1990).
It is desirable to conduct the post storm survey as soon after the event as practicable.
This often means that a few days may elapse before conditions at the site are
sufficiently safe for surveys to be conducted. Even then it is rarely possible to achieve
the desired seaward limit of profiles because of conditions. On this basis the upper
beach typically above mean low water neaps (MLWN), or perhaps higher, can be
surveyed and a limited range of characteristics can be measured. It is normal to relax
survey specifications in order to conduct surveys quickly after the storm event. Other
considerations for planning of post storm surveys are sample locations and the
performance of historical hotspot locations.

Shore parallel profiles of the beach structure intersection


Shore parallel profiles are extremely valuable tools for monitoring beach structure
interaction. Profiles are conducted by measuring elevations of the beach and structure
intersection (see Figure 3.12). The example presented shows the spatial and temporal
variability of the beach structure intersection level (Figure 3.21). Each elevation is
measured at predefined locations along the seawall; in this instance these are defined
using kinematic GPS and a data logger with a stakeout control file to ensure that each
point is precisely relocated on each survey.
The example presented covers a frontage distance of approximately 200 m. The time
series covers a period of about seven months. It demonstrates not only the vertical
extent of changes but also indicates how quickly these changes may occur. The
temporal intensity is extremely high at this location; this reflects the fact that the toe of
this structure was considered to be very vulnerable. The key variables plotted are the
foundation base elevation, bedrock elevation and tidal levels. The structure toe
becomes increasingly vulnerable as the beach elevation falls closer to the foundation
base.

Toe structures management manual


61
11/09/2008

Beach Elevations at Sea Wall Interface 12/09/2008

13/09/2008

4.500 14/09/2008

15/09/2008

4.000 16/09/2008

17/09/2008

3.500 20/09/2008

21/09/2008

3.000 22/09/2008

23/09/2008

2.500 24/09/2008
Elevation (mOD)

25/09/2008

2.000 26/09/2008

30/09/2008

1.500 06/10/2008

13/10/2008
1.000
MHWS 21/10/2008

27/10/2008
0.500 Foundation Base
03/11/2008

12/11/2008
0.000
18/11/2008

25/11/2008
-0.500
08/12/2008
Clay Level
18/12/2008
-1.000
15/01/2009
1
3
5
7

B 9
11
13
15
17
19
21
23
25
27
29
31
33
35
37
39
41
43
45
47
49
51
53
55
57
59
12/02/2009
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
B
B
B
B
B

B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
B
13/03/2009
Point Name
15/04/2009

Figure 3.21 Temporal and spatial distribution of beach and structure toe
intersection elevations (courtesy Channel Coastal Observatory)
An alternative approach that would achieve similar results can be achieved requiring
minimal equipment. Periodic measurements of the ‘dip’ or vertical distance from the top
of the wall to the beach using a tape can be made; these can be made repeatable by
marking predefined locations on the seawall.
An alternative representation of the evolution of a single point elevation at the toe of a
structure in Mablethorpe, Lincolnshire, is shown in Figure 3.22.

4.5

4
Beach level (mODN)

3.5

2.5

2
Level Trend
1.5
19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86

88

90

92

Year since 1900

Figure 3.22 Temporal distribution of beach and structure toe intersection


elevations (courtesy HR Wallingford)

Toe structures management manual


62
3.6.4 Monitoring of scour depth
As beach levels vary throughout a tidal cycle, together with varying wave conditions,
the development of toe scour is a dynamic process, highly dependent on the water
level at the wall as well as the incident wave conditions. In areas of varying tidal range
and wave climate, the development of a scour hole will be a variable process with
periods of erosion followed by infilling and perhaps even general accretion of bed
levels (Powell and Lowe 1994). The scour hole itself may therefore be a short-lived
feature with no obvious evidence of its existence after a storm has declined and infilling
has taken place as the tide recedes.
Significant damage can arise over a short period of time during which excursions of
beach level beneath critical levels may occur. Depending on the construction type,
falling beach levels over periods of just a few hours may have a significant effect on the
risk of structure failure or major damage. Those structures that are most at risk under
these circumstances are those constructed with an elevated foundation level (usually
above MLWS), where there is significant separation between the foundation base and
the underlying bedrock, and where a concrete or other hard outer structure shell
encases granular fill material. This occurs at many sites within the UK.
Hence, there is a need to be able to predict the maximum vertical excursion of the
scour hole during storms, as well as the more widespread and longer term processes
that cause the lowering of beach/shore platforms. This is important both in the design
stage of a coastal structure and in its subsequent monitoring if the risk to the future
integrity of the wall is to be fully understood and timely remedial action undertaken.
As scour is frequently short-lived, the biannual beach profile monitoring typically carried
out around the English coastline is unlikely to coincide with a major scour event. While
there are initiatives to carry out post storm event beach profiling (Section 3.6.2), this
will not capture the response at the peak of the storm. Evidence supplied by data from
bespoke scour monitors suggests that a significant fraction of a scour hole can fill in
within a few hours of the peak of a storm. Therefore even regular beach profiling with a
spacing of a few weeks, supported by profiles collected within a day or two of each
large storm, may not be enough to capture the transient phenomenon of toe scour in
the field.
The deployment of scour monitoring systems that remain on-site, just in front of a toe
structure operating at all water levels, for periods of weeks at a time may currently be
the only realistic way of assessing the site specific variability of a beach surface with
time. Figure 3.23 shows scour monitors deployed on a beach. Scour measurements
can be conducted using, among other things, the following devices (Sutherland et al.
2007) installed at locations where the beach at the toe of the structure is exposed at
low tide:
• HR Wallingford’s ‘Tell Tail’ scour monitoring system (Figure 3.23) is based
on a linear array of omni-directional motion sensors buried in the seabed
adjacent to the structure. Identifying the elevation of the lowest active
sensor places an upper limit on beach level. The system records the onset
of scour, the depth of scour reached and the in-filling of scour holes
following storm events. Tell-Tail scour monitors have been deployed in front
of seawalls at Teignmouth (Whitehouse et al. 2000), at Southbourne
(Pearce et al. 2006; HR Wallingford 2006a) and at Blackpool (Sutherland et
al, 2006a).
• Linear array of electrical conductivity meters (Ridd 1992; Cassen et al.
2005), which rely on the fact that seawater has a high electrical
conductivity, while dry sediment has a low conductivity and saturated

Toe structures management manual


63
sediment has an intermediate conductivity. Cassen et al. (2005) measured
erosion in the inter-tidal zone of a beach at Bicarrosse (France).
• Photo-electric erosion pin developed by Lawler (1991) which detects
daylight at an array of optical sensors and has been used in the swash
zone by Robinson et al. (2005).
• Sedimeter developed by Erlingsson (1991) which used an array of infrared
transmitters and backscatter detectors.
Installation and operation of scour monitors is a specialist activity and seems unlikely to
find its way into routine monitoring programmes. As scour monitors are likely to be
deployed for a period of weeks (or longer), a monitoring strategy could be implemented
that looks for the bed lowering to the point at which short-term fluctuation from the
slowly varying mean level could destabilise the defence asset. For this the likely scour
depth for a given storm would have to be estimated, plus the depth of scouring that
would create a risk of failure. The monitoring could then take place at least twice per
year and a more detailed study or remedial action undertaken should the data show
beach levels dropping below pre-determined values.

Figure 3.23 Scour monitors in operation on a shingle beach to obtain data on


within tide changes in bed levels (courtesy HR Wallingford)
The somewhat limited number of historical deployments of scour monitors means that
empirical statistics of performance are limited and estimation of scour is difficult. None
of the deployments have taken place during particularly severe conditions and the
application of results to date should be used with some caution. A reasonably typical
plot is shown (Figure 3.24), which indicates intertidal fluctuations of about 0.6 m, under
fairly benign wave conditions (Hs = 1.5 m). The type of pattern observed, on a sand
beach, seems reasonably typical of other sites, with maximum scour occurring over the
high water period and with a cyclic return over the low water period.

Toe structures management manual


64
3.5

Recovery
Lowering

Recovery
Lowering
3
Tide, beach levels (mCD) Hs (m)

2.5

1.5

0.5
Tide Height Scour monitor Hs
0
24/05/2005 00:00 24/05/2005 12:00 25/05/2005 00:00 25/05/2005 12:00
Date and Time GMT

Figure 3.24 Fluctuations in sand bed levels over a tidal cycle (from Pearce et al.
2006)
Complementary laboratory work (Pearce et al. 2006) suggests similar relationships
between scour depth, wave height and water depth to those shown in Figure 3.24.
While the science is not yet well advanced, the preliminary empirical framework
suggested by Pearce et al. (2006) appears to be supported by field scour
measurements.
It is suggested that, for the purposes of monitoring, a preliminary site-specific
assessment of scour potential is made using the following dimensionless formula. It is
suggested that maximum scour potential equates to:
St / Hs = 0.8
where:
dt/Lm = 0.015
St = maximum scour anticipated
Hs = significant wave height nearshore
dt = depth of water at toe
Lm = wave length.
These should be calculated for extreme conditions anticipated at the site. Although
based upon dimensionless laboratory tests, the significant wave height used in the
development of the formula was measured in deeper water than the toe. This might
typically be in the range 8–12 m depth.
Application of this framework to a reasonably typical south coast site under extreme
conditions of Hs = 4 m and Tm = 7 s suggests a maximum scour of 3.2 m should occur
with dt = 1.2 m. This suggests quite alarming results which should be considered in
context with other controlling variables at the site. In particular, the scour depth may be
restricted by the actual depth of mobile sediment above bedrock. The actual depth of
sediment may often be the governing limiting factor.

Toe structures management manual


65
It is recommended that both the theoretical formulation and the other practical
governing factors should be considered together to reach an assessment.
Actual scour monitoring has identified a maximum scour depth of about 0.9 m and this
should be considered a minimum expectation at most exposed sites, with the
theoretical scour potential also considered in context with the site specific geology.

Video system monitoring (based on CIRIA 2010a)


Video systems such as Argus may be used to identify changes to bed elevations in the
intertidal zone, while the beach is submerged. Shore-based video systems can provide
automated data collection over wide spatial and temporal scales, and during
unfavourable weather conditions such as storms. Furthermore, the measurements are
non-intrusive and therefore do not affect measurement results.
The Argus system (Holman et al. 1993) uses unmanned video stations, located at
conveniently elevated remote locations to monitor the nearshore hydrodynamics and
morphological processes, with high time and spatial resolution. The system offers cost-
effective monitoring of the beach evolution, erosion and sediment movement for
beaches up to 3 km in length. Measurements of morphology across the surf zone and
an estimation of surface currents and wave characteristics can also be derived. The
spatial resolution derived from the Argus cameras is similar to that of numerical
simulations, allowing effective comparison and evaluation.
More than 50 Argus systems have been installed on a worldwide basis including a
number in the UK. These cover a range of sites and systems including:
• barrier beaches (Slapton);
• surf reef performance and environmental assessment (Boscombe);
• nearshore ebb deltas (Teignmouth);
• beach structure interaction (Cleveleys);
• evolution of beaches in the presence of nearshore breakwaters (Sea
Palling).
Argus camera systems are particularly effective at capturing change within the intertidal
and shallow subtidal zone. This zone is the most challenging zone for data collection
using alternative techniques and is also the most significant zone for beach change,
when considering the toe of structures.
The images obtained by the Argus video system provide spatial and temporal
resolution that enables the study of sediment transport patterns in groyne fields and the
effect of storms. Profiles derived from an Argus system are compared with those using
traditional methods (Figure 3.25). The large variability and advantage of the extended
length of the Argus survey through the dynamic zone should be noted.

Toe structures management manual


66
Figure 3.25 Comparison of traditional profile with Argus profile envelope (from
Green and Illic 2009)
Argus is not able to detect changes on those sections of beach that are always
surface-emergent and its application is limited in this context when considering
changes to elevations of threshold conditions for storm damage in the subtidal zone.
As the system is based on photography it can only function during daylight conditions,
though this is typical of many monitoring techniques. Although the impression is given
in much published literature that these systems are automated, considerable calibration
and regular checking is required to ensure that the systems have not been affected by
external elements such as wind. A commitment to regular maintenance and
management is required.
Not all sites are suitable locations for installation of these systems. The primary
requirements are for suitable elevated camera mounting locations, an appropriate
power supply and access to a broadband network. High buildings may sometimes
provide suitable locations, but these need to provide a good view of an extended
section of beach. Towers are sometimes erected to mount cameras. Sometimes
valuable data collected during storms cannot be analysed due to the conditions
obscuring the images. Cameras suffer from salt spray on the enclosure lens and
require regular cleaning.
As access to the cameras may be difficult, the enclosures can be fitted with a
wash/wipe system to reduce the need to clean the lens manually. Enclosures made
from aluminium alloy suffer from corrosion, allowing water ingress into the enclosure.
Alternatively enclosures can be made from stainless steel and include clips for access
rather than screw fittings. The accuracy and resolution of the images depend on
number of cameras used, their height, angle and sampling frequency, and the
availability of field data. They also depend on environmental conditions such as
sunlight, water reflection, fog, and sea salt spray.
Successful applications of the system have been made in academic investigations.
Some of these have successfully transferred the outcomes to management
applications, but there is more scope for development of such automated analysis
techniques.

Toe structures management manual


67
3.7 Beach depletion and foreshore down-cutting
The performance of a beach depends largely on the volume of material present and the
limits to its plan and profile changes. Where there is a net loss of sediment, beach
recovery is an issue. Where there is clay beneath sand or shingle, then it is unlikely
that a beach will recover naturally once the clay layer is exposed. Figure 3.26 shows an
extreme example of beach depletion and foreshore down-cutting.
Erosion or changes to beaches are generally gradual (long-term) but significant erosion
and lowering can occur during ‘one-off’ storm events. In general, failure is a result of
depletion in the volume of the beach through increased longshore and/or cross-shore
transport of beach sediment, or a reduction in supply of sediment onto the frontage.
Changes in sediment transport are a result of changes to the wave conditions at a site
and can occur for various reasons.

Figure 3.26 An extreme example of beach depletion, foreshore erosion and


down-cutting – note the level of the base of the access steps in relation to the
level of the beach (courtesy HR Wallingford)
The underlying geomorphology influences the performance of a beach in the lower
foreshore and nearshore zones. The nature and properties of the seabed in these
zones can contribute to long-term changes to the beach profile. The seabed may be
formed from a variety of materials ranging from extremely durable rock to more easily
eroded limestones, chalk and clays. Where durable rock platforms exist, there will be
limited erosion and the impacts on the beach profile will be negligible. On softer rocks,
the presence of a thin layer of mobile sediment can gradually erode the platform
through abrasion. This lowering of the platform increases water depths and therefore
the size of waves that can reach the toe of the beach and that of coastal structures.
These conditions will lead to a loss of beach material and ultimately to failure unless
steps are taken to periodically replenish the beach.
In locations where the underlying substrate is clay, erosion can accelerate as the
condition of the beach deteriorates. Once the beach has effectively become a thin
veneer, the underlying clay is likely to be intermittently exposed during storms and
subject to erosion. The loss of clay beneath the beach profile results in a permanent
lowering of the beach and increased exposure to wave attack. This process is
commonly referred to as ‘clay down-cutting’ and can result in accelerated losses from
the beach prior to failure.
Various instruments and methods have been used in studies to assess down-cutting or
down-wearing of shore platforms (see Sutherland et al. 2007). One such monitoring
instrument is the traversing erosion beam (TEB) as shown in Figure 3.27.

Toe structures management manual


68
Figure 3.27 A traversing erosion beam instrument

3.8 Structure condition monitoring

3.8.1 Asset deterioration


Assessment of the defence should consider the condition and integrity of the structure
itself. The material of the structure may deteriorate over time to a point at which it can
no longer maintain its performance – even if external loadings do not change. A typical
example would be the oxidation (rusting) of sheet piles, resulting in loss of integral
strength due to loss of section thickness. As coastal structures are expected to perform
satisfactorily for decades, during which time conditions can vary extensively, the
assessment of condition and deterioration is an important issue. The performance of
different materials when used in sea defence is discussed in Section 2.3.
Once beach levels fall below the top level of a toe structure, it is exposed to the
elements of wave attack and abrasion by mobile sediments. In the case of concrete
and steel piles, this also means that chemical and biological processes may also
ensue.
Accelerated low water corrosion (ALWC) is a biological process of degradation that can
lead to loss of section (width) of unprotected steel in the tidal zone. Chemical reactions
in concrete caused by saline intrusion can contribute to crack expansion, spalling
(surface flaking) and oxidation of internal steel rebars and reinforcing mesh. Abrasion
of concrete by mobile sediments (Figures 3.28 and 3.29) can erode the structure and, if
left un-remediated, can cause a serious loss of its mass leading to a reduction in
function and performance and eventual failure (also see Section 4 on maintenance of
concrete structures).

Toe structures management manual


69
Figure 3.28 Abraded concrete toe and onset of undermining (courtesy of HR
Wallingford)

Figure 3.29 Exposed reinforcing mesh due to steel oxidation and resultant
cracking and flaking of concrete cover layer (courtesy HR Wallingford)
Box 3.1 presents an example from Norfolk of how estimates of asset deterioration can
be determined. Such figures can be used in strategic and forward policy and financial
planning. For example when capital works might be required for new structures, for
planning frequent or intermittent maintenance, or for demonstrating when a reduction in
the standard of service might ensue over time. Deterioration rates can be used at
scheme conception simply to help establish the potential whole life costs of an asset or
scheme option. Here ‘best’, ‘fastest’ and ‘slowest’ estimates are given to allow some
judgement-based assumptions to be included in the determination of the most realistic
rate of decline in condition.

Box 3.1 Site Description


This asset is located at Overstrand in north Norfolk.
The current defence consists of a 2.74 m high reinforced concrete wall with a 1.43 m
wide reinforced concrete apron and 4.3 m long piles as scour protection. The
average crest height of the wall is 4.50 mAOD. Behind the 5.00 m wide promenade
at the rear of the wall, the contorted glacial drift cliffs rise to a height of 23.6 mAOD
The defences were rebuilt in 1955. Observations are based on routine inspections.
The concrete wall has a condition grade of 2, tending to 3, and the steel piles have a
condition grade of 3. The beach is in very poor condition lowering at a mean rate of
70 mm per year. The beach has been assigned a condition grade of 5. The sea
breaks against the exposed steel piles at all high tides.

Toe structures management manual


70
Step 1: Identify the type of asset
This is a composite structure with two different types of assets:
• concrete wall
• sheet pile.
Step 2: Identify the factors influencing the asset life
• Coastal environment
• Aggressive wave action and abrasion
• Potential structural instability resulting from the lowering of beach levels
• No maintenance of the concrete wall
• No maintenance of the sheet piles
Step 3: Identify the appropriate deterioration curves
Three deterioration rates (best, fastest, slowest) in years are used to consider the
options:
• Vertical wall / Coastal / Concrete / Both / No maintenance
• Vertical wall / Coastal / Sheet piles / Both / No maintenance
Best estimate (y) Fastest estimate (y) Slowest estimate (y)
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Concrete wall 0 10 30 65 75 0 5 15 25 30 0 20 60 120 150
Sheet piles 0 8 30 43 50 0 4 12 25 30 0 10 44 60 70

The deterioration curve for the composite structure is obtained from the limiting
values of the two curves above:
Best estimate (y) Fastest estimate (y) Slowest estimate (y)
1 2 3 4 5 1 2 3 4 5 1 2 3 4 5
Composite 0 8 30 43 50 0 4 12 25 30 0 10 44 60 70
structure

Step 4: Determine the deterioration curve


The fastest curve is selected as it is assumed that the asset is under severely
adverse loading conditions.
Fast estimate (m)
1 2 3 4 5
Composite structure 0 4 12 25 30

Step 5: Assess the current condition grade


The condition grade (CG) of the composite structure is 3.
Step 6: Forecast the expected deterioration time
The time for the asset to deteriorate from its current condition grade (CG3) to
condition grade 4, which could be considered as the minimum condition grade
acceptable for that structure, is 13 years (that is, from 25 to 12 years).

Toe structures management manual


71
3.8.2 Visual condition assessment and indicators of
performance
Toe structure condition monitoring, as part of a ‘normal’ condition inspection regime,
can be hampered by the fact that toe structures are often obscured from view – they
are typically either submerged or below beach level.
Where the structure is not covered by sediment, an inspection can be scheduled for a
time and date when the tide is low enough for its inspection. If the structure toe is
permanently covered by sediment, then there is rarely a requirement to inspect it as
sediment provides a protective covering.
Inspection pits or trenches may be used if knowledge is required about the toe
structure or its configuration, especially for unknown foundations.
One of the most frequent problems is the lack of knowledge about the presence and
depth of toe structures, especially for older structures where engineering drawings
have been lost or do not exist.
Consideration should be given to the inspection of ground beneath revetted or stepped
toe revetments to assess any washout of fill material. Installation of inspection access
hatches, taking core samples or installing holes for small camera probes could be
prescribed for monitoring purposes.
A simple method of assessment of beach levels is to use a ‘Plimsoll’ type line painted
on a seawall, or by ‘dipping’ – measuring the beach level from the top of the structure
itself. A fixed line can visually indicate beach height at the wall in relation to the toe of
the structure. This can provide the asset manager with a datum to record information
on beach level variability over time in an inexpensive and straight forward way (see
Section 3.5). Pre-determined trigger levels for beach height can be measured to flag up
the need for intervention. Monitoring localised responses in this way allows beach
managers to be proactive in their maintenance programme and reduces the potential
for damage. Properly recorded, it also provides useful design information for future
schemes.
The assessment of the condition of assets is an important process for understanding
the particular state of the defence structures and the asset system as a whole. A series
of snapshots of condition at a particular point in time – by repeated assessment or
monitoring – can record changes in the assets over time and instigate intervention
where necessary to prevent unwanted deterioration in structures or in the level of
performance of the defence system.
The methods used to monitor asset condition and/or performance depend on the
nature of the specific asset types. The general types of inspection and monitoring that
can be applied to assets are:
• Automated – a monitoring system that provides feedback on asset
condition and/or performance without human intervention.
• Destructive testing – a method of inspection that determines the condition
of an asset by analysing a sample of the asset. This sample of the asset is
destroyed in the process of analysis.
• Non-destructive testing – methods of assessing the condition of an asset
without causing any damage to the asset or the removal of its components.
Non-destructive testing ranges from a purely visual inspection of an asset
to radiography, ultrasonic testing and a variety of other techniques.
• Remote sensing – a method of making detailed observations of an asset
from a distance (usually a large distance). The term often refers to

Toe structures management manual


72
observations made by Earth-orbiting satellites or low-flying aircraft. Remote
sensing is inappropriate for detailed monitoring of assets (for example, the
measurement of cracks and small deformations in structures) but provides
a highly efficient technique for general topographical data over large areas.
Visual inspection is the only form of inspection discussed here, being the simplest form
of non-destructive testing and the most widely used technique for monitoring assets –
and it does not require any specialist equipment. However, this method can only
assess the surface details and condition of an asset; changes to interior structure and
condition are not easily identifiable until they lead to changes to the surface of the
asset. Furthermore, toe structures are often designed to lie below the beach surface.
Thus these structures may often not be visible at the time of assessment and therefore
not observable by the inspector. These circumstances make visual assessment of toe
structures particularly opportunistic – only being observable when tidal and sediment
levels are sufficiently low enough.

3.8.3 Performance indicators, failure modes and performance


features
Performance indicators provide evidence (visual or measured) of asset performance (at
a point in time) under loading in relation to its designed or anticipated performance.
Performance indicators may also, in the case of natural structures such as beaches, be
referred to as ‘coastal state indicators’. Possible coastal state indicators include:
• the level of the beach at the toe of a seawall for example (for undermining);
• the beach level plus beach slope;
• the beach cross-sectional area above a set contour (for overtopping).
Flood and coastal defence structures or ‘assets’ can fail in several ways. There are a
number of well-known failure mechanisms, some of which have physical or statistical
models associated with them. These performance models can be used to determine
the likelihood of the type or ‘mode’ of failure occurring given a set of data relating to the
flood defence system being analysed.
Understanding failure modes is important to performance and reliability assessment for
two main reasons:
• So that the correct process-based models can be applied in the analysis of
fragility in the right circumstances.
• So that indicators of failure modes can be included in condition
assessments to elucidate the current condition of defences.
Failure modes display particular ‘features’ when they occur. For example, a sheet pile
wall that is overturning will move from its nominal vertical position, leaning one way or
the other. Washout of fines from beneath a structure will also result in its movement by
settlement – often pre-empted by ground settlement or depression behind the
structure. Such characteristics are referred to as ‘performance features’.
Problems with lowering beach levels may first become apparent from complaints about
access – either pedestrian or vehicular or both (Figure 3.30). Beach levels falling
dangerously below the lower extent of access steps or below the seaward end of
slipways (performance features) can be a handy early warning to monitor the beach
closely for what could be a relatively localised short-term issue or could be signs of a
longer term and wider problem.

Toe structures management manual


73
Figure 3.30 Beach lowering at access steps (courtesy HR Wallingford)
Washout of fill materials from behind and beneath seawalls and revetments can also
be exacerbated by lowered beach levels. Increased pore water pressures on the
landside of the defence can compound such problems. The latter may be caused by
overtopping water draining behind the wall, by water runoff from the hinterland, or by
seawater flushing under the revetment toe on each high tide.
Figure 3.31 illustrates this case where, over time, a stepped concrete revetment has
been effectively undermined – the fill material removed by water entering beneath the
toe of the structure leading to eventual collapse of the pavement behind (Figure 3.32).
However there are no visible signs of deteriorating condition of the revetment or the
seawall that may have contributed to this. This appears to be an extreme case but such
failures are alarmingly common. It illustrates the importance of maintaining design
beach levels. If this is not possible, there should be prudent consideration of structural
remediation or adaptation of the defences.
Avoidance of this type of failure is best conducted in conjunction with beach monitoring
and a clear understanding of the relationship between beach and foundation levels.
Further consideration of failure mechanisms is given in Chapter 5.
Although it would be rare to assess a toe structure in isolation from the rest of a
defence, the toe structure may be of a different structural form and materials from the
main body of the defence. For this reason the performance indicators that need to be
assessed for the toe may be different as well. A concrete toe beam may, for example,
show signs of cracking or settlement, whereas a toe protection built of rock armour may
have no such visible signs or indicators of movement. A rock armour toe may have
been intentionally designed to be ‘sacrificial’, that is, to settle into developing scour
holes. Assessment of condition and performance can therefore be complex and
requires an understanding of the purpose in design as well as structural integrity.
For the purpose of visual inspection, however, toe structures can be inspected
according to the relevant materials and structure inspection guidance (Figure 3.33) in
the Environment Agency’s Condition Assessment Manual (2006). The importance of
maintaining good design and construction records, as a source for comparison, is
emphasised.

Toe structures management manual


74
Figure 3.31 Schematic of backfill washout

Figure 3.32 Undermining and fill washout of stepped revetment and wave
return wall

Toe structures management manual


75
Figure 3.33 An example of a page from the ‘Seawalls’ section of the
Environment Agency’s Condition Assessment Manual (2006)
A particular problem with assessment of beach levels at the toe is that beach levels
can rise as well as fall over short timescales, even if there is a long-term trend of beach
lowering. Accounting for this in performance assessment is difficult. One option is to
record the condition as normal, that is, to report the condition exactly as it is at the
moment of the inspection/assessment. This level can then be assessed in the light of
beach data/knowledge and trends, for example:
• long-terms trends in beach platform lowering;
• seasonal trends, for example, winter season beach lowering followed by
beach level recovery in the summer;
• short-term beach processes, such as tidally induced scour, which fills again
on every tide, either partially or fully.
It is therefore wise, where appropriate, to undertake an analysis of beach level
variability to determine ‘critical’ beach levels against which beach condition
assessments can be made for flood defence and coastal erosion. To facilitate ease of
inspection, responsible authorities are encouraged to consider the installation of height
markers (relative to the toe) on seawalls or perhaps groyne piles from which beach
levels could be determined on a regular basis.

3.8.4 Detailed investigation and assessment


Visual inspection of defence structures and beaches can only be used to identify the
condition of features on the surface, although these can often identify the signs of
problems deeper beneath the surface or within the body of the defence (for example,
rust staining on concrete surfaces reflecting oxidation of rebar within). Sometimes it
may be necessary to undertake more detailed investigations to ascertain internal

Toe structures management manual


76
condition with more certainty. Such tests may be intrusive (for example,
boreholes/cores) or non-intrusive (that is, by remote sensing methods).
Table 3.1 indicates some of the techniques that might be proposed to assist in the
determination of the structural condition and performance of sea defence structures
and beaches. Note this table is not exhaustive.
The type of investigation required is usually a decision taken by specialist consulting
engineers who would provide advice on the specific method, which is best prescribed
on a case-by-case basis. Hence a full account of such techniques is not repeated here.

Table 3.1 Examples of detailed intrusive and non-intrusive assessment


techniques for defence structures and/or beaches
Data Purpose
Geotechnical surveying – window Investigation of geotechnical problems
sampling, trial pits, boreholes, cone highlighted during inspection
penetration test, etc.
Laboratory testing – moisture content, Define soil or material properties, i.e.
Atterberg limits, vane shear strength, resistance, shear strength, compressive
small strain behaviour from consolidated strength and consolidation
triaxial testing and/or chemical testing
In situ testing Investigation of concrete surface tensile
strength, resistivity, etc.
Core samples Condition assessment, e.g. timber
components
Thickness testing For example, indicates extent of damage by
corrosion or borers
To investigate load capacity
Carbonation testing Indicates extent of carbonated concrete
Corrosion testing Applicable where chloride contamination is
a problem
Dynamic load testing of piles Investigates dynamic resistance
Static load testing of piles Tests load/ settlement performance
Use of pressure cells and strain gauges Determine performance of bending
Installation of and monitoring with Records plastic deformation changes in soil
inclinometer to indicate slides in foundations
Used on piles to determine deflection and
bending moments
Installation of and monitoring with Measurements of pore water suction
tensometers
Installation of and monitoring with Measurements of pore water pressures to
observation wells and piezometers determine soil consolidation
Magnetic, acoustic or seismic tests, e.g. Traces anomalies in structure to infer
ground penetrating radar condition or used to determine distances,
e.g. pile length and decrease in concrete
cover

Toe structures management manual


77
Data Purpose
Ecological survey Determine animal presence for control
measures
Nuclear or electrical density Determine the quality of the foundation
measurements in bore holes layer underneath a revetment layer
Magnetic, acoustic or seismic tests, e.g. Traces anomalies in structure to infer
ground penetrating radar condition or used to determine distances,
e.g. detect local cavities and assessment of
revetment thickness
Topographic survey to provide crest level Overtopping analysis to determine current
and cross sectional geometry Standard of Protection
Current wave and water level conditions Overtopping analysis to determine current
and wave incident angle Standard of Protection

3.8.5 When to conduct surveys

Beaches
Seasonal variability in beach levels in front of a coastal structure will affect the results
of a beach monitoring programme. In practical terms, weather conditions make it
difficult to plan the timing of surveys precisely. Typical monitoring programmes make
provision for two regular equally spaced surveys per year, covering the summer and
winter months.
While post storm surveys provide responses that often depart significantly from the
typical long-term trend, they provide the most valuable data for the assessment of toe
vulnerability and should be conducted as soon as possible following the storm event.

Structures
Ideally surveys of coastal structures should be conducted when they are fully emergent
(that is, when the tide is below the lowest point (the toe) of the structure) so that they
can be readily observed and assessed by the inspector. However, this will not be
possible if the toe lies below the lowest tide level. In this case, other approaches should
be considered such as employing specialist divers or even remotely operated vehicles
to gather a visual record of the condition of the structure.
However, the toes of defence structures are typically (and ideally) covered by
sediment, making visual inspection impossible without removing the sediment. Coastal
managers should take any opportunity that may arise to enable the assessment of the
toe, for example, if a beach has ‘drawn down’ after a storm event to a level that
exposes the toe. The beach level may well recover quickly, and before the next
scheduled (defence) inspection, so advantage of the occurrence should be taken to
gather information on the condition (and type if not known) of the toe and its elements.
It may be necessary to conduct intrusive investigations (for example, trial pits) at the
toe to reduce uncertainty associated with structural condition and/or stability when
remedial or new works are being considered. Such activity can often be efficiently tied
in with maintenance works when excavators are available.

Toe structures management manual


78
Any landward developments that will introduce additional loadings on the defence
structure will need to consider the stability of the structure – including the nature of the
toe itself. If this is not known then investigations will be necessary. Similarly, if a
change in the seaward conditions is forecast or planned that will affect the amount of
sediment retained at the toe (for example, dredging of a nearby channel), then the
likely impact on the toe and the structure should be considered.

3.8.6 Risk assessment, defence reliability and determining


‘trigger levels' for action/intervention
There are two types of risk-based performance assessments that are influenced by the
performance of beach levels and structures at the toe of coastal defences:
• structural failure risk assessment;
• overtopping risk assessment.
In both cases the response of the structure is strongly dominated by the beach or
structure level in front of the main ‘wall’. Scour in front of the seawall increases water
depths, often leading to higher overtopping rates with the consequential impacts on
flooding, risk to life, damage to hinterland and properties, and on erosion. The
increased overtopping rates and associated wave impact forces also increase the
potential for failure of the defence, although in many cases the main impact of toe
scour on the failure or deterioration of defences is as a direct result of the loss of toe
support – due to beach lowering or to failure of the toe structure itself.
These risk assessments have been the subject of many previous reports. In particular:
• For information on the calculation of overtopping rates and comparison with
tolerable mean discharges, reference should be made to the European
Overtopping Manual (Pullen et al. 2007).
• Information on assessment of defence reliability is available in many
research reports, including those on the Performance-based Asset
Management System (PAMS) project. The most concise summary of the
issues is given by Simm et al. (2008).
A shortage of knowledge about how defences fail and variations in the characteristics
of defences means that the response of a defence can never be forecast with certainty.
The concept of fragility tries to capture the probability of a range of defence responses
to a given load. Fragility curves for vertical coastal defences such as anchored sheet
piles, cantilever walls and masonry, concrete or gabion walls contain a toe scour term
(Buijs et al. 2007, Table 8). Scour predictors can be used in the calculation of fragility
curves for coastal defences using the monitoring data described below.
• For sand beaches, the depth of scour can be predicted. This requires the
beach slope, the offshore significant wave height, the mean wave period
and the depth of water at the toe of the structure.
• For shingle beaches, the parametric scour plot of Powell and Lowe (1994)
can be used in form of a lookup table to predict the depth of scour. This
requires the offshore significant wave height, the mean wave period and
the depth of water at the toe of the structure.
Beach level at the toe of a coastal defence varies with short-term toe scour depth, but
also has a variation about a long-term mean. These variations in level systematically
alter the ratio between the water depth at the structure (from mean water level to un-
scoured seabed level) to the buried depth of seawall (from the un-scoured seabed level

Toe structures management manual


79
to the structure toe). Variations in the water depth and buried depth alter the forces on
the seawall and hence the elements of the failure limit state function (Buijs et al. 2007,
Sections 3.2–3.4).
The long-term trend in beach level can be obtained from extrapolation of the historical
data (possibly with seasonal variation). This will allow the variation with time of the best
estimate of the fragility curve to be predicted for a few years into the future, depending
on the prediction horizon. Using the predicted trend in mean beach level will allow the
change in the fragility curve with time to be calculated by again altering the water depth
and buried depths in the calculation of the limit state function. This procedure will assist
in calculating the deterioration of performance with time.

Toe structures management manual


80
4 Maintenance

Chapter 4 discusses the options for maintenance of different types


of toe protection structures and materials. A matrix response
summary is provided to identify potential actions to rectify common
issues.

Key links to other chapters:


• Chapter 2 – Toe structure types and materials
• Chapter 3 – Asset management

Who will be interested in this chapter?


• Asset managers
• Coastal engineers

4.1 Introduction
Analysis of existing scheme performance at many sites provides clear evidence that
small-scale maintenance treatment of early stage problems is far more cost-effective
than allowing problems to develop. While there are often economies of scale
associated with minimising mobilisation of plant and labour to conduct works, this
principal does not apply to management of structure toes. Damage can develop very
quickly on the toe of structures and small-scale problems can evolve rapidly to cause
major failures through the various failure mechanisms. In particular, failures involving
undermining of the foundations and loss of core material are particularly difficult and
expensive to deal with. Indeed, in the survey carried out by CIRIA in 1986, which
examined the performance of seawalls in England, Wales and Scotland, it was
concluded that erosion at the toe of structures represented, by far, the most prevalent
and serious form of damage to seawalls in the UK Over 12 per cent of all seawalls
reported erosion at the toe, which represented over a third of all damage reported
(CIRIA 1986).
The resultant cost of rehabilitation is usually extremely high. Notwithstanding these
observations, there are an alarming number of structure failures which could have been
avoided with timely maintenance. The fact that the toe underpins the coastal defence
superstructure means that in many cases it is just not practical to reconstruct the toe
without major modification to the whole defence structure.
Maintenance can be considered at two principal scales:
• minor maintenance without modification to the structure;
• toe modification by major reconstruction.
Minor maintenance might include such activities as structural maintenance of joints,
maintaining safety of the tops of steel piles and maintenance of beach levels above
defined trigger levels. Failure or partial failure of the toe usually requires construction of
a new or modified toe. In most cases where a new toe is installed, it is built to
seawards of the inadequate older structure. The design principles outlined in Chapter 5

Toe structures management manual


81
remain relevant, although there may be some additional requirements for details to tie
the modified toe to the old structure. The vast majority of toe protection works are not
‘new build’, but entail some form of modification to the existing structure.
Minor or regular maintenance is generally funded from the Environment Agency’s or
the local authority’s revenue budgets. ‘Capital’ maintenance is more appropriate to toe
modification which will probably require Flood Defence Grant in Aid (FDGiA) funding
and therefore need to go through a more rigorous approvals procedure.
Similarly, maintenance works to defences that do not alter the form, profile or footprint
of a defence are exempt from Marine Management Organisation (MMO) licence
requirements, while works that do modify the defence, such as addition of a rock toe,
do require a licence and associated consultations.
While it is normal for maintenance of the whole of the structure to be considered at the
same time, this section focuses on activities related to the toe elements only.

4.1.1 Definitions
The term ‘maintenance’ can be interpreted in a number of ways, each reflecting
different views on the scope and range of activities included. For example, the US
Army Corps of Engineers’ (USACE) Coastal Engineering Manual (CEM) (USACE
2012) defines ‘maintenance’ in accordance with Vrijling et al. (1995), as consisting:
‘of the following essential elements:
a. Periodic project inspection and monitoring of environmental
conditions and structure response.
b. Evaluation of inspection and monitoring data to assess the structure’s
physical condition and its performance relative to design
specifications.
c. Determining an appropriate response based on evaluation results.
Possible responses are:
• Take no action (no problems identified or problems are minor)
• Rehabilitate all or portions of the structure
• Repair all or portions of the structure.’
This definition introduces two further terms – ‘rehabilitation’ and ‘repair’. The CEM
(USACE 2012) goes on to distinguish these two levels of intervention by saying that
‘rehabilitation’ implies steps are taken to correct problems before structure functionality
is significantly degraded (for example, patching spalled concrete), while ‘repair’ implies
that damage has occurred and structural functionality is already significantly reduced
(for example, repairing a vertical wall).
Further distinctions can be made regarding the management approach to maintenance:
• Pro-active or preventative maintenance – (rehabilitation) maintenance
based on the observed condition of the project.
• Periodic maintenance – (rehabilitation) maintenance that occurs after a
prescribed time period or when a particular loading level is exceeded.
• Reactive maintenance – (repair) maintenance, undertaken in response to
the occurrence of actual damage.

Toe structures management manual


82
4.1.2 Scope of maintenance considered in this chapter
The term ‘reconstruction’ meaning the complete rebuilding or replacement of a
structure is not included in the list of definitions above. Arguably, reconstruction of just
the toe element could be regarded as a repair of a complete coastal defence structure.
However, given that these guidelines focus specifically on the toe, reconstruction is not
regarded as repair and is, therefore, not classed as maintenance. Reconstruction or
construction of the toe is dealt with in Chapter 5.
The maintenance considered here includes monitoring and evaluation of the outputs of
monitoring as the basis for the undertaking of rehabilitation or repair maintenance.
These aspects of asset management are dealt with in Chapter 3.
The present chapter is therefore confined to the practical aspects of maintenance of
toe structures by either rehabilitation or repair. Moreover, it is supposed that
maintenance involves attention to present toe structures together with topping-up or
additions to the host materials, but does not include replacement of a structure or the
substantial introduction of new materials (see Chapter 5).
The next section looks at aspects of maintenance that are particular to toe structures
such as limitation on access. Section 4.3 describes maintenance issues and remedies
for various materials and forms of construction used in toe structures. As these two
descriptors (materials and forms of construction) are closely allied (for example,
masonry is used in revetment type construction), the sections are ordered according to
material type in line with the description of materials given in Chapter 2.

4.2 Issues associated with maintenance of existing


toes

4.2.1 Designing for maintenance


Maintenance requirements should be considered as early as the Project Appraisal
Report (PAR) stage. Considerations at this stage will tend to focus on the financial
commitment and whole life costs. Subsequent development of ideas at the outline
design stage will provide important input to construction, design and management
(CDM) considerations. The need to recognise and carefully consider safety during
subsequent maintenance is likely to influence not only the design of the maintenance
operations per se, but also design of the capital works.
In practical terms, maintenance takes place on structures that are often several
decades old and thus pre-dating current safety standards, which should make
adequate provision for safe maintenance. Unlike the crest of a seawall and most of the
body of the wall, the toe is relatively difficult to access for maintenance purposes.
Where major maintenance to the structure toe is required, this can present an
opportunity to improve access for further maintenance. For example, when a rock toe is
added to replace or reinforce an existing toe, it may be possible to improve access by
construction of a rock berm of an appropriate width to support a tracked excavator; this
may facilitate improved access along the structure. Such works may involve additional
material to those required to deliver the hydraulic and stability requirements, but will
enable more cost-effective long-term repairs in the future that may not require
construction of expensive temporary haul routes. Once installed, the toe might be
mainly or entirely concealed from view under normal conditions of weather and beach
level. The times when the toe becomes exposed to new/accumulated wear (that is,

Toe structures management manual


83
during stormy weather accompanied by beach drawdown) is not usually the best time
to inspect or monitor its condition, let alone carry out maintenance. It is generally
preferable therefore to eliminate maintenance of the toe until such time as a more
major intervention might be required, for example, 30 years or more depending on the
circumstances and residual life of the main structure.
The design of capital works provides an opportunity for planning and minimising
maintenance commitment. For example, the inclusion of sacrificial allowance in the wall
thickness specification for steel sheet piling may reduce long-term maintenance
requirements. Clearly, such opportunities only exist in cases of new construction or
major reconstruction. Funding of maintenance is typically provided from revenue
budgets of the maintaining organisation, so this design stage consideration is
worthwhile.
Assuming they can still be accessed, much can be learnt from the original design
calculations for coastal structures. Design conditions may make allowances for
corrosion, abrasion, rounding/loss of rock mass, breakage, cracking and so on (that is,
damage arising from ageing and exposure to the marine environment).
Information gained from earlier records must be examined alongside any changed
conditions and limited inspection of the toe in its present condition, to reassess the
need for maintenance.

4.2.2 Changing loading conditions


Coastal loading by waves and currents can change over time. Wave climates are not
static and a review of these is suggested for each site on a rolling 10-year cycle. Data
to permit this assessment are generally becoming more widely available in the UK
through the regional coastal monitoring programmes.
This means that future conditions can be different to those adopted for initial design
purposes. Actual conditions can be very different to those evaluated for design simply
because of the approximations and assumptions made in determining the design wave
climate. Changes in the hydrodynamic effects of waves and currents will always result
in changes to a (non-cohesive) sediment regime including patterns and trends of beach
erosion or accretion.
Lowering of the beaches has a significant effect on the exposure of a coastal defence
to wave attack. A lowered beach level at the toe can mean that parts of the structure
that were previously buried become exposed to the action of the sea. In the more
severe cases, this can lead to instability of the toe – and hence the whole structure –
due to geotechnical pressure or undermining. The latter case could lead to the
requirement to reconstruct the toe. Before this stage is reached an exposed toe is likely
to require increasing levels of maintenance attention to counter the effects of direct
loading and, in particular, abrasion due to water and sediment movement. The extent
of abrasion and the scope for correcting it depends, among other things, on the
material used. For example, open stone asphalt is highly vulnerable to abrasion while
durable rock would be significantly less vulnerable. The maintenance commitment is
likely to increase once the toe becomes exposed, depending on the materials used.

4.2.3 Beach management


Beach management forms a significant element of structure toe maintenance at many
sites.

Toe structures management manual


84
The most common cause of failure of seawalls is undermining of the toe, which
generally arises as a result of falling beach levels. It is important to maintain the beach
above the trigger levels set for the site (see Section 3.5) and this requires regular
monitoring (Chapter 3). Maintenance operations may vary considerably in size, ranging
from the movement or addition of a few 100 m3 per year to more than 10,000 m3 per
year. Beach management activities may include either recycling from within the
immediate frontage or adjacent sections, or by topping up using additional imported
material
The approach using regular detailed baseline topographic surveys allows generation of
digital terrain models and calculation of beach volumes relative to trigger levels. The
approach adopted in Figure 4.1 highlights the volume necessary to maintain the
required design management levels at a site fronting a seawall. The approach locates
and quantifies volumetric shortfalls within groyne compartments, identified by survey.
Possible recycling sources are also shown within zones containing a surplus above the
required management profile. The number of truck loads required in each groyne bay is
highlighted, providing the basis for logistic planning of maintenance operations. The
maintenance undertaken at this site is at a very large scale relative to most UK sites.
Alternatively, material may be imported from offsite sources in order to achieve the
required beach levels.
On some occasions trigger levels are set on the basis of visual identification of key
features on the structure toe. The example shown in Figure 4.2 has an intervention
trigger when the interface between the bottom step and foundation shutter is exposed.
This structure has not been designed to allow direct wave attack at this elevation on
the toe.

Figure 4.1 Application of monitoring data to identify beach zones beneath trigger
levels (courtesy Worthing Borough Council)

Toe structures management manual


85
Figure 4.2 Beach that has reached trigger level prior to repair and recharge
(courtesy Poole Borough Council)
The crucial considerations to be made when planning beach maintenance are:
• Use projected rates of loss from monitoring to identify how much sediment is
required for each maintenance operation.
• Make adequate provision for accelerated losses at hot spots.
• Add more material to updrift zones to allow for sediment transport.
• Log quantities of material added or recycled to inform long term beach
management plan.
• Check the volume of material following significant storm events.
Figure 4.3 provides a valuable summary assessment of toe scour management using
beach recycling and small scale recharge. Trigger levels are shown. The experience of
previous toe failure at this site has been built into the maintenance plot. The
experiences and linkages between beach losses and storm events can be identified. A
series of small recharge operations have been conducted to maintain the beach above
trigger levels and there is a clear annual trend of loss which can be used to plan for the
requirements of further intervention. The vertical grey lines indicate interim recharges,
which are small in size (approximately 1,000–7,000 m3 each).

Toe structures management manual


86
Figure 4.3 Application of monitoring data to identify long term beach
performance in context with structure toe triggers (courtesy Channel Coastal
Observatory)

4.2.4 Re-use of materials


The reuse of materials is generally considered to have environmental benefits.
Exceptions to this might include situations where material retrieval might cause greater
damage to the donor environment than if new materials were to be used from a remote
source.
In the case of toe maintenance, materials for reuse are unlikely to have originated from
the toe itself but may have been derived from removal or reconstruction of other works
in the vicinity. Materials for reuse could include rock and steel sheet piling. In some
instances, failed concrete slabs arising from toe failure have been crushed and used to
provide fill material.
When reusing rock, attention should be paid to the condition of the salvaged material. If
it has spent some considerable time in an aggressive marine environment (for
example, tens of years), it may have become rounded, thus lessening its stability
properties if required for rubble-mound type of installation. If the stones have become
rounded in the marine environment but were only deployed previously for a limited
period (for example, for temporary works), this could indicate inadequate durability.
Provided its relative lack of internal friction is not a significant issue, partially degraded
armour can often, however, be used effectively in sublayers of multilayer construction.
Extraction of piling for reuse may be achieved by use of vibratory or jacking type
extractors. The Piling Handbook (Arcelor 2008) provides advice on pile extraction,
though this is largely directed at the extraction of temporary piles. Piles previously used

Toe structures management manual


87
in permanent works are likely to introduce a number of complications, including the
difficulty of relocating the original piling records, and in practical terms, the fixity caused
by walings, ties and so on. Although the removal and reuse of interlocking piles
provides a theoretical solution, the state of the piles is often inadequate for reuse.
Clutches are often extremely badly damaged, restricting the ability to interlock. In
addition, the tops of the piles are often damaged by the ends being bent or burred
under the combined action of waves and sediments (Figure 4.4). Reuse of piling for
maintenance implies the replacement of piles, which in turn implies the sourcing of
piles that will interlock with those left in place. These complications clearly limit the
reuse of piling for maintenance purposes.

Figure 4.4 Typical damage to tops of steel interlocking piles (courtesy AP


Bradbury)

Stockpiling
Materials may be stockpiled for one or both of the following principal reasons:
• to offset the relatively high cost of future mobilisation each time that
maintenance, using the relevant material, is carried out;
• to have materials readily available in the case of a breach or severe
damage to the defence.
Both these reasons could apply to toe protection maintenance, albeit that the second
one implies emergency restoration of a defence structure rather than maintenance.
Rock fill is a likely choice of material for sealing a breach. Given the circumstances
under which it might be deployed (that is, during or soon after a storm), it is imperative
that the stockpile is accessible using land-based plant. Moreover, the plant required to
recover, transport and place the materials must also be available at short notice.
While it is often not convenient to stockpile materials at a site, arrangements may be
made with nearby quarries to hold materials in reserve for emergency maintenance
operations. This avoids the need for sometimes lengthy mobilisation of production.

Toe structures management manual


88
4.2.5 Access
Access for maintenance may be especially problematic at the structure toe in cases
where the toe is buried, and inspection and maintenance may not be practical on a
routine basis. This may not be a major concern with new construction where well
planned and controlled design eliminates the need for maintenance for many years. In
these cases, intervention might be invoked on the basis of trigger levels being reached,
for example, lowering of the beach to a critical level (see also Chapter 3).
Where access for maintenance is necessary a number of factors need to be
considered:
• Materials to be used and/or the nature of the operation. These will
determine the type of plant needed (that is, cranes, pile driving equipment,
dump trucks and so on).
• Loads to be lifted and placed. This will determine the capacity and reach
of cranes and other plant.
• Capacity of the promenade to carry plant/constraints on loads and
reaches, and nearness to the edge of the promenade/seawall structure.
These considerations must take account of any excavation at the toe (and
hence reduced passive support to the wall).
• Access along the beach. This may be impeded by groynes, outfalls or
other structures.
• Access to the beach from the crest of the structure. Many aging
structures did not make suitable provision for plant access to the beach at
regular intervals.
As maintenance only applies to structures and defence systems that are already in
place, it follows that access must have been possible at some time when the defence
was originally installed or last rebuilt. Exceptions to this may result from:
• post defence construction infrastructure;
• especially in the case of very old structures, by the structure itself (for
example, access formerly having been gained from the land to form a
foundation, then working upwards to form a seawall);
• due to worsened (lowered) beach conditions limiting the tidal window for
working.
Access to the toe is a major problem at many sites and, where access is not possible
at the crest of the structure, access via temporary haul roads may be required
(Figure 4.5). In this instance, access is required over a number of closely spaced
timber groynes. The simplest solution is often reshaping of beach material to form a
temporary access way, although this approach is often fraught with the difficulties
arising from inadequate beach volume and wave action which may regularly destroy
the access.

Toe structures management manual


89
Figure 4.5 Temporary access road constructed from local beach material
(courtesy New Forest District Council)
Maintenance activities may provide the opportunity to enhance access to a site.
Addition of a rock toe (Figure 4.6) to support a vulnerable structure has provided the
opportunity to improve access along the toe of the same structure as shown in
Figure 4.5. The crest berm has been constructed to provide a suitable width to enable
a tracked excavator to move safely along the top of the rock toe at all states of the tide.
This approach is quite costly, since more rock is required for construction. In addition,
the structure has considerable hydraulic benefits in reducing overtopping by the
dissipative berm and reducing the volume of beach material required to recharge the
site following completion of the toe maintenance works.

Figure 4.6 Improvement of access by construction of rock toe with wide crest
berm suitable for plant access (courtesy New Forest District Council)

Toe structures management manual


90
An example of improvement of an existing dilapidated access along the interface
between a mid-19th century vertical wall and a toe revetment, at Llanfairfechan, North
Wales, is shown in Figure 4.7. The later shot (to the right) shows how the access was
upgraded for maintenance works that were carried out to the toe revetment.

Figure 4.7 Improvement of access by construction of concrete roadway suitable


for plant access (courtesy Alan Williams)

4.2.6 Temporary works


Maintenance works often need to be undertaken in difficult conditions. Working is often
required at extreme low water and any small wave activity is likely to reduce the safety
of operatives. Rock armour, which is eventually used to provide an additional toe, may
be used to provide a temporary bund to achieve safe working adjacent to the wall
(Figure 4.8).

Toe structures management manual


91
Figure 4.8 Temporary rock bund to provide safe working area for plant during
toe repairs (courtesy New Forest District Council)

4.2.7 Budget prioritisation


The budgets available for maintenance are often limited and decisions about where
monies are spent need to reflect the criticality of the each structure in the overall
defence system. Prioritisation is often required to decide which structures should take
precedence to be sustained at the required level of performance. This might be
achieved by a risk-based assessment of the flood or erosion defence system.
What is clear is that structure toes are extremely vulnerable and can suffer a rapid
brittle failure unless adequate maintenance of both the beach and structure is
undertaken. Any budget assessments should consider carefully the costs of routine
maintenance against the potential costs of emergency repairs following failure.
Although it is difficult to provide a generic comparison for widely varying structures,
recent experiences at several sites have suggested the cost of emergency repairs has
exceeded £25,000 per metre run (over typical distances of 20–50 m). These very high
costs reflect the fact that emergency maintenance requires:
• more complex safety arrangements;
• complex temporary works;
• dismantling of failed elements;
• working in a challenging environment to effect repairs.
It is suggested that adequate and regular planned maintenance will provide a better
cost-effectiveness ratio by a factor of at least 10 against emergency works.

4.3 Maintenance of toe structures

4.3.1 Concrete
Preventing deterioration of concrete is easier and more economical than repairing
concrete. Such prevention begins with construction by ensuring that:
• the proper and appropriate materials are selected;
• the mixture has the correct proportions;
• placement and curing procedures are correct for the purpose.
Particular attention should be paid to the correct selection and specification of ‘marine
grade’ concrete mixtures for all saline and coastal applications – including toe
structures. Guidance on such specifications can be found in the Maritime Concrete
Manual (CIRIA 2010b).
The most common types of maintenance for concrete include:
• repair of cracks and spalls;
• cleaning to remove unsightly material;
• surface protection;

Toe structures management manual


92
• joint restoration.
The Concrete Repair Manual (ACI 2008) contains a significant collection of concrete
repair information. Topics include:
• condition evaluation;
• materials for repair;
• surface preparation;
• application methods;
• corrosion management;
• structural strengthening;
• protection methods.
Contractual guidance for measuring concrete repair work is also included.
Methods of maintenance and repair should be considered to treat wear and tear of the
fabric. Such issues include:
• exposure of rebar;
• water ingress and oxidation of reinforcing bars;
• development of scour scars.
A general maintenance programme of surface treatments such as concrete, facing,
patching and concrete spraying should be considered to:
• replace abraded surface cover;
• prevent water ingress;
• replace lost material from concrete structures.
Left untreated, such processes will only accelerate the rate at which these structures
deteriorate.
Tables 4.1 and 4.2 give an indicative list of repair options and examples of where the
techniques may be used. It should be noted that some of these also apply to the repair
of masonry structures that occur frequently in the maritime environment. The tables
use the following key:
• UW: generally suitable for underwater elements or parts of the structure,
without specific cofferdams or limpet dams (subtidal zone);
• T: generally suitable for elements or parts of the structure in the tidal zone
(between MLWS and MHWS);
• S: generally suitable for elements or parts of the structure in the splash
zone (zone above highest astronomical tide);
• OW: generally suitable for overwater zones for which access is
constrained;
• D: generally suitable for elements or parts of the structure in the dry.
It is essential to check:
• the compatibility of all techniques and repair products with the environment;

Toe structures management manual


93
• the project programme;
• the required performance characteristics.

Table 4.1 Structure repair/restoration options and their applicability to the


maritime environment (based on CIRIA 2010b)
UW T S OW D
Restore structure condition and stability
Replace element by precast element     

Replace element by concrete bags  ?   


Cast in situ concrete ?    

Underpinning ? ?   

Pressure grouting of voids  ?   

Spray concrete or mortar  ?  ? 


Supplementary ground anchorages  ?   
Post-tensioned concrete elements  ?   

Relieving slabs  ?   
Restore structure performance
Restore drainage systems     

Restore transitions/joints ; create joints1  ?   

Recast levelling surface slab in situ ? ?   


Restore protection systems, e.g. fenders ? ?   

1
Notes: STRRES (2007a)
 = generally suitable; ? = challenging;  = generally not suitable

Table 4.2 Options for repair works related to defects in concrete and concrete
reinforcement and their applicability to the maritime environment (adapted from
BSI 2006)
UW T S OW D
Restore concrete
Applying mortar by hand  ?   

Recasting concrete ? ?   

Spraying concrete or mortar  ?   

Replacing element ? ?   

Injection of cracks1  ?   
Restoring reinforcement and reinforcement passivity
Restoring cover: replacement mortar/concrete ? ?   

Toe structures management manual


94
UW T S OW D
Replacing damaged concrete     

Re-alkalinisation: electrochemical   ?  
Re-alkalinisation: diffusion   ?  

Chloride extraction: electrochemical   ?  


Replacing or supplementing corroded rebar ? ?   

1
Notes: STRRES (2007b)
 = generally suitable; ? = challenging;  = generally not suitable

Box 4.1 provides a list of useful references on concrete protection and repair.
Box 4.1 Useful references on concrete protection and repair
• Principles and Practice of Galvanic Protection for Reinforced Concrete, Technical
Note 6, Corrosion Protection Association, 2004.
• Handbook of Coatings for Concrete, R. Bassi and K. Roy, Whittles Publishing,
2002.
• Guide to Surface Treatments for Protection and Enhancement of Concrete,
Technical Report No. 50, Concrete Society, 1997.
• Cathodic Protection of Reinforced Concrete, Technical Report No. 36, Concrete
Society, 1989.
• Guide to the Repair of Concrete Structures with Reference to BS EN 1504,
Technical Report No. 69, Concrete Society, 2009.
• Protection of Reinforced Concrete by Surface Treatment, Technical Note 130,
CIRIA, 1987.
• Guide FABEM 4 Protection des bétons, STRRES, 2007.
• Protection des bétons par application de produits à la surface du parement, LCPC/
SETRA, 2002.
• Mise en peinture des bétons de génie civil, LCPC, 1999.
• Choix et application des produits de réparation et de protection des ouvrages en
béton, LCPC, 1996.
• Méthodes électrochimiques appliquées au diagnostic et à la réhabilitation du béton
armé concerné par la corrosion, A. Rahanarivo, LCPC, 2005.
• Fluctuation du potentiel des aciers dans le béton et sous protection cathodique, I.
Pepenar, G. Grimaldi and A. Rahanarivo, LCPC, 1994.
• Concrete in Coastal Structures, R.T.L. Allen, Thomas Telford, 1998.

Underpinning of seawall between original foundation base and bed rock


Major repairs of concrete structures are often required following undermining failures.
Although concrete is weak in tension, it is possible for large areas of reinforced mass

Toe structures management manual


95
concrete walls to remain unsupported over fairly long stretches, at least on a temporary
basis (Figure 4.9). This occurs commonly where the structure foundations are perched
within the tidal limits and where the original foundation construction has not closed onto
bedrock. Support beneath the wall is required quickly to avoid cracking and failure
under tension. This may often be conducted by underpinning the wall using mass
concrete to provide a new foundation.

Figure 4.9 Unsupported section of reinforced mass concrete seawall following


major undermining (Chesil Cove 1962) © Stuart Morris
Maintenance operations which require underpinning are challenging and sometimes
require the bespoke development of specialist tools to achieve safe working and to
maximise the structural integrity of the repair. This often requires removal of failed or
loose material from beneath the undermined wall. A modified excavator bucket is used
to rake loose material from beneath the bridging wall while maintaining safety for
operatives (Figure 4.10). Excavations should ideally close onto competent bedrock
material, but this is not always possible due to the depth of beach sediments.

Toe structures management manual


96
Figure 4.10 Preparation of foundation excavation beneath wall for
underpinning, using bespoke plant to ensure safe working beneath seawall
(courtesy New Forest District Council)
The detail provided in Figure 4.11 shows an underpinned toe repair with the following
characteristics:
• The base footprint of the repair underpinning is wider than the original
foundation.
• Drainage is secured by regularly spaced high density polyethylene (HDPE)
pipe placed within the underpinned foundation.
• The void behind the wall has been filled with coarse free-draining material.
• Loose material has been removed from beneath the old foundation.
• The new foundation closes out onto bedrock (this is not always possible).
• The underpinning concrete is tied into the old structure by way of cast
interlocking upstands on either side of the underpinned foundation.
• Shuttering is provided by geotextile material held in place by rock armour
and shingle on the seaward face.
CROSS-SECTION OF PROPOSED ROCK REVETMENT & SEAWALL REPAIR

3m

0.65m ~1m

3.9m OD

3.4m OD

rock & geotextile


to acting as
shuttering
2 la
void to be M
filled with free
4.2m
draining
material 1.4m OD

seawall 0.3m*
0.45m
(variable) 0.55m OD

su
base of void
-0.8m OD

2m (variable)

3m

Toe structures management manual


97
Figure 4.11 Underpinning detail for seawall foundation repair (courtesy New
Forest District Council)
Shuttering the toe of underpinning foundation operations is often challenging since the
toe of the structure may be subject to wave loading even during construction. In
emergency situations, concrete may be pumped into voids, without manufactured
shuttering, in an attempt to provide a rapid solution. This can provide an extremely
expensive, untidy and not particularly strong repair since large losses of concrete may
occur. Use of beach material can be used to provide shuttering bunds in this situation,
although this would not be considered best practice. Ideally, shutters for repair should
close tightly on adjoining surfaces, although this may not always be possible because
of the safety aspects of their installation on a potentially unstable structure.
Promenade deck failures of the type shown in Figure 4.12 are fairly common and occur
as a direct result of toe failure. Historical construction of mass concrete walls typically
included a locally won backfilled material. This material is quite likely to have a
significant fines content. Repairs of the backfill zone, following structure toe failure,
should ideally be replaced with a coarse compacted granular fill that is free of fines.
This will assist with drainage of water from the landward side of the structure and
should prevent subsidence of the surface and formation of a void beneath decking,
which may often be a cast concrete slab – although the example shown is a simple
asphalt surface.

Figure 4.12 Typical collapse of promenade decking following undermining of


backfilled mass concrete wall and washout of fine core (Chesil Cove 1962) ©
Stuart Morris

4.3.2 Rock (including gabions)


While the definitions of toe repairs used within this section do not include the addition
of a new rock toe, this may often be undertaken as a large-scale maintenance option.
Such treatment is regularly chosen as management option when the toe has become
undermined. The details of the design of a modified rock toe are dealt with in Chapter

Toe structures management manual


98
5. It is clear, however, that rock toe structures are far better at energy dissipation than
vertical concrete structures (Figure 4.13).

Figure 4.13 Contrast between concrete toe and modified rock armour toe
Folkestone Warren (courtesy Bryan Curtis)

Reprofiling rock structures


Provided access is possible, Repositioning or reprofiling of dislodged armourstone in a
toe structure may be a relatively straightforward maintenance option for limited repairs
(Figure 4.14).

Figure 4.14 Manoeuvring armourstone blocks in a revetment (courtesy Dean


and Dyball Ltd)

Toe structures management manual


99
Access restrictions at the toe may be more challenging, especially if material has
moved from its original position seawards of the toe; in this case recovery of damaged
armourstones may be difficult. It is generally more economic to add new rock to the
existing profile, providing that a supply of suitable and economic of material is
available. Where armourstone must be reprofiled, it is often necessary to ‘unpick’ areas
of the structure to ensure that adequate interlock can be achieved from the repair.
Large-scale damage may require a more substantial commitment of resources and
could amount to a ‘rebuild’ rather than simply a repair.
When rock armoured toes have been constructed as part of the defence (see
Chapter 5), these may regularly suffer damage (Figure 4.15). A frequent problem
arises as the rock toe settles, particularly when this is placed on soft bedrock material.
Settlement may result in displacement and reduced interlock of some of the armour,
which may require reconstruction to maintain structure integrity. The benefit of the rock
armour at the toe is that it is self-healing at the point of bed settlement and will move
with the bed. This has the advantage that the structure will not become undermined.
The consequence of settlement though is that more material is generally needed to top
up the profile. Allowance for settlement of a falling toe is often built in to the design of a
rock toe.

Figure 4.15 Armour displaced from a rock toe, Llanfairfechan, North Wales
(courtesy Alan Williams)

Refilling/replacing gabions
Gabions placed in even relatively mild coastal conditions usually require extensive
maintenance and are not generally recognised as efficient toe structures. The contents
can settle within the gabions or may escape altogether, with contortion of the mesh or
breakage (Figure 4.16). Gabions that have simply ‘settled’, which remain intact and are
relatively in shape may just require topping up with fill material. Sometimes ‘leaked’

Toe structures management manual


100
contents may be reused if they remain close by. Replacement fill material should be
close to or larger than the size of the original. Gabions that have distorted badly or
collapsed altogether will usually require replacement of the whole gabion unit.

Figure 4.16 Weathered gabions at East Head, Chichester Harbour, West Sussex
(courtesy HR Wallingford)

4.3.1 Masonry

Regrouting/pointing
The regrouting or ‘pointing’ of joints in masonry structures is an important maintenance
task. Missing joint filler allows water into the structure. If the gap penetrates through to
the sublayer, then washout of fill material can ensue. The resultant voids in the
sublayer can affect the response to hydrodynamic forcing on the structure, which can
quickly weaken from repeated pressure changes. Loose blocks can simply be lifted or
sucked out of the structure under wave or surge action and lead to further, potentially
rapid, deterioration or even structural failure. Vegetation, once established in joints, can
aggravate the percolation of water into the fabric of the structure (Figure 4.17) and the
general weathering of the blocks.

Toe structures management manual


101
Figure 4.17 Vegetation growth on a revetment (courtesy ENBE)
If allowed to grow, vegetation such as bushes and trees can cause deformation and
movement in the structure – forcing blocks apart or out of the section (Figure 4.18).
Hence vegetation should not be allowed to take root in such structures and should be
cleared as necessary. Problems associated with vegetation are more likely to prevail in
the higher parts of the wall structure rather than at the toe.

Toe structures management manual


102
Figure 4.17 Vegetation colonising joints in a masonry coastal defence
(courtesy HR Wallingford)

Replacement of dislodged or lost masonry blocks


Figure 4.19 shows a masonry toe before and after repointing. A durable grout should be
used for pointing to ensure effectiveness and regular inspections made to identify any
damage or weaknesses, and to record general deterioration in joint grouting over time.

Toe structures management manual


103
Figure 4.19 Prior to and following maintenance of grouted masonry toe
(courtesy Alan Williams)
Dislodged masonry blocks at the structure toe may act as a catalyst to rapid toe failure.
Under severe storm conditions, failure of the whole defence could quickly ensue.
Figure 4.20 illustrates such a situation, which if left unaddressed could cause rapid
development of damage to the revetment, together with possible collapse. Major
damage arising from this initially small problem would result in significantly greater
costs of repair than replacing the single displaced block. Maintenance has to be
undertaken soon after damage occurred, if progressive damage is to be avoided. This,
in turn, implies regular inspection of defences (and inspection after storms) so that
damage can be quickly observed and acted on.

Figure 4.20 Dislodged masonry blocks from revetment toe (courtesy


HR Wallingford)
Collapsed, dislodged or missing blocks from toe revetments can quickly lead to more
major problems, with washout of fill material and further structural instability and
eventual failure (Figure 4.21).

Toe structures management manual


104
Figure 4.21 Washout and collapse of block revetment following block removal
(courtesy Canterbury City Council)

4.3.4 Steel
Sheet piles need replacement or repair where they have deteriorated over time either
through rusting, abrasion and erosion of section, or through accelerated low water
corrosion (ALWC). More generally, failure is manifested by holes in the piling rather
than failure of the pile as a structure (Figure 4.22). In the more extreme cases, the
thinning sections can become razor sharp. Remedial action should be taken
immediately where corroded exposed piles pose a hazard to people or animals.

Figure 4.22 Deteriorated sheet piles


While the extreme damage shown in Figure 4.22 is generally unsuitable for repair by
welding on steel plates, welding may provide a suitable repair method at locations
where the damage is localised and small holes or thinning of the piles has occurred.

Toe structures management manual


105
Damage to the tops of piles on toe structures is very common, especially when they
are subject to rapid abrasion in a sediment-charged environment. Where the piles
simply form a shutter to the mass concrete behind, this may not be a particularly
serious issue (Figure 4.23).

Figure 4.23 Complete loss of section of sheet piles in an area exposed to


sediment movement
Where damage is noted to the pile tops, reference should be made to the original
design to determine the role of the piling in the structure. In those instances where the
pile is simply forming a shutter, without a structural role, the piles can be maintained by
burning or cutting the tops to remove sharp edges or by covering with rock armour.
This is generally a safety measure (Figure 4.24).

Figure 4.24 Damaged interlocking steel piles requiring maintenance to make


site safe (courtesy A P Bradbury)

Toe structures management manual


106
The service life of sheet piles can be extended, albeit marginally, by applying protective
coatings. Cathodic protection systems can also be used but these are not frequently
encountered in coast defence works (unlike steel used in offshore applications,
pipelines and so on), presumably because of the aggressive environment and the
consequent risk of damage to the cathodes. In the toe structure especially, material
loss can be substantially due to abrasion by shingle or sand.

4.3.5 Asphalt
Asphaltic revetments with open joints or fissures extending to the full depth should be
repaired promptly, especially on the waterside slope. Loss of subsoil though such holes
leads to the revetment settling and deforming, and eventually failure. Ideally a
favourable time of year and weather conditions should be chosen to undertake such
repairs (Schönian 1999).

Sand mastic asphalt repair


Holes in sand mastic revetments should be overfilled so that the grout forms a cap
which binds well to the existing surface. The surface should be prepared by cleaning
away sand and plant growth and pre-treated or ‘tacked’ with a coat of hot bitumen. Thin
layers of repairs at the toe of the structure should be avoided where these are not
bound to an appropriate base of asphaltic concrete, stone base or sand mastic.
Damage to an open stone asphalt revetment should be repaired immediately. The
number of contact points between stones in the layer is limited and therefore integrity is
heavily reliant on the asphaltic binder. As the production of small amounts of open
stone asphalt for repair purposes is difficult and costly, stone fill grouted with sand
mastic asphalt is recommended. Importantly this does not change the permeability of
the revetment, although the viscosity should be well-adjusted to prevent any run out.
Only loss of material from the surface can be repaired by patches of sand mastic
spread over a prepared and cleaned old surface. If the damage is more serious, the
section may need to be cut out and refilled again with open stone asphalt, or the
existing surface may be cleaned, dried and tack coated before adding the new layer.
Figure 4.25 shows both concrete and neat bitumen being used to repair holes in an
open stone toe apron at Prestatyn, North Wales. While this is not best practice and has
potential impacts for the permeability of the structure if used extensively, it is clearly
cheaper especially for small repairs and has been effective particularly as a stop gap.
The concrete appears to have adhered reasonably well to the open stone asphalt,
probably due to the surface roughness of the open stone asphalt and good preparation
before the repair was effected. There has been no obvious change in the condition of
the repair at the example site over a five-year period.

Toe structures management manual


107
Figure 4.25 Repairs to open stone toe at Prestatyn, North Wales (courtesy Alan
Williams)
General guidance on asphalt repairs can be found in the Rijkswaterstaat publication
The Use of Asphalt in Hydraulic Engineering (van der Velde et al. 1985).
Methods of maintenance and repair for other types of asphaltic mixtures and structures
can be found in The Shell Bitumen Hydraulic Engineering Handbook (Schönian 1999).

Toe structures management manual


108
5 Toe structure design

Chapter 5 describes the design process with specific attention paid


to how particular issues surrounding the design of the toe to coastal
defences should be considered within it. New structures or
significant additions are examined and awareness is given to the
potential wider environmental impacts of any new toe structure.
Finally, construction issues such as the practicality of timing of
operations during spring tides are discussed.

Key links to other chapters:


• Chapter 2 – Toe structure types and materials

Who will be interested in this chapter?


• Contractors
• Coastal engineers

5.1 Introduction
Chapters 3 and 4 cover the management, monitoring and maintenance of the toe in the
context of gaining a detailed appreciation of the environment in which the toe has to
function. Chapter 3 identifies where intervention is necessary to continue to provide the
required standard of defence. This chapter guides the user through the situation where
the necessary degree of intervention is greater than that which can be carried out
under normal maintenance (Chapter 4). It covers a range of measures from the
implementation of significant repairs to designing the toe of a new sea defence.
Having established that there is a need for work, the design should progress through a
logical sequence which is likely to involve:
• identification of the problem;
• project appraisal – appraisal of options from technical, economic and
environmental points of view;
• design.
This manual does not give specific guidance on approvals, although guidance on
environmental aspects is introduced in Section 5.3.4. In addition, when planning works
to the toe the coastal manager should
• establish whether or not planning approval is necessary by consulting the
local planning authority;
• contact the Marine Management Organisation as to the procedure for
obtaining a marine licence, which will be required for works below high
water.

Toe structures management manual


109
5.2 Identification of the problem
Chapters 3 and 4 guide the user in the process of problem identification. These will
support the identification of the possible threats to the defence standard which might
include:
• flooding due to heavy overtopping;
• defence failure due to excessive overtopping;
• collapse of the defence;
• destabilisation (collapse) of the toe structure;
• undermining of the toe structure or geotechnical failure arising from low
beach levels, leading to failure of the defence.
Most of these threats relate to a threshold condition being exceeded (for example,
wave height). Moreover, the conditions that cause structural damage – be they impact
or overtopping and so on – will also generally lead to depressed beach levels (albeit
temporarily). Under these conditions larger waves can reach the defence. These
conditions are likely to worsen through climate change and, possibly, longer term
beach lowering. Hence, the probability of failure tends to increase with time.
In the case of geotechnical failure (undermining), the principal consideration is the
long-term beach lowering (or perhaps an initial inadequacy of the defence). Transient
loadings such as those due to storm action may be secondary, except in so far as they
lead to short-term beach lowering and possibly an increase in hydrostatic head across
the defence structure. Figure 5.1 shows an example of toe failure.

Figure 5.1 Toe failure example


It is important to identify all the possible modes of failure. The quantification of the
risks, whether they are expressed as a storm return period or as a residual life, will
usually be defined in the Project Appraisal or Strategy as appropriate.

Toe structures management manual


110
A reasonable understanding of the timing and extent of the problem is an essential part
of the process needed to define the scope of the subsequent design and to provide a
first indication of the timing of future intervention.

5.3 Project appraisal

5.3.1 Purpose and scope


A Project Appraisal Report (PAR) is carried out when a project, or the problem, has
been identified. Apart from identification of the problem, as outlined above, the PAR
may well have been preceded by a Strategy Study or plan covering a greater length of
coastline. In essence, a PAR is a feasibility report that sets out the technical,
environmental and economic arguments for investment in a given specific project.
In England and Wales, definitive advice on the preparation of PARs for flood and
coastal defence is presented in the Flood and Coastal Erosion Risk Management
Appraisal Guidance (FCERM-AG; Environment Agency 2010) together with supporting
documents. Note that several Welsh local authorities still use the predecessor to this
document, the Defra Project Appraisal Guidance note (PAGN).
The guidance given by the Environment Agency and Defra is aimed at flood and
coastal defence projects, in particular, those seeking central government funding. It
nevertheless provides an excellent basis for approaching other types of coastal
schemes (for example, for regeneration, marine and navigation purposes0.
The following paragraphs highlight some key points of the appraisal process as
set down in FCERM-AG for work related to flood and coastal defence projects.
The flood and coastal risk management (FCRM) appraisal process is summarised in
the Beach Management Manual (CIRIA 2010a), which states:
‘What all public-funded projects have in common is that they need to be
accountable and to provide a justified use of public money (demonstrating
that the return on investment is higher than the alternatives and at the very
least as high as might be expected from the wider basket of HM-Treasury-
funded projects). It is this requirement to demonstrate accountability for
investment of capital that necessitates a project appraisal. In the context of
a coastal flood or coastal erosion risk management strategy there will be
clear objectives that need to be taken into account within the appraisal
process, which may include, but are not restricted to the following, as listed
in the FCERM-AG (Environment Agency 2010):
• reducing the threat to people and their property from flooding and
coastal erosion;
• delivering the greatest environmental, social and economic benefit
consistent with the UK Government’s sustainable development
principles;
• working with natural processes;
• adapting to future risk and changes (for example, due to climate
change);

Toe structures management manual


111
• working with others to deliver better, more sustainable solutions that
can deliver wider objectives and maximise benefits for people,
businesses and the environment.’

5.3.2 Option appraisal


In the case of known or perceived falling beach levels, selection of the preferred option
in the PAR will require an understanding of future levels. This will require a thorough
understanding of the naturally occurring foreshore changes (see Chapter 3), together
with the influence of existing or proposed structures. Future projections must take
account of climate change as provided by the latest adopted advice, for example that
published by the Environment Agency (2011).

Do Nothing
Having identified and quantified the underlying problem (that is, the risk of failure or of
limited adequacy of the defence), the consequences of adopting a ‘Do Nothing’
approach are evaluated. This will usually entail prediction of defence failure and the
progression of erosion and/or flooding.

Do Something
The understanding of risks and consequences arising from ‘Do Nothing’ provides a
basis for preparing a long list of potentially viable solutions. Options for toe protection
should include schemes that counter the risk in each of two distinct ways:
i. By aiming to restore and maintain satisfactory beach levels. This might
include measures such as beach nourishment, recycling, control
structures and so on.
ii. By restoring and maintaining the stability of the coastal defence structure
against low or lowering beach levels. This might include measures such
as a piled toe, rock apron and so on.
An option might include a combination of different measures (for example, sheet piled
toe plus beach management plus control structures).
In the early stages of the PAR, it is necessary to set a number of primary objectives
and identify any significant constraints. In the first pass it should be possible to
eliminate a number of options on qualitative grounds because they clearly do not
satisfy these overriding criteria (for example, use of a material that is not allowed on the
given frontage).

5.3.3 Option shortlist


A shorter list of options should be appraised quantitatively in terms of the three
important criteria: technical, environmental and economic.
At this more quantitative stage in the appraisal process it is appropriate to introduce the
concept of ‘Standard of Protection’ or SoP. The SoP is expressed as the annual
probability (or equivalent return period) afforded by the defence.

Toe structures management manual


112
The SoP is usually assigned to the defence scheme (present or proposed) protecting a
given risk area as a measure of its flood defence performance. Flooding might be due
to overtopping or to breaching of the defence. Overtopping can be related to the
combined probability of occurrence of waves and sea level. Breaching might also be
related to severe overtopping (causing erosion behind the structure) or be due to some
other failure of the structure including undermining. Depending on the type of defence,
a breach failure of the structure body may also be calculable in probabilistic terms (for
example, exceedance of threshold of significant damage to the armour layer).
For the toe structure, such a probabilistic approach may be more difficult to apply,
failure being more often related to the beach level and hence the likely time to that
condition happening. Fragility curves and limit state design principles may be used in
the design stages of a project, but they would not normally be warranted for a project
appraisal. For appraisal purposes, it would be more usual to apply sensitivity analysis.
By way of example, based on a given rate of beach lowering, it might be predicted that
failure of a given defence would be likely to occur in, say, year 15; sensitivity analysis
could then examine the prospect and consequences of failure in years 10 and 20.

5.3.4 Environmental assessment


The weighting that the environmental impact of a proposal has on the appraisal
process is highly dependent on the status of the location in question and on local and
national planning policies.
Where key legislation – such as the appropriate national law relating to the Habitats
Directive 4 and designations such as special protection areas (SPAs), special area of
conservation (SACs) and Ramsar sites – are likely to apply to a site, it is vital that the
risk of having to prevent or compensate for damage to these designations is
incorporated into the project appraisal.
For example, a scheme that involves removing a designated wetland is unlikely to be
acceptable (unless covered by a Coastal Habitat Management Plan (CHaMP) which
has deemed otherwise) and the risk of having to protect the wetland should be
incorporated into the option in the appraisal process as early as possible.
Hence an environmental appraisal should be carried out in parallel with the economic
appraisal from strategy level down to detailed design. At the strategic level, this will
involve a strategic environmental assessment (SEA). At project (scheme) level, in
certain circumstances the environmental appraisal must be in the form of an
environmental impact assessment (EIA), which includes preparation of an
environmental statement.
The presence of an internationally designated site is also likely to require a habitats
regulations assessment (HRA). Guidance related to FCRM including sustainability,
biodiversity, heritage and landscape considerations can be found in FCERM-AG
(Environment Agency 2010).

5.3.5 Economic appraisal


Economic appraisal entails the evaluation of scheme costs and benefits (tangible and
intangible), the benefits being the value of damages avoided by the scheme over those
that would otherwise ensue in the Do Nothing case (and in the Do Minimum case for

4
Council Directive 92/43/EEC on the conservation of natural habitats and of wild fauna and flora

Toe structures management manual


113
capital schemes). In FCRM, future costs and benefits are discounted to present day
terms using discount factors advised in The Green Book (HM Treasury 2003).
As the toe of coastal defence structures is often buried or submerged, routine
maintenance can be problematic. Repairs and improvements are therefore likely to be
applied infrequently or only as capital works projects as and when a paramount need
arises. The timing of works can have a significant effect on the discounted costs and
hence the economic viability of a given scheme. Hence, the robustness of the
economic argument relies heavily on the assessment of future risk. Arguably, this is
more difficult to assess for the toe than for any other element of the defence structure,
not least because the toe may be concealed and, in the case of a historic structure,
poorly defined and understood. Sensitivity analysis should therefore form an important
part of the assessment.

5.3.6 Selection of preferred scheme


In addition to the formal quantitative evaluation of economics, there is a need to take a
broader view of the issues. For example, there might be overriding factors which, while
not expressed in financial terms, might suggest that a scheme is eliminated or short-
listed for consideration as a preference. This might include non-quantified
environmental enhancement, r the robustness of the perceived merits of a given option,
uncertainties and so on.

5.4 Design principles

5.4.1 Design criteria


There are some 2,935 km of built defences around the coast of England and Wales. A
few kilometres, at most, of new or replacement defences are built each year, that is,
significantly less than 1 per cent of the total stock of built coast defences. It is not
surprising, therefore, that there is a considerable demand for assessing, installing,
maintaining or extending toe structures to existing coastal defences. An important
consideration in many cases is, therefore, incorporation of new toe protection works
with an existing structure.
Figure 5.2 shows the seawall at Goodrington Sands, Paignton. Numerous additions to
the wall since original 19th century construction can be seen, the last being the
concrete/pile toe added by Torbay Council in 2007.
Design criteria can be separated into two main groups:
• those concerned with the functional purpose of the toe (for example;
avoidance of undermining of the coastal defence);
• those that relate to the interaction of the toe with its environment (for
example, heritage).

Toe structures management manual


114
Figure 5.2 Goodrington Sands Seawall, Paignton (courtesy of ENBE)

Functional and performance related criteria


The toe is just one component of a complete coastal defence structure. It should be
designed on the basis of a number of design criteria that will feature to varying degrees
depending on the nature and use of the coastal structure.
Design is aimed at producing a structure that avoids or mitigates problems and failures,
so it follows that much can be learnt from known or recurring shortcomings. CEM Part
VI, Chapter 2 (USACE 2012) provides a comprehensive catalogue of failure modes of
typical coastal structure types. Based on this reference, Table 5.1 summarises those
failure modes connected with the toe or leading edge of the main structure.

Table 5.1 Toe related failure modes of main structure types


Failure mode Main coastal structure type

Rubble Revetment Dike Gravity Sheet pile


mound wall /wall
(breakwater)

Sliding of armour/main
structure into scour   
hole – undermining
Subsidence of armour
blocks into fine  
material – liquefaction

Instability of toe
armour on a hard
substrate in breaking
 
waves

Toe structures management manual


115
Sliding of main
structure due to
geotechnical
 
imbalance

Overturning of main
structure due to
geotechnical
 
imbalance

Slip circle failure    


Foundation settlement     

Note  = referenced in USACE (2012)


 = added in Table 5.1

The failure modes outlined in Table 5.1 can be grouped together to arrive at the
following generic list of functional design criteria:
• to counter the effects of beach lowering and undermining;
• to counter the effect of liquefaction at the toe (this can induce geotechnical
imbalance at a vertical faced structure as well as result in subsidence of
armour at a rubble toe);
• to ensure or restore the geotechnical stability of the whole defence
structure – this objective includes mitigating against the risks of sliding,
overturning, slip circle failure and excessive settlement;
• resistance to wave and current loading including stability of toe armour on a
hard substrate.

Interaction of the toe with its environment


Impact, behaviour and environment related factors, both positive and negative, might
include the design criteria listed below. This list is by no means exhaustive and each
location should be carefully considered on its own merits:
• effects on hydraulic performance of the main structure;
• effects on coastal processes;
• durability – abrasion, corrosion and structural deterioration;
• public safety in construction and operation;
• effects on the natural and built environment;
• heritage and visual aspect;
• amenity - aspects concern both the beach and the structure itself;
• access both to and along the beach (pedestrian, vehicular and boats).

5.4.2 Toe structure types


Chapter 2 outlined a number of different toe structure types. It also pointed out that in
many cases the project entails restoration of an existing defence structure rather than

Toe structures management manual


116
all new construction. Consequently, there are many variants on the types of toe
existing and those that can be applied, utilising a range of techniques, often
conditioned by the need to integrate with existing features. It is neither practical nor
useful to describe the design principles of every structural permutation. Hence, for the
purpose of this exercise a number of toe structures are grouped under generic
headings as shown below:
• Rubble structures type:
− concrete unit or rock revetment
− tipped rock
− rock blankets.
• Mattress type:
− interlocking concrete armour
− gabion mattresses
− rock blankets.
• Concrete (gravity) type:
− concrete apron
− concrete foundation/underpinning
− steps integrated with toe.
• Sheet pile type:
− cut off wall
− sheet pile underpinning.
• Asphaltic construction:
− apron
− grouting for rock or stone.
• Cribwork type:
− timber and concrete cribs containing rocks
− gabion baskets.
Table 5.2 highlights the most relevant criteria applying to each of the generic structure
types. The remaining sections of this chapter outline the design principles in respect of
the key criteria listed in Table 5.2.

Toe structures management manual


117
Table 5.2 Relevance of criteria to toe structure types
Main defence Generic toe structure type
structure
criterion Cribwork Rubble Mattress Concrete Sheet pile Asphaltic

Undermining H H H H H H

Liquefaction at
L H H H H H
the toe
Geotechnical
H M M H H M
stability
Resistance to
wave and
H H H L M H
current
loading

Hydraulic
M H M H H M
performance
Effect on
coastal L M M M M M
processes
Public safety M H H H H H

Natural
L H H H H H
environment

Heritage and
M H H H H H
Visual Impact

Amenity L H M M H H

Note: Relevance: High, Medium, Low

5.5 Undermining

5.5.1 Beach lowering and scour


Fundamentally, there are two mechanisms to consider:
• Lowering of the beach due to coastal processes both in the long term (for
example, sediment loss through longshore transport) and the short term
(for example, storm-induced drawdown)
• Scour, induced by the presence of the coast defence structure itself (for
example reflection from a seawall). Appendix A of this report and
Sutherland et al. (2003) describe the processes of beach lowering in front
of coastal structures and the reader is referred to these for further
information on the processes.
In terms of the design of a new or replacement toe, it might reasonably be asked
whether the toe design itself can influence the extent or depth scour. In this respect the
arguments are similar to those that might be applied to the coastal defence structure as
a whole (for example, reducing reflectivity reduces scour potential). Clearly, however,

Toe structures management manual


118
for these factors to apply to the toe structure, it has to be exposed (or only moderately
covered).
In most cases, the critical condition for scour (that is, when the maximum depth of
scour is reached) is likely to occur at a high water levels when the largest waves can
reach the defence. Hence, unless the toe is quite exposed, it is unlikely to be a major
influence on scour compared with the upper face of the defence structure. Beach levels
can fluctuate widely and change rapidly under storm conditions. Previously covered toe
structures may then become influential in the scouring process. In cases where the toe
is substantially exposed, it should be examined according to the same principles as the
main body of the wall (looking at hydraulic performance – overtopping, wave reflection
and impact loads) and this should be checked during design.

5.5.2 Undermining failure


The term ‘undermining’ is not consistently defined in technical or non-technical
literature. Perhaps this is because other modes of failures are likely to have occurred
before a structure becomes truly undermined; see also USACE (2012). The case study
from Corton, Lowestoft, in Appendix C describes a case of true undermining of a piled
wall (also Figure 5.3). For discussion purposes here, and generally in these guidelines,
‘undermining’ is simply taken to mean the condition at which the beach is below the
bottom of the main defence structure such that further beach lowering would lead to
erosion beneath the structure.

Figure 5.3 Undermining of wartime coastal structures at Kilnsey, Holderness


(courtesy of ENBE)
Some gravity structures can tolerate a measure of undermining without collapse, but
the margin between some undermining and failure can be quite small and difficult to
predict. Apart from exceptional cases where some undercutting has been allowed for in
design, it should be assumed that undermining is not acceptable.
Figure 5.4 illustrates toe failure of a revetment due to undermining:

Toe structures management manual


119
Figure 5.4 Undermining failure of a revetment

5.5.3 Mitigation
Figure 5.5 illustrates a variety of toe structures as applied to revetments armoured with
rock or other proprietary type units.
Toe structures, including those depicted in Figure 5.5, mitigate against undermining of
the superstructure in one of two ways:
• as a static structure, that is, of sufficient depth and inherent stability to
avoid being undermined itself (for example, a stable rubble mound,
concrete toe or sheet piling);
• as a flexible mattress that adapts to the lowering bed level, thus preventing
undermining of the main structure (for example, various flexible mattress
types and asphalt).
A major determinant in the choice between these two fundamental options is the depth
of sediment and the geology at the toe.
Where a defence structure is underlain by rock or by rock beneath a shallow depth of
sediment, then there is an opportunity to found a toe structure on the hard substrate.
Where a stratum with limited resistance (for example, clay or weak rock) is located
within a manageable depth below the mobile deposits, this may provide founding for a
toe (see also the case study from Overstrand, Norfolk, in Appendix C). Episodic beach
lowering would be, or at least has previously been, confined within this limited depth.
However, longer term beach lowering, which might include the erosion of the
underlying soft rock such as clay or marl, needs to be checked.
When considering the underlying geology, factors other than undermining are also
important to the design.
A flexible toe would be unsuitable where:
• it is required to maintain ground level for reason of providing passive
support to the main structure (see Section 5.7);
• it is required to provide a rigid support for armour on the main structure face
(for example, single layer armour units), unless the mattress was of such
width and robustness as to eliminate the risk of lowering at its connection
with the main revetment;
• its width, depth or other properties make a compliant toe uneconomic
compared with a more rigid structure;

Toe structures management manual


120
• the apron width presents an unacceptable intrusion into the recreational
beach or causes damage to an area of conservation interest.
A flexible toe can, however, provide a more practical/cost-effective solution for the
avoidance of undermining in some cases such as:
• where the main structure is itself of a flexible type of construction (for
example, riprap slope protection – in some cases, the toe may simply be an
extension of the main revetment);
• at sites where more rigid forms of construction would be impractical;
• at sites subject to large but gradual bed variation/movement.

Toe structures management manual


121
Note: ‘d’s is anticipated scour depth and ‘1:m’ is structure slope.

Figure 5.5 Typical toe details (after McConnell, 1998)

Toe structures management manual


122
5.6 Liquefaction at the toe

5.6.1 Consequences of liquefaction


Wave-induced liquefaction at coastal structures is described in Appendix A.
At sites that are susceptible to this effect, liquefaction presents a problem for securing
the toe structure. Loss of shear strength of the bed material can lead to various
negative effects depending on the type of defence structure.
For rubble structures and shallow concrete toe structures, liquefaction can result in
rock or heavy concrete units sinking into the liquefied bed material – illustrated in
Figure 5.6.

Figure 5.6 Sinking of armour units into liquefied seabed


For sheet piling and deep vertical concrete toes, the loss of shear strength on the
seaward side of the structure leads to loss of passive resistance provided to the toe,
thus compromising its ability to resist geotechnical loading from the active, landward
side. Further to this, liquefaction can provide conditions whereby material behind the
wall can flow out beneath the structure, resulting in subsidence of the area behind the
wall. The tendency for this depends on the flow path, which will be shorter for certain
types of construction (for example, anchored walls that are less reliant on depth of
penetration than cantilevered walls).
Depending on the nature of the liquefaction, that is, whether it is momentary (and
localised) or residual (and widespread), mattresses and compliant toes might in the
former case offer a measure of protection to the main defence structure, or in the latter
case be susceptible to sinking into the liquefied ground.

5.6.2 Mitigation
Recommendations for dealing with wave-induced liquefaction are limited at the present
time. The process is not readily observed and hence, historically, it has been difficult to
link failure to liquefaction when other destructive mechanisms are also at play. The
following advisory notes are therefore given from a pragmatic perspective rather than
being based on robust scientific evidence:
• Existing defence structures for which a problem has been identified:
− For vertical wall structures where it is believed that fill is being lost
through flow beneath the structure, the situation might be relieved by
installing controlled and filtered drainage paths, thus providing a lesser
path of resistance for outflow of water, while still retaining soil within the
wall. Clearly, this is a more difficult construction operation when applied
to an existing structure than it would be for a new one.

Toe structures management manual


123
− For rubble mound structures where it is evident that the toe rocks/units
are sinking into the bed, mitigation might entail reconstruction of the toe
to install a bedding layer and scour apron in order to lessen the point
pressures of the individual armour units. Mattress or a shallow rock
apron might also be used to inhibit liquefaction at the face of a vertical
wall.
− The success of these mitigations (the latter example in particular)
depends on the type of liquefaction that might occur (that is, momentary
or residual). Better quantification of this can be derived from Sutherland
et al. (2007).
• New defence structures:
− For new structures, there is clearly more opportunity to allow for
liquefaction in the initial design. For example, the design of a piled wall
can include a pile length allowance to counter the passive pressure lost
to liquefaction. This might be preferable to installation of a scour
mattress that requires future maintenance. As with scour, avoidance of
liquefaction problems is likely to be preferable to cure.

5.6.3 Site Investigation


Sutherland et al. (2007) may be used to derive estimates of liquefaction depth and
extent. Predictive analysis such as that referred to above requires the input of various
site-specific factors as described in Appendix B.

5.7 Geotechnical stability


This section makes reference principally to British standards and practice. However, it
should be noted that ‘a new European suite of geotechnical design, testing and
construction documents will in due course largely replace British codes and standards’
(CIRIA 2008).
The European code, EC7-1, may be applied to new projects and stabilisation of
existing structures but it does not deal specifically with the assessment of existing
structures or reuse of existing foundations. Its application to toe protection works may
therefore be limited because many cases are concerned with the risks to existing
structures and/or their incorporation into new works. Nevertheless, new design may
consider the use of EC7-1, in particular where important distinctions may be made
between ‘ultimate limit states’ (states associated with collapse or with other similar
forms of failure) (see Section 3.1) and ’serviceability limit states’ (states beyond which
specified service requirements for a structure or structural member are no longer met).
Geotechnical stability problems often relate primarily to the overall defence rather than
the toe. However, they are often exacerbated or caused by problems at the toe and for
this reason they are considered in this manual.

5.7.1 Sliding and overturning of the main defence structure


These failure modes arise out of an imbalance between the active geotechnical load,
tending to move the structure seawards, and the passive resistance.

Toe structures management manual


124
Sliding failure is confined essentially to monolithic type structures (Figure 5.7) while
overturning failures are associated mainly with vertical or near vertical monolithic or
piled defence structures (Figures 5.8 and 5.9). See also the case study from Lowestoft
South Beach in Appendix C.

Figure 5.7 Sliding failure of gravity wall

Figure 5.8 Overturning failure of gravity wall

Figure 5.9 Rotation failure of sheet piled wall

Gravity walls
The potential for failure through sliding and overturning is of particular concern with
regards to old seawall structures, the following being relevant factors:
• factors of safety used in former times (for example, in the 19th century)
would not necessarily satisfy current standards;
• beach lowering (if prevalent) over many years or decades of service;

Toe structures management manual


125
• possible scour induced by the presence of the vertical wall itself.
The factor of safety for sliding is defined simply as the ratio of the sum of the resisting
forces compared with that of the disturbing force:
Factor of safety (sliding) = (Pp + Fb) (Eqn 5.1)
Pa
where:
Pa = sum of active geotechnical and hydrostatic forces (that is, on the landward side,
tending to push the seawall outwards) (Figure 5.10)
Pp = sum of passive geotechnical and hydrostatic forces (that is, on the seaward side,
tending to resist the active pressure and movement) (Figure 5.10)
Fb = the friction on the base of the wall, also tending to resist the active movement of
the wall (equals the weight of the wall multiplied by the friction coefficient of the wall
base on the ground beneath).

Figure 5.10 Simplified forces on a gravity wall


For overturning failure, the disturbing moment is that derived by taking the moments of
the active forces about, or close to, the seaward edge of the base. The resisting
moment is the sum of moments of the passive forces plus the moment of the wall
weight (Mw), about the same axis. As moments are taken at or close to the horizontal
line of action of friction on the base, the latter does not appear in the simple case.
Hence:
Factor of safety (overturning) = (Pp × passive moment arm + Mw × Lw) (Eqn 5.2)
(Pa x active moment arm)
BS 6349-2:2010 recommends minimum factors of safety of 1.75 for sliding and 1.50 for
overturning (BSI 2010).
The total geotechnical loads (both active and passive) are calculated by summing the
incremental forces attributed to each layer in the soil profiles on the landward and
seaward sides of the wall. The pressures imparted by each layer depend on:
• soil type, for example, fine sand, coarse sand, gravel and so on (hence,
angle of friction of the soil);
• whether cohesive, and hence the inclusion of the effect of cohesion
(negative on active force, positive on passive force);

Toe structures management manual


126
• weight of overburden (weight on the layer including the weight of the layer
above a given level), remembering that overburden may reduce over time
through erosion;
• specific weight of the soil in the layer;
• ground water pressure (allowing for buoyancy, but added on as a
hydrostatic pressure).
Each incremental soil layer has a moment arm about the base. Hence the total moment
(both active and passive) is actually the sum of the increment moments on either side
of the wall (the overall moment arm, as shown simplified in Equation 5.2 is therefore
the average value that would equate to the total moment/total force).
The calculation of active and passive pressures can be found in most standard text
books on geotechnical engineering, and is well described in the Piling Handbook
(Arcelor 2008). This is a comprehensive subject, which is dealt with specifically and at
some length by other texts and, as such, the details are not reiterated here. It should
be mentioned, however, that the example above has been deliberately simplified for
illustration purposes. In practical applications there are numerous possible complicating
factors. Examples include:
• ground surcharges due to point loads, line loads (for example, set back
wall), vehicular loads and so on;
• sloping ground;
• variable beach levels;
• variable sea levels;
• variable ground water levels;
• complex wall structures;
• inclusion of ties, struts or other supplementary supports;
• poorly understood (buried or rear face) seawall geometry;
• poorly understood properties of made ground behind a wall.
In practice, a number of simplifying assumptions have to be made. Experience of this
type of analysis is therefore necessary in order to apply appropriate simplifications.

Piled walls
For sheet piled walls, the analysis for overturning failure (and design) uses the same
basic principles for calculating the active and passive forces as used in the gravity wall
example above. However, there are significant differences in how the two types of
structure are designed or calculated to respond to the applied loads.
Whereas a gravity wall is considered to be rigid and monolithic, a sheet piled wall is
treated as elastic. For sheet pile wall analysis, two possible flexure models may apply:
• Free earth support – in this case the pile is modelled as a simply
supported vertical beam which, is assumed to be free to rotate at its toe.
• Fixed earth support – in this case the pile has greater penetration into the
ground such that the toe end of the pile is considered as fixed (not free to
rotate).

Toe structures management manual


127
In order to achieve fixity in the ground, a wall designed on the fixed earth support
principle is longer than that designed on the free earth support method but it carries a
reduced bending moment. The reduced bending moment means that a lesser pile
section is required, or might have been required in the retrospective case of an existing
structure.
Practical factors may indicate that one method might preferably be used in favour of
another, for example:
• Limited driving depth due to the proximity of a rock head might suggest that
the free earth support method be used.
• Ready availability of a low modulus section (lower bending capacity) but in
ready-cut long lengths (for example, recycled piling) might be suitable for
reuse if designed according to the fixed earth principle.
The free earth support method of necessity requires that the pile be propped or
anchored at or near the top. A true cantilever wall must, therefore, be designed on fixed
earth principles. These and related factors are important considerations to both the
designer and the analyst.
The integrity of a sheet pile wall depends on two basic conditions being satisfied:
• sufficient passive moment to resist the active moment for the chosen
flexure model;
• sufficient pile section capacity (known as section modulus) to sustain the
pile bending moment within the limits of permissible stress and deflection;
note that future loss of section through corrosion must be allowed for (see
Section 4.3.4).
This detailed subject is comprehensively covered in the Piling Handbook (Arcelor 2008)
to which the reader is referred. This reference includes several worked examples
including cases for cantilevered and tied walls, using both free earth and fixed earth
support principles. The Piling Handbook also contains design charts for simple cases
that can be used for initial estimation or concept design purposes. These should be
used with caution as there are many factors that can differ from ‘the norm’.
As with the gravity wall case outlined above, there are numerous complicating factors
involved in practical design. Experience is needed in the application of these design
methods. While there are several software packages available for the analysis of these
situations, it is important that they are applied by experienced practitioners in order to
appreciate their limitations and produce realistic results.
The position of the toe of the pile in relation to both the landward (active) side and
seaward (passive) side is a fundamental determinant in sheet pile wall design.
Whereas land levels will normally remain much the same with the passage of time,
clearly beach levels can vary both in the short and long term. Beach lowering would
tend to impair the load resisting capacity of an older structure.

Sliding and overturning – mitigation


The previous section outlined the crucial factors that determine the stability of gravity
and sheet pile defences as a result of geotechnical imbalance across the structure
section (for example, due to beach lowering). Whether designing a new structure or
restoring an older one, there are two possible fundamental approaches:
• retain/restore a sufficient beach level at the defence structure;

Toe structures management manual


128
• accepting that the desired beach level cannot be achieved and installing a
toe structure to support the main defence structure.

Maintaining beach level


Beach restoration might entail recharge, recycling and/or other beach management
measures to regain and/or retain the required beach level. Local scour, induced
principally by the seawall itself, might be mitigated using a scour apron, mattress or
similar. While these toe systems might prevent local scour, they will not prevent more
widespread beach lowering (for example, through long-term sediment loss), although
they may be designed to accommodate a measure of beach lowering in terms of their
own survival. See also Sections 5.5 and 5.6.
Where it is required to provide support to a deep structure such as a piled wall or a
deep founded gravity wall, the apron would need to be of sufficient width to avoid the
influence of reduced passive resistance resulting from scour occurring beyond the
mattress itself. The width of the mattress, W, to achieve this condition is given by CEM
(USACE 2012) for cohesionless beach deposits as follows:
W = de / tan(45 - Ø/2) or approximately 2de (Eqn 5.3)
where:
Ø = the angle of internal friction of the beach sediment

de = the depth of penetration of the structure on the beach side (allowing for beach
lowering as distinct from scour).
USACE (1995) also recommends that:
• for toe structures at sheet pile walls, W = not less than 2Hi (Eqn 5.4)
• for toe structures at gravity walls, W = not less than Hi (Eqn 5.5)
and
W = not less than 40 per cent of the water depth at the structure (Eqn 5.6)
where:
Hi is the incident wave height.
Comparison of the inequalities of Equations 5.4–5.6 with Equation 5.3 suggests that
the latter condition, relating to depth of penetration, will generally be the more onerous
in cases where it is required to provide geotechnical protection to a deep piled
structure. It also follows that the width W could, in this case, be rather large. For
example for piling with depth of penetration de = 6m and angle of friction of 30°, W
becomes 10.4. or 12 m using the approximation.
Thus, depending on the circumstances, it may be possible for geotechnical problems
due to scour to be pre-empted by the installation of a mattress or shallow blanket. The
last example indicates that there may be practical difficulties in providing effective
protection by these methods (for example due to size). Moreover, if a problem of
geotechnical instability is already present or is likely to arise due to general beach
lowering, then lightweight measures such as this are unlikely to work.

Toe structures management manual


129
Structural mitigation
The designer must carefully consider the circumstances and requirements of each
specific project.
A more engineered structure will probably be required where additional toe support is
required to preserve or reinstate the resisting forces on a superstructure and this
cannot be achieved by simply mitigating scour, and beach management is not feasible.
Options to achieve this may take the form of:
• a deeper toe to the superstructure;
• mass concrete blocks at the toe;
• a steel sheet piled toe;
• a combination of the last two measures;
• rock placed against the main structure wall.
These toe structures will be subject to the same threats as described elsewhere for the
main defence superstructure (that is, beach lowering, scour, liquefaction and so on)
and must be designed for accordingly. Important aspects of design are detailed below.
Deeper toe
The most fundamental option, albeit not without practical constraints (including
construction) in many cases, is simply to extend the toe to sufficient depth to avoid
geotechnical instability. However, this is a measure that principally applies to the
design of new structures. For existing structures, extending the toe to provide
additional passive support is not necessarily straightforward:
• Underpinning of gravity structures can provide protection against
undermining, but there may be difficulties in attaching sufficient mass to
effect satisfactory geotechnical support.
• Lengthening piles could be achieved by attaching new pile length to the top
and re-driving, but this will generally be complicated or impractical due to
the presence of pile caps, walings, tendons and so on within the existing
structure.
Mass concrete block
A concrete mass added to the wall may be designed as:
• an addition to a gravity wall, having the effect of improving both the sliding
resistance and overturning restoring moment of the wall itself;
• an independent mass that adds to the sliding resistance of a monolithic wall
and, in the case of a deep wall, provides a fixed overburden pressure and
hence improved passive resistance.
Where the main wall is of sheet piling, a concrete block may similarly be used to
provide friction (thrust) resistance and to improve passive resistance at the pile through
increased overburden.
Steel sheet piled toe
In essence this method entails construction of a new wall to seaward of the existing
structure. As such it will be subject to the same design principles as a main wall
constructed in steel sheet piling. The distance between the walls will determine the
interrelation of load transfer between the two – that is, the closer they are the greater

Toe structures management manual


130
the transferred active loading will be to the new wall. The surface gap between the
existing wall and the new wall/toe piling will need to be armoured in some way; this
may take the form of a slab possibly coupling up as a pile cap for the new construction
(see also below).
Combination of concrete mass and sheet piled toe
A concrete mass placed between an existing wall and a new sheet piled wall/toe
structure can also provide part of the resistance to sliding and overturning of the main
wall. If the beach level is already close to the base level of the existing wall, it may be
undesirable to build the new wall to a much higher and, therefore, obtrusive level. In
this case, a concrete block can act both as resistance and as a thrust block to transfer
active load through to the new pile. Note that in this case, the position of the thrust and
its magnitude will have a sensitive effect on the new pile design – see Figures 5.11 and
5.12, and the case study from Lowestoft South Beach in Appendix C.

Figure 5.11 Example of thrust block and piled toe

Toe structures management manual


131
Figure 5.12 Geotechnical support for the concrete seawall at Lowestoft
(courtesy Waveney District Council)
Rock mound
A mound or fillet of rock can be used to provide passive support, both directly to the
existing wall and indirectly by way of a ground surcharge on the passive side. A
properly designed rock mound incorporating appropriate sublayers/geotextile can also
provide scour resistance. Rock used in this way is sometimes applied as an interim
measure pending installation of, or incorporation into, a long-term solution (for
example, a beach recharge and management programme). Where very large rock is
used, boulders should be placed against the existing wall in a manner that provides a
good bearing (and is secure) while avoiding impact or pressure damage to the existing
wall face.
The design of the rock should follow the principles set out in Section 5.4. The design
should, moreover, make due allowance for the limited permeability of the section (that
is, due to proximity of the solid interface with the existing wall). Note that short-
term/interim/emergency measures may well be required and these are likely to be
carried out at minimal cost or using simplified design processes, and this could
translate into a minimal cross-section being used.
It is wise to design for a longer service life than the immediate urgencies might suggest
(for example, 10 years rather than 2 years), as the very act of providing some relief can
have the effect of changing priorities. An example is provided in Figure 5.13.

Figure 5.13 Example of rock used to provide passive support

5.7.2 Slip circle failure


Slip circle failure is a form of slope instability. Where the failure is deep seated, it can
pass beneath a coastal defence structure including a piled wall (Figure 5.14).

Toe structures management manual


132
Figure 5.14 Slip circle failures
The subject of slope stability is covered by most geotechnical text books and is dealt
within the context of coastal structures in CEM (USACE 2012).
Instability can occur with weak soils or when structures are placed over weak soil
strata. Groundwater is a significant factor and, with a coastal structure, this can be
aggravated by tidal action, landward ground being saturated by overtopping, or
drainage problems.
Slope instability and slip circle failure are serious issues, the significance of which must
be assessed at the outline design stage and mitigated through appropriate detailed
design of the whole defence structure. Retrospective correction of slip circle failure is
likely to involve major engineering and probably rebuilding of the whole defence
structure. Pre-emptive measures to reduce the risk of slip circle failure may sometimes
be possible but are likely to be major endeavours, for example surcharging with a deep
and extensive apron, or very substantial beach nourishment. They are also likely to
include measures applied to the main part of the defence (for example, re-grading the
slope).
This subject is not described further in this manual and the reader is advised to refer to
the reference texts for further details of the soil mechanics and analysis.

5.7.3 Foundation failure


Settlement is a function of the defence structure (type, materials, density, load
distribution and so on) and the soils on which it is founded. Soft soils including weak
clay, silty sand and mud are most likely to be problematic, with vertical or differential
settlement of heavy structures being possible.
Settlement takes place over varying time periods:
• Instantaneous (applies to high and low permeability soil). This refers to
settlement that occurs during or upon completion of construction.
• Primary (applies to low permeability soil). This relates to the gradual
consolidation of the soil due to the dissipation of excess pore water and
takes place over a long time period – possibly the life of the structure.
• Secondary (applies to high and low permeability soil). This long-term creep
of the soil material is usually less than the combined effect of the
instantaneous and primary settlements.
Although the toe of a given structure is located at or about the foundation level, there is
little scope for mitigating settlement through the toe detail. This is because settlement
relates to weight distribution of the whole structure and not just a small part of it. Where
a structure is to be founded over poor bed materials, the significance and implications
of settlement need to be examined on a structure-wide basis. Situations where

Toe structures management manual


133
particular consideration would need to be given to the toe would, however, include the
following:
• If the toe is of a markedly different form of construction to that of the main
defence (for example, a mass concrete block compared with a shallow
revetment). then differential settlement relative to the main construction
could occur.
• If the toe is required to support the leading edge of a primary armour layer,
especially a single layer armour system placed on a steep slope, then
settlement of the toe should be checked and evaluated.
Further advice on evaluating settlement can be found in CEM (USACE 2012).

5.8 Resistance to wave and current loading


Table 5.1 identified rubble mound and revetment type coastal structures as being
prone to toe damage through wave or current action. This is because they are the
structure types most likely to have toes of similar construction (that is, of rubble). This
perceived vulnerability is also reflected in the related Table 5.2, which identifies toes of
rubble construction, together with mattresses and asphaltic construction, as being most
at risk.
Monolithic concrete and piled wall construction are identified as being at less risk of
damage because they naturally possess a strong degree of internal strength.
Nevertheless, for some applications, the evaluation of wave loads could be significant
(for example, see CEM (USACE 2012) for wave forces on vertical walls) but these
considerations will tend to relate to the larger superstructure. This topic is, therefore,
not dealt with in further detail here.

5.8.1 Rubble mound toe structures


The key advantages of a toe structure consisting of rock are:
• flexibility, that is, the potential to accommodate changes in beach level, and
scour holes;
• potential to dissipate wave energy, thus reducing wave loads on the toe
and the main coastal defence structure, and also reducing the tendency for
scour.
Selection of the most appropriate design concept will depend on:
• the phase in the life cycle of the coastal defence structure (for example, is it
design of a new structure or an emergency repair?);
• the soil type of the beach or foreshore (rock, gravel, sand, silt, clay and so
on);
• the hydraulic load conditions (waves, currents, tidal levels).
Readers are recommended to consider the design methodology for rock structures as
described in The Rock Manual (CIRIA et al. 2007) as the state-of-the-art methodology
(as of 2012). Typical toe details are also given by McConnell (1998), see Figure 5.5.
When designing a rock toe for stability under wave attack two principal failure
mechanisms may be considered:

Toe structures management manual


134
• displacement of rock;
• loss of bed material through the rock matrix.
The following summarises the design methodology for these most common failure
mechanisms.

Failure mechanism: displacement of rock


For rock slopes under wave attack, formulae for rock stability are treated in detail in
The Rock Manual (CIRIA et al. 2007), being those due principally to Hudson, van der
Meer and van Gent. Empirical formulae developed by van der Meer are also presented
for typical submerged toe structures, where the crest of the toe is well below the trough
of the design wave. These formulae have been developed for the design of rubble
mound breakwaters, but they may be applied appropriately and with caution for
submerged rock toes in front of coastal defence structures as well. Some 65 pages of
The Rock Manual are given over to describing the background, formulations and
limitations of the various equations and methods that might be applied; the reader
wanting to use design formulae is therefore advised to refer to The Rock Manual. The
key features of practical coastal defence toe design are summarised below.
The relevant design parameters are:
• water depth (under the design event) at the rock toe and the elevation of
the toe in relation to design sea level;
• wave conditions (wave height or statistical distribution of wave height,
possibly affected by bathymetry of foreshore, wave period, and wave
direction relative to the shoreline);
• current loading;
• main structure geometry (sloped structure or vertical wall, structure with or
without berm, structure crest height and toe crest height);
• relative density of rock;
• expected or allowable damage level of the rock toe.
The interaction between a rock armoured toe and the incident wave field depends
crucially on the relative depth of the toe, ht/h (Figure 5.15).

Figure 5.15 Relative depth of toe


If the ht/h ratio is low (<0.5), that is, the toe is relatively close to the surface in relation
to the ambient water depth, then the design tends towards that of an armoured slope
that is partially or substantially submerged. Corrections derived by Vermeer (1986) and

Toe structures management manual


135
van der Meer (1990) may be applied that allow for different slopes in a composite
structure but essentially the rock sizing design follows the general principles derived by
van der Meer (1988), van Gent et al. (2004) and others for rock slopes.
If the ht/h ratio is high (>0.5), then different principles can be applied to justify a smaller
rock size due to the greater water depth between the toe and the position of greatest
wave impact. The methods described in The Rock Manual (CIRIA et al. 2007) relate
the stability number Hs/∆Dn50 to the ratio ht/h where:
Hs = incident significant wave height
∆ = the relative density of the rock = (ρs – ρ)/ρ
Dn50 = characteristic stone size
The rock sizes derived from The Rock Manual methods vary massively with the ratio
ht/h; for example, Hs/∆Dn50 = 6.5 for ht/h = 0.8, while Hs/∆Dn50 = 3.3 for ht/h = 0.5, thus
implying a factor of nearly two-fold on Dn50 or 7.6 on rock mass. Moreover the values
thus derived relate to 0–10 per cent damage. Based on the work of van der Meer, The
Rock Manual develops more sophisticated equations that relate rock size to the
number of units displaced. The Rock Manual also contains relevant formulae for rock
toes in front of vertical structures being derived primarily for vertical wall breakwaters
and large caisson type structures.
The Rock Manual tends to focus on larger breakwater type structures where the toe is
more often in deeper water and other practical considerations (for example, placing)
might favour more optimal rock sizing. However, this manual on the management of
toe structures advises extreme caution in applying the latter approaches (for example,
those that formulate a reduction in rock size with water depth) for general coastal
defence applications, for the following reasons:
• In most situations, the toe rocks will have a much reduced depth of water
covering them at some stage of the tide (thus implying a low ratio of ht/h).
• In practical terms, the toe structures of coastal defence are generally small
compact features; the loss of any rock can seriously affect the integrity of
the structure and can be difficult or relatively expensive to rectify.
• In many cases, the rock forming the toe will simply be an extension of
armour layer comprising the upper revetment, and it may be desirable to
actually increase the toe armour size or take other steps to secure the
vulnerable leading edge (see below).
In cases where the toe structure is founded on a hard substrate, the stability of the
leading edge of the armour is likely to be reduced due to reduced friction, especially
under the action of breaking waves (Figure 5.16). These rocks are more vulnerable to
movement due to the reduced friction resistance at the bed and the absence of the
mutual support of rocks on the seaward side.

Toe structures management manual


136
Figure 5.16 Sliding of toe armour on hard substrate
The following alternative mitigations apply to armour on a hard substrate. The first two
apply equally to armour on a sedimentary substrate:
• use of oversized toe rock to counter the reduced interlock between units
and reduced bed friction;
• excavation of a trench to secure the leading edge of the armour;
• use of piles (if practicable) to form a crib to secure the leading edge of the
armour;
• creation of a concrete toe beam cast onto the rock bed and secured with
dowels;
• anchor bolts to hold the leading armour blocks in place – see CEM (USACE
2012).
For more sheltered sites, the factors listed above would still apply to some degree.
In cases where wave activity is minor, stability considerations may be dominated by the
current. Bed and slope protection for current-induced erosion has been studied over
many years, the earlier significant contributions being attributed to Shields and Izbash.
The Rock Manual (CIRIA et al. 2007) describes the historic research and more recent
developments, in particular: those due to Pilarczyk (1995). In cases where turbulence
may be high (for example, near to culverts or other shore structures), the formulae of
Escarameia and May (1992) are given. Formulae by Maynord (1995) take into account
the thickness of the stone blanket.
Physical model tests may be carried out to optimise a given design, especially in cases
where the structure geometry or load conditions deviate from the valid ranges of the
available design formulae.

Failure mechanism: erosion of seabed material through voids in the rock


layers
Loss of bed material through the voids in rock layers can lead to a lowering of the bed
beneath the rocks with consequent lowering of the rock itself until it becomes
embedded – an effect akin to that of liquefaction. Unless an allowance is made for this,
as for example in the case of a falling toe, it should be prevented. It can also be the
case that rock armour will entrain beach sand, but under storm conditions, this will
usually be washed out and sediment loss through the armour from the bed will ensue
unless prevented. Erosion of bed material through a rock layer can be avoided by
installing granular filters or suitably designed geotextile of appropriate specification
(Figure 5.17).

Toe structures management manual


137
Figure 5.17 Geotextile being incorporated as part of a toe structure
Granular filters may be applied where depth permits. In many cases, coastal defences
require a rock size that is large relative to the depth of the structure profile, with the
result that it simply becomes impractical to install a multi-layered rock system at the toe
without creating an (otherwise) unnecessarily deep excavation to accommodate it. In
addition, construction considerations might point to the use of larger rocks that can be
placed individually, rather than small (filter) gravel sized material that has to be placed
and formed to a specified depth (for example, stability of material subject to wave
action). These and other factors will generally point to the need to incorporate a
geotextile either as part of a layered filter system (including granular layers), or in some
cases as the only separator between the primary armour and the bed.
Whether granular filters or geotextiles are used there are two fundamental criteria to
satisfy for correct functional design. These are filter stability and filter permeability.
Filter stability
This relates to the perceived problem, that is, the prevention of the migration of fine
sediment particles through the filter and hence through the overlying rock voids.
Using conservative principles, granular filters may be designed to be ‘geometrically
tight’, requiring that the pores in the filter are too small to allow the finer bed sediment
grains through. This approach is safe but can be onerous in terms of the grading of the
filter, and possibly requiring two or more layers to achieve the required succession of
grade ratios.
More recent research, which takes into account the actual hydraulic load on the bed
layer, has facilitated the design of ‘geometrically open’ granular filters, which can be
more economical. As, under normal circumstances, a geotextile cannot pass through
the pores of an overlying rock layer, the requirement for filter stability only really applies
to the geotextile/bed interface which can, therefore, be economically designed
according to the geometrically tight principle.

Toe structures management manual


138
Filter permeability
This relates to flow of water through the filter and hence the avoidance of excess pore
pressure.
For granular filters, the criterion can be designed for either by evaluation of the pore
water pressure head or, more commonly, the direct application of safe geometric ratios
for successive gradings between bed, filter and rock layers.
The design of geotextile filters for permeability follows similar principles to those for
granular filters except that greater consideration must be given to the longer term
permeability of the materials and how this might reduce due to clogging, in particular
where the bed material is silty.
The Rock Manual (CIRIA et al. 2007) details the methods and formulae for designing
according to the above principles, together with the related topics of heave and piping.
The manual also gives guidance on application of geotextiles in rock structures and
selection of the appropriate geotextile properties (extensibility, puncture resistance,
thickness and durability).

5.8.2 Rock blanket, concrete armour mattresses and asphaltic


apron
This section outlines the design principles of three types of toe protection system under
the action of wave or current loading:
• rock blanket or apron;
• concrete armour mattress;
• asphaltic apron.

Rock blanket or apron


An apron can limit the effect of scour induced by the main structure but cannot prevent
naturally occurring beach lowering due to sediment starvation, adverse longshore drift
gradient, or cross-shore movement and reprofiling. Rock size and the depth and extent
of the blanket depend on the exposure (for example, open coast or sheltered estuary)
and the likelihood of future beach lowering.
For a rock apron that can be exposed to open sea conditions, wave attack is likely to
be the dominant design factor. The design of a blanket in this case would follow the
same principles described for a rock toe (see Section 5.8.1). It is essential that the
depth of the apron allows for:
• depth to accommodate sufficient number of stone diameters, and hence to
be sufficiently energy absorbent, to be stable under the incident wave
action;
• depth to accommodate any allowable localised scour within the apron itself;
• sufficient reserve of material to accommodate bed deformation due to
general beach lowering – to this end the blanket must be designed to
anticipate the most disturbed bed profile plus an allowance for the loss of
some material displaced in the process of lowering;

Toe structures management manual


139
• a margin of safety in the rock size to cater for the fact that, as it adjusts to
accommodate beach lowering, the slope of the toe is likely to become
steeper than that at which it was initially placed.
Figure 5.18 shows an example of a rock apron:

Figure 5.18 Example of rock apron (courtesy Jersey States Government)


The discussion above provides only an overview of this topic. For design purposes, the
reader is advised to refer to the Coastal Engineering Manual (USACE 2012) and The
Rock Manual (CIRIA et al. 2007).

Concrete armour mattress (or flexible armoured revetment)


These systems consist of concrete blocks joined together with cables, or attached to a
geotextile sheet, to form closely spaced pattern of armour (Figure 5.19).
When placed as a revetment, they may be filled with granular material either artificially
or through natural beach action – this improves the interaction between the blocks and
the weight of the mattress but lessens their energy absorbing properties. When placed
as a toe or buried, infill with sediment is certain.
A mattress may be laid on a sublayer of gravel or small rock, or placed over a
geotextile to prevent sediment loss through the interstices.

Toe structures management manual


140
Figure 5.19 Example of a flexible armoured revetment
The stability of the mattress depends on retention of its integrity. Some designs ignore
the effect of the interconnecting tendons, taking the conservative approach that the
stability of the system, ultimately, is governed by that of the individual armour units. In
any case, any unanchored edges of the mattress are clearly more vulnerable to
displacement than the inner area, and this is likely to be the case at the toe.
Some key advantages of these systems are as follows:
• They can be placed quickly – advantageous when working within brief tidal
windows (if appropriate they can be lifted and reused elsewhere).
• They can accommodate a measure of deformation (for example, due to bed
lowering) and still retain their integrity. Consequently, in certain cases they
can provide an economic alternative to a rock apron where the latter would
have to be of a greater volume to cater for bed lowering and hence some
loss of material.
• If they are only required temporarily (for example, if they are subsequently
covered permanently as a result of beach nourishment), they can be lifted
out and re-laid elsewhere.
A general description of flexible armour systems may be found in Seawall Design
(Thomas and Hall 1992). An important aspect of design concerns the permeability of
the bed or embankment over which a flexible armoured system is placed. Generally,
the mattresses are of low permeability and this may be further impaired by infilling with
sand and so on.
If the ground over which mattress is placed is clay or silty sand (that is, of lower
permeability than the mattress), then water pressure below the mattress can dissipate
through it. If, however, the mattress is placed over a more permeable sediment or a
granular layer, then pressure on the mattress may cause it to lift under wave action.
Lift, as just described, requires there to be a pressure differential due to wave action,
as would be the case on the sloping part of a revetment. The effect would be much
reduced at the toe if submerged throughout the wave passage, although piping through
the underlayers or embankment could pose a similar threat.

Toe structures management manual


141
Design rules in respect of wave loads are described by Pilarczyk (1995). The basic
equation for critical stability of a semi-permeable cover layer is:
Hs = F (Eqn 5.8)
∆D ζ2/3
where:
Hs = incident significant wave height
∆ = relative density of the mattress material = (ρs – ρ)/ ρ
D = depth of mattress
F = stability coefficient – depends on system type
ζ = breaker parameter = tanα / (Hs/1.56Tp2)
Tp = peak wave period
α = ground or structure slope
However, this formula does not apply well to toe structures as it assumes there to be a
realistic structure slope α, whereas a scour mattress is laid on a near horizontal bed.
Pilarczyk (1995) advises the use of formulae for slope protection in which the toe is
schematised by way of a mild slope of between 1:8 and 1:10. Design must also allow
for some displacement of the mattress and the possibility that a more tangible slope
develops as a result. The stability coefficient F must be obtained from the manufacturer
of the system.

Asphaltic apron
An overview of various asphaltic materials for use in coastal defence is given in
Chapter 2. For toe construction, asphaltic mastic or grouted rock is usually used.
Like a flexible armoured system, an asphaltic apron can accommodate a measure of
deformation due to bed lowering. At coastal sites subject to dynamic beach behaviour,
asphaltic materials may not be appropriate as rapid beach drawdown may exceed the
rate at which the asphalt will deform without breaking. If appropriately applied,
however, asphaltic materials can provide an economic alternative to a rock.
For design of an asphaltic toe, the reader is referred to:
• The Shell Bitumen Hydraulic Engineering Handbook (Schönian 1999);
• The Use of Asphalt in Hydraulic Engineering (van der Velde et al. 1985).
Figure 5.20 shows an early (1979) application of bitumen to a large coastal structure in
the UK, which was initially constructed with a falling apron as shown in the diagram.
Following beach lowering, it was discovered that there were various buried obstacles
and debris which fouled the apron. To counter this, subsequent design (1987–1988)
adopted a more robust design which used a 6 m long horizontal toe slab of grouted
stone on a geotextile and a seawards sheet pile cut-off with some loose rock placed in
front of it.

Toe structures management manual


142
Figure 5.20 Caister on Sea, UK 1979: cross-section phase I with open stone
asphalt revetment and a sand mastic asphalt toe slab (from Schönian 1999)

5.9 Hydraulic performance


Hydraulic performance refers to the reflective properties of a wall together with run up
and overtopping. If the toe is covered by beach material, changes to its structure
negligible impact on hydraulic performance. If the toe is uncovered under normal or
extreme conditions then there could be some impact on hydraulic response which will
depend, obviously, on the nature of the toe construction.
The nature of the impact will depend on the physical extent of any additions or changes
to the toe.
For structural additions that are physically small in relation to the incident waves (for
example, small rock fillet or sheet pile set just forward of the existing wall), the changed
hydraulics will depend on the properties of the toe itself; for example, if energy
absorbing, then reduced reflection and run up would likely result.
For structures that are physically large in relation to the incident waves (for example, a
wide concrete platform), then the effect that this has on wave characteristics could be
significant and needs to be considered in design – in addition to the properties of the
toe structure itself. The effect on overtopping of locally shallower water at a seawall
may be calculated using the methods described in the European Overtopping Manual
(Pullen et al. 2007).

5.10 Effects on coastal processes


The term ‘coastal processes’ includes:
• the movement of sediment along the shore (longshore transport);
• the movement of sediment across the beach profile (cross-shore transport).
These movements can occur over a short timescale (for example, one tide or storm –
especially the cross-shore movement) or develop over many years of accumulated
change (especially the longshore movement). Gradients in longshore transport give
rise to patterns of erosion or accretion while cross-shore transport during a storm can
result in significant drawdown of the beach (perhaps greater than two metres’
lowering).

Toe structures management manual


143
As with hydraulic performance, the impacts of new toe construction on coastal
processes depend on the extent to which the toe is exposed under normal or extreme
weather conditions. Generally, for toe protection works to linear defences, the effect of
the toe on coastal processes will tend to be marginal and usually confined to cross-
shore impacts. The introduction of an energy absorbing toe will tend to encourage
beach retention while more reflective structures may well aggravate the scour.

5.11 Public safety


According to the HSE (2007), The Construction (Design and Management) Regulations
2007 (CDM2007):
‘are intended to focus attention on planning and management throughout
construction projects, from design concept onwards. The aim is for health
and safety considerations to be treated as an essential, but normal part of a
project’s development – not as an afterthought or bolt-on extra’.
Designers have a duty to avoid foreseeable risks ‘so far as reasonably practicable’,
taking due consideration of other relevant design considerations. The responsibilities of
designers extend beyond the construction phase of a project. Designers also need to
consider the health and safety of those who will maintain, repair, clean, refurbish and
eventually remove or demolish all or part of a structure as well as the health and safety
of users of workplaces.
Every project must be considered in terms of its specific circumstances. The following
identifies some generic design issues:
• ‘Softer’ defences have the potential to fail under very extreme conditions.
The damage caused can create a major danger to construction or
maintenance workers, and to the public.
• Construction will often entail exposing the existing wall down to its base or
thereabouts. This could pose a threat of collapse of the existing wall in the
period between it being excavated and the new toe being backfilled. The
design and specification of the construction run lengths must be carefully
selected with a view securing the stability of the defence at all times.
Generally, construction should only be undertaken in short runs. Similarly,
the depth of the toe and hence excavation must also be checked in terms
of the stability of the existing wall. The time during which the wall remains
exposed should be kept to a minimum.
• Further to the matter of stability of the main wall structure, there is a safety
risk during construction, concerning exposure to intertidal conditions. The
structure is vulnerable to collapse at this point, particularly during high tides
aggravated by wave action. The risks can be reduced by excluding the
public from the working area and by utilising simple design that can be
constructed in short runs, thus enabling construction to progress
incrementally within short time windows.
• Simplicity of design (including avoiding concealed complications) will
eventually also facilitate ease in demolition of the structures, thus reducing
the time taken and the risks associated with this process.
In terms of public safety during service, the following general points are highlighted –
note that this list is not exhaustive:

Toe structures management manual


144
• The introduction of hard engineering structures may introduce slips, trips
and fall hazards. Detailed design should carefully consider this and
endeavour to reduce these risks by maintaining these features at a low
level such that, under fair weather conditions, they should normally be
buried. In addition, the design should endeavour to maintain the continuity
of level(s) of the concrete aprons and so on, thus avoiding trip edges.
• The public are unlikely to be aware of rock or other hard structures installed
below normal beach level. These can present a hazard when the beach is
low and, in particular, when it reaches the critical level at which the
structures are just subsurface – thus presenting a trip hazard while not
being visible, or of presenting zones of unsafe ground where beach
materials bridge over underlying voids in the toe. Mitigation might take the
form of signs to warn the public of buried obstacles and/or of emergency
response preparedness. Local emergency services would need the
capability of lifting large rock in short timescales.
• Installation of a hard apron at the toe of a vertical wall increases the risk of
injury in the case of a fall compared with a soft beach landing. Again,
appropriate signage and possibly improved hand-railing would be desirable.

5.12 Natural environment

5.12.1 Introduction
The construction of toe structures in the marine or estuarine environment creates an
artificial rocky or hard substrate within that environment. On a soft muddy or sandy
shore, this is likely to represent the introduction of a habitat that is currently absent. On
a rocky shore, it may represent the introduction of a substrate with different chemical
properties compared with the natural rocks. While this may enhance local biodiversity,
it may result in some loss, damage or change of existing habitats and species. It must
be remembered that many coastal areas are covered by statutory nature conservation
designations such as Special Protection Areas (SPAs), 5 Special Areas of Conservation
(SACs) 6 and Sites of Special Scientific Interest (SSSIs), 7 and the introduction of a new
substrate within a designated site may not accord with the site’s conservation
objectives.

5.12.2 Enhancement
Toe structures can offer opportunities for environmental enhancement including, for
example, the provision of habitats for marine life. Guidance is given by Jensen et al.
(1998), who discuss habitat creation, present suggestions to encourage colonisation of
species naturally attracted to hard surfaces, and identify the types that may be
attracted.

5
Established under Directive 2009/147/EC of the European Parliament, and of the Council, on
the conservation of wild birds (the codified version of Council Directive 79/409/EEC as
amended) (‘the Birds Directive’).
6
Established under Council Directive 92/43/EEC on the conservation of natural habitats and
wild fauna and flora (the ‘Habitats Directive’).
7
Designated under the Wildlife and Countryside Act 1981, as amended.

Toe structures management manual


145
The colonisation of plants on toe structures is dependent on the plants’ ability to
survive at various levels of hydration (normally linked to height above low water) or light
penetration when underwater or buried by beach material. Structures at, or below, high
water level may be colonised by seaweed, and the density and range of seaweeds is
likely to increase at lower levels and attract a wide range of animal communities for
shelter or feeding.
Under normal conditions, surfaces of concrete or quarried rock structures in the marine
environment are rapidly colonised by naturally occurring micro-organisms that
consume many of the dissolved and suspended substances in water. Settlement of
larger organisms, such as barnacles and mussels, which can directly filter suspended
matter for their food, can also occur. Grazing and browsing organisms living on rock
structures devour many of the plants and animals living on the hard surfaces, creating
a dynamic community with continued re-colonisation.
Fish and crustaceans can use crevices between stones and concrete blocks to avoid
predators, lay eggs, or feed on organisms growing on the structure. If the structure is
permanently submerged, shelters for crabs (crevices on the outside of the structure)
and lobsters (galleries within the structure), and shelter for fish species such as
wrasse, lumpsuckers and conger eels can all be incorporated.

5.12.3 Potential loss, damage or change


Sandy or muddy seabeds and foreshores contain a multitude of organisms (worms,
crabs, molluscs and so on), many of which are important to the food chains of
commercially fished species and birds (particularly in the intertidal zone). Specialised
plant communities, such as saltmarshes and drift line communities, also grow above
the level of mean high water neap tides. When a structure is constructed, there is
inevitably a direct loss of habitat and, with it, associated species. In addition, the
presence of a hard structure can modify the local wave and current climate which may
give rise to sediment and, therefore, habitat changes.

5.12.4 Designated sites


Where the toe works may affect a designated nature conservation site, it is necessary
to consult the statutory government’s advisor on the natural environment, that is,
Natural England, Countryside Council for Wales, Scottish Natural Heritage or, in
Northern Ireland, the Council for Nature Conservation and the Countryside (CNCC).

5.13 Heritage and visual impact


Heritage and visual impact considerations may include the appearance of a toe
structure when it becomes exposed. The relevant planning authority should be able to
advise on such issues during the appraisal stage of a project and on the suitability of
design alternatives during the outline design stage (Figure 5.21).

Toe structures management manual


146
Figure 5.21 Heritage interest – the flint-faced seawall at Lowestoft South Beach
The prospect of using rock on the beach receives mixed reactions from different
authorities and councils. In some cases, rock is accepted as part of coastal landscape,
while at others there is a presumption against the use of rock. The use of asphaltic
materials might also be resisted at locations where it has not been used before. The
same applies to the use of other materials. Each project and site is unique and must be
approached in terms of the specific circumstances and with regard to the wishes of
local residents and the planning authority.

5.14 Amenity
Large exposed aprons or other obtrusive toe structures have a negative impact on
amenity by eliminating part of the otherwise available beach area. In some cases,
though, the installation of steps or an apron can improve amenity. If the apron is too
high above average beach levels, however, then there may be demands for hand rails
which are very liable to storm damage and corrosion. The issue is very site- and
project-specific (Figure 5.22).

Figure 5.22 Toe apron proves walking platform – Paignton, Devon (courtesy of
ENBE)

Toe structures management manual


147
5.15 Construction issues
Thomas and Hall (1992) discussed construction and maintenance in the context of both
new defence structures and renovation; substantially, this discussion relates equally to
toe construction works. The following sections highlight the salient points.

5.15.1 Toe working in general


Works to the toe of a sea defence are often subject to tidal influence. Tidal and wave
action makes such works potentially dangerous to personnel. Access to safe ground
can be difficult. They are in a location that is physically aggressive (for example, due to
corrosion by saline water and abrasion by wave-borne material).
While high quality workmanship and materials are often paramount to achieve a
durable result, the achievement of high quality can be uniquely difficult in this
environment. In these circumstances, it is crucial that those planning, managing and
undertaking works at the toe are experienced in tidal and coastal working such that
they are:
• aware of what is and is not achievable;
• capable of specifying and carrying out works in an appropriate manner;
• aware of the hazards;
• use plant and techniques that are suitable for the environment;
• aware of the need to keep continuously informed about forecast wind,
water level, wave and other meteorological conditions, and how to access
and interpret these data;
• able to plan for and manage the inevitable risks to which they and other
beach users will be subject.

5.15.2 Tidal working


Tidal conditions limited the time available for in situ construction work. While the
astronomical tides are predictable before any construction takes place, sea level
increase or decrease due to meteorological factors can only be reliably predicted a day
or so before the event. The long-term and short-term planning of operations must
therefore allow for the limitations on predictions.
The danger in not completing a section in time is that the incomplete works are more
vulnerable to damage by high tides and wave action. The risks of this can be reduced
by limiting incremental construction to short runs and by limiting the number of different
operations that must be undertaken in a tidal window. As described elsewhere, simple
design usually makes for simple and speedy construction.

5.15.3 Seasonal working


The time of year that works are carried out has a major impact on the planning and
logistics of construction operations. Moreover, the summer holiday period, bird
migratory patterns and other seasonal factors can determine when operations may or
may not be carried out.

Toe structures management manual


148
Construction during the summer normally offers a number of positive benefits including:
• longer daylight hours;
• higher healthier beaches, which can either provide easier working
conditions or sometimes make the work more difficult due to the need for a
greater amount of excavation;
• reduced risk of storms and beach drawdown;
However, for seaside resorts, works on the seafront are usually avoided during the
busy summer holiday period. Apart from the disturbance to the beach during toe
protection works, construction also requires the use of noisy plant and can have an
impact on access roads. It is often possible to fence off part of the beach and limit site
traffic, which may go some way to minimising the impacts to users. Construction
windows may also be restricted by ecological issues such as nesting birds.
Construction during the winter months reduces the impacts on tourism and recreational
use connected with seafront use, but clearly introduces a number of limitations in terms
of timing of works due to shorter daylight hours, higher risk of storm damage and so on.
Careful consideration therefore needs to be given to matters such as:
• the effect of low temperature on certain construction materials, in particular
asphalt and in situ concrete;
• contingencies for curtailing sections of work early should weather or other
circumstances change unexpectedly;
• planning of work stages to avoid vulnerable exposed parts of the
construction should access be made difficult between shifts (for example,
this would apply to layered rock revetment where it is preferable to avoid
leaving secondary armour exposed);
• using standard formwork, or expendable formwork that obviates the need
for recovering formwork – thus speeding up operations and reducing the
risk of damage over high water;
• working short sections at a time;
• restrictions imposed by special working hours agreements (to avoid
disturbance to local residents or wildlife).
Should it be necessary to gain access at very low tide to perform toe protection works,
then the contractor may want to take advantage of exceptionally low tides such as
those occurring at the autumn and spring equinoxes.
The depth of beach and hence exposure, access and excavation needed at the toe can
vary massively with the seasons. This may be reflected in major differences between
beach surveys undertaken at different times. The effect of this on contract terms and
price needs to be considered carefully. Available working times can have a major
impact on:
• the construction techniques that are feasible;
• the quality of the finished job;
• the price.
Use of construction expertise in preparing the designs and specification for the works
can thus increase quality and reduce risk, and also identify areas where the temporary
relaxation of constraints needs to be considered.

Toe structures management manual


149
5.15.4 Defence standard
The undertaking of toe restoration necessarily implies disruption to, and possibly part
removal of, an existing defence. Clearly, it is important to avoid any compromise on
existing defence standard during the undertaking of works. This is essential during the
winter when storm surges are more prevalent. Contingency measures need to be
considered and provided if necessary (for example, provision of a stockpile of materials
for emergency backfilling or rock and/or beach material for protection in advance of the
position of the works).

5.15.5 Sequencing construction


Construction work at the toe of a defence probably requires more considered
programming than any other part of the defence structure. Elsewhere in this section it
has already been mentioned that speed and ease of construction are advantageous.
Depending on conditions, the use of toe sheet piling as a cofferdam can extend the tide
free period considerably and equate to considerable time and cost savings. The
sequencing of staged works is therefore important. The following summarises some
relevant factors:
• The designer/contractor needs to consider design limits of each stage of
construction, for instance, upon excavation of the toe of an existing wall.
What is the residual factor of safety prior to it being restored?
• For anchored or tied sheet piling, it is clearly important to determine the
sequence of backfilling and tying to avoid overloading the pile section.
• Incorporation of existing features into a new construction can save time and
may serve the additional purpose of proving support or as expendable
temporary works.
• With reference to ambient weather conditions, the designer/contractor
should specify the striking times for formwork for concrete and bituminous
materials in order that they gain sufficient strength prior to next stage.
• The use of precast units can save time, but this may not be a viable option
for the toe where the detail has to cater for irregular ground conditions.
• Would the use of a cofferdam be feasible and provide advantages?

5.15.6 Access for materials and plant


The characteristics of a site and its environs, and the impact this has on access, differ
greatly from site to site. For example, access through a built-up area may present
certain advantages in terms of distance but have limitations on road widths or weight
limitations 8 and so on, thus limiting the size or types of vehicle that can be used.
Access to a site at the foot of a cliff may present the obvious problem of height
difference, which might be overcome by use of cranes or by taking a longer route along
the beach. Beach access and working time is more often than not limited by the tide.
Moreover, the operation of plant from cliff tops needs to be carefully and conservatively
examined in respect of stability of cliff edges to the loads imposed by the plant – the
same applies to promenades and other elevated working platforms.

8
Access through built-up areas may be limited by environmental health officers.

Toe structures management manual


150
At the construction site, access from an existing seawall crest may facilitate the placing
of concrete using pumping and, within the limits of crane reach, the placing of
armourstone. Access along the beach will almost certainly be needed for certain
operations including excavation and placing of rock delivered to the beach. Tide
permitting, access along the beach might otherwise be impeded by groynes.
Depending on the circumstances, it may be necessary to create an access over or
through a groyne(s), to be reinstated when no longer required.
Bulk materials such as rock may be delivered by sea but this still may require a short
haul delivery by vehicles through a built-up area (if delivered to a port) or by vehicles
on the beach. Suitable restrictions on public access to the beach would therefore need
to be implemented during operations for health and safety reasons, and staff briefed on
any specific health and safety concerns regarding safe operating conditions (for
example, tidal state, weather/sea conditions). 9
Access may also be a determinant in the choice of toe construction material. For
instance, asphaltic materials can be prepared and delivered to sites that might
otherwise be difficult to access with a sheet piling rig.
Access to land from the works in times of difficulty (storms and high water levels) is
essential to mitigate the risk to beach users, workers and plant.

5.15.7 Placing materials


The circumstances of a given site, in particular the level of the toe works in relation to
groundwater and the tide heights, can be an important criterion in the choice of
construction method and material. Materials that require precise placing are less suited
to situations where installation is completely underwater (this would include flexible
armoured mattresses and pattern placed armour), while certain asphaltic materials
cannot be placed under water due to damage to the binder.

5.15.8 As built drawings


For toe construction, it is especially important to survey completed works and prepare
as built drawings. In a situation where further works may well be required in the future,
it is essential that information on the structure is readily available. While in many cases
the time available to prepare as built drawings can be limited by the urgency to backfill
excavations, or by natural covering by the beach, nonetheless there are numerous
examples where future work becomes excessively costly or risky due to lack of
information on the existing structure.

9
For further guidance on construction safety see Cruickshank and Cork (2007).

Toe structures management manual


151
List of abbreviations
ALWC accelerated low water corrosion
AMS asset management system
AONB Area of Outstanding Natural Beauty
BSS Briar skill score
CDM construction, design and management
CEM Coastal Engineering Manual [published by USACE]
CG condition grade
CHaMP Coastal Habitat Management Plan
EIA environmental impact assessment
FCERM-AG Flood and Coastal Erosion Risk Management Appraisal Guidance
[Environment Agency]
FCRM flood and coastal risk management
FDGiA Flood Defence Grant in Aid
FEPA Food and Environmental Protection Act
FOS factor of safety
HRA habitats regulations assessment
HWMOT high marks of ordinary tides
Hs significant wave height nearshore
LAT lowest astronomical tide
LiDAR light detection and ranging
Lm wave length
LWMOT low water marks of ordinary tides
MAFF Ministry of Agriculture, Fisheries and Food
MHW mean high water
MHWS mean high water spring
MLW mean low water
MLWN mean low water neaps
MLWS mean low water spring
MMO Marine Management Organisation
mODN metres above Ordnance Datum Newlyn
OPC ordinary Portland cement
OS Ordnance Survey
PAMS Performance-based Asset Management System

Toe structures management manual


152
PAR Project Appraisal Report
PFA pulverised fuel ash
RTK GPS real-time kinematic global positioning system
RMS root mean square
RMSIE RMS interpretation error
RMSSE root mean square source error
RMSVE root mean square variability error
SAC Special Area of Conservation
SCAPE Soft Cliff and Platform Erosion [model]
SEA strategic environmental assessment
SMP Shoreline Management Plan
SoP Standard of Protection
SPA Special Protection Area
SSSI Site of Special Scientific Interest
TEB traversing erosion beam
Tp peak wave period
tx zero-crossing period
ULS ultimate limit state
USACE US Army Corps of Engineers
UV ultraviolet

Toe structures management manual


153
References
ACI, 2008. Concrete Repair Manual, 3rd ed. Farmington Hills, MI: American Concrete
Institute.
AMOS, C.L., DABORN, G.R., CHRISTIAN, H.A., ATKINSON, A. AND ROBERTSON,
A., 1992. In situ erosion measurements on fine-grained sediments from the Bay of
Fundy. Marine Geology, 108, 175-196.
ANNANDALE, G.W., 1995. Erodibility. Journal of Hydraulic Research, 33 (4), 471-494.
ANNANDALE, G.W., 2006. Scour Technology: Mechanics and Engineering Practice.
New York: McGraw-Hill.
ARCELOR, 2008. Piling Handbook, 8th ed. Esch-sur-Alzette, Luxembourg:
ArcelorMittal.
BEACH EROSION BOARD,1954. Shore protection planning and design. Office of the
Chief of Engineers, Corps of Engineers, Department of the Army, Technical Report No.
4. Washington DC: US Government Printing Office.
BRAY, M.J. AND HOOKE, J.M., 1997. Prediction of soft-cliff retreat with accelerating
sea-level rise. Journal of Coastal Research, 13 (2), 453-467.
BREW, D.S., 2004. Understanding and predicting beach morphological change
processes associated with the erosion of cohesive foreshores – scoping report. Joint
Defra/Environment Agency Flood and Coastal Erosion Risk Management R&D
programme, R&D Technical Report FD1915/TR. London: Defra.
BRUUN, P., 1962. Sea level rise as a cause of shoreline erosion. Journal of the
Waterways and Harbors Division, 88, 117-130.
BSI, 2006. BS EN 1504:2006. Products and systems for the protection and repair of
concrete structures. Definitions, requirements, quality control and evaluation of
conformity. London: British Standards Institution.
BSI, 2008. PAS 55-1:2008. Asset management. Specification for the optimised
management of physical assets. London: British Standards Institution.
BSI, 2010. BS 6349-2:2010 Maritime works. Code of practice for the design of quay
walls, jetties and dolphins. London: British Standards Institution.
BUIJS, F., SIMM, J., WALLIS, M. AND SAYERS, P., 2007. Performance and reliability
of flood and coastal defences. Joint Defra/Environment Agency Flood and Coastal
Erosion Risk Management R&D programme, R&D Technical Report FD2318/TR2.
London: Defra.
CARTER, T.G., LIU L.-F.P. AND MEI, C.C.,1973. Mass transport by waves and
offshore sand bedforms. Journal of Waterway, Port, Coastal, and Ocean Engineering,
99 (WW2), 165-184.
CASSEN, M., ABADIE, S., ARNAUD, G. AND MORICHON, D., 2005. A method based
on electrical conductivity measurement to monitor local depth changes in the surf zone
and in depth soil response to the wave action. In Coastal Engineering 2004,
proceedings of 29th international conference (Lisbon, 2004), edited by J. McKee-Smith,
pp. 2302-2313. Singapore: World Scientific Publishing.
CIRIA, 1986. Seawalls: survey of performance and design practice. Technical Note
125. London: CIRIA.

Toe structures management manual


154
CIRIA, 1992. Old waterfront walls – management, maintenance and rehabilitation. B13.
London: CIRIA.
CIRIA, 2004. Potential use of alternatives to primary aggregates in coastal and river
engineering. C590. London: CIRIA.
CIRIA, 2007. Management of accelerated low water corrosion in steel maritime
structures. C634. London: CIRIA.
CIRIA, 2008. EC7 – implications for UK practice. Eurocode 7 Geotechnical design.
C641. London: CIRIA.
CIRIA, 2010a. Beach management manual, 2nd ed. C685. London: CIRIA.
CIRIA, 2010b. The use of concrete in maritime engineering – a good practice guide.
C674. London: CIRIA.
CIRIA/CUR/CETMEF, 2007. The Rock Manual. The use of rock in hydraulic
engineering (2nd ed.). C683. London: CIRIA.
COOPER, J.A.G. AND PILKEY, O.H., 2004a. Alternatives to the mathematical
modelling of beaches. Journal of Coastal Research, 20 (3), 641-644.
COOPER, J.A.G. AND PILKEY, O.H., 2004b. Sea level rise and shoreline retreat: time
to abandon the Bruun rule. Global Planet Change, 43, 157-171.
CROSSMAN, M. AND SIMM, J. 2004. Manual on the Use of Timber in Coastal and
River Engineering. London: Thomas Telford.
CRUICKSHANK, I. AND Cork, S., 2007. Construction health and safety in coastal and
maritime engineering. London: Thomas Telford.
DEAN, R.G., KRIEBEL, D.L. AND WALTON, T.L., 2002. Cross-shore sediment
transport processes. In Coastal Engineering Manual EM 1110-2-1100, Part III,
Chapter 3. Washington DC: US Army Corps of Engineers.
DEFRA, 2006a. Flood and coastal defence appraisal guidance (FCDPAG3). Economic
appraisal: supplementary note to operating authorities – climate change impacts.
London: Defra.
DEFRA, 2006b. Shoreline management plan guidance. Appendix D: Shoreline
interactions and response. London: Defra.
DE GROOT, M.B., BOLTON, M.D., FORAY, P., MEIJERS, P., PALMER, A.C.,
SANDVEN, R., SAWICKI, A. AND TEH, T.C., 2006a. Physics of liquefaction
phenomena around marine structures. Journal of Waterway, Port, Coastal, and Ocean
Engineering, 132 (4), 227-243.
DE GROOT, M.B., KUDELLA, M., MEIJERS, P. AND OUMERACI, H., 2006b.
Liquefaction phenomena underneath marine gravity structures subjected to wave
loads. Journal of Waterway, Port, Coastal and Ocean Engineering, 132 (4), 325-333.
DE VRIEND, H.J., 2003. On the prediction of aggregated-scale coastal evolution.
Journal of Coastal Research, 19 (4), 757-759.
DE VRIEND, H.J., CAPOBIANCO, M., CHESHER, T., DE SWART, H.E., LATTEUX, B.
AND STIVE, M., 1993. Approaches to long-term modelling of coastal morphology: a
review. Coastal Engineering, 21, 225-269.
DE WIT, P.J.,1995. Liquefaction of cohesive sediments caused by waves, PhD thesis,
Technical University of Delft., The Netherlands.

Toe structures management manual


155
DICKSON, M.E., WALKDEN, M.J.A. AND HALL, J.W., 2007. Systematic impacts of
climate change on an eroding coastal region over the twenty-first century. Climatic
Change, 84 (2), 141-166.
DODD, N., 1998. Numerical model of wave run-up, overtopping and regeneration.
Journal of Waterway, Port, Coastal, and Ocean Engineering, 124 (2), 73-81.
DOUGLAS, B.C. AND CROWELL, M., 2000. Long-term shoreline position prediction
and error propagation. Journal of Coastal Research, 16 (1), 145-152.
ENVIRONMENT AGENCY, 2006. Managing flood risk – Condition Assessment
Manual. Document Reference 166_03_SD01. Bristol: Environment Agency.
ENVIRONMENT AGENCY, 2010. Flood and coastal erosion risk management
appraisal guidance. FCERM-AG. Bristol: Environment Agency.
ENVIRONMENT AGENCY, 2011. Adapting to climate change: advice for flood and
coastal erosion risk management authorities. Bristol: Environment Agency.
ERLINGSSON, U., 1991. A sensor for measuring erosion and deposition. Journal of
Sedimentary Petrology, 61 (4), 620-623.
ESCARAMIA, M. AND MAY, R.P.W., 1992. Channel protection: turbulence
downstream of structures. Report SR 313. Wallingford: HR Wallingford.
ESRIG, M.I. AND KIRBY, R.C., 1977. Implications of gas content for predicting the
stability of submarine slopes. Marine Geotechnology, 2 (1-4), 81-100.
FOOKES, P.G. AND POOLE, A.B. 1981. Some preliminary consideration on the
selection and durability of rock and concrete materials for breakwaters and coastal
protection works. Quarterly Journal of Engineering Geology and Hydrogeology, 14, 97-
128.
FOWLER, J.E., 1992. Scour problems and methods for prediction of maximum scour at
vertical seawalls. Technical Report CERC-92-16. Vicksburg, MS: US Army Corps of
Engineers, Waterways Experiment Station.
GENZ, A.S., FLETCHER, C.H., DUNN, R.A., FRAZER, L.N. AND ROONEY, J.J.,
2007. The predictive accuracy of shoreline change rate methods and alongshore beach
variation on Maui, Hawaii. Journal of Coastal Research, 23 (1), 87-105.
GREEN, C. AND ILLIC, S., 2009. The application of ARGUS video monitoring for
assessing the impact of coastal structures on beaches. In Proceedings of 44th annual
Defra conference on flood and coastal risk management, FCRM09 (Telford, 2009).
GRIGGS, G.B., TAIT, J.F., SCOTT AND PLANT N., 1991. Field observations over 4
years at Monterey Bay, California. In proceedings of Coastal Sediments ’91 (Seattle,
1991), edited by N.C. Kraus, 1871-1885. New York: American Society of Civil
Engineers.
GRIGGS, G.B., TAIT, J.F. AND CORORA, W., 1994. The interactions of seawalls and
beaches: seven years of field monitoring Monterey Bay, California. Shore & Beach, 63
(4), 32-38.
HALCROW, 2011. National Coastal Erosion Risk Mapping – statistics, Environment
Agency, dated 21 January 2011.
HANSON, H., AARNINKHOF, S., CAPOBIANCO, M., JIMÉNEZ, J.A., LARSOM, M.,
NICHOLLS, R.J., PLANT, N.G., SOUTHGATE, H.N., STEETZEL, H.J., STIVE, M.J.F.
AND DE VRIEND, H.J., 2003. Modelling coastal evolution on yearly to decadal
timescales. Journal of Coastal Research, 19 (4), 790-811.

Toe structures management manual


156
HARRIS, J. M., WHITEHOUSE, R. J. S. AND BENSON, T. 2010. The time evolution of
scour around offshore structures. Proceedings of the Institution of Civil Engineers:
Maritime Engineering, 163 (1), 3-17.
HEIBAUM, M. AND TRENTMANN, J., 2010. Partial grouted riprap for enhanced scour
resistance. In Proceedings 5th International Conference on Scour and Erosion, ISCE-5
(San Francisco, 2010), edited by S.E. Burns, S.K. Bhatia, C.M.C. Avila and B. Hunt, 1-
10. Reston, VA: ASCE Publications.
HM TREASURY, 2003. The Green Book: appraisal and evaluation in central
government. London: The Stationary Office.
HOLMAN, R.A. AND SALLENGER, A.H., 1985. Set-up and swash on a natural beach.
Journal of Geophysical Research, 90, 945-953.
HOLMAN, R.A., SALLENGER, A.H. JR., LIPPMANN, T.C. AND HAINES, J.W., 1993.
The application of video image processing to the study of nearshore processes,
Oceanography, 6, 78-85.
HR WALLINGFORD, 2006a. Scour monitor deployment at Blackpool. Technical Note
CBS0726/04. Wallingford: HR Wallingford.
HR WALLINGFORD, 2006b. Assessment of beach lowering and toe scour. Technical
Note CBS0726/03. Wallingford: HR Wallingford.
HR WALLINGFORD, 2006c. Wave-induced liquefaction of sediment in front of coastal
structures. HR Wallingford Technical Note CBS0726/07. Wallingford: HR Wallingford.
HR WALLINGFORD, 2006d. Beach lowering and recovery at Southbourne (2005). HR
Wallingford Technical Note CBS0726/01. Wallingford: HR Wallingford with University of
Southampton
HR WALLINGFORD, 2006e. Integrating scour research into reliability analysis of
coastal structures. HR Wallingford Technical Note CBS0726/05. Wallingford: HR
Wallingford.
HR WALLINGFORD, 2006f. Understanding the lowering of beaches in front of coastal
defence structures. Phase 2 – Improved predictors for wave-induced scour at seawalls.
HR Wallingford Technical Note CBS0726/09. Wallingford: HR Wallingford.
HSE, 2007. Managing health and safety in construction. The Construction (Design and
Management) Regulations 2007. Approved code of practice. Sudbury: HSE Books.
HSU, J.R.C. AND SILVESTER, R., 1989. Model test results of scour along
breakwaters. Journal of Waterway, Port, Coastal, and Ocean Engineering, 115 (1), 66-
85.
HUGHES, S.A., 1992. Estimating wave-induced bottom velocities at vertical wall.
Journal of Waterway, Port, Coastal and Ocean Engineering, 118 (2), 175-192.
HUGHES, S.A. AND FOWLER, J.E., 1991. Wave-induced scour prediction at vertical
walls. In Proceedings of Coastal Sediments ’91 (Seattle, 1991), edited by N.C. Kraus,
1886-1900. New York: American Society of Civil Engineers.
HUGHES, S.A. AND FOWLER, J.E., 1995. Estimating wave-induced kinematics at
sloping structures. Journal of Waterway, Port, Coastal, and Ocean Engineering, 121
(4), 209-215.
IPCC, 2007. Climate Change 2007: The Physical Science Basis – summary for
policymakers. Contribution of Working Group 1 to the Fourth Assessment Report of the

Toe structures management manual


157
Intergovernmental Panel on Climate Change. Geneva: Intergovernmental Panel on
Climate Change.
IZBASH, SV AND KHALDRE, K.Y., 1970. Hydraulics of river channel closure.
Butterworths London
JENSEN, A.C., HAMER, B.A. AND WICKINS, J.F. 1998. Ecological implications for
developing coastal protection structures. In Proceedings ICE conference, Coastlines,
Structures and Breakwaters (London, 1998), edited by N.W.H. Allsop, 70-81. London:
Thomas Telford.
JENG, D.-S., 1998. Wave-induced seabed response in a cross-anisotropic seabed in
front of a breakwater: An analytical solution. Ocean Engineering, 25 (1), 49-67.
KAMPHUIS, J., 1987. Recession rates of glacial till bluffs. Journal of Waterway, Port,
Coastal, and Ocean Engineering, 113 (1), 60-73.
KRAUS, N.C. and McDOUGAL, W.G., 1996. The effects of seawalls on the beach: Part
1 – an updated literature review. Journal of Coastal Research, 12 (3), 691-701.
LARSON, M., CAPOBIANCO, M., JANSEN, H., RÓŻYŃSKI, G., SOUTHGATE, H.N.,
STIVE, M., WIJNBERG, K.M. AND HULSCHER, S., 2003. Analysis and modelling of
field data on coastal morphological evolution over yearly and decadal time scales. Part
1: Background and linear techniques. Journal of Coastal Research, 19 (4), 760-775.
LAWLER, D.M., 1991. A new technique for the automatic monitoring of erosion and
deposition rates. Water Resources Research, 27 (8), 2125-2128.
LEATHERMAN, S.P., 2003. Shoreline change mapping and management along the
U.S. East Coast. Journal of Coastal Research, Special Issue 38, 5-13.
LEE, E.M. AND CLARK, A.R., 2002. Investigation and Management of Soft Rock Cliffs.
London: Thomas Telford.
LIN, M.C., WU, C.T., LU, Y.C. AND LLIANG, N.K., 1986. Effects of short-crested
waves on the scouring around the breakwater. In Coastal Engineering 1998
Proceedings, 20th international conference (Taipei, 1986), edited by B.L. Edge, 2050-
2064. New York: American Society of Civil Engineers.
LONGUET-HIGGINS, M.S., 1953. Mass transport in water waves. Philosophical
Transactions of the Royal Society of London, Series A, 245, 535-581.
LONGUET-HIGGINS, M.S., 1957. The mechanics of the boundary layer near the
bottom in a progressive wave. In Proceedings 6th international conference on coastal
engineering (Gainesville, FL, 1957), edited by J.W. Johnson, 184-193. Richmond, CA:
Council for Wave Research.
LOVELESS, J.H., GRANT, G.T. AND KARLSSON, R.I., 1996. The effect of
groundwater on scour near structures. In Coastal Engineering 1996, proceedings 25th
international conference (Orlando, FL), edited by B.L. Edge, 2152-2165. New York:
American Society of Civil Engineers.
MAENO, S. AND TSUBOTA, U., 2001. Flow out limit of back-filling sand behind
revetment under cyclic loading of water pressure. In Proceedings 29th IAHR World
Congress (Beijing, 2001), Theme E, edited by G. Li, Z. Wang and A. Pettigean, 43-48.
Beijing: Tsinghua University Press.
MARKLE, D.G., 1989. Stability of rubble mound breakwater and jetty-toes: survey of
field experience. Technical Report REMR-CO-01. Vicksburgh, MS:US Army Engineer
Workshop Experiment Station.

Toe structures management manual


158
MAYNORD, S.T., 1995. Corps riprap design for channel protection. In River, Coastal
and Shoreline Protection: Erosion Control using Riprap and Armourstone (ed. C.R.
Thorne, S.R. Abt, F.B. Barends, S.T. Maynord and K.W. Pilarczyk), pp. 41-52.
Chichester: John Wiley & Sons.
McCONNELL, K., 1998. Revetment Systems against Wave Attack: A Design Manual.
London: Thomas Telford.
McDOUGAL, W.G., KRAUS, N.C. AND AJIWIBOWO, H., 1996. The effects of seawalls
on the beach: Part II, numerical modelling of SUPERTANK seawall tests. Journal of
Coastal Research, 12 (3), 702-713.
MOORE, L.J., 2000. Shoreline mapping techniques. Journal of Coastal Research, 16
(1), 111-124.
MORY, M., MICHALLET, H., ABADIE, S., PIEDRA-CUEVA, I., BONJEAN, D., BREUL,
P. AND CASSEN, M., 2004. Observations of momentary liquefaction caused by
breaking waves around a coastal structure. In Coastal Engineering 2004, proceedings
29th international conference (Lisbon, 2004) edited by J. McKee Smith, 4104-4214.
Singapore: World Scientific.
MORY, M., MICHALLET, H., BONJEAN, D., PIEDA-CUEVA, I., BARNOUD, J.M.,
FORAY, P., ABADIE, S. AND BRUEL, P., 2007. A field study of momentary
liquefaction caused by waves around a coastal structure. Journal of Waterway, Port,
Coastal, and Ocean Engineering, 133 (1), 28-38.
MULDER, J.P.M., HOMMES, S. AND HORSTMAN, E.M., 2011. Implementation of
coastal erosion management in the Netherlands. Ocean & Coastal Management, 54
(12), 888-897 [Special issue: Concepts and Science for Coastal Erosion Management
(Conscience)].
NAIRN, R.B., PINCHIN, B.M. AND PHILPOTT, K.L. 1986. Cohesive profile
development model. In Proceedings of the symposium on cohesive shores (Burlington,
Ontario, 1986), edited by M.G. Skafel, 246-261. Ottawa: National Research Council,
Associate Committee for Research on Shoreline Erosion and Sedimentation.
NAIRN, R.B. AND WILLIS, D.H., 2002. Erosion, transport and deposition of cohesive
sediments. In Coastal Engineering Manual EM 1110-2-1100, Part III, Chapter 5.
Washington DC: US Army Corps of Engineers.
NISHIMURA, H., WARANOKE, A. AND HORIKAWA, K., 1978. Scouring at the toe of a
seawall due to tsunamis. In Coastal Engineering 1978, proceedings of 16th
international conference (Hamburg, 1978), pp. 2540-2547, New York: American
Society of Civil Engineers.
O’DONOGHUE, T. AND SUTHERLAND, J., 1999. Random wave kinematics and
coastal structures. Proceedings of Institution of Civil Engineers: Water, Maritime &
Energy, 136, 93-104.
OUMERACI, H., 1994. Scour in front of vertical breakwaters – review of problems. In
Proceedings of International Workshop on Wave Barriers in Deep Waters (Japan,
1994), pp. 281-307. Yokosuka, Japan: Port and Harbour Research Institute.
OWENS, J.S. AND CASE, G.O., 1908. Coast Erosion and Foreshore Protection.
London: St Brides Press.
PEARCE, A.M.C., SUTHERLAND, J., MÜLLER, G., RYCROFT, D. AND
WHITEHOUSE, R.J.S., 2006. Scour at a seawall- field measurements and physical
modelling. In Coastal Engineering 2006, proceedings 30th international conference
(San Diego, 2006), edited by J. McKee Smith, 2378-2390. Singapore: World Scientific.

Toe structures management manual


159
PILARCZYK, K.W., 1995. Simplified unification of stability formulae for revetments
under current and wave attack. In: C.R. Thorne, S.R. Abt, F.B. Barends, S.T. Maynord
and K W Pilarczyk (eds), River, Coastal and Shoreline Protection: Erosion Control
using Riprap and Armourstone. John Wiley & Sons, Chichester.
POWELL, K.A., 1990. Predicting short term profile response for shingle beaches.
Report SR219. Wallingford: HR Wallingford.
POWELL, K.A. AND LOWE, J.P., 1994. The scouring of sediments at the toe of
seawalls. In Proceedings of the Hornafjordur International Coastal Symposium
(Reykjavik, 1994), edited by G. Viggosson, 749-755. Reykjavik: Icelandic Harbour
Authority.
POWELL, K. AND WHITEHOUSE, R.J.S., 1998. The occurrence and prediction of
scour at coastal and estuarine structures. In Proceedings 33rd MAFF Conference of
River and Coastal Engineers (Keele, 1998).
PULLEN, T., ALLSOP, W., BRUCE, T., KORTENHAUS, A., SCHUTTRUMPF, H. AND
VAN DER MEER, J., 2007. Eurotop – wave overtopping of sea defences and related
structures: assessment manual. Available from:
http://www.overtopping-manual.com/eurotop.pdf [Accessed July 2012].
RENDEL PALMER AND TRITTON LIMITED, 1996. History of coastal engineering in
Great Britain. In History and Heritage of Coastal Engineering (ed. N.C. Kraus), pp. 214-
263. New York: American Society of Civil Engineers.
RIDD, P., 1992. A sediment level sensor for erosion and siltation detection. Estuarine,
Coastal and Shelf Science, 35, 353-362.
ROBINSON, C., BALDOCK, T.E., HORN, D.P., GIBBES, B., HUGHES, M.G.,
NIELSEN, P. AND LI, L., 2005. Measurement of groundwater and swash interaction on
a sandy beach. In Coastal Dynamics 2005, proceedings of international
conference(Barcelona, 2005), edited by A. Sanchez-Arcilla, 12 pp. Reston, VA:
American Society of Civil Engineers.
ROYAL HASKONING, BRITISH GEOLOGICAL SURVEY, UNIVERSITY OF SUSSEX
AND NEWCASTLE UNIVERSITY, 2007. Understanding and predicting beach
morphological change associated with the erosion of cohesive shore platforms. Joint
Defra/EA Flood and Coastal Erosion Risk Management R&D Programme, R&D
Technical Report FD1926/TR. London: Defra.
SCHÖNIAN, E., 1999. The Shell Bitumen Hydraulic Engineering Handbook. London:
Thomas Telford.
SHIELDS, A., 1936 . Anwendung der Adenlichkeitsmechanik und der
Turbelenzforschung auf die Geschiebebewegung [Application of similarity principles
and turbulence research to bed-load movement] Mitteilungen der Preussischen
Versuchsanstalt fur Wasserbau und Schiffbau, Berlin, no 26
SILVESTER, R. AND HSU, J.R.C., 1997. Coastal Stabilisation. Advanced Series on
Ocean Engineering, Volume 14. Singapore: World Scientific.
SIMM, J., GOULDBY, B., SAYERS, P., FLIKWEERT, J., WERSCHING, S. AND
BRAMLEY, B., 2008. Representing fragility of flood and coastal defences: getting into
the detail. In Proceedings European conference, Flood Risk Management: Research
into Practice (FLOODrisk 2008), edited by P. Samuels, S. Huntington, W. Allsop and J.
Harrop, 621-631. London: CRC Press.
SLADEN, J.A., D’HOLLANDER, R.S. AND KRAHN, J., 1985. The liquefaction of sands,
a collapse surface approach. Canadian Geotechnical Journal, 22, 564-578.

Toe structures management manual


160
SOUTHGATE, H.N. AND BRAMPTON, A.H., 2001. Coastal morphology modelling: a
guide to model selection and usage. HR Wallingford Report SR 570. Wallingford: HR
Wallingford.
STIVE, M.J.F., 2004. How important is global warming for coastal erosion. Climatic
Change, 64, 27-39.
STRRES, 2007. Guide FABEM 2 Traitement des fissures par calfeutrement – pontage
et protection localisee – creation d’un joint de dilatation. Paris: STRRES.
STRRES, 2007. Guide FABEM 3 Traitement des fissures par injection. Paris:
STRRES.
SUMER, B.M. AND FREDSØE, J., 2002. The Mechanics of Scour in the Marine
Environment. Singapore: World Scientific.
SUNAMARA, Y. AND KRAUS, N.C., 1985. Prediction of average mixing depth of
sediment in the surf zone. Marine Geology, 62, 1-12.
SUTHERLAND, J. AND O’DONOGHUE, T., 1998a. Wave phase shift at coastal
structures. Journal of Waterway, Port, Coastal, and Ocean Engineering, 124 (2), 90-98.
SUTHERLAND, J. AND O’DONOGHUE, T., 1998b. Characteristics of wave reflection
spectra. Journal of Waterway, Port, Coastal, and Ocean Engineering, 124 (6), 303-311.
SUTHERLAND, J., BRAMPTON, A., MOTYKA, G., BLANCO, B. AND WHITEHOUSE,
R., 2003. Beach lowering in front of coastal structures. Research scoping study. Joint
Defra/Environment Agency Flood and Coastal Erosion Risk Management R&D
Programme R&D Technical Report FD1916/TR1. London: Defra.
SUTHERLAND, J., PEET, A.H. AND SOULSBY, R.L., 2004. Evaluating the
performance of morphological models. Coastal Engineering, 51, 917-939.
SUTHERLAND, J., BRAMPTON, A. AND WHITEHOUSE, R., 2006a. Toe scour at
seawalls: monitoring prediction and mitigation. In Proceedings of the 41st Defra Flood
and Coastal Management Conference, pp. 03b.1.1-03b.1.12. London: Defra.
SUTHERLAND, J., OBHRAI, C., WHITEHOUSE, R.J.S. AND PEARCE, A.M.C.,
2006b. Laboratory tests of scour at a seawall. In Proceedings 3rd international
conference on scour and erosion (Gouda, 2006) [CD-ROM]. Amsterdam: A.A.
Balkema.
SUTHERLAND, J., BRAMPTON, A.H., OBHRAI, C., DUNN, S. AND WHITEHOUSE,
R.J.S., 2007. Understanding the lowering of beaches in front of coastal defence
structures, stage 2. Joint Defra/Environment Agency Flood and Coastal Erosion Risk
Management R&D Programme, R&D Technical Report FD1927/TR. London: Defra.
THOMAS, R.S. AND HALL, B., 1992. Seawall design. B12. London: CIRIA.
TRENHAILLE, A.S., 2009. Modelling the erosion of cohesive clay coasts. Coastal
Engineering, 56, 59-72.
USACE, 1995. Evaluation and repair of concrete structures. Engineering and design
manual. Report No 1110-2-2002. Washington DC: US Army Corps of Engineers.
USACE, 2012. Coastal engineering manual. Report No 110-2-1100. Washington DC:
US Army Corps of Engineers.
VAN DE GRAAF, J., 2004. Seawalls and revetments in coastal engineering practice. In
Coastal Engineering 2004, proceedings of 29th international conference (Lisbon,
2004), edited by J. McKee-Smith, 3839-3851. Singapore: World Scientific Publishing.

Toe structures management manual


161
VAN DER MEER, J.W., 1988. Rock slopes and gravel beaches under wave attack.
PhD thesis, Delft University of Technology, Delft. Also Delft Hydraulics publication no.
396.
VAN DER MEER, J.W., 1990. Stability of low-crested and composite structures of rock
under wave attack [in Dutch]. Report no M1983-V, WL. Delft: Delft Hydraulics.
VAN DER VELDE, P.A., EBBENS, E.H. AND VAN HERPEN, J.A., 1985. The use of
asphalt in hydraulic engineering. Rijkswaterstaat Communications 37. The Hague:
Rijkswaterstaat.
VAN GENT, M.R.A., SMALE, A.J. AND KUIPER, C., 2004. Stability of rock slopes with
shallow foreshores. In Coastal Structures 2003, proceedings 4th international coastal
structures conference (Portland, OR, 2003), edited by J.A. Melby, 100-112. Reston,
VA: American Society of Civil Engineers.
VAN RIJN, L.C., 1998. Principles of Coastal Morphology. Blokzijl, The Netherlands:
Aqua Publications.
VAN RIJN, L.C., 2005. Principles of Sedimentation and Erosion Engineering in Rivers,
Estuaries and Coastal Seas. Blokzijl, The Netherlands: Aqua Publications.
VAN RIJN, L.C., WALSTRA, D.J.R., GRASMEIJER, B., SUTHERLAND, J., PAN, S.
AND SIERRA, J.P., 2003. The predictability of cross-shore bed evolution of sandy
beaches at the time scale of storms and seasons using process-based profile models.
Coastal Engineering, 47, 295-327.
VERMEER, A.C.M., 1986. Stability of rubble mound berms and toe constructions.
Report on literature survey and model investigation [in Dutch]. Report no M2006, WL.
Delft: Delft Hydraulics.
VRIJLING, J.K., LEEUWESTEIN, W. AND KUIPER, H. 1995. The maintenance of
hydraulic rock structures. In River, Coastal and Shoreline Protection; Erosion Control
Using Riprap and Armourstone (ed. C.R. Thorne, S.R. Abt, F.B. Barends, S.T.
Maynord and K.W. Pilarczyk), pp. 651-666. Chichester: John Wiley & Sons.
WALKDEN, M.J.A. AND HALL, J.W., 2005. A predictive mesoscale model of the
erosion and profile development of soft rock cliffs. Coastal Engineering, 52, 535-563.
WALKDEN, M.J. AND ROSSINGTON, S.K., 2009. Characterisation and prediction of
large scale, long-term change of coastal geomorphological behaviours: proof of
concept modelling. Environment Agency Project Record – SC060074/PR. Bristol:
Environment Agency.
WHITEHOUSE, R., 1998. Scour at Marine Structures: A Manual for Practical
Applications. London: Thomas Telford.
WHITEHOUSE, R.J.S., HEARN, S., WATERS, C.B. AND J. SUTHERLAND, 2000.
Data report on measurements by HR Wallingford at Teignmouth UK (1998–1999). HR
Wallingford report.
TR105, November 2000.WHITEHOUSE, R., BALSON, P., BEECH, N., BRAMPTON,
A., BLOTT, S., BURNINGHAM, H., COOPER, N., FRENCH, J., GUTHRIE, G.,
HANSON, S., NICHOLLS, R., PEARSON, S., PYE, K., ROSSINGTON, K.,
SUTHERLAND, J. AND WALKDEN, M., 2009. Characterisation and prediction of large-
scale, long-term change of coastal geomorphological behaviours. Science Report
SC060074/SR2. Bristol: Environment Agency.
XIE, S.-L., 1981. Scouring patterns in front of vertical breakwaters and their influence
on the stability of the foundations of the breakwaters. Department of Civil Engineering,
Delft University of Technology.

Toe structures management manual


162
ZEMAN, A.J., 1986. Erodibility of Lake Erie undisturbed tills. In Proceedings of the
symposium on cohesive shores (Burlington, Ontario, 1986), edited by M.G. Skafel,
150-169. Ottawa: National Research Council.

Toe structures management manual


163
Appendix A: Scour processes
This appendix provides guidance on how to assess the lowering of beaches at the toe
of coastal defence structures. It discusses the processes that control beach levels,
including the form of the different processes and their relative importance.

A.1 Introduction
The coastlines of the world are constantly assailed by winds, waves, tides and surges,
which cause coastal sediments to be transported and cliffs, shore platforms and other
rocks to be eroded. As a result, our coastlines are continually changing. Under the
pressure of development, mankind has for centuries attempted to stabilise our dynamic
coasts through the construction of coastal defences (Rendel Palmer and Tritton 1996),
often on naturally eroding coastlines. There is then a complex interaction between the
coastal defence structure and the beach at which it was constructed.
Beach levels in front of coastal defence structures are continually changing, with (in the
UK) a general trend for lowering, rather than accretion. Beach lowering is caused by a
number of processes that take place at a range of different spatial scales and
timescales, and which combine cross-shore and longshore sediment transport. The
individual timescale and form of transport should not be considered in isolation. For
example, storm-induced toe scour may not be a problem if the beach level at a coastal
defence structure is high, so that the scour that does take place can be accommodated
within the design limits of the structure, whereas if the same beach experiences a long-
term beach loss, the storm-induced scour may in the longer term become serious.
The overall performance of a coastal structure therefore depends on morphological
changes over a broad range of scales, as detailed below and illustrated schematically
in terms of the beach profile response in Figure A.1:
• Toe scour (Figure A.1a) – beach levels in front of the coastal defence
structure often dropping and recovering completely during the course of a
single tide. Toe scour occurs over a cross-shore length of a few metres but
may extend considerably further in the longshore direction. Scour at a
coastal seawall or similar coastal defence structure is often referred to as toe
scour because it occurs at the intersection of the beach and the structure,
even though that may be at a point well above the actual toe of the structure
itself.
• Storm response (Figure A.1b) – lasting for a few tides and causing toe
scour, beach lowering and recovery over cross-shore scales of up to a few
hundred metres and rather longer distances in the longshore direction. The
coherence of the longshore response will depend on how long the coastal
defence structure is, and how the nearshore bathymetry, beach profile and
sediment characteristics vary.
• Inter-storm recovery (Figure A.1c) – the beach will respond to the
changing forcing conditions after a storm and variations in beach level can
be observed. Recovery from storm action can take tens of tides to occur
and will affect a similar longshore area to the storm response.
• Seasonal variability (Figure A.1d) – commonly it is observed that beach
levels draw-down more in winter (due to storm-induced erosion) and build
up during summer, leading to a seasonal variation in beach profiles and
hence levels at the toe of a structure.

Toe structures management manual


164
• Inter-annual variability (Figure A.1e) – the wave climate varies from year
to year, altering the nett magnitude (and possibly direction) of longshore
drift and generating erosion or accretion of the beach. The annual wave
climate affects the whole coastline so its effects are felt over the scale of
the sediment cell, say tens of kilometres alongshore and by of the order of
one or two kilometres cross-shore.
• Coastal evolution (Figure A.1f) – changes in the coastal profile are driven
by sea level rise and wave climate, and dominated by longshore transport
controlling coastal evolution. Coastal erosion occurs over longer timescales
and even larger spatial scales than beach changes due to variations in
annual conditions. Coastal erosion is often associated with the lowering of
ground levels caused by the erosion of a rock platform.

Toe structures management manual


165
Figure A.1 Conceptual model of beach profile response in front of a seawall

A.1.1 Time and space scales


In general, the spatial scale of beach changes increases with the timescale, and
longshore sediment transport processes increase in importance compared with cross-

Toe structures management manual


166
shore transport processes as the timescale increases. However, a storm surge event
moving down the North Sea basin from north to south will produce a time-varying
response at each stretch of coastline that it impacts upon; the integrated effect may
extend for tens or hundreds of kilometres along the coast.
Figure A.2 relates to scour and erosion caused by sediment transport. Another
process, wave-induced liquefaction, may also be important in the assessment of the
performance of the toe of coastal defence structures as it can reduce the bearing
capacity of the seabed in front of a structure. Generally liquefaction is associated with
shorter length scales and smaller spatial scales than toe scour and occurs in finer
sediments; it would not be expected to occur in a permeable shingle beach
experiencing wave loading.

Figure A.2 Beach responses to natural forcing, indicating associated length-


scales and timescales (from Sutherland et al. 2007)

A.1.2 Influence of a coastal defence structure


The lowering of beaches and/or shore platforms in front of coastal structures at
different time scales and spatial scales is caused by a number of mechanisms. Some
of these reflect the characteristics and geomorphological processes of the coast where
the structure has been installed. They are associated with the longer time scales and
larger spatial scales illustrated in the upper right quadrant of Figure A.2. The
mechanisms are largely independent of the type of structure, that is, they occur
whether the structure is permeable or impermeable, whether it is steep-faced and
reflects waves, or whether it is more gently sloping and dissipates wave energy.
Other effects though are dependent on the characteristics of the coastal structure. For
the most part, these have primarily localised effects, are short-lived and are reversible
– at least on sandy beaches. They can lead to local beach lowering. In these
circumstances it is necessary to assess and understand the range of beach levels in
front of coastal structures in order for those assets to be managed.

Toe structures management manual


167
It is convenient to develop a conceptual framework and scenario models and tools
based on a breakdown into the following different time and space scales:
• toe scour of the beach sediment over tides and days;
• liquefaction of the beach sediment over seconds to minutes;
• beach variability over weeks and seasons, which includes storm
response, storm recovery, seasonal and inter-annual variations;
• coastal erosion (long-term beach lowering) over years and decades.
There is some overlap in these categories, but it has nevertheless proved to be a
useful classification as there are different processes (see Sections A.2–A.6) and
modelling approaches (Appendix B) that can be broadly associated with these
categories which are defined below.

A.1.3 Definitions

Scour
In the context of toe structures, scour can be defined as:
‘the process of sediment erosion from an area of seabed in response to the
forcing of waves and currents as modified (enhanced) by the presence of a
structure’.
Although this mechanism affects many types of marine structure, scour processes tend
to be very similar. Scour can be caused by the following processes (Whitehouse, 1998;
Sumer and Fredsøe 2002):
• reflection of waves from the coastal structure, leading to increased wave
action in front of the structure;
• wave breaking in front of or over the structure.
Further significant influences can arise from:
• contraction of currents along the front of a breakwater or seawall;
• generation of wave-driven currents by oblique incidence waves.
And where the structure has an end:
• diffraction of waves around the coastal structure;
• formation and shedding of vortices at the heads of coastal structures.
Thus the extent and type of scour process is dependent on:
• the wave climate and water level;
• the beach and nearshore shore profile;
• the design and position of the seawall on the shore profile.

Toe structures management manual


168
Liquefaction
In common usage, the term ‘liquefaction’ refers to the loss of strength in saturated,
cohesionless soils due to the build-up of pore water pressures during dynamic loading
leading to a loss of effective stress. The definition of liquefaction given by Sladen et al.
(1985) as is follows:
‘Liquefaction is a phenomenon wherein a mass of soil loses a large
percentage of its shear resistance, when monotonic, cyclic or shock loading
is applied, and flows in a manner resembling a liquid until the shear
stresses acting on the mass are as low as the reduced shear resistance.’

Beach variability
Variations in the patterns and rates of sediment transport are very common when
sediments are susceptible to erosion, for example when either fine (sand) or coarse
materials (shingle) are subject to wave and/or current action. These processes may
lead to natural cycles of erosion and accretion irrespective of the existence, position or
configuration of a coastal defence structure.
Processes such as the drawing down of beach material from the top of a beach during
a storm and the gradual recovery of the beach level after the peak of the storm lead to
bathymetries in front of coastal defence structures (and in natural undefended
beaches) that vary in space and time. This phenomenon leads to beach variability.
Figure A.3 illustrates this phenomenon where a storm event removed sediment locally
to a level below the toe of the defence (right image); the subsequent image (on the left)
which was taken after a calm period of weather over several months shows sediment
levels have recovered considerably once more (without active intervention).

Figure A.3 Storm event sediment scour (right) and post event recovery (left)
(courtesy of Peter Frew, NNDC)

Beach lowering
Although often linked, beach lowering is not necessarily the same as coastal erosion.
Coastal erosion is the long-term and systematic loss of sediment from the coastal zone
that occurs over periods of years. It is commonly associated with the irreversible
erosion of rock, whether in the shore platform or in cliffs. In contrast, ‘beach lowering’ is

Toe structures management manual


169
more precisely related to the short- or long-term loss of beach materials (mainly shingle
or sand) from foreshores.
Typically, however, ‘beach lowering’ refers to scour over spatial extents much greater
(that is, across the whole shore platform) than ‘localised’ scour associated with
particular structures.

A.1.4 Problems caused by beach lowering


Both localised scour and the more widespread beach lowering can lead to problems
such as increased rate of deterioration of exposed toe structures, and the undermining
of the foundations along the seaward toe of structures which can lead to partial or total
collapse.
A comprehensive survey published by CIRIA (1986) concluded that scour at the toe of
coastal defence/protection structures represented the most prevalent and serious form
of damage to seawalls in the UK. It accounted directly for 12 per cent of the seawall
failure case histories studied and was linked indirectly to a further 5 per cent of cases.
Similar conclusions were drawn by Markle (1989) in the USA for rubble-mound
structures. The causes of failure included:
• the removal of supporting beach material from in front of a coastal defence
structure;
• a gradual dislocation of the rubble mound or blockwork foundation;
• the washing out or winnowing of granular ‘fill’ from behind the face of the
structure;
• a modification of the wave and flow conditions in front of the structure which
may, for example, increase the rate of overtopping, which in turn can lead
to the erosion of the rear face of a coastal defence structure.
In addition to these, prolonged exposure of toe sheet piles (for example, as a result of
scour or beach lowering) will increase the rate of corrosion and abrasion of the metal
piles – thereby increasing deterioration and reducing structural strength of the toe.
Exposed concrete toe beams will also be subject to wave impact forces, sediment
abrasion, weathering (such as freeze–thaw cycles where these occur), and other
chemical and biological processes. Damage could also be caused by boat collision or
other anthropogenic means. Deterioration is discussed further in Section 3.8. Failure
modes are discussed further in Section 3.3 and in Chapter 5.
The second most prevalent form of damage to coastal defence and protection
structures in the UK was outflanking, where erosion occurs at the end of a seawall,
allowing the removal of material from behind the structure (CIRIA 1986). Other
problems associated with beach lowering and scour include the following:
• Access to the beach by steps and ramps can be made more difficult (for
example as shown in Figure A.7).
• Beach lowering increases the water depth in front of the structure, allowing
larger waves to reach it and potentially increasing wave forces and wave
run-up on the structure. The possibility of reduced beach level, with
increased water depth and potentially increased forces should be included
in the design of structure toes (see Chapter 5). The potential for increased

Toe structures management manual


170
overtopping can be calculated (Pullen et al. 2007) and, if significant, may
need to be designed out or accommodated. 10
• The potential for increased agitation of beach sediments in front of the wall
or increased rate of erosion of the shore platform caused by increased
wave action, and perhaps by faster tidal currents resulting from greater
water depth at the toe of the structure. The erosion of cohesive shore
platforms is not dealt with here, but has been the subject of research by
Brew (2004) and Royal Haskoning et al. (2007).

A.1.5 Appendix contents


The rest of this appendix describes the occurrence of bed level changes seen in front
of coastal defence structures:
• evidence for scour, liquefaction, beach variability and coastal erosion;
• description of processes controlling toe scour;
• description of processes controlling liquefaction
• description of processes controlling beach variability;
• description of processes controlling coastal erosion.

A.2 Evidence for scour, liquefaction, beach variability


and coastal erosion

A.2.1 Evidence for toe scour


Toe scour is blamed for the failure of many coastal structures (CIRIA 1986) but toe
scour holes have been infrequently observed in the routine monitoring of beach profiles
(Griggs et al. 1994). Recently, however, evidence has been collected at structures in
the intertidal zone (Sutherland et al. 2006a, 2007; Pearce et al. 2006), showing that
scour holes can develop and substantially, or completely, fill in again during a single
tide, demonstrating that simple topographic beach surveys may well miss the more
severe cases of scour. Some of this evidence is presented in Box A.1.

10
The calculation and assessment of overtopping is not discussed in depth in this manual as
extensive guidance can be found elsewhere. For further information on overtopping, the reader
is referred to the Eurotop Manual (Pullen et al. 2007).

Toe structures management manual


171
Box A.1 Scour monitoring and analysis at Southbourne, Bournemouth
Two scour monitors, each consisting of eight motion sensors, were deployed at
Southbourne (Bournemouth) in 2005. Under non-eroding conditions, the sensors
remain buried in the beach and did not move. When a scour hole began to develop, the
sensors were progressively exposed and each began to oscillate in the flow. Figure A.4
shows the elevation of the lowest oscillating motion sensor, which indicates an upper
limit to the possible beach level, plotted with water level and offshore significant wave
height, Hs (m) measured in approximately 10 m water depth (Sutherland et al. 2006a,
2007).
Figure A.4 shows that, as the wave height and water level rose during the morning of
the 24 May, the beach level dropped by at least 0.60 m. The bottom monitor became
exposed, so there is no record as to exactly how far the beach level dropped below this
level. However, as water levels fell during the afternoon, the beach recovered to its
previous low-tide level. The beach level fell again as water levels rose during the
afternoon of 24 May, even though wave heights were lower. The bottom scour monitor
again became exposed and again the beach recovered fully by low tide. There was
only a small change in bed level during the next high tide as water levels were lower
and wave heights were smaller.
The results from Southbourne and extensive analysis of scour monitor data collected at
Blackpool between 1995 and 1998 (HR Wallingford 2006a) showed that beach levels
frequently drop and recover to, or close to, their original level within a single tide,
providing the water levels and wave heights are high enough. This beach lowering and
recovery could not have been detected from beach profiles conducted at low tide, even
if the profiles had been collected at successive low tides before and after the tide in
question, as the beach levels recovered partially or completely during the falling tide.

3.5
Recovery
Lowering

Recovery
Lowering

3
Tide, beach levels (mCD) Hs (m)

2.5

1.5

0.5
Tide Height Scour monitor Hs
0
24/05/2005 00:00 24/05/2005 12:00 25/05/2005 00:00 25/05/2005 12:00
Date and Time GMT

Figure A.4 Scour monitor data showing beach lowering and recovery during a
tide measured at Southbourne
Storm response is the residual change in bed elevation, when the beach does not
recover fully to the elevation it was at before the tide came in. The same location in

Toe structures management manual


172
Southbourne is shown in Figure A.5 towards the end of a storm in 2006, when the
beach had been drawn down and a scour trough had formed at the toe of the seawall
exposing rock previously placed at the toe.

Elliptical beach
shape

Figure A.5 Local toe scour at Southbourne on 11 January 2006 (courtesy of


Andrew Pearce)

A.2.2 Evidence for liquefaction in front of seawalls


The authors are not aware of any engineering problems at UK seawalls that have been
attributed to wave-induced liquefaction. Liquefaction is rarely observed due to its
temporary nature and because it can only occur underwater.
Mory et al. (2004, 2007) used arrays of pressure sensors to identify instances of
momentary liquefaction of the seabed in front of a near vertical concrete wall in the
inter-tidal zone on a beach in southwest France. Liquefaction occurred to a depth of
over 0.3 m over a fraction of a wave period with wave heights between around 0.7 m
and 1.7 m in water depths of 1.0–1.4 m. The gas content of the pore water in seabed
was also measured. The thin top layer of the seabed, disturbed by wave action, had a
low gas content, but the pore water in the seabed below that down to about low tide
level showed a higher gas content. This meant the pore water was compressible, which
played an important role in allowing liquefaction to occur. Below low water level the
seabed had a much lower gas content, so was less liable to liquefy (see Sections A.4
and B.2).
The liquefaction of a beach due to seepage, whether through a natural beach or from
under a coastal defence structure, may be more commonbut is not dealt with in this
section.

A.2.3 Evidence for beach variability


A time series of beach levels measured at a single point in front of the seawall at
Mablethorpe Convalescent Home at approximately monthly intervals between 1959
and 1991 is shown in Figure A.6 (Sutherland et al. 2007). The trend line is the best-fit
straight line through the points, which fell at an average rate of 23 mm per year during

Toe structures management manual


173
this period. Figure A.6 shows that there is a significant amount of variability about the
best-fit straight line and that this variability occurs over different timescales.

4.5

4
Beach level (mODN)

3.5

2.5

2
Level Trend
1.5
19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86

88

90

9
Year since 1900

Figure A.6 Time series of beach levels in front of a seawall at Mablethorpe in


Lincolnshire

A.2.4 Evidence for coastal erosion


The lowering of ‘ground levels’ in front of seawalls, revetments or other coastal
structures is a common phenomenon not only in the UK but also around the world. In
some circumstances, the beach becomes flatter and lower over a wider area in front of
the structure, sometimes with the sand or gravel being largely removed to reveal the
underlying rock of the shore platform (Figures A.7 and A.8). The erosion of the shore
platform has been monitored for a range of coastal sites, as listed in Table A.1, where
under ‘source’ L&C refers to Lee and Clarke (2002) while RH refers to Royal
Haskoning et al. (2007). The latter measured temporal variation in the rate through a
year and the average figures are quoted here. As an indication based on the results at
one site, Warden Point, in one specific year, Royal Haskoning et al. (2007) found the
greatest downwearing to occur in the period February to May.

Toe structures management manual


174
Table A.1 Measured shore platform erosion rates (from Lee and Clark 2002)
Location Rock type Lowering Source
rate
(m/year)
Saltburn-Ravenscar Jurassic limestones, 0–0.18 L&C
sandstones and shales
Isle of Thanet, north Kent Chalk 0.025 L&C
Lyme Regis Jurassic clays 0.1 L&C
South Glamorgan Lias limestone 0.064 L&C
Warden Point upper platform London Clay 0.031 RH
Warden Point upper middle London Clay 0.014 RH
platform
Warden Point lower middle London Clay 0.008 RH
platform
Easington Glacial till 0.042 RH

Notes: L&C = Lee and Clark (2002); RH = Royal Haskoning et al. (2007)

Figure A.7 Beach and shore platform lowering, Shakespeare Cliff, Thanet

Toe structures management manual


175
Figure A.8 Removal of beach sediments from shore platform, Penrhyn Bay
The removal of sediment may lead to the exposure and potential failure of the
structure’s toe, as shown in Figure A.9, which shows exposed sheet piling, some of
which has been removed from the toe as a result of erosion and is lying on the beach.

Figure A.9 Example of structural damage to a seawall at Milford-on-Sea (courtesy


Andrew Bradbury, New Forest District Council)

Toe structures management manual


176
A.3 Description of processes controlling toe scour

A.3.1 Introduction to toe scour


This section outlines scour mechanisms and lists the environmental parameters that
control toe scour. Some of these parameters are further investigated in other sections
of the manual.
Seawall toe scour occurs when the base of the wall can be acted upon by waves,
either directly, when the sea level is higher than the bottom of the wall, or through wave
run-up. The presence of a structure in relatively shallow water, for example, abruptly
breaks the wave and the energy is dissipated within a much smaller zone than on a
natural, unimpeded beach profile. This sudden release in energy is converted into
turbulence and wave reflection. The extra kinetic energy released around the toe of the
seawall induces lowering of the beach at the bottom of the wall by:
• increasing local shear stress on the bed to levels exceeding the threshold
for sediment motion;
• generating shock waves through the impact of waves breaking on the
seawall. The pressure waves set up in the water column are transmitted to
the bed and away from the wall. These high pressure gradients disturb the
sediment and make it more vulnerable to erosion. Wave-induced
liquefaction of bed sediments may become a contributing process at this
time.
• increasing removal of the suspended sediment by longshore currents as
the extra turbulence sustains sediment motion and allows it to be
transported by currents; and,
• reducing sedimentation as the greater water velocity close to the seawall
reduces the rate of settlement of sediment brought into the area from
longshore drift.
The process of toe scour can be self-sustaining. For example, consider the situation
where the beach level at the base of the seawall is above the mean high water spring
tide level and therefore not vulnerable to scour under normal conditions. Once a
sufficiently large storm (that is, surge water level plus storm waves) produces initial
scour, a greater range of wave/water level conditions can reach the seawall and the
beach level in front of the seawall may then lower progressively.
As the beach lowers further, the water table is closer to the surface, pore pressures
increase and the sand can be fluidised and this increases the degree of sediment
removal through backwash (Powell and Lowe 1994). Periodically, conditions may allow
a recovery of the beach level if there is a sufficient sediment supply, but for narrow
beaches with a sediment deficit, it may never accrete to the pre-scour level.
Further discussion about the processes of scour can be found in van Rijn (1998, 2005),
Whitehouse (1998), Kraus and McDougal (1996) and Sumer and Fredsøe (2002)
among others.
The following wave/water level characteristics dictate the extent and type of scour at a
seawall:
• wave height;
• wave period;

Toe structures management manual


177
• water depth at the toe of the seawall;
• beach slope;
• seawall slope;
• seawall type;
• sediment size;
• storm duration;
• angle of wave approach;
• wave overtopping.
The incident wave height, wave period, water depth at the toe of the structure, beach
slope, seawall slope and seawall type determine the wave kinematics in front of a
coastal defence structure and in particular the way that waves break on the structure.
The sediment size is important in determining whether sediment transport occurs and,
if so, whether the sediment moves as bedload within the wave boundary layer or as
suspended load in the water column.
The storm duration determines whether the scour can reach an equilibrium depth.
The angle of wave approach has an effect on the wave kinematics and in-particular the
cross-shore position of wave nodes and anti-nodes (Figure A.10) and also on the
wave-generated longshore currents.
The rest of this section consists of:
• an introduction to sediment transport in a non-breaking reflected wave;
• an introduction to the kinematics of waves in front of a reflecting seawall in
relatively deep water;
• a description of scour caused by random waves at a vertical wall;
• a description of scour caused by random waves at a sloping wall;
• commentaries on the effect of storm duration, the angle of wave attack and
overtopping on potential scour depths.

A.3.2 Sediment transport under a non-breaking wave reflected


off a vertical wall
A regular non-breaking wave with wavelength L reflecting off a vertical wall generates a
standing wave, as shown in Figure A.10. The incident and reflected components are in
phase at the wall, so an anti-node (zone of maximum vertical amplitude) is formed.
This is an area with a relatively high root mean square (RMS) surface elevation but
zero horizontal velocity. On moving away from the wall, the incident and reflected
components move out of phase until they are completely out of phase (when the
incident and reflected surface elevations cancel each other out, so there is no surface
movement but a maximum in horizontal velocity) and a node is formed. This occurs at
a distance of a quarter of a wavelength in front of the wall. On moving further out, the
incident and reflected components move back into phase and another anti-node occurs
at a distance of half of a wavelength from the wall.

Toe structures management manual


178
The standing wave generates steady streaming (a small net current) in the thin wave
induced bottom boundary layer (Longuet-Higgins 1953, 1957). This streaming is
manifested as a slow recirculating current from anti-node to node at the bottom of the
bottom boundary layer and from node to antinode at the top of the bottom boundary
layer – as shown in Figure A.10for the two-dimensional (2D) normal incidence case.
The current at the top of the boundary layer drives a counter-rotating recirculating cell
in the (much thicker) body of water above the boundary layer. This work was extended
to oblique-incidence by Carter et al. (1973).

Figure A.10 Recirculating currents due to streaming under regular


standing waves on a horizontal bed
If the sediment in the bed is coarse and travels close to the bottom, it will be most
influenced by the horizontal movements in the bottom boundary layer, which are
towards the node. The result is scouring midway between anti-node and node, and
deposition under the node, as shown in Figure A.11 (lower panel). This is known as
bedload or ‘N-type’ scour (Xie 1981).
If the sediment is small and is maintained in suspension, it will be most influenced by
the current above the bottom boundary layer, so the net movement is away from the
nodes towards the antinodes, as shown in Figure A.11 (upper panel). This is known as
suspended or ‘L-type’ scour (Xie 1981). Thus the pattern of sediment erosion and
accretion varies with the mode of sediment transport; bedload transport gives a
different pattern from suspended load transport.
Table A.2 indicates the type of scour (‘L’ or ‘N’ type) that may be predicted for a depth-
limited incident wave at a coastal structure. For simplicity, the wave height has been
set to 0.8h, where h = water depth and a constant wave period of 8 s has been chosen
to represent a storm wave in UK coastal waters.

Table A.2 Indicative scour pattern


Water Sediment grade Typical grain Scour type for depth limited wave of
depth (m) size (mm) eight seconds
2 Fine sand 0.1 L-type (suspended transport)
2 Medium sand 0.25 L-type (suspended transport)
2 Coarse sand 1 Borderline N/L type
5 Fine sand 0.1 L-type (suspended transport)

Toe structures management manual


179
5 Medium sand 0.25 L-type (suspended transport)
5 Coarse sand 1 L-type (suspended transport)

These definitions are indicative of the behaviour that can be expected, although the
mode of scour and pattern will be different with random waves and with wave breaking,
and with a sloping beach – as opposed to a horizontal bed.

Figure A.11 Scour and deposition patterns on a horizontal bed over half
a wavelength in front of a vertical seawall: Upper panel: Suspended sediment
transport (‘L-type’; Xie 1981); Lower panel: bedload sediment transport (‘N-type’;
Xie 1981)

Toe structures management manual


180
A.3.3 Random wave kinematics in front of a coastal defence
structure
Waves incident upon a coastal structure are reflected from it to some extent. The
interaction of incident and reflected waves sets up a partial standing wave pattern in
front of the breakwater. It is common in coastal engineering studies involving random
waves to consider the random sea as the linear sum of a large number of incident
component waves, plus the reflected components (O’Donoghue and Sutherland 1999).
This approach ignores nonlinear interaction (such as clapotis where waves interact to
form a breaking wave) but allows solutions to random wave problems to be formulated
relatively easily. Each component has a reflection coefficient, given by the ratio of
reflected wave amplitude over incident wave amplitude (Sutherland and O’Donoghue
1998b) and a phase shift, which relates the phase of the incident and reflected waves
at the toe of the structure (Hughes and Fowler 1995; Sutherland and O’Donoghue,
1998a).
If there is a random sea state, all reflected components will be in phase with the
incident component of the same frequency at a vertical wall; hence an anti-node is
formed at the wall. On moving further out from the wall, each component will move out
of phase at a different rate, as each has a different wavelength. Therefore the next anti-
node out from the wall will have a lower RMS surface elevation than the one at the
wall, and the one out from that will be lower again. Hughes (1992) showed how the
spatial variation in RMS surface elevation depends on relative depth kh with k=2π / L
being the wave number, L the wavelength and h the water depth. It also depends on
the narrowness of the wave spectrum. The narrower the wave spectrum, the larger the
cross-shore distance over which the partial standing wave pattern will be apparent.
Moreover, the shallower the water, the greater the cross-shore distance over which this
phase-locking will be apparent.
The variation of RMS orbital velocity with distance from the toe of a vertical wall is
shown as a function of relative depth in Figure A.12 where:
urms = root mean square horizontal wave orbital velocity at the seabed (ms-1)
g = gravitational acceleration (ms-2)
kp (and Kp) = linear theory wave number based on spectral peak wave period (m-1)
Tp = spectral peak wave period (s)
Hmo = spectral significant wave height (m)
x = cross-shore distance from seawall (m).

Toe structures management manual


181
Figure A.12 Non-dimensional plot of RMS wave orbital velocity on the
bed as a function of cross-shore position for different relative water depths (after
Hughes and Fowler 1991)
For a sloping seawall, the incident and reflected components are already out of phase
at the structure toe and reflection coefficients are lower than for a vertical wall. Both
factors mean that the partial standing wave pattern generated in front of a sloping
seawall is less obvious than that in front of a vertical wall (Hughes and Fowler 1995).
Hughes and Fowler (1995) and O’Donoghue and Sutherland (1999) have shown how
to combine linear theory expressions for the kinematics with reflection and phase shift
spectra to determine the velocities and surface elevations in front of reflecting coastal
structures. These are analytical models that do not include wave breaking due to
shoaling.

A.3.4 Scour in sand at a vertical wall under random waves


The pattern of scour under random waves reflects the kinematics of the waves that
generate it. Bed level changes (final minus initial elevation) at the end of four laboratory
experiments (after 3,000 waves) conducted with random waves, fine sand and a
vertical seawall are provided in Figure A.13 (data from Sutherland et al. 2007). The
wavelength in these tests was such that the scour plotted in Figure A.13 is over a
distance of half of a wavelength (at a water depth of 0.4 m) to almost one wave length
(at a depth of 0.1 m) – to compare with Figure A.11 – and the scour type in the model
was a mixture of bedload and suspended load type. This also covers a distance of kpx
to a value of 3 (at a water depth of 0.4 m) to 5.9 (at a depth of 0.1 m), which means
that it covers the most intense part of the variation in wave orbital velocity plotted in
Figure A.12.
In Figure A.13 negative values represent scour while positive values represent
accretion. These four tests had the same initial bed profile (a 1:30 slope), the same
wave period (Tp = 3.24s) and the same offshore incident wave height (Hs ≈ 0.2m), but
different initial water depths at the toe of the seawall(ht = 0.0, 0.1, 0.2 and 0.4m
respectively). These tests show the controlling influence of water depth. A comparison
has been drawn between these four tests as they resulted in very different breaking
wave conditions at the wall and hence different bed profiles. Thus the scour profiles
shown in Figure A.13 are for random, breaking waves in sand on a sloping beach.

Toe structures management manual


182
0.1
ht =0.4
0.05
Bed level change [m] ht =0.0m
0

-0.05
ht =0.1m Hs ≈ 0.2m, Tp = 3.24s
-0.1
Test 7, ht=0.0 Test 12, ht=0.1
-0.15
ht =0.2m Test 4, ht=0.2 Test 11, ht=0.4
-0.2
0 0.5 1 1.5 2 2.5 3
Distance from seawall [m]

Figure A.13 Variation in final scour depth with water level for a vertical
wall measured in laboratory experiments with a sand bed (from Sutherland et al.
2007)
During Test 7 (ht = 0.0 m), the waves broke offshore and the wave energy was largely
dissipated before the waves reached the wall in the swash zone. As a result there was
a slight accretion at the wall but a general lowering throughout the rest of the profile.
The vertical seawall was situated within the surf zone during Test 12 (ht = 0.1 m) and
some breaking occurred onto it, although most of the larger waves had already broken
by the time they reached the seawall. The resulting scour profile includes a small dip at
the toe of the seawall caused by turbulence and a deeper scour hole at about 0.5 m
from the structure toe.
However during Test 4 (ht = 0.2 m), the waves tended to break onto the structure and
the impacts sent water high up above the seawall. In these cases, water plunging down
the face of the seawall to the bed resulted in suspended sediment transport at the toe
and this mechanism generated the deepest scour depths. Figure A.13 shows that the
maximum scour occurred at the wall (0.158 m), with significant accretion (0.056 m)
occurring 1.3 m offshore.
In deeper water (Test 11, ht = 0.4 m), most waves did not break onto the seawall as
plunging breakers, but tended to reflect from the seawall without breaking. Erosion still
occurred at the seawall toe but was deepest at about Lp/16 from it, where Lp is the
linear theory wavelength at the spectral peak wave period, calculated using the water
depth at the seawall. The maximum scour of 0.117 m was significantly less than for
Test 4, the plunging breaker case where the toe scour was 0.158 m.
The peak in accretion occurred in Test 11 at around Lp/4 or kpx ≈ π/2, where Figure
A.12 indicates there will be a peak in RMS horizontal velocity. Figure A.11 indicates
that accretion at Lp/4 from the seawall would be expected for bedload transport, yet
almost two-thirds of incident waves were expected to meet Komar and Miller’s criteria
for suspended sediment transport, increasing to around 75 per cent when wave
reflection was taken into account. This indicates that the tests were conducted in the
suspended sediment transport regime, although with a greater percentage of bedload
than would be expected in field cases.
The variations in scour depth with distance from the seawall are different in Figure A.11
(top, for suspended load) and Figure A.13, as Figure A.11 is for an idealised case of a

Toe structures management manual


183
flat seabed with regular non-breaking waves while Figure A.13 is for a 1:30 sloping
seabed with breaking irregular waves.

A.3.5 Scour in sand at a sloping wall under random waves


Bed level changes (final minus initial elevation) at the end of three laboratory
experiments (that is after 3,000 waves) conducted with random waves, fine sand, and a
1:2 (V:H) smooth sloping seawall are provided in Figure A.14. Negative values
represent scour, while positive values represent accretion. These three tests had the
same initial bed profile, wave period (Tp = 3.24s) and similar offshore incident wave
height (Hs = 0.19m to 0.24m) but different water depths (ht = 0.0m, 0.2m and 0.4m
respectively). A comparison has been drawn between these three tests as they
resulted in very different breaking wave conditions at the wall and hence different bed
profiles.
In Test 27 the wave down-rush reached the sediment bed and caused the greatest
scour, whereas in Test 30, the water depth was sufficient to ensure that the down-rush
did not reach the bed and a lower scour depth occurred. In Test 32, the water depth at
the toe of the structure was initially zero – the still water and sand beach intersected at
the seawall. In this case there was accretion at the toe of the seawall, caused by
sediment transport in the swash zone.

0.1

0.05
Bed level change [m]

-0.05
Hs ≈ 0.2m, Tp = 3.24s
-0.1
Test 32, ht = 0 Test 27, ht=0.2
-0.15
Test 30, ht = 0.4

-0.2
0 0.5 1 1.5 2 2.5 3
Distance from seawall [m]

Figure A.14 Variation in final scour depth with water level for a 1:2
sloping wall (ht is initial toe water depth in m)
(Sutherland et al, 2007).

A.3.6 Scour in shingle under random waves


The scour response in shingle beaches tends to be confined more closely to the
seawall than is the case with sand beaches, based on the result of laboratory tests
(Powell and Lowe 1994), as shingle is relatively less mobile than sand.
As with sand, scour of a shingle beach in front of a seawall will be controlled by water
depth and wave conditions. The scour profiles in Figure A.15 show how the response
of the beach changes from accretionary in Types I and II, with a minor formation of toe

Toe structures management manual


184
scour at the seawall in the latter case, to scouring in Types III and IV. The toe scour in
Type III is maximised owing to the particular combination of water depth and wave
height occurring, and in Type IV the toe scour is reduced owing to the deeper water
depth.

Figure A.15 Schematic diagram of beach profile and scour response


under wave attack in shingle – based on model tests of Powell and Lowe (1994)

A.3.7 Angle of wave approach


When a single obliquely incident wave is reflected from a seawall, a short crested wave
field is formed, which is characterised by a diamond-shaped pattern of ‘island crests
and troughs’ (Hsu and Silvester 1989). Lines of island crests and troughs occur at
regular intervals in front of the structure. The reflection coefficient of the seawall
determines the magnitude of the island crests and troughs as a function of incident
wave height, while the phase shift on reflection gives the distance from the structure to
the first line of crests and troughs.

Toe structures management manual


185
The angle at which the wave front hits the seawall has also been suggested as a factor
affecting toe scour (Hsu and Silvester 1989). The depth of toe scour is expected to be
greater if waves hit the wall obliquely, because the incident and reflected wave trains
interfere with each other constructively, producing an interference pattern of short
crested waves. Consequently, the wave height and hence scour potential at the base
of the wall should be larger if the angle of incidence is oblique rather than
perpendicular to the seawall. In addition, oblique waves may induce local currents
parallel to the seawall, (Lin et al. 1986; Oumeraci 1994), which enhance sediment
removal at the toe of the structure.
For further discussion on the effect of oblique waves the reader is referred to Section
B.1.6 in Appendix B.

A.3.8 Overtopping
A further factor of importance may be the extent of any overtopping of the seawall. It is
reasonable to expect that seawalls that experience heavy wave overtopping will offer
less scour because the proportion of energy reflected or dissipated as turbulence at the
wall will be reduced. This effect has probably not been taken into account in previous
studies of toe scour for which the majority of walls appear to have been of sufficient
size to limit the extent of any wave overtopping. Thus, most empirically based methods
for the prediction of toe scour may be conservative if applied to low crested structures
that experience regular overtopping.
There are no design relationships to take into account the overtopping influence on
scour depth, although the following description is informative. Nishimura et al. (1978)
studied the scour at seawalls caused by an incident tsunami (effectively a very long
period wave). In this case the overtopping water returned down the face of the
structure and much of the scour was caused by the flow return. They noted that:
• scour depth decreases with decreasing wave height and increasing crown
elevation (as there is less return flow), although the area of serious
scouring is displaced towards the seawall in this case;
• scour increases (and it occurs at the toe precisely) when the face slope is
mild;
• scour decreases markedly when the water depth at the seawall increases,
as less turbulence reaches the bed;
• when waves are applied repeatedly, much less scouring is induced by each
successive wave.
To date most numerical models can only simulate overtopping by reducing the
reflection coefficient for a given seawall profile (see Pullen et al. 2007).
However, developments in phase-resolved modelling of non-linear shallow-water
waves (following, for example, Dodd 1998) have allowed wave-by-wave overtopping
events to be modelled. Such models could be coupled with sediment transport and bed
updating models to investigate the effect of overtopping on scour, although such work
is in its infancy. Few, if any, numerical models are able to simulate accurately the
turbulent dissipation occurring at the seawall.

Toe structures management manual


186
A.4 Description of processes controlling liquefaction
In soil mechanics, liquefaction starts to occur when the effective stress of the seabed
becomes zero. A useful introduction to the liquefaction of non-cohesive seabeds is
given by Sumer and Fredsøe (2002, Chapter 10). Seabed liquefaction may be caused
by the passage of waves (Jeng 1998), earthquakes and other shocks (de Groot et al.
2006a) or the rocking of coastal structures subjected to wave action (de Groot et al.
2006b). Two types of liquefaction have been observed in laboratory test and field trials,
namely residual liquefaction and momentary liquefaction. Liquefaction can lead to the
reduction in bearing capacity of the soil adjacent to the foundation of a coastal
structure. The potential consequences of this include reduced resistance to the slipping
of a coastal structure and settlement of armour stones into the seabed.
Residual liquefaction occurs in loose sand beds due to the progressive increase of
residual excess pore pressure. Under a wave crest, the pressure at the bed is greater
than hydrostatic and so the bed is compressed. Under a wave trough, the pressure at
the bed is less than hydrostatic and so the bed is dilated. This creates shear stresses
in the soil, which will lead to some rearrangement of the grains and a building up of the
pore pressure (dissipated by draining). If the pore water pressure builds up to such an
extent that it exceeds the overburden pressure, the soil will liquefy. Residual
liquefaction hardly occurs in dense seabeds due to the high shearing resistance, which
prevents the excessive build-up in pore pressure. Therefore the occurrence of residual
liquefaction in front of a coastal defence structure is unlikely as coastal structures are
usually founded on a medium dense to dense sand layer. However, in loose sand fills,
such as may be present immediately after beach re-nourishment or excavation works,
there is a risk of residual liquefaction which may need to be examined. Residual
liquefaction can persist for several wave periods and can extend over larger areas of
the seabed than momentary liquefaction.
Definitions
Pore pressure: the pressure experienced by the fluid between the sand grains within
the seabed.
Excess pore pressure: the difference between pore pressure and hydrostatic
pressure.
Overburden pressure: pressure caused by weight of sediment and water above

Momentary liquefaction usually occurs in dense seabeds due to the damping of


amplitude and the development of phase lag between the pressures at the seabed
surface and lower in the bed. Under the wave trough, the pressure at the bed is less
than hydrostatic. This pressure difference decays with depth through the seabed,
creating a pressure gradient. If the pressure gradient is sufficiently large, it can
generate more lift than the submerged weight of the soil above, resulting in momentary
liquefaction, which will occur for a fraction of a wave period only (Sumer and Fredsøe
2002) and only affects a limited area of the seabed.
The presence of a coastal structure will affect the potential for liquefaction in two main
ways:
• Wave reflection from the structure will increase the effective wave height in
front of the structure by up to a factor of two, thereby increasing the
potential range of the excess pore pressures experienced within the
seabed. Moreover the setting up of a partial standing wave pattern (see
Section A.3) will create larger horizontal as well as vertical pressure
gradients in the seabed. A method for predicting whether residual

Toe structures management manual


187
liquefaction is likely to occur in front of a coastal structure is presented in
Section B.2 in Appendix B.
• Flows may be induced under the toe of the structure, which can affect the
pressure in the seabed. Maeno and Tsubota (2001) carried out scale model
experiments investigating the flow out of back-filling sand behind
revetments due to wave loading. They showed that the cyclic seepage
force which occurs around the revetment plays an important role in the flow
out of the sand. Loveless et al. (1996) also demonstrated the effect that
groundwater flows under a seawall could have on the scouring of a model
gravel beach, even without inducing liquefaction. In the most extreme case,
an onshore flow (beach de-watering) produced accretion near the seawall
toe and an offshore flow (under the seawall from the landwards side)
produced more erosion than occurred during the no flow test.
In cohesive material, de Wit (1995) found that the potential for fluidisation of a soft clay
cohesive bed was linked to the magnitude of the wave induced stresses and the
undrained shear strength of the bed. Typically if the wave-induced stresses exceeded
the shear strength of the bed, the bed will fluidise. The depth of fluidisation will be
controlled by the variation in depth of the forcing and the strength. For given wave
conditions and soil strength profile, there will be an equilibrium fluidisation depth.
Results of surveys at Warden Point with a London Clay platform (Royal Haskoning et
al. 2007) showed the presence of thin fluid mud layers overlying more consolidated
deposits at various times and locations on the foreshore. Observations of extensive
areas of fluid mud were found on the upper shore platform during July (2005) and
these were greatly reduced in the succeeding autumn and winter. From that dataset, it
is not clear that the process of mud fluidisation directly led to formation of the observed
layers since the clay is not soft. It is more likely that erosion of clay took place which
subsequently re-deposited through flocculation as a layer of unconsolidated mud on
the platform (Royal Haskoning et al. 2007).

A.5 Description of processes controlling beach


variability
The variability in beach morphology over timescales of tides to seasons depends on
both longshore and cross-shore sediment transport processes. The changes in seabed
level observed are the residual changes from the smaller scale processes that occur
during each tide.
Sediment is moved in the cross-shore direction by wave action, with onshore motion
occurring as bedload where waves become skewed (with sharp crests and long flat
troughs) during shoaling and asymmetric (with steep front faces and more gentle back
slopes) after breaking. The breaking waves also create setup (a local increase in water
level at the shore) which drives undertow – an offshore directed return flow in the
central part of the water column between the top of the wave boundary layer and the
wave troughs. Sediment that has been lifted into suspension by the speed of the near-
bed wave velocities and/or turbulence generated by breaking is transported offshore as
suspended sediment transport by the undertow.
During a storm, offshore-directed suspended sediment transport tends to dominate and
the beach is drawn down with sand being deposited near the breaker line, which may
cause the formation of, or increase the size of, a breaker bar. During calmer conditions
and particularly when there are relatively long low waves, such as swell, the beach
material is brought back towards the shoreline as bedload.

Toe structures management manual


188
It is therefore common to find beach levels lower in winter than in summer due to the
increased occurrence and severity of storms during winter (Figure A.2d). It also follows
that beach levels may show a greater variation about their seasonal mean during
winter. Both these phenomena can be detected in repeated surveys of beach profiles.
Sediment is also moved in the longshore direction by waves and currents. Waves
approaching the shoreline at an angle generate a net longshore current when they
break, which can transport sediment along the beach. Wave action in the swash zone
also drives sediment in the longshore direction as the uprush occurs at an angle to the
beach while the downrush occurs more directly down the beach.
Longshore transport, also known as longshore drift, affects beach levels in front of
coastal defence structures when there is a gradient in the longshore transport rate.
Where the volume of sand leaving a cross-shore profile is on average greater than the
volume entering it, erosion occurs and beach levels at a structure are likely to show a
trend of lowering. Conversely, when the volume of sand leaving a cross-shore profile is
on average less than the volume entering it, accretion occurs. This is not a common
long-term trend in front of coastal defence structures as there is generally no need for a
coastal defence on an accreting beach. Shorter term reversals of drift direction may
lead to accretion, even when the long-term (decadal) trend is for erosion. The
controlling effect of gradients in longshore transport as a cause of increasing or
decreasing the sediment volume at any one coastal section conditions the beach
response to storms causing cross-shore exchanges of sediment (van de Graaff 2004).
The initial losses of sediment at a beach section due to longshore sediment transport
processes may occur on the lower beach, possibly below low water and remote from
the influence of the seawall, which means the beach has increased susceptibility to
erosion due to cross-shore losses during storm events.
The effect of longshore drift on beach levels in front of defences can probably be seen
most clearly in a groyned beach, where changes in wave direction can cause the
longshore transport direction to change. The result is that sand builds up first against
the side of one groyne with a corresponding reduction in beach level at the other
groyne, then against the other side when the direction changes. The beach levels at
both ends of the groyne bay are likely to vary substantially more than in the centre of
the groyne bay as the plan-shape of the beach changes around that point.

A.6 Description of processes controlling coastal


erosion
Many coastlines around the world are eroding as a consequence of the continual and
damaging effects of winds, rainfall, waves and currents. This tendency is further
strengthened by the gradual rise in sea level, allowing larger waves to travel further
inshore, hence increasing their effects on the seabed and on beaches. This section
considers first the erosion of non-cohesive beaches (predominantly sand) and then the
erosion of shore platforms.

A.6.1 Erosion of non-cohesive beaches


Where for example a seawall is built to protect an eroding shoreline, as sketched in
Figure A.16, it will not directly prevent erosion of the adjacent sections of that coastline.
In this hypothetical situation, it is assumed that there is no longshore sediment
transport and that the material eroded from the coastline and the shore platform is so

Toe structures management manual


189
fine-grained that it is transported out into deep water. This situation is similar to that
found on the coastline of the Great Lakes in North America.

Figure A.16 Conceptual sketches of the effect of a seawall on coastal


erosion
The principal cause of coastal recession in this situation is the continual erosion of the
shore-platform where the wave-induced water velocities are high, particularly in the
breaker zone and just outside it.
As the erosion of the coast either side of the seawall continues, the ground level either
side of the toe of the wall will fall. Now assume that the seawall affects the
hydrodynamic/ geomorphological processes in front of it ‘beneficially’ in the short term,
for example, by reducing the height of the reflected waves compared to those in front
of, say, adjacent cliffs. It might then be hoped that the ground levels in front of the wall
would not lower as quickly as those in front of the cliffs.
However, this would require an increasingly steep longshore slope in the foreshore
levels (in this case the shore platform levels) as time passes, that is, at either end of
the seawall. Experience indicates that this does not occur. It is a reasonable
approximation therefore to assume that the ground level at the toe of the wall will be (at
best) equal to that of the ground level on either side of it. This was the advice provided
in the USA by the Beach Erosion Board (1954). Griggs et al. (1991) refer to this
process as ‘passive erosion’. In this simple situation, therefore, the ground levels in
front of the coastal structure depend on the ground levels on either side of that
structure. If these continue to fall, beach lowering in front of the structure will occur.
Where groynes are built out from the shoreline, along the seawall and perhaps along
the adjacent unprotected coastline as well, they can reduce the tendency for levels
immediately in front of the wall to match those at either side of it, at least in the short
term. Even so, the ground level at the seaward end of the groynes will be similar to that
on either side of the groyne field. To maintain higher levels than would be expected

Toe structures management manual


190
without the groynes would therefore require a steeper bed/beach gradient within the
groyne bays than outside.
The simple example sketched in Figure A.16 for a short length of seawall may not
apply for situations, such as at Blackpool or Bournemouth, where seawalls stretch
along many kilometres of coastline. A further complication arises when different types
of seawall are present along a stretch of coastline, since the lowering of the beach in
front of one section of an energy-dissipating seawall may be altered by the effects of
adjacent, more reflective structures.
Many coastal erosion problems are a result of the interruption or alteration of the rates
of sediment transport along the coastline (that is, littoral drift). If we consider the simple
situation above, but now assume that there is a beach and a nett longshore drift rate,
then the seawall will interfere with that sediment transport (Figure A.16B). Experience
has shown that the normal effect of a seawall on longshore drift is to reduce its rate in
front of the wall. Van de Graaf (2004) suggested that, as the water depth in front of the
seawall deepens, the longshore transport rate will decrease. As a consequence of the
differences between the drift rates in front of the wall and beyond either end of it, there
tends to be an accumulation of beach sediments ‘updrift’ of the wall and corresponding
erosion ‘downdrift’. On the updrift end, the accumulation of beach sediments will tend
to compensate for any trend of long-term recession of the shoreline and indeed may
prevent this from occurring (for example the situation at the western end of the
promenade at Sheringham in Norfolk – see Figure A.17).

Figure A.17 Accumulation of shingle at updrift end of seawall,


Sheringham, Norfolk
Conversely, the interruption of longshore drift by the seawall results in greater downdrift
erosion problems, at least locally, than would have occurred otherwise (for example the
situation at the end of the seawall shown in Figure A.18 – Zanzibar). This localised
erosion problem will be often reflected in the beach/shore/platform levels just downdrift
of the wall, and the ground level contours in front of the wall can be expected to be
lower at its downdrift end than at its updrift end (see Figure A.16B).

Toe structures management manual


191
Figure A.18 Erosion at downdrift end of seawall, Zanzibar
Where a seawall prevents erosion of material from cliffs or dunes that would otherwise,
by receding, have provided sediment to the beaches, there will be a further deleterious
effect on the downdrift coast. Many ‘promenade’ seawalls built over the last 200 years
around the coast of the UK have not only caused this situation (for example at
Bournemouth – see Figure A.5 where the cliff is situated on the right of the figure) but,
using more emotive language, have ‘impounded’ (Griggs et al. 1991) or ‘imprisoned’
(Owens and Case 1908) a considerable amount of potential beach sediment.
A photograph taken during the construction of the promenade at North Beach,
Llandudno, for example (from Owens and Case 1908) seems to show that the shingle
ridge that once ran along the beach was used as fill material for the new wall. This
practice may have been commonplace in that century. In some cases, structures may
have been filled with sand from the beach in front of them, hence leading to the
likelihood of an ‘instant’ lowering of the beach in front of and downdrift of the wall.
Thus, it is important to note that coastal erosion at coastal defence structures
constructed to mitigate against coastal erosion may experience continued erosion
despite the presence of the structure, which neither adds nor removes sediment but
prevents it from entering the coastal sediment transport system. As previously
discussed, the interaction of waves with seawalls can cause local scour at the toe
during storms, but there is no evidence to show that coastal defence structures delay
the recovery of beaches if there is sufficient sediment available in the vicinity to rebuild
the beach.

A.6.2 Erosion of shore platforms


A shore platform and coastal defence is similar to a shore platform and cliff system.
Indeed many shore platform and defence systems used to be shore platform and cliff
systems. Defences may have been added when cliff recession threatened valuable
assets and while these defences can halt the erosion of the subaerial cliff, they do not
stop the erosion of the foreshore, as illustrated in Figure A.19.

Toe structures management manual


192
Figure A.19 Erosion of a cohesive shore platform at the toe of a seawall
Thorough reviews of the mechanisms for the downwearing of shore platform and cliff
systems have been provided by Nairn and Willis (2002), Walkden and Hall (2005),
Royal Haskoning et al. (2007) and Trenhaille (2009).
The following list of relevant erosion and weathering processes is based on that in
Royal Haskoning et al. (2007), which contains more information and references on
each process:
• Abrasion by mobile non-cohesive sediment. Waves and currents can
move sand and gravel across the surface of a shore platform, leading to
abrasion. Abrasion will cease if the sediment becomes deep enough or
the fluid velocities low enough to ensure that there is no motion of the
particles in contact with the shore platform.
• Mechanical wave erosion. The surface of a shore platform will erode if
the shear stress generated at its surface is higher than its threshold shear
stress for erosion. This can occur under non-breaking waves, but will be
enhanced by wave breaking, particularly under plunging waves when the
turbulence reaches the seabed. Rapid rates of erosion have been noted
in the Canadian Great Lakes just in front of cliffs and seawalls where
reflections enhance the turbulent energy dissipation.
• Biological processes. Boring organisms can weaken the surface layers
of a shore platform (for example, upper 10,mm), making it more
susceptible to mechanical erosion. This is perhaps more the case where
the physical rates of erosion are low, as this allows extensive colonisation
to take place. It is also affected by the tidal range as this affects the
duration of submergence.
• Desiccation and wetting. Repeated wetting and drying can cause the
upper layer of the shore platform to desiccate, which may lead to the
upper surface desiccating and cracking.

Toe structures management manual


193
• Physico-chemical effects in clay. Salt seeping and diffusing into a clay
shore platform from seawater may improve its resistance to erosion.
• Freeze–thaw cycle. Shore platforms may freeze when exposed to air
temperatures below zero, but then rapidly thaw when the tide comes in
causing frost damage (assuming that the sea remains unfrozen). It is rare
for the freeze–thaw cycle to cause much damage in the UK, but it is an
important mechanism in colder countries.
• Softening following the removal of overburden. Shore platform that is
newly exposed by the erosion of a cliff is subject to a lower load than it
was previously and may expand and lose strength.
• Softening due to pressure fluctuations caused by waves. The
pressure at the surface of the shore platform will vary as waves pass
overhead and this repeated increase and decrease in pressure may
soften the surface of the rock.
Consider the case of a vertical seawall with its toe in the shore platform, but where a
sand beach is normally present, as sketched in Figure A.20. If a storm were to occur
when the beach level was already low (probably in winter), then it may be possible for a
section of shore platform in front of a seawall to be exposed by the scouring away of
the sand beach. The exposed section of shore platform may then be eroded through
abrasion or excess shear stress, causing a local lowering of the level of the shore
platform. When the storm subsides, the scour hole is likely to fill up with sand (see
Section A.2.1) leaving little trace of the presence of a scour hole in the sand and none
of the erosion of a shore platform. If the non-cohesive beach were to be completely
lost, the same location may be eroded more than nearby locations due to the increased
levels of turbulence.
The interaction also occurs as a three-dimensional process due to the formation of a
patchy veneer of sand and gravel on the platform, migration of sand bars and spatial
variations in biological engineering.

Figure A.20 Seawall founded in cohesive shore platform with beach

Toe structures management manual


194
Appendix B: Predictive methods
B.1 Methods for predicting toe scour

B.1.1 Introduction
This section describes the process by which scour in front of seawalls can be predicted for
sand and shingle beaches. Previous studies have reviewed a range of existing scour
prediction methods and refined and developed various aspects of them (see Sutherland et
al. 2003, 2007; Royal Haskoning et al. 2007). Although numerical cross-shore profile models
have been used to model wave induced toe scour (Powell and Whitehouse 1998), they still
lack some of the main processes involved and so the most common approach to predicting
toe scour is through the use of empirical predictors fitted to experimental data. This section
contains the most up-to-date:
• predictor for toe scour depths caused by irregular waves at a vertical seawall in a
sand seabed;
• predictor for toe scour depths caused by irregular waves at a vertical seawall in a
shingle seabed;
• commentary on how to compensate for oblique angle waves and sloping
seawalls.

B.1.2 Prediction of toe scour depths at vertical seawalls in sand


seabed
The new empirical equations for the depth of scour at the toe of a vertical seawall presented
below have been developed using an extensive database of new and previously published
laboratory data (HR Wallingford 2006b, Sutherland et al. 2006b) (see Figures A.13 and A.14
in Appendix A for typical scour profiles).
All the laboratory data was collected in wave flumes using irregular waves with a TMA or
JONSWAP spectral shape and a constant water depth through each test. Each test lasted
for 3,000 waves or until equilibrium was reached. In the majority of cases, a fine sand with
median diameter d50 = 0.13 mm (Fowler 1992) or d50 = 0.11 mm (HR Wallingford 2006b) was
used and the incident offshore wave height was approximately 0.20 m. The initial bathymetry
for each test was a smooth sloping sand bed. The scaling of the tests was designed to
ensure the key processes were represented in the modelling.
Field data on beach lowering and recovery during a tide has also been collected (HR
Wallingford 2006a, 2006c). These data were collected in situations with constantly varying
water levels and wave heights. The field data have been plotted with the laboratory data in
Figure B.1, which shows that the field data generally have lower scour depths than the
laboratory data. This difference is believed to have been caused by the use of wave height,
wave period and scour depth measured at high tide and which are therefore valid for only a
small duration. The field scour did not have sufficient time to reach an equilibrium scour
depth. Nevertheless, in some cases the field measurements of scour did approach the
values measured in the laboratory.
Also plotted on Figure B.1 is Equation B.1 (Equation 12 of Sutherland et al. 2007), which is
intended to serve as a conservative predictor of scour depths and which may be used in the

195
absence of site-specific information on beach slope. The scour depth is scaled with, and less
than, Hs which is the unbroken offshore significant wave height:
St max
Hs
= 4.5e-8π (ht Lm + 0.01)
(1 − e - 6π ( ht Lm + 0.01)
) [-0.013 ≤ ht / Lm ≤ 0.18] (Eqn B.1)

where:
Stmax = maximum toe scour depth (m)
ht = water depth above the beach level at the toe of the structure (m)

Lm = g Tm2 2π = linear theory wavelength based on mean wave period Tm (s) and
acceleration due to gravity g (assumed to be 9.81 m/s2).
The range of validity of Equation B.1 is given in the square brackets in terms of ht and Lm,
which is the wavelength.

1.0
Vertical Wall, 1:30 beach
0.8 Sloping Wall, 1:75 beach
Vertical Wall, 1:75 beach
Fowler 1992, 1:15 beach
0.6 Xie, flat beach
Supertank, 1:23 beach
0.4 Eq. B.1
3.1
Blackpool
Southbourne
St/Hs

0.2

0.0

-0.2

-0.4

-0.6
-0.02 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14 0.16 0.18
h t /L m

Figure B.1 Measured and predicted laboratory and field data of relative toe scour
depth in sand (Sutherland et al. 2007)

HR Wallingford (2006f) showed that the relative toe scour depth from the laboratory
experiments depends on the beach slope and is given by:

St
Hs
= 6.8(0.207 ln (α ) + 1.51)e -11.7π ht
*
Lm
(1 − e − 6π ht* Lm
)- 0.137 [-0.015 ≤ h t Lm ≤ 0.12] (Eqn B.2)

where:
St = scour depth at the toe of the structure (m)
Hs = incident significant wave height (m)
α = beach slope (radians)

196
ht* = water depth (m) above the beach level at toe of structure including effect of wave setup
calculated using the equation of Holman and Sallenger (1985)
ht = water depth above the beach level at the toe of the structure (m).
The range of validity of Equation 3.2 is given in the square brackets in terms of ht and Lm,
which is the wavelength: Lm = g Tm2 2π = linear theory wavelength based on the spectral
average wave period.
Equation B.2 is plotted with the measured data in Figure B.2, where ‘O 1:N’ and ‘P 1:N’ are
the observed and predicted scour depths with a beach slope of 1:N (with N = 15, 30 or 75)
respectively. Equation B.2 has zero bias and systematic error and predicts the highest toe
scour depths relatively well. In Figure B.2, Equation B.2 is plotted for the range of ht*/Lm for
which data was obtained at that particular beach slope. The water depth ht* includes a
correction for wave set up which will be most influential at low value of ht.

1.0 O 1:75 P 1:75


O 1:30 P 1:30
0.8
O 1:15 P 1:15
0.6

0.4

0.2
St/Hs

0.0

-0.2

-0.4

-0.6

-0.8
-0.02 0.00 0.02 0.04 0.06 0.08 0.10 0.12 0.14
*
h t /Lm

Figure B.2 Measured and predicted (Equation B.2) relative toe scour depths as a
function of relative toe depth in sand (Sutherland et al. 2007)

The best-fit straight line (fitted to that predicted against measured data) has a slope of 0.999
and an intercept of zero. This indicates that the relationship between relative toe scour and
relative toe depth has been represented accurately. Moreover, there are relatively low errors
for the high relative scour depths, which are likely to be the most important, while the largest
errors in the predictions occur for negative observed scour depths (that is, accretion at the
toe of the structure). However, these cases are relatively unimportant – at least as far as the
stability of a structure is concerned.

B.1.3 Storm duration


The duration of the wave/water level conditions is also an important control on toe scour
development. Scour is not an instantaneous process – the trough deepens over a number of
waves. Powell and Lowe (1994) used physical wave flume models of a coarse grained
beach to demonstrate how scour develops until a quasi-equilibrium is obtained within about
3,000 waves. There was rapid initial scour that declines exponentially towards the
equilibrium depth. In Type I and II and Type III scour (Figure A.15 in Appendix A)

197
approximately 55 per cent and 45 per cent respectively of the quasi-equilibrium scour depths
formed in the first 100 waves; the development was slower for Type IV, requiring about 500
waves.
Similar trends are also apparent for sand beaches, though results from model studies
(McDougal et al. 1986) suggest that the scour hole is slower to develop, with equilibrium
unlikely to be achieved within a realistic storm/water level duration. This is supported by the
result contained in Powell and Whitehouse (1998). The experimental tests of Sutherland et
al. (2007) indicated that the average timescale of the scour was such that 95 per cent of the
equilibrium scour depth would be reached after about 2,500 waves, although there was
considerable scatter in the timescales derived. For typical storm mean wave periods of 6–8
seconds, this would take between about 4 and 5.5 hours to achieve.
The use of Equation B.2 is therefore recommended for predicting potential scour depths in
the field. If the duration that the environmental conditions are expected to hold for is less
than 3,000 wave periods, the expected scour depth may be reduced by a factor determined
from Equation B.3 for the time variation of scour depth.
S(t) = Se(1 – exp(-t / Ts) (Eqn B.3)
where:
S(t) = scour depth at time t (m)
t = time since start of scour process (s)
Se = equilibrium scour depth (m)
Ts = timescale for scour (s).
McDougal et al. (1996) suggest Ts = 3100T, with T the wave period. Xie (1981) suggested
that, for fine sand in suspension, the equilibrium scour depth would be reached in 6,500–
7,500 wave periods for H/L >0.02 and in 7,500–10,000 wave periods for H/L <0.02. Powell
and Lowe (1994) found that for a shingle beach the equilibrium scour depth was reached in
about 3,000 waves.

B.1.4 Prediction of toe scour at vertical seawalls in shingle beaches


Scour depth in shingle beaches can be predicted using the parametric plot of Powell and
Lowe (1994) reproduced as Figure B.3 (as also used by Whitehouse 1998), which was
based on an extensive set of laboratory tests conducted with normally incident irregular
waves that broke on a 1:7 slope shingle beach, with a vertical impermeable seawall. The
method is valid for 5 mm <d50 <30 mm (modelled at 1:17 scale). (Figure A.15 in Appendix A
shows some schematic scour profiles.)
Figure B.3 shows contours of S/Hs plotted on a graph with axes of relative depth, ht/Hs and
relative wave steepness, Hs/Lm, where:
• ht/Hs: relative water depth
• ht is the initial water depth at the wall
• Hs is the extreme unbroken deep water wave height
• Hs/Lm: wave steepness
• Lm is the mean wavelength of the unbroken wave (using T2g/2π)
• S: scour depth after 3,000 waves.

198
Figure B.3 Prediction diagram for contours of S/Hs scour (erosion) and accretion at
vertical seawalls with shingle beaches (after Powell and Lowe 1994)

To select the worst possible scour, look at the dimensionless scour values for all ht/Hs values
below the maximum relative water depth corresponding to the wave steepness, Hs/Lm, and
select the greatest relative scour height, which can exceed Hs. The plot gives the scour after
3,000 waves; a correction must be used to predict scour for time intervals other than 3,000
waves – see Section B.1.3.

B.1.5 Effect of a sloping or permeable seawall


The effect of a sloping wall on scour depths has been investigated by several authors,
including:
• Powell (1990) noted that, for impermeable sloping structures of 1:1.5 to 1:2,
there is no significant reduction in scour depth compared with that at a vertical
wall. However, reducing the slope of an impermeable structure to 1:3 reduced
the scour hole depth by 25 per cent to 50 per cent. Powell also noted that rock
armour revetments generally showed less susceptibility to local scour and may
even show accretion.
• Powell and Lowe (1994) showed a reduction in scour depth of almost 65 per cent
in a shingle beach when a vertical wall was replaced by a sloping wall of 1:1.25.
The scour depth was reduced by about 80 per cent for a 1:2 slope and there was
accretion at the structure toe for a 1:3 slope. A rubble mound coastal defence
showed no scour at its toe.
• Sumer and Fredsøe (2002, Figure 7.17) quantified the effect of wall slope in the
non-breaking case (d/L >0.05) and showed that scour was reduced by about 80

199
per cent (60 per cent) for a wall slope of 30° (40°) above horizontal (compared
with the scour from a vertical wall).
• Sutherland et al. (2006a) compared the maximum scour depths and the toe
scour depth at a 1:2 (27° above horizontal) sloping impermeable wall to those at
a vertical impermeable wall for four different offshore wave conditions and water
depths with Hsi/ht = 0.5–1.0, where Hsi is the incident significant wave height and
ht the toe water depth. The results are shown in Figure B.4 and show no
systematic reduction in scour depth with wave height. In these cases, the
downrush from the highest waves was reaching the seabed in some cases,
which caused scour to occur.

0.12

0.10
Scour at sloping wall [m]

0.08

0.06

0.04

0.02
Toe scour Maximum scour
0.00
0.00 0.02 0.04 0.06 0.08 0.10 0.12
Scour at vertical w all [m]

Figure B.4 Comparison of scour depths in sand at a 1:2 sloping wall and at a
vertical wall for the same offshore wave conditions (Sutherland et al. 2006a)

In shallow water, the depth of scour is controlled by waves breaking on the wall and
turbulence reaching the seabed. Under these circumstances the effect of reducing the
seawall slope can be insignificant. It is only when water depths at the toe of the structure are
sufficient to prevent turbulence reaching the seabed that a systematic reduction in scour
depths with wave height can be expected. Moreover, for a sloping seawall, the antinodes will
not occur exactly at the seawall, as there is a phase shift on reflection (Sutherland and
O’Donoghue 1998a) so the position of deepest scour may change.

B.1.6 Effect of oblique waves


Following on from the discussion on the angle of wave approach in Section A.3.7, waves
that are incident at an oblique angle to a structure will reflect from it causing a pattern of
island crests and troughs that, for regular waves, will propagate along the front of a seawall.
The steady streaming in the wave boundary layer may transport bedload along the front of
the structure. If the coastal structure is in sufficiently shallow water that the waves are
breaking, a longshore current will be generated which can also transport sediment that has
been mobilised by the wave action along the face of the structure. Both circumstances may
substantially increase the scour depth, particularly if a tidal current also flows along the face
of the structure. Under these circumstances, a physical model or possibly a coastal area
numerical model may be needed to assess the potential for scour.

200
B.2 Methods for predicting liquefaction
An analytical solution for the wave-induced pore pressure response in an isotropic infinite
thickness seabed in front of a breakwater has been developed by Sutherland et al. (2007)
based on the approach of Jeng (1998). It has been used to study the liquefaction potential of
the seabed in front of coastal defence structures subjected to various wave loadings. The
results can be used to indicate whether liquefaction of the seabed in front of a coastal
structure is likely to occur. If the results are considered significant to the structure, a more
detailed study should be carried out.
The liquefaction potential was determined by calculating the minimum total wave height to
depth ratio that will cause the momentary liquefaction of the top 0.05 m of a sandy seabed in
front of a vertical seawall. The analytical solution for an infinite and homogenous seabed
was implemented into a Mathcad calculation sheet to determine the wave heights required to
cause liquefaction to the soil. Details of the assumptions used to derive the analytical
solution, the developed equations and coefficients are given in Appendix A of HR
Wallingford (2006c).
The effects of wave height (H) and degree of saturation of the seabed (Sr) on the occurrence
of liquefaction to a fine sand bed in a water depth of 5 m is shown in Figure B.5 for a typical
storm wave period of 8 s. The degree of saturation of the seabed ranges between 0.9 and
1.0. Esrig and Kirby (1977) reported that the in situ values of the degree of saturation Sr for
marine sediment normally lie in the range 90–100 per cent. Mory et al. (2007) observed from
their field data that the Sr value of sand bed on the Atlantic coast of France ranged from 94
per cent to 100 per cent for the top 0.5 m of the sand bed.
The wave height, H, presented in the figures is the wave height of the combined incident and
reflected waves. Liquefaction occurs in the seabed when both wave condition and seabed
condition fall into the area above the line. For instance, the medium fine sand bed, with Sr
value of 0.98, starts to liquefy under the wave condition with wave height greater than 1.3d
(d = 2 m, see Figure B.5).
Table B.1 shows the minimum fully reflected wave height required to liquefy the seabed to a
depth of 0.05 m for different water depths and with degrees of saturation of 90, 95 and 100
per cent. For the unsaturated fine sand seabed with a sea depth of 2 m, the seabed could
liquefy with a wave height as small as 0.4 m. For the deep water case (d = 15 m), the
unsaturated fine sand seabed could liquefy under the wave condition with a wave height of
0.9 m. No wave-induced momentary liquefaction can possibly occur in a fully saturated
seabed in shallow water (d ≤5 m).

201
1.6
d =5m
1.5 T = 10 s
G = 10 GPa
1.4
γ s = 18 kN/m3
1.3 γ w = 10 kN/m3
kx = ky = kz Liquefaction
1.2
ν = 0.3
1.1 n = 0.3
K 0 = 0.5
1

0.9
H/d

0.8
Fine Sand (k = 1E-4 m/s)
0.7

0.6 Medium Fine Sand (k = 1E-3 m/s)

0.5 Coarse Sand (k = 1E-2 m/s)

0.4

0.3

0.2

0.1
Non-Liquefaction
0
0.9 0.91 0.92 0.93 0.94 0.95 0.96 0.97 0.98 0.99 1
Sr

Figure B.5 Wave-induced liquefaction potential around a marine structure founded


on a saturated/unsaturated seabed subjected to various standing wave loadings with
a water depth of 5 m (Sutherland et al. 2007)

Table B.1 Minimum wave height required to cause the occurrence of liquefaction
to seabed
Wave water depth (m)
Sand type
2 5 10 15
Coarse (Sr = 90%) 3.4 4.3 6.1 8.3
Medium fine (Sr = 90%) 1.1 1.4 2.0 2.7
Fine (Sr = 90%) 0.4 0.5 0.6 0.9
Coarse (Sr = 95%) 4.4 5.8 8.3 11.4
Medium fine (Sr = 95%) 1.6 2.0 2.8 3.8
Fine (Sr = 95%) 0.5 0.6 0.7 1.2
Coarse (Sr = 100%) – – – –
Medium fine (Sr = 100%) – – – 24.0
Fine (Sr = 100%) – – 14.2 18.7

B.2.1 Simplified approach for assessment of liquefaction


Figure B.6 presents the minimum height of standing wave required to induce momentary
liquefaction to a depth of 0.05 m in a seabed with an Sr value of 0.95 at various water
depths. This figure can be used to estimate the minimum wave height required to induce
liquefaction to seabed. An Sr value of 0.95 was selected in the plot because the typical air
content of an inter-tidal sand bed is approximately 5 per cent (Mory et al. 2007).
To assess liquefaction potential with Figure B.6, first select the water depth, d, and then
determine the combined wave height, H, from the incident wave height, Hi (that is, H = 2Hi).
If the value of H is greater than 1.6d then H is limited to H = 1.6d. Select the most

202
representative bed sediment grading and, if the value of H is equal to or greater than the
value of H on the y-axis, then momentary liquefaction can occur.

12

11

10
Coarse sand: k = 1E-2 m/s
9 Medium fine sand: k = 1E-3 m/s

8 Fine sand: k = 1E-4 m/s

7
H (m)

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
d (m)

Figure B.6 Wave heights required to liquefy three different types of sand bed with
given permeability, k, and with degree of saturation of 0.95, at various water depths
(Sutherland et al. 2007)

B.2.2 Guidelines and recommendations on liquefaction


The following conclusions about wave induced momentary liquefaction can be drawn from
this study:
• The likelihood of the occurrence of momentary liquefaction increases with a
decrease in seabed permeability, which is associated with a decrease in grain
size. A seabed of fine sand is therefore more likely to experience momentary
liquefaction than a seabed of coarse sand.
• The likelihood of the occurrence of momentary liquefaction increases with a
decrease in the degree of saturation of the seabed.
• The wave height required to liquefy a fine sand seabed increases significantly
when the degree of saturation of the seabed increases higher than 0.995.
• In the absence of a site-specific study, an Sr value of 0.95 is recommended for
the estimation of the minimum wave height required to liquefy the seabed.
• Figure B.6 can be used to provide a quick check on the potential for momentary
liquefaction of the top 0.05 m of the seabed. If the potential for momentary
liquefaction exists, a more detailed, site-specific study can be carried out by
adapting the Mathcad code developed within this project or using another
liquefaction model (de Groot et al. 2006a).

203
B.3 Methods for assessing erosion of soft rock
Scour in sand beaches front of reflective coastal defences has been studied extensively in
recent years (Sumer and Fredsøe 2002; Sutherland et al. 2006a, 2007). Beach lowering in
cohesive foreshores has been studied and management guidance given by Royal
Haskoning et al. (2007). Little or no attention has been paid to the possibility of scour in
cohesive seabeds in front of reflective structures. The possibility of this is explored below.
First, a way of establishing spatially varying pattern of wave kinematics in front of reflective
coastal structures is discussed; this follows the approach discussed in Section A.3.3. The
mechanisms for eroding soft and hard cohesive seabeds are then summarised.

B.3.1 Constant erosion above threshold shear stress


HR Wallingford has developed a simple method for predicting the erosion in a seabed in
front of a reflecting structure. This takes the reflective wave kinematics method of
Section A.3.3, driven by wave conditions, and calculates time series of the spatial
distribution of bed shear stress in front of the structure. The erosion rate is set at a constant
rate for every hour when the bed shear stress was greater than the threshold value.
Example time series of bed shear stress and scour depth at a point are shown in Figure B.7
for a soft clay bed and relatively long period waves. The shear stress distribution varies
spatially in front of the structure due to variations in the velocity field. It follows that the
erosion profile varies spatially as well, as shown in Figure B.8, where the time taken to reach
scour depths of 0.5, 1.0 and 2.0 m is plotted against the distance from the structure toe.
The advantage of this approach is its simplicity and speed of operation. A significant
disadvantage is that it is almost certainly unrealistic to assume a constant rate of erosion
once the threshold value has been passed. An erosion formulation more like that given in
Section B.5.3 is likely to be more realistic.

Figure B.7 Time series of bed shear stress and scour depth assuming a constant
erosion rate above the threshold (Whitehouse, 2006)

204
Figure B.8 Time taken to scour to given depth for reflection case

B.4 Erodibility index approach


Annandale (1995, 2006) proposed and presented a methodology that compares the stream
power, P, to the ability of the soils to resist scour defined by an erodibility index, K.
Calculation of the scour depth in a multi-layer soil (seabed) consists of five steps:
1. Calculation of the erodibility index, K.
2. Calculation of the stream power required for erosion to occur, PR, which is a
function of the erodibility index.
3. Calculation of the available stream power at the undisturbed (zero scour
depth) seabed, Pa.
4. Calculation of stream power at the base of each layer of the seabed, Pn, by
applying a reduction factor to Pa based on the relative scour depth, S/Smax,
where S is the depth of the base of each layer (0 < S ≤ Smax) and Smax is the
maximum scour depth (independently determined for non-cohesive fine sand
– for example using Equation 3.1). The form of the reduction factor can be
expressed as Pn = a.Pa.exp(-b[S/Smax]) where a is the amplification of the
stream power at the seabed level caused by the presence of the structure, b
controls the rate of decay of stream power with depth and both a and b have
been fitted to data. An alternative form, Pn = Pa.exp(b[1-S/Smax]) has been
proposed but not tested, with b another fitted constant and 0 <= S/Smax <= 1.
5. Comparison between Pn and PR for all layers down to Smax. Scour can occur
for all layers where Pn ≥ PR.

205
In practice, if Pa > PR (the stream power required for the erosion of the top layer of the
seabed) then the entire seabed will erode. In this case we assume that the base of the scour
hole will fall with the eroding seabed. The results produced can be summarised for generic
soil types as shown in Figure B.9. This shows the variation of relative scour depth on the y-
axis with wave forcing quantified on the x-axis, the following plausible behaviour is observed:
• relative scour depth is less than the scour depth that would be experienced in
non-cohesive fine sand;
• relative scour depth increases with wave forcing;
• relative scour depth decreases with increasing soil stiffness.

0.45
0.4
0.35

0.3 very soft


S/Smax

soft
0.25
firm
0.2 stiff
0.15 very stiff

0.1
0.05

0
0 2 4 6 8 10

a.Pa

Figure B.9 Curves of reduction in scour depths in soft rock

This approach provides a good screening tool for evaluation of scour risk in cohesive soils
and soft rock. The limiting factor in its application is the ready availability of standard
geotechnical data for the soil at the toe of the structure. The case studies in Appendix C
were not found to have comprehensive information on soil conditions. The application of this
method requires careful consideration in the light of case study information where soil
parameters are well known; this leads to the conclusion that scour risk evaluation for new
design work or maintenance work would benefit from the addition of appropriate soil
parameters such as vane shear strength and soil structure during site investigation. This
method has been used for offshore marine structures (Harris et al. 2010)

B.5 Shore platform erosion models


A shore platform and coastal defence is similar to a shore platform and cliff system. Indeed
many shore platform and defence systems used to be shore platform and cliff systems.
Defences were added when cliff recession threatened valuable assets and, while these
defences can halt the erosion of the subaerial cliff, they do not stop the erosion of the shore
platform.

206
There is still considerable debate over the relative contributions of bed shear stress,
abrasion and turbulence to the erosion of cohesive clay coasts. Thorough reviews of the
mechanisms for the downwearing of shore platform and cliff systems have been provided by
Nairn and Willis (2002), Walkden and Hall (2005) and Trenhaille (2009). Potentially useful
methods include the ones outlined below.

B.5.1 COSMOS
The model of Nairn et al. (1986) related downcutting to wave-induced shear stress and to
the intensity of wave breaking (which effects the turbulence and velocity fluctuations at the
bed). Observations indicate that downcutting increases towards the shore and so cannot be
due to shear stress alone.

B.5.2 SCAPE
The Soft Cliff and Platform Erosion (SCAPE) model of Walkden and Hall (2005) models the
erosion of the shore platform based on the wave power and a vertical shape function. The
erosion model is a variation of model developed by Kamphuis (1987) of the rate of erosion
as:
3.25 1.5
F H b T tan α
E= = (Eqn B.4)
R R

where:
Hb is the breaking wave height
T is the wave period
α is the average slope across the surf zone
R represents material strength and some hydrodynamic constants.
Walkden and Hall (2005) modified Kamphuis’ approach to give:

dy F H b3.25T 1.5
= ( )
f w - z tan α = f 1 (w - z ) tan α (Eqn B.5)
dt R 1 R

where:
y is the retreat distance
F the erosive forces under random waves
f1 is a shape function that describes the variation in F with elevation, z below the time-
varying water level, w
α is the platform slope, which is a function of elevation.
The volume of material eroded is calculated by integrating the erosion over the vertical
shape function at each time-step, then time-stepping through a tide. The shape function f1
was determined by dividing the erosion rate of some physical model tests by the beach
slope and interpolating.
An example application of SCAPE to the platform elevation change in front of a coastal
structure is shown in Figure B.10. This figure demonstrates that (at least for this specific
application, which is for a cliffed coast) continued erosion of the cliff leads to the release of
sediment onto the foreshore which maintains the platform elevation above 1 m (line labelled

207
‘Natural coast’). Where the scenario of a seawall is introduced (line labelled ‘Engineered
coast’) the platform is predicted to undergo a progressive reduction in elevation, in this case
by 0.7 m over a period of 100 years. The average rate of 0.007 m/year is in the same
ballpark as the observed rates in Table B.1. The model results were averaged between 10 m
and 20m from the cliff toe or seawall toe, and not at the toe itself.

Figure B.10 Downcutting of shore platform predicted by SCAPE model (Walkden


and Rossington 2009)

B.5.3 Trenhaille
Trenhaille (2009) presented a model of the erosion of soft rock coasts based on his
experience of modelling hard rock coasts. He included erosion by three main mechanisms;
• Erosion of bare clay by excess shear strength
• Erosion by abrasion
• Erosion by wave impact.
These mechanisms are described below.

Erosion of bare clay by excess shear strength


Trenhaille (2009) relates the erosion of bare clay surfaces to the excess shear stress using:

E ss = N o K ss (τ - τ c ) p (Eqn B.6)

where:
Ess is the erosion (ma-1) by a single wave type at a single level

208
No is the number of waves of that type at each tidal level
Kss is a calibration constant, the dimensions of which depend on the value of p (when p = 1,
the dimensions of Kss are m2kg-1s)
p is a calibration coefficient (assumed dimensionless) with typical values of 0.81 (Amos et al,
1992) or 1 (Zeman 1986)
τ is the bed shear stress (Nm-2)
τc is the critical bed shear stress (Nm-2), which depends on the clay content and the shear
strength
Trenhaille (2009) noted that for the Canadian Great Lakes, τc varies between 0.5 Nm-2 and
20 Nm-2 and performed model runs with 5 Nm-2 and 20 Nm-2. The exponent p was set to 1
and Kss was set to 2.4 ×10-7 m2kg-1s.

Erosion by abrasion
Clay surfaces with sediment on them can be abraded. The mechanisms for abrasion are not
well understood – it is not even clear if abrasion is more effective under a thinner, more
mobile layer or a thicker less mobile layer, although abrasion will stop when the layer is
sufficiently thick to become immobile at the seabed.
Trenhaille (2009) relates abrasion to the ratio between the sediment thickness, ζt, to the
maximum thickness of sediment that can be moved by a given wave, ζtmax as follows:
ζ t max
Ea = N o K a (Eqn B.7)
ζt

where:
No is as defined above
Ea is the abrasion (ma-1) achieved by a single wave at a single tidal level
Ka is a coefficient to convert the sediment thickness ratio to abrasion.
A minimum value of ζt = 0.01 m is set to prevent excessive erosion as ζt tends to zero.
Trenhaille (2009) uses the equation of Sunamura and Kraus (1985) for the maximum
thickness of sediment that can be moved by a given wave in the surf zone, namely:
ζ t max = 81.4d 50 (θ b - θ cr ) (Eqn B.8)

where:
d50 is the median sediment grain size
θb is the Shields parameter at the breakpoint
θcr is the critical (threshold) Shields parameter, taken to be 0.04.
The conversion constant for abrasion, Ka was set to 1 × 10-6. Note that the only relationship
between the erosion rate by abrasion and wave conditions is through the Shields parameter
at the breakpoint. As wave heights increase, it can be assumed that the Shields parameter
at the breakpoint will increase, so the maximum thickness of sediment that can be moved
will increase and so the abrasion rate will increase.
Equation B.8 is useful for exploring the role of sediment veneer thickness in protecting the
shore platform. Typically under storm conditions, the Shields parameter in Equation B.8 can

209
reach values of order one and probably more; this is much larger than the threshold value of
0.04. This means the sediment is very mobile and Equation B.8 predicts that the maximum
thickness of sediment that can be moved by the wave is of the order 10–100 times d50. For
typical values of d50 on beaches of fine sand (0.1 mm) and coarse sand (1 mm), this
indicates a depth of movement of 1–10 mm for the fine sand and 10–100 mm for the coarse
sand.
By way of comparison we refer to the data collected on beach veneer variability by Royal
Haskoning et al. (2007) at Warden Point in Kent. In July 2005 the beach was found to be
very thin and formed from a mixture of sediment; overall it comprised approximately a
100 mm thickness of sand sized sediment, with shell fragments and pebble sized material. A
follow up survey in February 2006 showed that the veneer had been largely removed and
the platform was exposed between individual pebbles. While there was no measurable
change in platform elevation in this period, it does confirm the thickness of beach material
that can protect the platform and which can also be removed leading to exposure of the
platform surface.

Erosion by wave impact


Mechanical erosion of clay by wave impact was calculated by Trenhaille (2009) as:
E bf = N o K bf (S F - S Fcr ) (Eqn B.9)

where:
Ebf is the recession (ma-1) from a single wave and water level condition
Kbf is a wave erosion calibration coefficient
SF is the stress exerted by the surf (kgm-2).
The surf stress is given by:

 H 
S F =  0.5γ b e - χS w  sin 2 ϕ (Eqn B.10)
 0.78 
where:
γ is the specific weight of water (kgm-3)
Hb is the wave height at breaking (m)

χ is a dimensionless surf attenuation coefficient (set to 0.01)


Sw is the width of the surf zone (m)
φ is the local beach slope.
The term sin2φ reduces the surf stress for lower slopes, with φ limited to a maximum of 50°.
Initial calibration runs indicate a threshold for excess surf stress of between SFcr = 50 kg m-2
and 500 kgm-2 depending on the resistance of the material. A conversion factor of Kbf = 1 ×
106 was found to give suitable erosion rates.
This model is expected to be able to give similar outputs to the time series graphs shown in
Figure B.10.

210
B.6 Methods for predicting beach levels over timescales
of a tide to seasons
The variations in beach levels near coastal structures at timescales of the order of the tide to
a year are the accumulation of the residual changes that occur during each tide. These
changes can occur at a range of timescales, as shown by Figure A.2, of beach levels at the
toe of a seawall at Mablethorpe (Lincolnshire). For example, it is common to find beach
levels lower in winter than in summer due to the increased occurrence and severity of storms
during winter. It also follows that beach levels may show a greater variation about their
seasonal mean during winter. This will affect the optimum timing of beach surveys.
This section is concerned with the analysis of beach levels close to the toe of a structure at
seasonal and shorter timescales. The prediction of beach levels on these timescales is
important as they provide the initial conditions for the calculation of toe scour during a tide or
storm.
Process-based numerical models of cross-shore beach evolution have been used for a
number of years to predict the (generally) short-term cross-shore response of beaches to
storms (van Rijn et al. 2003). Cross-shore profile models assume longshore uniformity and
model the cross-shore hydrodynamics, sediment transport and bed level changes. These
models have often been used to model the short-term cross-shore beach profile response to
storms, but are generally less capable of modelling the recovery of beaches after a storm.
Recent advances in the understanding of skewness and asymmetry in the surf zone,
incorporation of swash processes, the development of phase resolving nearshore numerical
wave models and the improvement of coastal sediment transport models all hold out the
possibility of improving coastal sediment models to be able to model beach recovery. If this
can be done then, for example, profile models may be able to model periods between a tide
and a few weeks where there is presently a shortage of understanding of beach behaviour
due to a shortage of data and model skill.
Until it happens, however, a more common approach to assessing the variability of beach
levels at these timescales is through the analysis of beach monitoring data. Box B.2
demonstrates how an inter-annual and seasonal trend may be fitted to time series of beach
level data collected at a point in front of a coastal structure.

Box B.2 Mablethorpe – investigating long-term beach trends and residual levels
The best-fit line of the form given in the equation below was fitted to time series of
measured beach levels in front of coastal structures at seven locations in Lincolnshire
stations (HR Wallingford 2006b, Section 3.1.5) including the time series from
Mablethorpe shown in the figure below.

Z (T ) = a − bT + c sin(2π T ) + d cos(2π T ) (Eqn B.11)

where Z(T) is the best-fit beach level at the toe of the structure, T = time (in years)
since 1900 and a, b, c and d are the fitted variables. The latter two terms can be
combined to give the amplitude and phase of the best-fit seasonal trend, represented
as a sine function.
Figure B.11 shows the best-fit seasonal trend for the seven stations calculated. Six out
of seven stations had seasonal trends between 0.1 m and 0.2 m in amplitude, which
had their highest values in August or September. The other profile, from the
convalescent home, has a much lower amplitude (22 mm) and peaked in October. The
average profile had an amplitude of 110 mm and peaked in September, with its lowest
value coming in March.

211
0.15 0.40

0.10 0.35
Seasonal trend (m)

Standard error (m)


0.05 0.30

0.00 0.25

-0.05 0.20

-0.10 0.15

-0.15 0.10
Jan Mar May Jul Sep Nov
Month

Average trend 3-month average std error

Figure B.11 A best-fit seasonal trend from Lincolnshire stations


The residual level was calculated by subtracting Z(T) from the measured values. For
four of the stations, the mean residual level was calculated for each month (noting that
the annual average residual level is zero). The standard error (standard deviation of the
residual) was calculated for all stations.

16
surveys
14
normal distribution
12

10
percentage

0
-1.2 -0.9 -0.6 -0.3 0 0.3 0.6 0.9
residual level (m)

Figure B.12 Measured and Gaussian distribution of residual beach levels at


Mablethorpe convalescent home
Residual beach levels are obtained when the long-term trend is removed from a time
series of beach levels. The seven Lincolnshire datasets were de-trended by subtracting
the best-fit straight line from the time series. The probability distribution of residual
beach levels was then calculated and plotted with a Gaussian distribution, which had
measured average and standard deviation (HR Wallingford 2006b, Section 3.1.3). The
plot for Mablethorpe is shown in Figure B.12.

212
B.7 Methods for predicting coastal erosion
The first places to look for an indication of whether there is a long-term problem of coastal
retreat at a location are as follows:
(a) Local Shoreline Management Plan (SMP). If this is from the second round of
SMPs, it should contain predicted changes for three epochs: 0–20 years,
20–50 years and 50–100 years for no active intervention and with present
management scenarios.
(b) FutureCoast 1 CD ROMs, which contain the analysis of historic Ordnance
Survey map tidelines, as well as statements on coastal ‘behaviour systems’
and local scale ‘shoreline response', which describe the future evolutionary
tendency.
(c) Local strategy studies, which may have modelled the coastal evolution of a
smaller stretch of coastline in more detail than the SMP.
(d) National Coastal Erosion Risk Mapping project.
(e) Long-term records of beach levels in front of a coastal defence.
If a long-term record of beach levels in front of a structure is available, such as the
Environment Agency’s biannual beach surveys carried out in Anglian Region for the last 10
years, then long-term trends in mean beach level in front of the structure and in intertidal
beach volume should be calculated. If these values show a statistically significant decrease
in mean beach level with time, existing trends should be projected forwards to identify when
the structure may become vulnerable to the additional effect of local toe scour, should recent
trends continue.
If there is a systemic problem of long-term coastal erosion at the location of a coastal
defence, beach levels at the structure are almost certain to have a long-term trend towards
lower values. This will have implications for the stability of coastal defences. Bed levels at
the toe of structures are not generally calculated at a timescale of years and decades. It is
more common to try and predict the behaviour or position of the shoreline. Methods for doing
this are discussed in the guidance for producing Shoreline Management Plans (Defra 2006b)
and Sutherland et al. (2007, Chapter 3).
Changes in shoreline position can be related to beach level at the toe of a structure through
knowledge of the beach slope. The SMP guidance (Defra 2006b) includes a comparison of
the following methods for analysing shoreline interactions and responses:
• extrapolation of historical data (covered here in Sections B.7.2 and B.7.3);
• numerical modelling (covered here in Section B.7.6);
• geomorphic extrapolation (covered here in Section B.7.7);
• parametric equilibrium models (covered here in Section B.7.8).
Intrinsic limits to knowledge mean that predictions of future shoreline position over a
timescale of years to decades will never be definitive, particularly when considering the
effects of climate change. Therefore it is useful to take an approach based on a range of
available methods and data to improve confidence in shoreline position and to determine the
most likely position. To obtain more site-specific data or data for shorter periods than in (a)–
(d) above, the following methods may be used.

1
Developed by Halcrow (2000–2002) on behalf of Defra and the National Assembly for Wales.

213
B.7.1 Shorelines
Ordnance Survey (OS) maps have shown tidelines since the introduction of the first OS one
inch to the mile maps in 1801. OS maps therefore provide up to about 200 years of shoreline
positions collected at different known epochs (although the practice of recording the date, or
at least the year, of a survey of tidelines has ended with the move to digital mapping, as the
date of survey is not an attribute stored with the resulting tideline in OS digital maps).
Previous maps tend to be less reliable but can still be useful for indicating the form of the
geomorphology. The shoreline position is mapped more accurately on larger scale maps,
however, so historical trend analysis (for example, Whitehouse et al. 2009) often starts with
the first County Series of OS maps published between 1843 and 1893 at scales of 1:2,500
and 1:10,560. The County Series was also the first set of maps to include high and low water
marks of ordinary tides (HWMOT and LWMOT) – earlier maps had included high and low
water marks of spring tides. Subsequent map series have continued to use HWMOT and
LWMOT, so the use of County Series maps onwards ensures a consistency in the definitions
of tidelines used to analyse shoreline change.
The representation of tidelines in OS maps is discussed in some detail in HR Wallingford
(2006d), which contains an error analysis that can be used to assess the reliability of the
trends identified (see also Sutherland et al. 2007). This error analysis is summarised in
Box B.3.
Aerial photographs have been used by Ordnance Survey and in some SMPs to illustrate
geomorphologic features and how they change with time. Beach profiles can also be
obtained from photogrammetry, as can a detailed topographic map. They are not, however,
maps and offsets may be apparent between overlapping images which can necessitate the
use of automated software to correct the distortion (Moore 2000; Leatherman 2003).
Georeferenced orthorectified aerial photographs can be incorporated within a geographical
information system (GIS) to provide the basis for displaying features. Overlaying
photographs from different periods allows the changes in identifiable features to be plotted. It
is also possible to use satellite photographs for the same purpose, particularly since the
launch of more accurate satellite photographic services such as IKONOS in 1999 and
Quickbird in 2001.
Care should be taken when combining shoreline positions from different sources as some
may be proxy-based (that is, measure a discernible feature on the beach) while others may
be based on a vertical datum (that is, measure the position of a fixed contour).

Box B.3 Error analysis of OS tidelines


Estimates of the total uncertainty in shoreline position are made up from a combination
of source uncertainty, interpretation uncertainty and natural variability (HR Wallingford,
2006d, Section 3.5; Sutherland et al. 2007, Section 3.2).
Source uncertainty reflects the errors involved in the measurement of any point and
includes errors in triangulation, the resolution of and type of corrections applied to
aerial photos and GPS errors. A suitable root mean square source error (RMSSE) for a
tideline is 3.3 m for the OS County Series and 2.8 m for National Grid maps, including
the digital Mastermap series.
Interpretation uncertainty represents the error in turning the data into a shoreline. This
includes the difficulty of determining the shoreline from an aerial photo and the error in
determining the mean high water position from a single visit. Four components of the
interpretation uncertainty have been identified:
1. Truncation of levels in Admiralty Tide Tables to nearest 0.1 m.

214
2. Surveys can be taken when predicted high or low water is within ±0.3 m
of the target level.
3. Surveys can be taken within ±0.5 hours of high tide.
4. The root mean square (RMS) vertical error in determining the
instantaneous position of the tideline, which should have been surveyed
in calm conditions.
The four errors are assumed to be independent, so are combined to give typical values
of RMS interpretation error in level of 0.23 m for high tide and 0.29 m for low tide
(although these values are likely to increase with tidal range). HR Wallingford (2006d)
demonstrates how the calculations can be made for a specific site. The vertical RMS
errors can be converted into a horizontal RMS interpretation error (RMSIE) using the
beach slope.
Natural variability reflects the dynamic changes in the shape of the beach that occur in
response to changes in waves and water levels. The root mean square variability error
(RMSVE) for this figure should be obtained for each site by analysing beach profiles.
The sources of error are summarised below:
1. RMSS for 1:2500 scale mapping decreases from 3.3 m for County Series
maps to 2.8 m for National grid maps. Mastermap mapping is taken to
have the same error as National Grid mapping.
2. RMSIE is given approximately by 0.23/tan(α) m for MHW and 0.29/tan(α)
m for MLW where α is the beach slope at MHW/MLW. Similar values
apply for County Series, National Grid and Mastermap. Regional
differences are probably larger than differences between map series.
3. RMSVE can be determined from beach profiles. As an example, in
Lincolnshire between 1959 and 1991, the RMSVE at MHW varied
between 0 m and 8 m, while that at MLW varied between 10 m and 23 m.
Beach profiles were relatively steep, being around 1:30 at MLW. Larger
errors may be anticipated on flatter beaches or on flatter beaches with
topographic features such as a ridge or runnel.
These values are not necessarily applicable outside the areas they were derived for
and local values should be estimated in all cases. If the different errors are
independent and have normal distributions, as we have assumed, then the total RMS
error, RMSTE, is given by this equation:

RMSTE = RMSSE 2 + RMSIE 2 + RMSVE 2 (Eqn B.12)

The range of expected values will then be about four times the RMS total error (at 95
per cent confidence level). A number of examples from Lincolnshire are set out below:
• MHW on a National Grid map with a 1:25 slope would have a RMS total
error of 6–10 m.
• MLW on a National Grid map with a 1:30 slope would have a RMS total
error of 14–24 m.
• MLW on a National Grid map with a 1:100 slope would have a RMS total
error of 31–37 m.
So, for example, two surveys of MLW (if on a 1:100 slope) could be up to 150 m apart,
with the differences being caused by the survey methods used and the natural

215
variations in the beach morphology. No net erosion or accretion need have taken
place. The above examples are not the worst-case scenarios as there are obvious
problems in determining MLW in cases where there are sandbanks (if the inshore
channel level is about MLW) and ridge and runnel beaches. In the former case, the
channel bed may be above MLW and MLW will run at the seaward side of the
sandbank or it may be below MLW and the MLW will run along the beach side of the
channel. In the latter case, the position of low water will depend on the configuration of
ridges and runnels. Estimates of the error in MLW assume that MLW was surveyed,
whereas in practice this was not always the case. Trends from MLW are therefore less
reliable than trends from MHW.

B.7.2 Extrapolation of historical data


The prediction of future shoreline positions or beach elevations at the toe of a coastal
defence by the extrapolation of a historical trend is one of the most common methods in use
today. The main components of the method are:
1. Collect historical data from different times.
2. Determine the coefficients of a best fit trend of shoreline position or elevation
against time.
3. Extrapolate the best fit line into the future.
Methods for determining the coefficients of a best-fit line are discussed next, followed by
examples of extrapolation.

B.7.3 Determination of the best-fit trend


The most common form of historical trend analysis involves fitting a simple linear trend to
data. Douglas and Crowell (2000) have shown that simple regression is superior to end-point
rate and complex statistical methods for calculating shoreline erosion rates. Genz et al.
(2007) reviewed methods of fitting trend lines, including using end-point rates, the average of
rates, ordinary least squares (including variations such as jack-knifing, re-weighted least
squares, weighted least squares and weighted re-weighted least squares) and least absolute
deviation (with and without weighting functions). Genz et al. (2007) recommended that
weighted methods should be used if uncertainties are understood, but not otherwise. The
ordinary least squares, re-weighted least squares, jack-knifing and least absolute deviation
methods were preferred (with weighting, if appropriate). If the uncertainties are unknown or
not quantified then the least absolute deviation method should be preferred.
Confidence limits can be calculated to provide a measure of the reliability of the erosion or
accretion rate. They provide a range for the calculated erosion or accretion rate and depend
on the variance of the data, the number of samples and the desired level of confidence.
They strictly apply only to the time period the data was collected in. The extrapolation of
trends and confidence limits into predictions assumes that the future hydrodynamic climate
will be statistically similar to the climate during the period the measurements are made.
There are a number of advanced linear and nonlinear data analysis methods that can be
used to analyse the long-term prediction of beaches. The linear methods include using
correlation, Fourier series, random sine functions, wavelets, empirical orthogonal functions,
canonical correlations and principle oscillation patterns (Larson et al. 2003; HR Wallingford
2006e, Section 3.2). Non-linear analysis methods include singular spectrum analysis, multi-
channel singular spectrum analysis and fractals (HR Wallingford 2006e, Section 3.3). The
more advanced data analysis methods rely on having a large quantity of regularly-sampled,

216
accurate data. The use of such analyses will become more common as the amount of data
collected by organised regional coastal observatories increases, but in the meantime they
are mainly research tools.
Beach level time series data can be statistically analysed to give an indication of the rate of
change of elevation and hence of erosion or accretion. Measured rates of change are often
used to predict future beach levels by assuming that the best-fit rate from one period will
continue into the future. The historical trend is then extrapolated to give predictions of future
beach levels, which can be used by a coastal engineer to predict when a trigger/alert or
damage level may be reached. Alternatively or in addition, long-term shoreline rates of
change can be determined using statistical analysis of cross-shore position versus time
data. Box B.4 outlines a range of methods that can be used to undertake a standard linear
analysis of beach level data.

Box B.4 Linear analysis of beach level data


The linear analysis of beach level data is demonstrated here using a set of beach
profile measurements carried out at eight locations along the Lincolnshire coast
between 1959 and 1991, as described in HR Wallingford (2006c, Section 3.1).
Locations backed by a seawall were chosen and time series of beach levels were
output at points near the seawall toe. An example of a time series has been given in
Figure B.13.

4.5

4
Beach level (mODN)

3.5

2.5

2
Level Trend
1.5
19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19

19
58

60

62

64

66

68

70

72

74

76

78

80

82

84

86

88

90

92

Year since 1900

Figure B.13 An example of time series beach levels


A linear trend fitted to a time series of beach levels gives an indication of the rate of
change of elevation and hence of erosion or accretion. The measured rates of change
are often used to predict future beach levels by assuming that the best-fit rate from one
period will be continued into the future. Alternatively, long-term shoreline change rates
can be determined using linear regression on cross-shore position versus time data.

Confidence limits can be calculated to provide a measure of the reliability of the erosion or
accretion rate. They provide a range for the calculated erosion or accretion rate and depend
on the variance of the data, the number of samples and the desired level of confidence.
They strictly apply only to the time period in which the data was collected.

217
B.7.4 Extrapolation of trend to future dates
Once a trend of position against time has been established, the equation and its fitted
coefficients can be used to extrapolate the trend beyond the date of the last data point and
into the future. Any such extrapolation depends on future conditions being similar to past
conditions. The results of an extrapolation must be interpreted in light of the underlying
principles of geomorphology and sediment transport (Whitehouse et.al. 2009).
The extrapolation of trends and confidence limits into predictions assumes that the future
hydrodynamic climate will be statistically similar to the climate during the period the
measurements are made. The use of confidence limits and their limitations are illustrated in
Box B.5.

Box B.5 Example of the extrapolation of beach survey data with confidence
limits
The use of an extrapolated trend to hindcast beach levels is illustrated using data
collected at Boygrift Outfall between 1970 and 1990. A linear trend in beach level was
fitted to the data from 1970 to 1980 and the 95 per cent confidence limits were
calculated on the assumption of a Gaussian distribution of residual beach levels (see
Box 3.2). Figure B.14 shows that only three out of the 92 measured beach levels
between 1970 and 1980 fell outside the 95 per cent confidence limits. The linear trend
between 1970 and 1980 was then extrapolated between 1980 and 1990, as were the
confidence limits. Over a quarter of the measured beach levels from 1980 to 1990 were
outside the extrapolated 95 per cent confidence limits.

1.5

1.0
Elevation (mODN)

0.5

0.0

-0.5

-1.0
1970 1972 1974 1976 1978 1980 1982 1984 1986 1988 1990
Year

fitted levels (from 1970 to 1980)


1970 to 1980 trend
1970 to 1980 95% confidence limit
extrapolated trend
extrapolated 95% limit
measured levels (1980 to 1990)
Figure B.14 Linear trend in beach levels at Boygrift outfall from 1970 to 1980
and extrapolated trend from 1980 to 1990 plotted with measured beach levels

218
and 95 per cent confidence limits

The usefulness of an extrapolated best-fit trend in beach levels as a predictor for future
beach levels has been examined using 30-year long datasets of beach levels at the toe of
the seawalls from four locations in Lincolnshire (Sutherland et al. 2007). The data were
divided into 10-year long sections starting from 1960. At each location, a least-squares best-
fit straight line was fitted to each 10-year section and the rates of change in elevation are
shown in Table B.2. For a 10-year trend to be useful as the basis for predicting beach levels
over the following 10 years, the rates of change from successive decades should be similar
and should ideally be close to the 30-year average rate of beach level change, which is also
given. Generally, in Table B.2 they are not. Only one of the 10-year rates of change is within
±100 per cent of the previous one and only three are within ±200 per cent (out of eight
combinations). Only five of the 12 decadal rates were within ±100 per cent of the 30-year
rate.
Table B.2 Rates of change in elevation in front of seawalls for different periods
Rate of change (m/year)
Period
Convalescent Home Bohemia Point Boygrift Outfall Chapel Point
1960–1990 -0.025 -0.021 -0.030 -0.028
1960–1970 -0.017 -0.001 0.010 0.069
1970–1980 -0.063 0.010 -0.035 -0.028
1980–1990 0.047 -0.061 -0.051 -0.186

In this example, the 10-year rates of change in beach level provided little predictive
capability for estimating the change in elevation for the following 10-years, let alone for the
planning horizon that might need to be considered for a coastal engineering scheme.
However, they may still provide a useful prediction over a shorter time interval. In order to
determine how far ahead a trend can be extrapolated and still provide a useful prediction of
future beach levels, its prediction horizon can be calculated.
The prediction horizon is the length of time over which a predictive technique produces on
average a better prediction of future beach levels than a simple baseline prediction. The
quality of a prediction is determined using a skill score (Sutherland et al 2004), which is a
non-dimensional measure of the accuracy of the prediction relative to the accuracy of a
baseline prediction of future beach levels. The most commonly used skill score in
morphodynamics modelling is the Brier skill score (Sutherland et al. 2004, 2007) described
in Box 3.6. A common baseline prediction of future elevations is that they will not change.

Box B.6 Brier skill score


The Brier skill score (BSS) is a non-dimensional measure of accuracy of prediction
compared to a simple ‘baseline prediction’ and is determined by:
n

∑ (Measured − Pr edicted )
2
i
BSS = 1 − 1 (Eqn B.13)
n

∑ (Measured − Baseline )
2
i
1

A resultant BSS score of 1 is a ‘perfect’ prediction of the extrapolated data. A score of 0


means the prediction is the same as the baseline prediction. A score less than 0
indicates that the prediction is worse than the baseline prediction. This skill score is
reduced by errors in the prediction of amplitude, phase and mean. It provides an

219
objective measure of a model’s performance (Sutherland et al. 2004).
The BSS has been used to calculate the skill of coastal profile and coastal area models
(Sutherland et.al. 2004) by comparing measured and predicted bathymetries at one
point in time, using the baseline assumption that the final bathymetry was the same as
the initial bathymetry. It has also been used to compare measured and predicted time
series of beach levels at a point in space (Sutherland et.al. 2007) where the BSS was
calculated as a function of the duration of the prediction, then averaged in bins of equal
duration.

B.7.5 Procedure to establish an average prediction horizon


The concept of the prediction horizon was derived from meteorology, where it is used to
assess how far in advance weather forecasts can be made. It was applied to the prediction
of future beach levels using a linear trend fitted to historic data by HR Wallingford (2006c,
Section 3.1.2) where the Brier skill score at each point in time was calculated as a function of
the duration of the prediction (see Box B.7). The skill scores were ordered by the duration of
the prediction and sorted into bins of equal length of time. The BSS were averaged for each
bin to give the mean skill score as a function of the duration of the prediction.
An example of this for beach level data at the toe of a seawall in Lincolnshire is shown in
Figure B.15, where the Brier skill score is plotted against duration of prediction for seven
locations. The prediction horizon is the duration at which the average BSS decreases to
zero.

1.0

0.5
Brier Skill Score

0.0

-0.5

-1.0

-1.5
1 2 3 4 5 6 7 8
Prediction length (years)

Bohemia Point Trusthorpe Outfall Bohemia Point

Sutton Pullover Boygrift Outfall Chapel Point

Trunch Lane Mean

Figure B.15 Brier skill scores versus time for Lincolnshire profiles based on linear
trends fitted to 10 years of data (Sutherland et.al. 2007)

The extrapolation of the best-fit trend in historical beach profile time series will act as a better
predictor of future beach levels than the average beach level for time differences where the
average skill score remains above zero.

220
The use of extrapolated beach level data to predict future beach levels should therefore be
limited to periods of a few years only. As already noted in Section B.7.4, this duration is
shorter than the timeframes normally considered for the precautionary approach to coastal
management. However this duration is likely to be suitable for a managed/adaptive policy of
tracking risk, informing toe management and performing multiple interventions.

Box B.7 Determination of prediction horizon


The procedure for determining the average prediction horizon given by a trend line
fitted to M years of a time series of beach levels at a point is given in detail in HR
Wallingford (2006e, Section 3.1.2) and is summarised below.
Fit a straight line to the first M years’ data, starting from the first point.
For each data point beyond the data used in the fitting, extrapolate the fitted line to that
point and record the following three values together:
1. The duration of the extrapolation (time between last point used in fitting and
data point)
2. The difference (x – y) between the measured elevation, x, and the
extrapolated, y
3. The difference between the measured elevation, x, and the baseline
prediction of the elevation, B, which is the average elevation of the data
used in the fitting
Repeat the above procedure, only starting from the next point each time until the fitted
time series extends to the end of the time series.
Sort the results by duration of extrapolation into bins of, say, one year (that is, all
results with duration between 0 and 1 year, 1 and 2 years, and so on).
Calculate the Brier skill score for each bin, i, using the equation below.

(x − y )2
BSS(i ) = 1 − (Eqn B.14)
(x − B)2

B.7.6 Numerical modelling


A considerable amount of research has been carried out over the last two decades to
develop predictive numerical models of coastal evolution covering periods of up to 20 years
or more. These models are based on representations of physical processes and typically
include forcing by waves and/or currents, a response in terms of sediment transport and a
morphology-updating module. However, there are still major gaps in our understanding of
long-term morphological behaviour (de Vriend et al. 1993, Southgate and Brampton 2001;
de Vriend 2003; Hanson et al. 2003), which mean that modelling results are subject to a
considerable degree of uncertainty. Their use requires a high level of specialised knowledge
of science, engineering and management.
Southgate and Brampton (2001) provide a guide to model usage which considers the
engineering and management options and the strategies that can be adopted, while working
within the limitations of a shortfall in our scientific knowledge and data. An introduction to the
following model types can be found in HR Wallingford (2006e):
• one-line models for sand beaches;

221
• coastal profile models for sand beaches;
• coastal area models for sand beaches;
• systems model SCAPE for soft cliff and platform erosion (with a sand beach) –
see Section B.5.2 for example application.
The approximate limitations and applicability of the types of existing numerical models are
illustrated in Figure B.16. Coastal tract models are based on sediment budgets. Figure B.16
shows that the numerical models attempt to describe fewer and fewer processes in detail as
the spatial and temporal scale over which they are deployed increases.

Figure B.16 Indication of spatial scale and length of prediction for different
numerical model types

B.7.7 Geomorphic extrapolation


Geomorphic extrapolation is an expert-led, feature-focussed assessment of morphological
behaviour and response, such as those provided for some scenarios by FutureCoast
(Halcrow, 2002). Although a consistent methodology can be applied, this approach relies on
expert judgement and so a range of outcomes is possible. Further information is available in
Whitehouse et.al. (2009).

B.7.8 Parametric equilibrium models


Parametric equilibrium models represent the shape of the coastline or its response to forcing
through simple equations that have been derived through a mixture of curve fitting and
theoretical considerations. They are necessarily simplistic, but quick to apply.
The two main equilibrium beach concepts commonly used to predict coastal morphology are:

222
• the Bruun rule for coastal retreat (Bruun 1962);
• log-spiral coastlines (Silvester and Hsu 1997).
Bruun (1962) proposed Equation B.11 for the equilibrium shoreline retreat, R, of sandy
coasts that will occur as a result of sea level rise, S.
L
R=S (Eqn B.15)
h+B

where:
L is the cross-shore width of the active profile (that is, cross-shore distance from closure
depth to furthest landward point of sediment transport)
h is the closure depth (maximum depth of sediment transport)
B is the elevation of the beach or dune crest (maximum height of sediment transport).
The equation balances sediment yield R(h+B) from the horizontal retreat of the profile with
sediment demand, S×L, from a vertical rise in the profile (Dean et al. 2002). The magnitudes
of h and B are difficult to determine, however, and the actual seabed will need time to
respond to a change in sea level.
The Bruun rule does not depend on a particular coastal profile, but does assume that no
sediment is lost from the coastal system (which is likely to happen if there are fines in the
area eroded). It assumes a coast of unconsolidated sediment, mainly sand, with (originally) a
coastal dune and makes no allowances for gradients in the longshore or cross-shore
transport of sand. However, the Bruun rule has been extensively modified, developed and
used (see Dean et al. 2002 for a summary). An example of how the Bruun rule can be used
to calculate potential changes in shoreline retreat rates is given in Box B.8.

Box B.8 Relative shoreline retreat rates using Bruun rule


In the coastal regions where the Bruun rule can be said to apply, the rate of shoreline
retreat (dR/dt) is directly proportional to the rate of sea level rise (dS/dt). It follows that
the ratio of future shoreline retreat rate to present day shoreline retreat rate (the
shoreline retreat rate multiplier) will be the same as the ratio of future sea level rise rate
to present day sea level rise rate. The future rates of sea level rise, including the
effects of isostatic change, can be obtained from Defra (2006b) while the present day
rate is given by adjusting the global rate IPCC (2007) for regional isostatic changes
(Defra 2006b). These were combined to give the shoreline retreat rate multipliers
shown in Table B.3. The Bruun rule indicates that shoreline retreat rates could increase
significantly – in some cases by a factor of 13 – during the 21st century.
Table B.3 Shoreline retreat rate multipliers for different time spans

Region Shoreline retreat rate multipler


1961–2003 1990–2025 2025–2055 2055–2085 2095–2115
East and 1.0 1.5 3.3 4.6 5.8
south-east
England
South-west 1.0 1.5 3.5 5.0 6.3
England
and Wales
North-west 1.0 2.5 7/0 10.0 13.0
and north-
east
England

223
These results should be treated with some caution, however, as the Bruun rule is a
very simplistic analysis tool and difficult to validate. Bray and Hooke (1997) adapted it
to look at the erosion of soft cliffs by adding sediment exchange and considered it
particularly suitable for assessing the sensitivity of eroding soft cliffs to future climate
change. However, both Cooper and Pilkey (2004a, 2004b) and Stive (2004) cautioned
against its use due to its simplicity and restrictions. The Bruun rule is likely to be
particularly inadequate in regions where there is a significant variability in the longshore
transport rates (Dickson et al. 2007).
Log-spiral curves have been fitted to characterise the equilibrium plan-shape of a sandy
beach between two hard control points with a dominant wave direction (Silvester and Hsu
1997). The control points may be headlands or beach control structures. If, in particular, new
structures are planned, the equilibrium beach shape should be calculated to see how close it
comes to any coastal defences at the back of the beach.

224
Appendix C: Case studies
Case Location Structure type Issue
study
C1 Ael-Y-Bryn, North Rock armour toe • Erosion at the toe of
Wales revetment a cliff and slippage of
cliff face
• Reduction of beach
volumes
C2 Corton, Suffolk Rock armour toe • Beach lowering
protection and leading to
revetment undermining
• Structural failure
C3 South Beach, Sheet pile wall and • Beach lowering
Lowestoft, Suffolk concrete thrust
block/beam
C4 Holme Dunes, Beach drainage • Erosion of dunes due
north Norfolk to a change in
coastal processes
C5 Overstrand, north Sheet pile wall and • Failure of apron due
Norfolk stepped concrete to corrosion, leading
apron to cliff erosion
C6 West End, Open stone asphalt • Shoreline retreat
Dovercourt, Essex and geotextile leading to potential
erosion of
embankment and
release of landfill
C7 Teignmouth to Concrete stepped toe • Loss of beach
Dawlish Railway, beam material leading to
Devon undermining of
structure
C8 Felixstowe, Suffolk Sheet piling, beach • Failure of groynes,
recharge and control leading to rapid
structures reduction in beach
levels, resulting in
undermining of
structure
C9 Clayton Road, Rock armour toe and • Beach lowering and
Selsey, West geotextiles toe scour, caused by
Sussex wave reflections,
leading to risk of
structural failure

Toe structures management manual


225
Case Location Structure type Issue
study
C10 Fort Wall, Encasement and rock • Beach lowering
Canterbury, Kent armour toe leading to structural
failure (overturning)
C11 Prestatyn, North Beach recharge, rock • Wave and tidal
Wales groynes and sloping induced scour
apron
C12 Colwyn Bay, North Rock berm and rock • Beach lowering,
Wales groynes leading to heavy
overtopping and
foundation instability
C13 Rhos-on-Sea, Rock armour • Beach lowering
North Wales breakwater
C14 Penrhyn Bay, North Beach recharge and • Beach lowering and
Wales fish-tail groynes overtopping
C15 Sandbanks Rock groynes and • Beach lowering
Peninsula, Poole, rock toe protection caused by groyne
Dorset failure and complex
tidal current regime
C16 Seaford, East Beach recharge • Beach lowering
Sussex leading to
undermining and
structural failure
C17 Selsey Bill, West Concrete armour • Beach lowering
Sussex units and a rock berm
C18 Sidmouth, Devon Offshore breakwaters • Beach lowering
and a rock groyne leading to wave
overtopping
C19 Portobello Beach, Beach recharge and • Beach lowering
Edinburgh timber groynes leading to wave
overtopping

NB: Case studies C11–19 were sourced directly from Appendix 1 of Beach Lowering in
Front of Coastal Structures – Research Scoping Study (Sutherland et al. 2003) and are
thus presented in a different format to case studies C1–10.

Toe structures management manual


226
C1 CASE STUDY: Ael-Y-Bryn, North Wales
Courtesy of Conwy Council and Coastal Engineering UK Ltd

C1.1 Identification of the problem


The site is located just to the east of the Little Orme headland, Llandudno, North Wales
(see Figures C1.1 and C1.2).

Figure C1.1 Site location


This case study concerns coastal protection to the small development known as Ael-y-
Bryn. Reference is also made to the adjacent development of Craigside (Figure C1.2).
Properties at Craigside were built much earlier than those over the shorter 170 m
frontage of Ael-Y-Bryn.
The coast here consists of cliffs of about 15–20 m in height, fronted by a beach
consisting of sand, shingle and rounded limestone boulders derived from earlier cliff
erosion. The beach at the toe of the cliff face has a marginal depth of covering over the
underlying clay that forms the cliff. Table C1.1 gives some data extracted from Coastal
Engineering (2001).

Table C1.1 Relevant levels abstracted from Coastal Engineering (2001)

Frontage TP Top of Beach Base/clay Beach material above


wall level level level clay horizon
(mODN) (mODN) (mODN)
Craigside 1 6.790 5.690 4.140 Coarse sand and shingle
Craigside 6 6.790 4.130 3.800 Shingle
Craigside 11 6.800 4.370 3.340 Shingle, small boulders
Ael-Y-Bryn 14 - 2.715 2.315 Sand, gravel and small
boulders
Ael-Y-Bryn 15 - 2.775 2.175 Coarse sand, cobble and
small boulders

The Craigside properties were protected from the erosive action of the sea by a vertical
masonry retaining wall. Most of the Ael-Y-Bryn properties were built between 1974 and
1980, the six houses closest to the cliff edge being built subsequently after 1983. As a

Toe structures management manual


227
condition for building these later properties, the developer was required by the (then)
local authority, Aberconwy Borough Council, to provide coast protection. These
measures entailed regrading the cliff and constructing a revetment. According to local
observers, the revetment was formed by pushing the indigenous boulders from lower
down the foreshore up to the toe of the cliff (Figure C1.3). Larger boulders were set at
the bottom of the mound, with smaller boulders at the top.

Figure C1.2 Location plan showing Ael-Y-Bryn, Craigside and the paddling
pool area

Toe structures management manual


228
Figure C1.3 Coast protection measures constructed as a condition of
the development
Llandudno Bay (the bay between the Little Orme and the Great Orme) is shielded from
some directions of wave attack, but exposed to waves from northwest through to
northeast. The incident wave climate gives rise to both easterly and westerly migrations
of beach sediments between the two headlands. Coastal Engineering (2001) describes
how the bay is effectively a closed cell in terms of drift. In spite of the cellular nature of
the bay, comparison of historic and recent plans showed there had been a depletion of
sediment, and that the cliffs at Ael-Y-Bryn had been slowly eroding. Even after the
installation of the developer’s revetment at Ael-Y-Bryn, continuing erosion was evident,
being manifest by way of erosion at the toe of the cliff and slippage of the cliff face.
Examined in the 1990s, records showed that there had been a significant reduction in
the beach volume at Craigside. This was, at the time, incorrectly perceived to be due to
coast protection measures to the west, including construction of a groyne. Subsequent
monitoring demonstrated that other factors were at play; in particular, the shallow
revetment, having a crest level of 5–6 mODN was readily overtopped by storm waves
on high tides (1 in 1 year sea level was calculated to be 4.63 mODN). It is also possible
that the alleged removal of boulders from the lower shore had reduced the capacity of
the foreshore to attenuate wave energy, thus resulting in greater impact higher up the
cliff face.
As part of the initiatives for this frontage, Coastal Engineering UK recommended an
extension of a monitoring programme which had begun for other parts of the frontage
in 1997. Monitoring identified no real problems with the vertical walled defences at
Craigside. The wall was adequate and groundwater monitoring showed no variation
with tide level. The principal concern remained with Ael-Y-Bryn and, in particular, how
to protect the cliff.

C1.2 Appraisal
The Shoreline Management Plan (Sub-cell 11a, Great Orme’s Head to Formby Point)
published in 1999 advised a ‘hold the line’ policy for the whole of Llandudno Bay. The
outcome of a review reported in the project appraisal, Coastal Engineering (2006),
stated that the then existing policy for the specific Ael-Y-Bryn frontage was one of ‘no
active intervention’. However, the residents had been expressing concerns over the
adequacy of the defences for over 15 years; moreover, the distance between the
boundaries of the properties and the cliff edge was by now only 15–20 m.
Given the nature of the infrastructure at risk, the indicative standard of protection for
the frontage was 0.5–2 per cent when expressed as annual probability. By comparison,
the actual risk of serious erosion was calculated to have an annual probability of 20 per
cent. The stated objective of the project appraisal was, therefore, ‘to provide an
appropriate level of coastal defence to cliff top properties at Ael-Y-Bryn, Llandudno,
threatened by erosion’.
Following the breaching of defences at Towyn and elsewhere on the North Wales coast
in 1990, a number of initiatives involving combinations of private and public sector
funding were considered; this included a possible scheme to restore the standard of
coast protection afforded to the cliff top houses at Craigside and Ael-Y-Bryn. The
prospect of 50 per cent of the funding being provided by the residents, with the balance
coming from the Welsh Assembly Government , enabled a financial means by which a
coast protection scheme could be promoted.
In addition to the ‘do nothing’ option, as required to determine scheme benefits, three
main options were examined:

Toe structures management manual


229
• linear defence consisting of a re-armouring of the cliff face;
• linear defence consisting of a vertical wall, similar to that at Craigside;
• beach recharge and management.
Apart from being significantly the most costly option, beach recharge was rejected
partly due to concerns over who would be responsible for its continued management.
Of the two linear defence options, re-armouring of the revetment was the preferred
one, yielding a benefit/cost ratio of around 3.
Consultation was carried out with each of the six residents benefitting from the project,
and with the Countryside Council for Wales (CCW). Though initially hesitant in
endorsing the use of rock, CCW agreed to the scheme, given that the locally occurring
rounded boulders would be reused in the works.

C1.3 Outline design and consents


A ground investigation was carried out including trial pits (Figure C1.4) and boreholes.
However, the latter proved problematic due to the percussion tool hitting boulders.

Figure C1.4 Position of foreshore trial pits


The capital works scheme would consist principally of cliff regrading and a rock
revetment laid over geotextile. Subject to maintenance, the proposed scheme would
provide a Standard of Protection against erosion damage of 1–4 per cent annual
probability and it would have a scheme life of 50 years.
Consents were obtained as follows:
• planning approval (Town and Country Planning);
• approval by CCW in respect of works being carried out within a SSSI (the
Little Orme had been designated as a SSSI noted for terrestrial features,
which was extended in 2001 to the foreshore at Ael-Y-Bryn in respect of the
boulders on the lower foreshore);

Toe structures management manual


230
• from Conwy County Borough Council in respect of the Coast Protection Act
1949;
• approval from the then Department for Environment, Transport and the
Regions (DETR) in respect of the Coast Protection Act, Section 34
(interaction with navigation);
• Food and Environmental Protection Act (FEPA) licence.
The last two consents were dealt with under a single application to Defra’s Marine
Consents Unit.

C1.4 Detailed design


The detailed design was prepared by the local authority. Figure C1.5 shows a typical
cross-section which comprised:
• re-grading and reseeding of the cliff face;
• a crest of level of +6.8 mODN;
• a revetment consisting of 1–3 tonne rock rip-rap laid at a 1 in 2 slope on
geotextile placed on the cliff face;
• a toe, buried into the beach to about 1 m depth or less if underlying clay
was encountered.

Figure C1.5 Typical cross-section

Toe structures management manual


231
C1.5 Construction issues
The capital works scheme was procured through competitive tender. Further to the
reuse of the local rounded boulders, rock was obtained from local quarries. This
limestone had been used elsewhere where it was known to be performing well after 30
years’ service. Given its position at the foot of the cliff, access to site needed careful
consideration. To facilitate access, a temporary haul road was created from the
paddling pool (see Figure C1.2), the latter area also being allocated for the contractor’s
compound. Figure C1.6 shows the regraded cliff face.

Figure C1.6 Regraded cliff face


The construction works were at a high elevation in the beach profile but still required
shift working in line with the tidal window and weather conditions (see Figure C1.7).

Toe structures management manual


232
Figure C1.7 Construction in progress

C1.6 Post construction Issues


The completed scheme is shown in Figures C1.8 and C1.9. Beach surveys and an
inspection of the site are carried out twice a year; there is an annual report. No
problems have been encountered with the toe of the revetment. Public perception of
the scheme is good.

Toe structures management manual


233
Figure C1.8 The completed scheme

Figure C1.9 Aerial view of the project (note Craigside vertical wall to the
west of the revetment)

C1.7 Lessons learnt


Working with the residents at Ael-Y-Bryn yielded a satisfactory outcome. The residents
made a significant contribution to the cost of the works, amounting to some £130,000
(roughly £10,000 per property), the remaining 50 per cent of the cost being met by the
Welsh Assembly Government.

Toe structures management manual


234
It is acknowledged that the crest level might need to be topped up in the future to
sustain the standard of protection but the adaptive nature of the works allows for doing
this. The toe is not expected to be a problem.

C1.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Dyfed
Rowland of Conwy County Borough Council and Alan Williams of Coastal Engineering
UK Ltd in the preparation of this case study.

C1.9 References
Coastal Engineering, 2001. Craigside/Ael Y Bryn frontages Llandudno, Coastal
Engineering Consultancy Services for Conwy Council, September 2001.
Coastal Engineering, 2006. Ael Y Bryn, Llandudno, proposed coast defence
improvements: project appraisal report. Coastal Engineering UK Ltd for Conwy County
Borough Council, August 2006.

Toe structures management manual


235
C2 CASE STUDY: Corton, Suffolk
Courtesy of Waveney District Council

C2.1 Identification of the problem


Corton village is located on the Suffolk coast, about five miles north of Lowestoft
(Figure C2.1). Up to about 20 m in height, the soft coastal cliffs that back the narrow
beach at Corton are rich with fossil remains and are of geological importance. Corton
was the 19th century home of the Colman (mustard) family at a time when the local
economy was based around rural activities; Corton has since become a popular
seaside holiday resort.

Figure C2.1 Location of Corton


In 1870 the Colman family had concrete and timber coastal defences built to protect
their property between the middle and the southern end of Corton. Relics from these
early defences are still evident today (Figure C2.2).
The seawall built prior to the works described in this paper originated from 1960 and
1967 and consisted of a steel sheet piled toe behind which was a mass concrete berm
and a raked concrete slab covering the lower cliff face (Figure C2.3). In addition to the
new seawall, timber framed/steel sheet faced groynes were installed along the
frontage.

Toe structures management manual


236
Figure C2.2 Detached concrete relics of 1870 Colman coast defences

Figure C2.3 Typical 1960–1967 toe and revetment structure (note date
of photograph)
In 1986, 50 metres of the 1960s seawall failed due to undermining. By this time, the
groyne field was also in poor condition due to abrasion of the steel piles. The failed
section of seawall was rebuilt together with two groynes; otherwise only routine
maintenance was applied. However, the shoreline continued to retreat and further
collapses of the seawall occurred in November 1999 and April 2000 (Figure C2.4). It
appears that the beach level dropped below the toe of the sheet piles but that the metal
ties restrained them close to the top, resulting in the piles kicking out at the toe
(Figure C2.5). Where scour penetrated behind the piles, the whole structure collapsed
with the concrete berm articulating over the piling.
Uncertainties as to the form and scale of the long-term solution led to a decision to
carry out immediate holding measures, thus giving the necessary time to properly
consider and implement more major reconstruction work. The holding measures,
carried in the autumn 2000, consisted of a double layer of 3–6 tonne rock placed along
the failed defence line. In spite of these measures, further collapses occurred at other

Toe structures management manual


237
part of the frontage in the 2000–2001 winter requiring further holding repairs in the
autumn of 2001.

Figure C2.4 Failure of the seawall in winter 2000–2001

Figure C2.5 Rotation of the toe piling (photograph taken after the 1999
failure)

C2.2 Appraisal
The Lowestoft Ness Shoreline Management Plan 3B (1996) had advised a coastal
defence policy of ‘Hold the Line’, although the economic case was reported to be

Toe structures management manual


238
marginal. The subsequent Strategy Study (Halcrow 1998, adopted 1999) had
concluded that Hold the Line at Corton was not sustainable in the long term but might
be viable in the short term. The Strategy Study did not, however, consider the
intangible benefits of defence such as recreation and access to the beach.
The Flood Hazard Research Centre at Middlesex University was therefore
commissioned to carry out a specialised study on intangible benefits – the Corton
Village Study. Taking the findings of the Corton Village Study into account, the Project
Appraisal Report prepared by Halcrow (completed 2002) examined a range of options
based principally around two defence methods, rock revetment and beach recharge
with groynes, in combination with variants on the level of intervention and period for
holding the line. Beach management options were significantly more expensive than
those using a rock revetment. The recommended preferred option was to Hold the Line
to year 20 using a rock revetment, to be followed by management of an eroding coast;
the scheme included beach access to be maintained for 20 years. The preferred option
had a benefit/cost ratio of 1.11.
The PAR enjoyed a high level of consultation with locally elected members and the
local community. The understanding and awareness brought about by this liaison
resulted in the preferred option being generally accepted as a reasonable compromise
between the competing objectives. The use of rock was not challenged during
consultation.
Further important consultation took place with Natural England with particular reference
to the geological SSSI. This consultation influenced the design of the rock revetment
such as to allow sufficient wave overtopping to prevent vegetation of the lower cliff face
where it was required to keep geological features exposed.

C2.3 Outline design


Figure C2.6 shows a schematic of the outline design.

Figure C2.6 Schematic of the outline design

Toe structures management manual


239
C2.4 Detailed design
The original design concept supposed that the toe would extend to -1.8 mODN.
Between the tender and procurement stage, however, there was a worsening of
coastal conditions which suggested that better protection was needed for the cliff part
of the works. Consequently, in rebalancing the design to retain a similar overall cost,
the toe level was raised to -1.0 mODN though, understandably, this raised concerns
about increased future risk of toe damage. To counter this, future costs for subsequent
maintenance repairs were included in the strategic budget. Further to this, steps were
taken to secure the leading edge of rocks as safeguard for the time when the toe
became exposed; the geotextile was wrapped around the toe rocks and held in position
by the next line of rocks (a so-called Dutch Toe). However, wave attack whilst the toe
was still exposed resulted in the fabric being torn and this idea was therefore
abandoned.
Figure C2.7 shows two design cross-sections through the revetment. The upper
section shows the cross-section of the new defence where the original construction
was totally destroyed (see Figure C2.4); the lower section shows the typical section to
the south of this where, partly for cost reasons, the original cliff apron slab survived and
was incorporated into the section.

Full reconstruction over northern part

Revetment fronting remaining 1960s structure


Figure C2.7 Design cross-sections

C2.5 Construction issues


The construction works were put out to tender in August 2002 and J T Mackley & Co.
Ltd was appointed to undertake the construction in February 2003. In the interim period

Toe structures management manual


240
they worked with Waveney District Council (WDC) in the development of cost saving
design modifications.
At the time of works construction, the beach level had fallen significantly below that
anticipated at the time of tendering. This required a new plan for undertaking rock
placing and resulted in an extended period for completing the works.
Rock was delivered to the beach by barge (Figure C2.8). Figure C2.9 shows rock
recovery operations. During rock recovery, an excavator sank into an isolated area of
extremely low-bearing capacity ground; it remained there for 17 days by which time it
was a write off (Figure C2.10).
There were concerns regarding the topping up of the rock revetment installed in 2000.
The project costs did not allow for removal and replacement of all the rock; moreover,
removal of the rock presented a serious risk to safety as it was providing support to the
failed wall behind. Various options were considered, it being concluded to concentrate
on discrete areas where voids could be filled with imported rock, and individual rocks
replaced as thought necessary.
The completed scheme is shown in Figure C2.11.

Figure C2.8 Rock delivery

Figure C2.9 Rock recovery

Toe structures management manual


241
Figure C2.10 Excavator loss

Figure C2.11 Finished works

C2.6 Post construction issues


A storm in December 2003 damaged the upper concrete slab over about 200 m of the
southern frontage. The rock revetment did not suffer damage. The repairs to the
existing slab part of the structure had, knowingly, been minimal in this area owing to
the need to contain the scope of works within the available finance. The damage was
subsequently funded and repaired.

C2.7 Lessons learnt


The project dealt with the potentially emotive topic of the managed withdrawal of
coastal defence, albeit that continued defence and maintenance of beach access was
assured for 20 years. This sensitive issue was appropriately handled through high level
consultation and co-operation between the stakeholders and the public.

Toe structures management manual


242
C2.8 Acknowledgments
We would like to acknowledge the invaluable advice and assistance provided by Paul
Patterson, Waveney District Council, in the preparation of this case study. The case
study is substantially derived from the paper by Patterson et al. (2004).

C2.9 References
Halcrow, 1998. Gorleston to Lowestoft Ness Strategy Study, for Waveney District
Council, 1998.
Halcrow, 2002. Corton village coast protection project appraisal. September 2002.
Patterson, P., Glennerster, M. and Millar, G. 2004. Corton coast protection.
Presentation to 39th Defra Flood and Coastal Management Conference, July 2004.

Toe structures management manual


243
C3 CASE STUDY: South Beach, Lowestoft
Courtesy of Waveney District Council

C3.1 Identification of the problem


By 2005, Waveney District Council (WDC) had become increasingly concerned about
falling beach levels at Lowestoft South Beach (Figure C3.1). The area of particular
concern was 200 m long in an area known as ‘Children’s Corner’ (Figure C3.2). The
lowered beach levels presented a threat to the seawall and jeopardised the amenity
value of the site. Given the urgency of the situation, WDC was keen to study and
implement short-term remediation works (minimum five years’ service life) in advance
of an eventual longer term strategic solution.

Figure C3.1 Location map – Lowestoft, Norfolk


There were two types of seawall in the area of concern (Figure C3.3):
• a concrete wall (90 m northern half) dating from 1922 (actually an
encasement of the original old flint wall);
• an old flint wall (110 m, southern half) dating from c.1880, with a concrete
core of variable quality, faced with grouted flint cobbles.
For the purposes of this case study, only the old flint wall is discussed, though the
same approach and fundament design solution was applied to both.

Toe structures management manual


244
Figure C3.2 Location of the seawall and toe protection works,
Lowestoft South Beach

Figure C3.3 Old Flint Wall and concrete wall

C3.2 Appraisal
An appraisal study was commissioned whose principal objectives were to:
• review the existing assumptions on risk and the nature of potential defence
failure;

Toe structures management manual


245
• identify and assess viable response options leading to the selection of a
preferred solution.
Children’s Corner and particularly the northern corner had historically suffered low
beach levels. This was due, in part, to the more intrusive aspect of the promenade at
its root with the harbour’s South Pier, coupled with the concentration of incident and
reflected wave energy. Other factors, including the influence of offshore banks on
nearshore wave climate and the impact on the harbour on littoral drift, were examined
in the appraisal study.
The risk of wall failure was evaluated in accordance with standard geotechnical
procedures for a gravity wall. The required factors of safety (FOS) for overturning and
sliding were defined as follows:
FOS (overturning) = Resisting moment about wall toe
Disturbing moment about wall toe

FOS (sliding) = Resisting horizontal force


Active horizontal force
BS 6349 (1988) recommended minimum factors of safety of 1.5 for overturning and
1.75 for sliding.
The principal variable in the analysis was the assumed level of the beach. Sea level
had a lesser effect on the result. For the purposes of the appraisal, two beach levels
were examined:
(i) The level of the bottom of the Old Flint Wall (0.36 mAOD), that is,
the level at which the wall would become undermined
(ii) The height of beach needed to provide sufficient passive resistance
to ensure stability in accordance with the factors of safety shown
above.

Site investigation showed that the Old Flint Wall was rather variable in cross-sectional
geometry (Figure C3.4). The results of the geotechnical analysis are shown in
Table C3.1.

Figure C3.4 Old Flint Wall sections

Toe structures management manual


246
Table C3.1 Factors of safety for the Old Flint Wall

Beach level (mODN) FOS FOS sliding


overturning
Sections A and B
Case (i) 0.36 0.60 0.64
Case (ii) 2.17 1.50 1.75
Section C
Case (i) 0.36 0.46 0.54
Case (ii) 2.25 1.50 Not calculated

In the case that the beach level was drawn down to the base of the wall (that is, to
0.36 mODN), the wall would fail in respect of both overturning and sliding. This low
beach level is unlikely to be reached by 2010 given the outlook for beach lowering.
Nevertheless, levels close to this critical point could be encountered.
The calculations showed that the wall needed a beach level above 2.17 mODN (that is,
approximately 1.81 m above the base of the wall) to ensure stability (to required factors
of safety) in respect of overturning. In the case of resistance to sliding, the lowest
acceptable beach level was less onerous at 1.15 m above the base of the wall. In
conclusion, and being mindful of the expected beach trends, the risk of instability (to
2010) was considered to be moderate. However, it was also considered prudent to
allow for the possibility that any short-term mitigation might be required to secure the
wall for a longer period (for example, 15 years rather than five). This prospect
significantly increased the likelihood of a failure and possibly a breach of the
unprotected wall within the given time horizon.
The appraisal examined a range of options for mitigating the risk to the Old Flint Wall
including measures directed at maintaining a higher beach, alongside measures
directed at securing the wall in the case that the beach continued to fall:
• beach nourishment (recycling);
• beach nourishment (imported);
• short timber groyne;
• Armorflex revetment;
• stepped concrete seawall with piled toe;
• concrete surcharge;
• sheet piled toe;
• rock toe;
• concrete armoured toe;
• ground anchors.
Having given due consideration to the technical merits, environmental factors (including
amenity) and financial issues, the preferred option was the use of a sheet piled toe to
secure the wall against geotechnical instability and undermining.

Toe structures management manual


247
C3.3 Outline design
The fundamental design consisted of a six-metre long steel sheet piled wall (length
selected for efficient cutting and use of delivered pile lengths), together with a concrete
thrust block (apron) cast in situ between the toe of the Old Flint Wall and the newly
driven sheet piled wall. The outline design was based on the conventional methods for
cantilever sheet pile design. The loadings are listed in Table C3.2.

Table C3.2
Type of pressures Loading
Active • Lateral load transmitted through the thrust block, being
the net load after friction losses from the old wall and the
thrust block were removed.
• Active geotechnical loads induced by the vertical loadings
of the thrust block and the Old Flint wall.
• Active geotechnical loadings induced by the soil profile
on the landward side.
• Water differential.
Passive • Passive geotechnical resistance induced by the soil
profile on the seaward side.

The outline design provided the basis for consultation. Consultations were undertaken
as follows:
• Environment Agency re need for a notice under the Coast Protection Act
1949 and consent under Water Resources Act – neither needed in this
case;
• Marine and Fisheries re FEPA licence – not needed in this case;
• planning consent with particular reference to heritage and amenity;
• land ownership.
The scheme was classed as maintenance and the consents were very straightforward
given the nature and extent of the works.
Both outline and detailed design were carried out in compliance with CDM objectives
and regulations. Among other matters this identified the potentially increased risk due
to falls from the promenade edge onto the concrete thrust block (apron) when it
becomes exposed.

C3.4 Detailed design


Figure C3.5 shows the detailed design cross-section. Particular issues and areas of
detail considered within the detailed design included the following.
The outline design included a pile cap (a concrete wrap over including steel
reinforcement) which was eliminated in the detailed design due to:
• potential frailty;
• eventual corrosion of reinforcement;

Toe structures management manual


248
• difficulty in forming it in areas of detail (for example, around the circular
ramp sections).
In the simplified design, the concrete thrust block was either simply made to fill the void
up to the level of the pile, or it was shuttered back from the pile edge at a higher level in
order to yield a sufficient mass of concrete (and hence contributory friction resistance
to the Old Flint Wall).
Drainage was included by way of holes cut in intermittent pile pans, augmented by
allowing every fifth pile to be curtailed at 3 m depth (instead of 6 m) – the reduced
passive resistance was allowed for in the design.
The plan included details of complicated features. In particular the toe detail around the
circular ramps was examined with particular reference to the radius of curvature
(actually the angular compliance at clutches) of alternative pile types.

C3.5 Construction issues


Figures C3.6 and C3.7 show the new toe detail during construction. Particular issues
that arose during construction included the following.
Excavation of the old wall toe revealed considerably greater variability in the toe depth
than had been anticipated on the basis of the earlier trial pits; the problem was
overcome by varying the size and hence mass and friction resistance of the thrust
block, keeping the piling the same. This apparently straightforward alteration required
careful consideration as the block’s depth, width and top level were all constrained
within practical limits.
Flowing groundwater emerged over a small section of excavation; this was countered
by increasing the drainage across the pile section as noted above.
One of the ramp sections was damaged (but did not collapse) during excavation; here
the priority was to continue to secure the toe and deal with the ‘topside’ damage
subsequently.

C3.6 Post construction issues


Following construction there has been a need to manage the increased hazard of a fall
from the promenade edge onto the concrete apron when it is exposed. This has been
achieved by banking up the sand to cover the apron. Although the prospect of an
increased risk was recognised earlier in the project, the issue was aggravated by the
need to increase the depth of the concrete apron in places to counter the
(subsequently revealed) shorter depth of the old seawall.

C3.7 Lessons learnt


Possibly the most significant lesson to emerge from this project was the need for early
(during detailed design stage) and more comprehensive trial pits to better determine
the variability of the depth of the existing wall. Discovery of the extent of the toe
variations during construction led to the need for rapid decision-making and re-analysis
of the design section while the works were in progress. In the event, once the likely
extent of the variations was realised, a number of additional pits were dug in order to
pre-empt further discovery and to enable the ongoing works’ details to be planned
accordingly.

Toe structures management manual


249
Further to the above point, when working with old and poorly understood structures
such as the Old Flint Wall, it would be worth using enhanced factors of safety. This
would then offer greater flexibility in any adaptive engineering that might ensue
following discoveries made during construction.

Toe structures management manual


250
Figure C3.5 Old flint wall design cross-section

Toe structures management manual


251
Figure C3.6 Toe piling being installed (at concrete wall section)

Figure C3.7 Shuttering for concrete thrust block (apron) – old flint wall
section (note water ingress and exposure of the toe of the existing wall)

Toe structures management manual


252
C3.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Paul
Patterson, Waveney District Council, in the preparation of this case study.

Toe structures management manual


253
C4 CASE STUDY: Holme Dunes, Norfolk
Courtesy of the Environment Agency

C4.1 Identification of the problem


Holme Dunes are located about 5 km to the east of Hunstanton on the Norfolk Coast
(Figure C4.1). The dunes provide a natural flood defence to a number of properties,
agricultural land and the Holme Dunes National Nature Reserve. The nature reserve,
managed by Norfolk Wildlife Trust, is a designated SSSI, SPA, Ramsar site and
Biosphere Reserve noted for its dune system and freshwater marshes. Figure C4.2
shows the site photographed in 2009 (note the more recent brushwood dune
protection).

Figure C4.1 Location of Holme Dunes


In the mid-1990s, concern was expressed about the perceived erosion of the dunes in
the area referred to as the Firs (named after the large property just inland of the
dunes). A report 11 by the Department of Marine Sciences and Coastal Management at
the University of Newcastle commissioned by the Environment Agency suggested that
erosion was due to larger waves reaching the toe of the dunes, noting that this
appeared to be linked to recent (1991–1996) evolution of the ebb delta at Thornham. It
was thought that loss of tidal volume of the inlet had resulted in a weakening of the ebb
delta which, in turn, had resulted in a reduction in protection to the shore due to
changes in wave refraction and shoaling.

C4.2 Appraisal
The North Norfolk Shoreline Management Plan had recommended that property should
be protected in the short term by strengthening the eroding dune system using
construction that was acceptable environmentally and financially. The economic
benefits of coast protection at the site were very limited, thus implying that any scheme
would have to be low cost. There was, nevertheless, a statutory requirement to protect

11
Holme Dunes: coastal Processes and Geomorphology Study

Toe structures management manual


254
the designated conservation interests. A formal detailed appraisal was suspended
pending the identification of suitable options that might, at the outset, satisfy these
fundamental criteria. Options considered included:
• recycling of sand and shingle from a nearby bar (Gore Point), but this was
not allowed because of potential damage to habitats there;
• beach recharge, as well as detached breakwaters and groynes, but high
costs relative to the limited benefits precluded these more expensive
options.
The Environment Agency worked with English Nature and Norfolk Wildlife Trust in
formulating a potentially viable option. The concept of beach drainage as a means of
shore stabilisation came to light following discovery of the principle in Denmark.
Invented by the Danish Geotechnical Institute, the system was licensed, designed and
marketed by MMG Beach Management Systems (UK). The prospect of a workable
system at Holme was attractive, not least because it appeared to offer a solution that
entailed no structural interference with the dunes. However, should a permanent
installation to be installed there would be royalty payments to make in respect of the
patent rights held by MMG. Given that the method was still highly innovative with only a
limited history of applications at this stage (there had been only one other application in
the UK), it was decided to carry out a trial for which MMG kindly agreed to waive the
royalties.

Figure C4.2 The site in 2009 – note the recently installed brushwood
dune protection

C4.3 Outline design


The design was prepared by MMG for the size of perforated pipe, geotextile wrapping,
depth of pipe and pump requirements. Its ‘Beach Management Systems’ brochure
described the theory of beach drainage thus:
‘with a sandy beach, lowering the water table in the beach face eliminates
buoyancy factors and vastly reduces the lubricating effect between the
grains, thus restoring the frictional characteristics of the sand. Furthermore,

Toe structures management manual


255
the percolation of ‘swash water’ into the relatively dry beach encourages
sand to settle out at the beach face’.
No further details were available (for this study) of the bespoke design considerations
or workings.
A grading curve of a beach sample taken from the site (Figure C4.3) shows it to
comprise predominantly sand with some gravel.

Figure C4.3 Example grading curve

Basically, the scheme comprised a 200 m shore parallel drainage pipe set about 30 m
from the toe of the dunes. In fact two drainage pipes were used to yield the overall
capture width of 200 m. These two pipes drained into a sump which contained a pump.
The pump discharged the collected water back to sea through a higher outfall pipeline.
Figure C4.4 shows cross-sections for the west limb together with a schematic section
taken through the pump sump.

Toe structures management manual


256
Figure C4.4 Beach sections

C4.4 Detailed design


The drainage lines consisted of 200 mm high density polyethylene (HDPE) perforated
pipes, 100 m long in each direction running broadly parallel to the beach surface, as
indicated in Figure C4.4.
For the initial trial, the shore parallel parts of the pipes ranged in level from +0.9 mODN
at their seaward ends (100 m east and 100 m west) to +0.7 mODN at 50 m east and
50 m west. These levels resulted in a depth of sand cover of about 1.5–2.0 m
depending on beach level and pipe gradient.
In an attempt to make the beach drainage more effective, an additional pipe was
installed after a few weeks of operating the system with a reduced depth of cover of
between 0.5 and 1.0 m as illustrated in Figure C4.4. The levels shown in Figure C4.4
can be compared with the tide levels for Hunstanton (posted in 1996) of MHWS +3.7
mODN, MHWN +1.9 mODN, MLWN -1.2 mODN and MLWS -2.8 mODN. Hence, the
drainage pipe inverts were installed at a level between mean tide level and MHWN,
while the beach (at the time of illustrated survey) was at a level between MHWN and
MHWS.
The drainage pipes fed seawater back to the 4.5 m deep sump (Figure C4.5). A pump
located in the bottom of the sump extracted the collected seawater and returned it to
sea via the outfall pipe. The pump was rated at 55 l/s (against a 5 m head) and was
powered by a generator located on the dunes.

C4.5 Construction issues


The most significant construction issue was thought to be a problem with floatation of
the sump chamber. This was corrected by ballasting the structure using concrete rings
attached near the top (Figure C4.5).

Toe structures management manual


257
C4.6 Post construction issues
A beach monitoring programme was set up and maintained for about three months
(February to April 1997) – the period of the trial. This was based around a series of
beach profiles located at the centre line and at ±50, ±100 and ±200 m from the centre
line. The outer profiles were taken as a reference check on the ambient beach level to
distinguish changes that might be linked to the beach drainage scheme from natural
changes happening along the whole frontage. Visual observations were also made.
Initial observations, following installation of the first drainage pipes at 0.7–0.9 mODN,
showed little or no effect in terms of drying of the beach. A second pipe was installed at
a higher level (see Figure C4.4). After installation of the second pipe, the beach surface
in the vicinity of the pipe appeared to be drier but, over the trial period, there was no
change to beach levels that could be confidently attributed to the drainage system.
During construction of the sump and operating the system, the water levels in the
beach were noted to be very high at all times and states of the tide, so much so that
the beach drainage system would be working by siphon effect even at low tide.
By the end of the formal trial a decision was needed whether to adopt a permanent
system, which would have required payment of the licence royalty or curtail the project.
Given the results up to that point, the latter was opted for, although it is understood that
MMG funded the system’s continued operation for a short further period.

C4.7 Lessons learnt


Based on the formal trial, the system did not appear to afford an improvement in the
beach level. It should also be noted that three months is a very short period over which
to conduct such a trial, as perceived performance depends not just on efficacy of the
system but also on the pressures applied to it in terms of wave and tidal factors.
Nevertheless, upon completion of the trial, as there was no noticeable effect on the
frontage and given that a permanent scheme would have attracted royalty payments,
the scheme was abandoned.
Before embarking on any new permanent scheme, a considerable amount of technical
data need to be gathered and evaluated before proceeding to detailed design

C4.8 Acknowledgments
We would like to acknowledge the invaluable advice and assistance provided by Paul
Miller of the Environment Agency in the preparation of this case study.

Toe structures management manual


258
Figure C4.5 Sump detail

Toe structures management manual


259
C5 CASE STUDY: Overstrand, north Norfolk
Courtesy of North Norfolk District Council

C5.1 Identification of the problem


Overstrand is located 3 km to the east of Cromer on the north Norfolk coast
(Figure C5.1). The site in question is located at E3 on Figure C5.2, at a mild turning
point in the frontage which is defended both to east and west. The long-term average
rate of undefended cliff retreat here is about one metre per year. The sandy foreshore
falls away at a mild slope of about 1 in 50 and covers chalk some 2 m below the sand
level at the seawall.

Figure C5.1 Location of Overstrand, north Norfolk

Figure C5.2 Site location

Toe structures management manual


260
Prior to construction of the present structure, the coastal defence consisted of a sheet
pile toe behind which there was a lower concrete apron, concrete retaining walls and a
higher concrete apron, all containing the coastal cliff to a height of about 25 mODN.
The old sheet piles were badly corroded and there were holes in some of the pile pans.
The whole structure was believed to date from c.1908. Figure C5.3 shows part of the
wall that did not fail.

Figure C5.3 The original wall


During December 1997, a 25 m length of the lower apron collapsed leading to part
failure of the main seawall and promenade. This initial failure put the whole of the west
promenade in jeopardy. It was believed that the old piled wall failed through corrosion
and apron failure leading to loss of backfill.
Further failure was expected to follow, which would progress to erosion of the cliff. In
order to make the structure safe in the short term, 100 m3 of concrete were placed in
the void behind the wall. This provided a short-term solution while a permanent scheme
was planned and constructed.

C5.2 Appraisal
The engineers report, Overstrand Coast Protection Scheme 973 (1999), served the
purpose of the Project Appraisal. It was recommended that the project to restore the
Overstrand seawall be undertaken as an emergency scheme under section 5(6) of the
Coast Protection Act 1949. The scheme was to comply with a number of limiting
factors, that is:
• works were to be contained within the financial constraints and have a
minimum benefit/cost ratio of 1:5;
• works were to be designed to take account of the degree of difficulty in
working below the high water mark in a hostile marine environment;
• the scheme was to comply in all respects with the criteria required by the
then Ministry of Agriculture, Fisheries and Food (MAFF) including that of
being environmentally acceptable.

Toe structures management manual


261
To expedite the process of option identification, discussions were held with several
experienced contractors to elicit their views and recommendations on design and
construct schemes. The various options considered included:
• do nothing (baseline);
• reconstruct the existing structure – this was rejected due to the inherent
residual shortcomings of the original structure (for example, inadequate
foundations) together with the increased pressures due to beach lowering
making reconstruction major and costly endeavour;
• construct the lower apron at a lower level than before – this was rejected on
the grounds of practicality because of the need for substantial excavation
below the existing wall together with the downwards extension of the wall
and foundation, thus requiring a lot of intertidal working;
• construct a rock revetment – this was examined but considered to be
uneconomic in this case because of the short length of the works (about
100 m) and hence the relatively high mobilisation cost;
• sloping or stepped concrete apron – this was a preferred option, the choice
between in situ concrete or pre-cast units being one of practicality.
Consent was required from The Crown Estate. English Nature was notified as the
western 20 m of the site fell within the Overstrand Cliffs SSSI and the land beyond
Overstrand village was an Area of Outstanding Natural Beauty (AONB). There were no
significant environmental issues to consider other than rectification of the failed
frontage and public access.
The chosen scheme delivered a benefit/cost ratio of 2.2 and thus qualified for MAFF
grant aid.

C5.3 Outline and detailed design


North Norfolk District Council was keen to use pre-cast concrete in this location to
overcome the problems associated with the placing and curing of concrete in this highly
exposed location. Four contractors were invited to discuss the Council’s design
proposals before the final design was prepared and the contract was put out to tender.
McKinnon Construction won the tender.
The final design (Figure C5.4) comprised a sheet piled toe with pre-cast concrete
superstructure. The components were:
• steel sheet piles driven into the chalk – these cantilevered piles were
considered to be secure, even in the case that the beach lowers to the
chalk level as their stability did not rely on having a depth of beach to
provide support;
• a concrete beam located behind the piles;
• pre-cast concrete step bridges;
• pre-cast concrete steps.

Toe structures management manual


262
Figure C5.4 Design cross section drawing of the Overstrand defence
(extract)

C5.4 Construction Issues


The use of substantially pre-cast construction minimised in situ works and hence
exposure to action of the sea during installation.
The end details, where the section had to tie into existing remaining details
(Figure C5.5), were most difficult both in terms of design and construction.

C5.5 Post construction issues


Figures C5.5 and C5.6 show the finished construction.
The front edges of the steps have become rounded (being sharp edged when cast). In
retrospect, it would have been better to allow for corner chamfers in the shuttering.
The new construction has extended the life considerably of the defence. The
Overstrand Foreshore Study ET 4252(2008), carried out for the Council by St La Haye
Consulting Ltd, assigns the new section a residual life of 35–40 years allowing for
beach degradation. However, the adjacent wall sections have considerably shorter
estimated residual lives.

Toe structures management manual


263
Figure C5.5 End detail

Figure C5.6 Finished toe and wall

C5.6 Lessons learnt


The scheme provides amenity benefits and has been successful. The use of pre-cast
concrete provided a quick economic solution in this case. Better attention to the
detailing of the steps (edges) in future should be noted.

Toe structures management manual


264
C5.7 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Brian
Farrow, North Norfolk District Council, in the preparation of this case study.

Toe structures management manual


265
C6 CASE STUDY: West End Dovercourt, Essex
Courtesy of Tendring District Council

C6.1 Identification of the problem


Figure C6.1 shows the location of Dovercourt, just to the south of Harwich on the
Essex coast. The site had been used as a landfill before Tendring District Council
(TDC) took possession. The waste material was isolated from the sea only by a clay
covering and embankment. Shoreline retreat presented a growing risk of erosion of the
embankment containing the fill. Had this been allowed to continue then penetration of
the embankment would have resulted in the release of landfill waste into the marine
environment.

Figure C6.1 Dovercourt site location


The site fronts a conservation area for ground nesting birds and is immediately
adjacent to Hamford Water SSSI. It was essential therefore to prevent contamination of
the shore.
Earlier attempts were made to arrest erosion of the spoil site by placing interlocking
mattresses on the embankment (Figure C6.2). However, these were inadequate for the
given exposure and were soon destroyed by storm wave action (Figure C6.3). An
alternative low-cost revetment was sought.

C6.2 Appraisal
No formal appraisal was carried out but a number of mitigation options were
considered. Although a comparatively new technique to the UK, an asphaltic revetment
appeared to offer good value. It would be sufficiently robust for the comparatively
benign wave climate and, for the Dovercourt location, its (sometimes undesirable)
appearance was not a significant issue.

Toe structures management manual


266
Figure C6.2 Flexible armoured revetment (earlier defence structure)

Figure C6.3 Failed flexible armoured revetment

C6.3 Outline design


The first phase of the protection works were carried out in the early 1990s, being
extended in 2001. Figure C6.4 shows a plan layout of the revetment. The revetment
extended to the edge of the SSSI designated area.
Specialist contractor, Hesselberg Hydro, was commissioned to design and construct
the works.
The toe was to be buried and backfilled with disturbed clay. Given the nature of the
site, there was no sand covering to the clay. The exposed clay (Figure C6.2) was

Toe structures management manual


267
susceptible to continued lowering. The toe structure was therefore designed to rotate
once the underlying ground level was reached.

C6.4 Detailed design


A design drawing is reproduced in Figure C6.5. The construction consisted of:
• regrading of the shore to adopt the revetment slope and deeper excavation
for the toe;
• a 200 mm layer of open stone asphalt laid over a 150 mm layer of lean
sand asphalt on a 1:2.5 slope (150 mm layer of open stone asphalt used in
the more sheltered sections);
• a 350 mm thick grouted stone slab placed over a geotextile on a 1:15 slope
to form the buried toe (the main revetment section extended into the
excavation to form buried toe in the more sheltered areas).

C6.5 Construction issues


Beach excavation and toe construction were undertaken using plant on the foreshore.
Plant operated from the crest of the embankment for placing asphaltic materials on the
revetment. Construction of the works was straightforward except for some damage to
the top landscaped area due to plant movement; see Figures C6.6 and C6.7.

C6.6 Post construction issues


The revetment is subjected to wave attack which, together with shingle and debris,
results in abrasion of the asphalt. The interface between the shore and the open stone
asphalt revetment suffers most wear. Erosion tends to happen in pockets and seems to
return to the same places.
Monthly monitoring of the beach is carried out to identify problems and repairs are
usually carried out once per year. Repairs are made either by appointed contractors or
by TDC’s in-house resources – typically one repair is needed per year. Maintenance is
an important management consideration but is not a problem providing that a regular
monitoring and repair regime is adhered to.
Recently, TDC experimented with the use of a polyurethane binder (instead of bitumen)
for repairs. The advantages of this are:
• it is basically clear;
• there is no need for a hot pot;
• the material is not affected by water except for some discolouration if the
binder gets wet too soon.
To date, the buried toe has not been exposed and so its ability to articulate with falling
beach levels has not yet been tested.

Toe structures management manual


268
C6.7 Lessons learnt
The use of asphaltic construction provided a good economical solution in this case. It is
accepted that, due to its appearance, the use of asphalt could be location sensitive but
this was not a problem at the Dovercourt site.
A good inspection and repair regime is essential.

Toe structures management manual


269
Figure C6.4 Plan extent of revetment

Toe structures management manual


270
Figure C6.5 Design drawing of asphalt revetment

Toe structures management manual


271
Figure C6.6 Revetment construction

Figure C6.7 Toe construction

C6.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Bob
Howell, Tendring District Council, in the preparation of this case study.

Toe structures management manual


272
C7 CASE STUDY: Teignmouth to Dawlish sea wall
Courtesy of Network Rail

C7.1 Identification of the problem


During the early months of 1986, the sand level at a section of the sea wall supporting
the main Paddington–Penzance railway disappeared completely between Smugglers
Cove and Sprey Point on the line between Dawlish and Teignmouth (Figure C7.1).
Although fluctuation in sand levels can be quite significant throughout the length of the
sea wall, total loss at this location had not been experienced within the memory of
those involved with the wall at the time.

Figure C7.1 Location of Dawlish and Teignmouth


The sea wall at this location is a vertical masonry wall, with a small toe section of
approximately 1.5 m2 in cross-section probably constructed at some time after the main
wall. The bedrock onto which the wall is founded is comprised of soft red sandstone
interlaced with bands of breccia. This sea wall was originally constructed in about 1840
by Isambard Kingdom Brunel as part of the South Devon Railway and has undergone
improvements to various sections throughout its history.
With the loss of the beach material in 1986, erosion quickly occurred in three areas to
the extent that the wall became undermined for the full cross-section, resulting in the fill
material which supported the railway line being ‘sucked out’ by successive tides.
Consequently, the line was closed for a period of approximately two weeks, and it was
touch and go as to whether these three sections of wall would be lost.

C7.2 Appraisal
The first objective facing British Rail was to secure the wall and restore rail services.
Block stone from Meldon Quarry was delivered to site by rail on the ‘Up’, London-
bound line, which had not become undermined. This was placed into the voided areas
in concrete and the breached wall face repaired with concrete filled sand bags, allowing
the supporting ballast to be replaced, and the service restored.

Toe structures management manual


273
It then became clear that the interface between the soft bedrock and the wall toe
required underpinning and sealing to secure the wall and to mitigate against future
sand loss. This was achieved by providing an extension to the toe in wet sprayed
concrete (shotcrete) of approximately 1.5 m2 in cross section. The principle behind this
approach was that should total sand loss occur in the future, the bedrock would have to
erode beneath the sprayed concrete, generally 1.5 m wide, before the main wall was
affected, by which time the sand level would hopefully restore naturally – affording
protection to the toe once more.
A length of wall of approximately 850 m was treated and the solution provided
protection for a period of about 10 years. During this period, the sand level continued to
fluctuate, with total loss occurring in some areas for a short period. For the majority of
the length of wall, the erosion of bedrock was approximately 100–200 mm, but in one
particular area, the loss was over 800 mm. By this time, ownership of the network had
passed to Railtrack under privatisation of the rail industry, and a more lasting solution
was being considered, with the principal objective of securing a solution ensuring a
safe and reliable rail service. The options under review are summarised in Table C7.1.

Table C7.1 Possible solutions

Option Advantages and disadvantages


Rock armour • Would not have found favour with the local
authority due to loss of beach amenity.
• Costly due to poor access.
• Concern about loss of fine material through
holes opening in main wall, which could not be
properly accessed after the placement of rock
armour. This could lead to difficulty in
maintaining track geometry.
Sheet piled toe • Difficulty in driving in bedrock.
• Limited longevity.
• Poor access for size of plant required.
Concrete toe • Preferred solution.
construction
Groyne field • Already in place and would be maintained.

The area where rock erosion had been greatest was selected for a trial construction.
The term sea wall maintenance contractor was commissioned to design and construct
a concrete toe solution, utilising a consultant experienced in coastal engineering.

C7.3 Outline design


The principal design comprised a mass concrete toe, cut 1.5–2 m into the bedrock,
approximately 1 m wide, with three mass concrete steps each approximately 1 m ×
1 m, as shown in the cross-section (Figure C7.2). The principal considerations were:
• to provide long-lasting protection against the effects of bedrock erosion;
• to improve passive resistance against sliding and overturning;
• to minimise sand loss;
• to allow continued access to the masonry wall for inspection and repair.

Toe structures management manual


274
Consultations were undertaken with organisations including the Environment Agency,
The Crown Estate and Defra, and planning consent was obtained.

C7.4 Detailed design


After successful completion of the trial, the project was extended to cover the majority
of the sea wall between Dawlish Warren and Teignmouth. The more vulnerable areas,
namely between Smugglers Cove and Teignmouth (850 m), and a section at
Rockstone between Dawlish Warren and Dawlish (28 m) received the full strengthening
profile as shown in Figure C7.2, while most of the remaining sections, totalling 1,690 m
received the partial strengthening profile as shown in Figure C7.3. The decision to
adopt the lesser profile was based on the relative stability of the sand level and
exposure of the length concerned. Other areas of the sea wall either did not require a
strengthened toe, had been strengthened by other means, or the existing arrangement
or position did not lend itself to the design.
Locally, provision in the detailed design was made to extend existing steps to the
beach downward, to provide ramped sections where necessary, to negotiate features,
and to make provision for watercourse outlets. Figures C7.4 and C7.5 show the full
foundation arrangement at Smugglers Cove and Rockstone respectively. Figure C7.6
shows partial foundation strengthening.

C7.5 Construction issues


The main issues at the construction stage were:
• de-watering during the outgoing tide period;
• avoiding undermining the existing wall during excavation;
• trimming the bedrock to a suitable constancy and achieving a clean cut into
the bedrock for the deep foundation;
• transporting large volumes of concrete from delivery point to site;
• constructing formwork able to withstand the incoming tide.

C7.6 Post construction issues


No significant issues were identified.

C7.7 Lessons learnt


The most important lesson learnt was the vulnerability of vertical sea walls to sand loss
and toe erosion. This prompted an ongoing need for frequent inspection, monitoring of
sand levels, an adverse weather mitigation plan and consideration of suitable methods
of keeping beach sand in place.

C7.8 Acknowledgements
We would like to acknowledge the invaluable assistance of Peter Haigh, Network Rail,
for the provision of this case study.

Toe structures management manual


275
Full Foundation Strengthening

Typical cross section with full foundation strengthening to be found at :


205m 36c to 205m 51c
207m 46c t0 208m 09c (Smugglers Cove to Spray Point)

Mileage A B C D E Bedrock AOD Walkway AOD Track AOD


205m 48c 2.56 3.56 3.08 4.00* 2.41 N/A 5.82 5.75
208m 04c 2.19 3.31 2.97 4.00* 2.14 N/A 5.46 N/A
* Approximate
Mileages are from Paddington Station

Figure C7.2 Typical cross-section: full foundation strengthening (not to


scale)

Toe structures management manual


276
Partial Foundation Strengthening

Typical cross section with partial foundation strengthening to be found at :


205m 00c to 205m 36c
205m 70c to 205m 74c
208m 31c to 208m 55c

Mileage A B C D E Bedrock Walkway Track


AOD AOD AOD
205m 25c 2.40 3.31 5.0* 1.00* N/A N/A N/A N/A
205m 71c 2.06 3.16 5.70 1.20* 4.00* N/A 5.90 6.08
208m 39c 2.40 2.72 4.86 1.00* N/A N/A 5.93 N/A
* Approximate
Mileages are from Paddington Station

Figure C7.3 Typical cross-section: partial foundation strengthening


(not to scale)

Toe structures management manual


277
Figure C7.4 Full foundation strengthening, Smugglers Cove to Sprey
Point

Figure C7.5 Full foundation strengthening – Rockstone footbridge

Toe structures management manual


278
Figure C7.6 Partial foundation strengthening – Sprey Point to
Teignmouth (note only outer edge of single step visible at the time photo was
taken)

Toe structures management manual


279
C8 CASE STUDY: Southern Felixstowe sea
defences, Suffolk
Courtesy of Black & Veatch Ltd

C8.1 Identification of the problem


The town of Felixstowe is located on the Suffolk coast between the estuaries of the
River Deben and River Orwell (Figure C8.1). The area is low lying and there is a risk of
flooding due to high surges and tides. During the 1953 floods, 39 people died and 700
homes were damaged at Felixstowe.
The area that extends from the war memorial, just north of the pier, to Landguard
Common is referred to as Felixstowe South, forming part of the 11 km coastal defences
protecting the flood risk area of southern Felixstowe. This 2.7 km section is the most
exposed part of the Felixstowe coastline; failure of the defences here could result in
inundation of the entire southern Felixstowe flood risk area. The locations of the key
points featuring in this case study are shown in the Black & Veatch ‘Figure 1 Site Area’
map provided as an annex to the case study.

Figure C8.1 Site location


The Felixstowe South frontage can be further considered in two parts, which are
demarked by their respective jurisdictions. The northern part of the frontage is
principally the responsibility of Suffolk Coastal District Council (SCDC) under the
provisions of the Coastal Protection Act 1949, while the southern part is the
responsibility of the Environment Agency under the Land Drainage Act 1991.

SCDC frontage
The earlier coastal defences consisted of a mass concrete seawall and promenade,
constructed in 1903. The seawall was fronted by a mixed shingle–sand beach, with
timber groynes encased in concrete, spaced at approximately 30 m centres
(Figure C8.2). The beach was in a poor condition, sediment loss having been
accelerated by wave reflection from the exposed vertical wall. Severe overtopping
occurred regularly, damaging the promenade and causing localised flooding.

Toe structures management manual


280
Figure C8.2 Seawall and groynes at the SCDC frontage, 2005
Approximately 20 m landward of the promenade, there is a 1983 floodwall of a sheet
piled construction; this is clad in concrete with facing bricks (Figure C8.3). This
floodwall was in good condition in 2006 and provided a standard of protection of 1 in
200 years. However, its performance as a flood defence depended on the closure of
floodgates and, in any case, was conditioned by the frontline seawall which was
calculated to have a standard of protection of less than 1 in 1 year based on critical
overtopping for damage. Furthermore, the seawall was at imminent risk of collapse due
to undermining. It was estimated that following failure of this seawall, failure of the
crucial floodwall would occur within five years.

Figure C8.3 Floodwall and gate


The seawall, promenade and groynes are the responsibility of SCDC, while the
floodwall and gates are the responsibility of the Environment Agency.

Environment Agency frontage


The Environment Agency frontage consists of a sheet pile toed concrete seawall.
There were two variants on the type of structural defence:
• the Manor Terrace (northern) section of the wall was constructed in 1981
and has a concrete block revetment (Figure C8.4);
• the Landguard Common (southern) section of the wall was constructed in
1985 and has a stepped concrete apron.

Toe structures management manual


281
The beach was partially controlled by a series of old timber and concrete groynes. The
standard of protection of the Environment Agency floodwall was estimated to be 1 in
200 years, though this was calculated to reduce to 1 in 10 years in 100 years’ time.
Moreover, because of regular exposure above the beach level, the residual life of the
steel piled toe was considered to be shortened through deterioration. Were the toe to
fail then this would initiate failure of the whole flood defence structure.

Figure C8.4 Environment Agency seawall and flood defence, 2005

Recent problems
Failure of some groynes in 2003 on the Environment Agency frontage resulted in a
rapid fall in beach levels (Figure C8.5), requiring urgent repair works to the groynes
and blockwork.
Again, due to very low beach levels, the SCDC section of the frontage was undermined
in May 2006 resulting in collapse of 150 m of the seawall (Figure C8.6). SCDC
undertook emergency works in response to the damage, costing around £900,000 in
total (Figure C8.7). Had action not been taken, the loose backfill behind the seawall
would have washed away leading to rapid retreat, thus putting the floodwall at risk and
indeed the low lying land behind. This temporary restoration was replaced with a
permanent structural repair as part of the later main works.

Figure C8.5 Exposed toe piling at Environment Agency frontage, 2003

Toe structures management manual


282
Figure C8.6 Rotating seawall, 2006

Figure C8.7 Emergency repairs to the seawall, 2007

C8.2 Appraisal

Chronology
The Lowestoft to Harwich Shoreline Management Plan (1998) recommended a policy
of ‘Hold the Line’ for the southern Felixstowe frontage. Subsequent to the SMP, SCDC
partnership with the Environment Agency, commissioned a Strategy Study (Halcrow
2003), together with a Strategic Environmental Assessment. The Strategy Study
confirmed the Hold the Line policy, identifying the highest priority as the 2.7 km long
frontage at South Felixstowe.

Toe structures management manual


283
In 2004, Black & Veatch Ltd was appointed by the Environment Agency–SCDC
partnership to further investigate options for the Southern Felixstowe frontage and to
produce a Project Appraisal Report (PAR) (Black & Veatch 2007).
Scheme construction had been scheduled to begin in May 2006, however the PAR
could not be approved as the earlier Strategy had not been approved by Defra.
Consequently, the Strategy needed to be reviewed and updated in line with the then
latest best practice guidance (Black & Veatch 2008). The coastal defence strategy
review began in January in 2007 and approval from the Environment Agency was
obtained in February 2008.

Technical considerations
Coastal studies were carried out as part of the PAR preparation (Black & Veatch 2005).
The main conclusions of the studies were as follows:
• the shoreline (mean sea level) had been retreating landward by as much as
1.8 m per year (based on data from 1991 to 2003), equating to a rapid rate
of beach lowering;
• the frontage is exposed to two significant wave sectors – from the south
and from the north-east;
• between Cobbolds Point and the war memorial (part of the central
Felixtowe frontage), the net annual longshore drift is from north to south;
• between the war memorial and Landguard Common (southern Felixstowe)
longshore drift prevails in both directions – these drifts are highly variable,
but generally the net drift is from south to north;
• day-to-day cross-shore sediment movement is relatively small except close
to the pier, being at the focus of drifts from both central and southern
Felixstowe – however, the beach is susceptible to significant drawdown
during storm events;
• with limited sediment input, the sediment volume is reducing due to drift
towards the pier, coupled with offshore loss due to cross-shore movement –
the net result is an ongoing trend of beach retreat across the southern
Felixstowe frontage.
While the overall standard of protection against coastal flooding in 2006 was high (1 in
200 years), the main concern was that the defences would fail due to lowering beach
levels.
From a longer list of options, the following were taken forward for further consideration:
• timber groynes and beach recharge;
• rock groynes and beach recharge;
• fishtail rock groynes with visible timber root and beach recharge;
• fishtail rock groynes with buried root and beach recharge.

Environmental considerations
Consultation was undertaken throughout the development of the scheme with
Environment Agency specialists, statutory consultees, interest groups and the public.

Toe structures management manual


284
The scheme would reinstate the beach at Felixstowe and remove the existing health
and safety hazards associated with the existing structures.
Landguard Common including the foreshore is designated as a SSSI, mainly for its
vegetated shingle interest. It was therefore important to demonstrate to Natural
England that the coastal processes would not be adversely affected by the works.
There is also a local nature reserve immediately behind the promenade, close to the
Martello Tower, designated for its vegetated shingle. Natural England provided a
comfort letter giving its support to this scheme. An Appropriate Assessment was not
required. However, under the Town and Country Planning (Environmental Impact
Assessment) Regulations 1999, the proposed scheme was required to have a statutory
EIA. An Environmental Statement (including an Environmental Action Plan) was
submitted with the planning application. Planning permission was received in
September 2005.
The Martello Tower located behind the frontage approximate mid-way along the
frontage is a Scheduled Monument. The Landguard Peninsular is also a designated
Scheduled Monument due to the presence of Landguard Fort and the associated field
works. Access across this area was therefore prohibited during construction. A pre-
construction walkover archaeological survey was also undertaken to record any
features – none were found.
The identified environmental impacts of the scheme were mostly limited to the
construction phase. Disruption to the local community and tourism industry of the town
were to be mitigated as much as possible through careful planning and liaison with the
local community. The promenade (a public right of way) was to be kept open during the
construction phase with manned plant crossings.

Preferred option
The preferred option was a scheme designed to raise beach levels and to reduce
beach lowering. It would provide a minimum standard of defence of 1 in 100 years in
100 years’ time allowing for sea level rise. This was raised to a standard of 1 in 150
years, being justified by the reduction in the risk to life. This scheme would comprise
beach recharge and a series of rock fishtail groynes.
Consultees were asked to respond to the question of whether the fishtail groynes
should have exposed or buried roots. The exercise clearly showed that consultee
preference was for groynes with buried roots. The feedback to the preferred option was
positive. The configuration of the groyne field would lead to a more open beach,
allowing continued access along the beach for both pedestrians and maintenance
plant.

Economic considerations
Economic considerations examined the benefits of providing defence. The benefits
consisted of the protection to a range of assets at risk which included 960 residential
properties, 428 non-residential properties and the Port of Felixstowe (the largest
container port in the UK). Although the scheme also provides protection to other assets
(such as tourism, Landguard Common SSSI, Landguard Fort and associated Field
Works Scheduled Monument and the Martello Tower Scheduled Monument), the
economic benefit was not considered as there was ample justification from the port and
properties.

Toe structures management manual


285
The final economic analysis yielded a benefit/cost ratio of 29 as summarised in
Table C8.1.

Table C8.1 Results of economic analysis


Do Nothing Rock fishtail, buried root and beach recharge
PV costs £33,017,000
PV damage £967,044,000 £5,422,000
PV benefits £961,622,000
Net present value £928,515,000
Benefit/cost ratio 29.05

C8.3 Outline design and consents

Outline design
The preferred option was to construct a new groyne field comprising fishtail rock
groynes together with beach recharge. This option was found to be technically the most
effective in terms of retaining beach material. It was also the option most favoured
during consultation and was the most economic.
The proposed scheme involved the construction of 21 rock fishtail groynes, 200,000 m3
of beach recharge material and 150 m of sheet piling to reinstate the failed section of
seawall. Figure C8.8 (taken from the PAR) illustrates the basic form of the groyne
cross-section.

Figure C8.8 Fishtail rock groyne with buried rock root

Consents
The following consents were required:
• Agreement to the project was secured with Natural England. Section 28
consent under the Countryside and Rights of Way Act 2000 was granted by
Natural England for working within Landguard Common SSSI.

Toe structures management manual


286
• Scheduled Monument consent was obtained for the emergency access
route within the Landguard Fort and Associated Fieldworks Scheduled
Monument site.
• A FEPA license and consent under the Coast Protection Act 1949 for works
below MHWS were required for elements of the scheme below MHWS.

C8.4 Detailed design


Figures C8.9, C8.10 and C8.11 show extracts from the detailed designs drawings of
the layout and groyne structures. Figure C8.12 shows a design cross-section for the
permanent repair over the 150 m failed seawall.

C8.5 Construction issues


The works were procured through the Environment Agency’s framework agreement
with Team Van Oord – identified as the preferred contractor following a mini-
competition.

Undermining of the seawall


The scheme was designed through discussions with the contractor so that the beach
recharge works would be undertaken before construction of the control structures. This
would improve the buildability of the groynes and maximise access along the frontage
under the tidal conditions.
This meant that the existing groynes and temporary rock protection needed to be
removed ahead of beach recharge and without any beach control structures in place,
thus exposing the seawall to increased risk of undermining. To manage this risk, the
project team agreed a methodology that involved this work being undertaken in a
staged approach so that at no time was the seawall in this vulnerable state for more
than 12 hours (that is, one tide). The construction team also monitored the weather
forecast carefully and delayed unnecessary exposure of the seawall if adverse
conditions were likely. If a storm was expected or a section of wall was to be exposed
for longer than one high tide, temporary protection was placed in front of the seawall at
the toe.
There was also the risk that the wall would fail if the temporary rock armour protection
was removed. Although this risk was assessed as low (the wall had been standing up
before the rock protection was put in place), the works to remove the first section of
rock protection were carried out under close control to minimise any collapse and to
enable the team to react quickly. This work was all undertaken from the shore to
minimise any loading of the promenade area behind, which had to be closed off to
public access during these operations. The operation to remove the rock protection
proved successful and no movement of the seawall structure was encountered.
Rock protection that had been placed further south around Manor End several years
previously was left in situ as little was known about these works or the condition of the
seawall behind. The risk of failure during the rock protection removal operation was
therefore assessed as high.

Toe structures management manual


287
New seawall
For the seawall reinstatement works, it was required that sheet piles were driven in
front of the existing seawall. It was known that the seawall toe ‘kicked out’ and that over
the failed section it was leaning slightly seawards. These details therefore needed to be
determined on-site and the design refined to suit. When the seawall was fully exposed,
it was surveyed to determine its actual cross-section and any deviation from the vertical
so that the required line of the sheet piling could be determined.

Rock groynes
At the landward end, the groyne section extended down to the foundation of the
seawall. Excavation below the foundation level was therefore minimised in design
(Figure C8.11). Although the form of construction of the seawall beneath the
promenade and behind the seawall face was not known, it was known from the
behaviour of the seawall that it did not rely purely on material on the seaward side for
support. The team were confident that excavation at the seawall would not lead to
instability providing that the wall was not undermined. No issues were encountered in
the groyne construction.

C8.6 Post construction issues – performance


The frontage has now been subjected to sufficient wave action to develop a
dynamically stable beach profile and plan shape. The scheme performed well over its
first winter.
Due to the very dynamic nature of this part of the coast, beach nourishment will be
required every ten years to maintain a healthy beach and protection to the seawall.
Figure C8.13 shows the completed scheme.

C8.7 Lessons learnt


The PAR benefits assessment included consideration to the avoidance of risk to life
using recently published guidance (HR Wallingford et al. 2006; Defra 2008). This
consideration should be included in future PARs, in particular where flood risks are
prevalent and the risks to people are significant.
The decision as to whether to recharge the beach first or install control structures
always presents a logistical problem. Installation of the control structures first can
create a draw on the existing depleted beach, thus aggravating erosion in some areas.
However, the placing of recharge first can result in sediment loss until such time as the
control structures are installed.
Having considered all the relevant factors at Felixstowe, it was decided to recharge the
beach first. Contrary to the arguably more popular approach (control structures first),
this method of working had a number of advantages which were realised in practice
and which are worth noting here:
• This method allowed the works to be constructed essentially as a land-
based operation, thus avoiding the use of marine plant.
• Health and safety risks were alleviated due to the substantially reduced
requirements for wet working and tidal working.

Toe structures management manual


288
• The improved scope for access across the new beach avoided having to
provide an emergency access route across the SSSI vegetated shingle and
the Landguard Fort scheduled monument.
• The improved access increased the working window, reducing the
programme time (and hence less impact on community) and the cost.
Measures were taken to ‘over recharge’ the beach to take into account sediment losses
incurred before the control structures were put in place.
On a more general note, the project highlighted the need to manage the risks carefully
during both planning and construction. This is a continuous process that requires both
forethought and the flexibility to deal with rapidly changing conditions, for instance the
wall collapse that occurred in 2006.

C8.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by
Alexandra Schofield, Black & Veatch Ltd, in the preparation of this case study.

C8.9 References
Black & Veatch, 2005. War Memorial to Landguard Common, Felixstowe Strategy
Implementation Plan, Coastal Processes Report, June 2005, for the Environment
Agency and Suffolk Coastal District Authority.
Black & Veatch, 2007. Felixstowe South FAS, Project Appraisal Report, for the
Environment Agency and Suffolk Coastal District Authority.
Black & Veatch, 2008. Southern Felixstowe Coastal Strategy (update on Halcrow
Strategy of 2003).
Defra, 2008. Defra Flood and Coastal Defence Appraisal Guidance. Social appraisal.
Supplementary notes to operating authorities: assessing and valuing the risks to life
from flooding for use in appraisal of risk management measures. London: Defra.
HT Wallingford, Flood Hazard Research Centre and Risk & Policy Analysts Ltd, 2006.
R&D outputs: flood risk to people, Phase 2. Joint Defra/Environment Agency Flood and
Coastal Defence R&D Programme, FD 2321/TR2 Guidance Document. London: Defra.

Toe structures management manual


289
Figure C8.9 Extract from detailed design drawings – layout

290
Figure C8.10 Extract from detailed design drawings – sections through rock groynes

Toe structures management manual


291
Figure C8.11 Extract from detailed design drawings – root connections with existing seawall/toe

292
Figure C8.12 Extract from detailed design drawings – detail of permanent wall restoration

Toe structures management manual


293
Figure C8.13 Southern Felixstowe – the completed scheme

294
Annex: Black & Veatch Figure 1 Study area
C9 CASE STUDY: Clayton Road – Selsey
Courtesy of Royal Haskoning

C9.1 Identification of the problem


Selsey Peninsula is a low=lying headland situated in the lee of the Isle of Wight, just to
the east of Chichester Harbour (Figure C9.1). Selsey is popular with holidaymakers
who are mainly attracted to the caravan parks in the area. At West Beach, the subject
of in this case study, the frontage is occupied by private properties that extend to the
waters’ edge.

Figure C9.1 Location map


Selsey Peninsula is exposed to waves from the English Channel together with waves
originating in the Atlantic Ocean. The sheltering effect of the Isle of Wight means that
severe wave attack due to oceanic waves is very sensitive to wave direction. In
addition to a complex wave regime, there are strong currents that sweep around the
end of the peninsula. Sediment is driven onto Selsey Bill from an offshore bank.
However, the orientation of the shoreline in relation to the complex and sensitive
hydraulic regime is such that most of the sediment is driven up the east side of the
peninsula, with West Beach receiving only intermittent feed. The natural consequence
has been for the west shore, where the private properties are located, to gradually
retreat.
There has been a sea defence at West Beach for some considerable time. In response
to falling beaches, the seawall was reconstructed during the 1950s and 1960s. The
defences originating from this time are made up of a steel sheet pile toe surmounted by
concrete steps and/or a battered concrete apron with a wave return wall.
It was concluded at the time of these earlier works that it would be necessary to realign
the wall to produce a more sustainable longer term solution (Lewis and Duvivier 1950).

296
However, there were a number of lengths of substantial private defences; for the sake
of expediency these were left in place, that is, to seaward of the then newly adopted
defence line. These discreet protrusions subsequently formed local promontories
(Figure C9.2).
The predicted continued beach lowering, aggravated by wave reflection from the
vertical seawall, resulted in scour at the toe of the protuberant wall sections. There
were no failures of the wall as such but the promontories required a high level of
maintenance. Moreover, the risk of the seawall being undermined was increasing with
time.
The principal threat to the seawall was toe scour. Failure of the toe would have led to
failure of the whole seawall and ensuing land erosion. In view of the impending threats
to infrastructure and consequent economic loss, it was concluded in the late 1980s to
carry out coastal defence studies and prepare a case for improved coast protection at
Clayton Road.

Figure C9.2 Site plan showing the promontories created by selectively


holding the line

C9.2 Appraisal
During the appraisal process, a number of options were considered for continuing
protection to the Clayton Road frontages. On the basis that retreat of the defence line
would not be acceptable, the options tended to focus on alternative means of securing
a hard defence line along the existing, albeit convoluted, alignment. These options
therefore included extending the apron to a new lower level steel sheet piling and the
(then) comparatively novel use of a rock revetment applied to the toe.
Technically, a major consideration was how to fit a given type of construction to suit the
complicated seawall toe alignment. This called for a design that was suitable in its
application and detail.
From an environmental perspective, a rock revetment was preferred as it reduced
wave reflections from the wall, thus providing an improved environment for the natural
build up of beach levels (or more precisely, reduced scour). Despite common concerns
regarding the placing of rocks on the beach, the public generally accepted the scheme.
In economic terms the principles of best value applied. In essence this meant targeting
low cost works that would be funded by under the Coast Protection Act 1949.
Alongside this, there were opportunities to combine the works as part of a larger
groyne refurbishment scheme, thus yielding economies of scale. A more formalised
economic appraisal, project prioritisation and procedures that are nowadays required
for grant qualification did not apply.
Consultation on the appraisal was principally with/through Chichester District Council.
The preferred solution was arrived at on the basis of:
• relatively low cost;
• flexibility;
• environmental acceptability (or improvement).
A rock revetment was found to offer the best overall solution in respect of the above
objectives.

C9.3 Outline design


Input parameters to the outline design process comprised:
• the existing seawall cross-sections;
• newly surveyed beach levels;
• tide levels and wave heights.
Royal Haskoning’s predecessor in the UK, Posford Duvivier (and previously Lewis and
Duvivier), had worked on the frontage for some 40 years and so was well acquainted
with the ground conditions and previous engineering at the site.
Further to the basic parameters needed for technical design, other important
considerations included site access and the availability of suitable rock. There were few
access points to the beach within reasonable distance of the works and vehicles would
have to negotiate groynes to get to the site. Rock would be obtained from Frome in
Somerset, the individual limestone boulders being delivered on flat top transporters by
road. Cost considerations, due to the small scale of the project, ruled out the prospect
of procuring more durable rock from more distant sources (for example, granite from
Norway).
Figure C9.3 shows a typical cross-section through the revetment. The essential
principles of the design concept were as follows:
• the revetment had to have low wave reflectivity in order to reduce bed
scour which had previously been aggravated by the presence of the vertical
seawall – this was achieved by using a double layer of armourstone;

298
• the rock mound was required to provide passive support to seawall toe –
this was achieved by setting the rock profile into the beach at the face of
the toe pile;
• the revetment armour was to be very stable in order to prevent movements
against the existing seawall steps;
• the structure cross-section had to be sufficiently compact to fit into the
limited space available.
Consents for the scheme were obtained from:
• The Crown Estate as the land owner;
• FEPA licence administrators;
• planning authority.

C9.4 Detailed design


The scheme was designed by Posford Duvivier (now Royal Haskoning). The design
was developed and fully detailed for construction tender purposes. In addition to the
rock structure, the detailed design included intermittent concrete blocks (‘Dragon’s
Teeth’) cast onto the upper existing concrete apron (certain sections only). These
provided a back-stop to the boulders at the crown of the rock mound.
Four different cross-sections were prepared which encompassed the variations in the
existing wall profile and beach variations. The designs were of sufficient detail that
there was effectively no requirement for the contractor to adapt or develop the design
either before or during construction.
The first construction contract was awarded to Z Peskett & Sons Ltd of Littlehampton in
1990. Since then there have several additions to the toe defence works such that now
some 50 per cent of the whole defence length has toe protection using rock.

Figure C9.3 Typical cross-section through the revetment


C9.5 Construction issues
The construction was entirely land-based including the use of plant from the beach.
Some of the limestone rocks split during handling and placing. Usually, damaged rocks
were discarded. Rocks with hairline cracks that were well embedded within the
structure were left in place where it was felt that this did not compromise the security of
the structure.
Despite the complexity of the site geometry, no changes from the tendered design
were found to be necessary.
Careful placing of the boulders together with protection provided by the Dragon’s Teeth
meant there was no damage to the existing concrete steps or apron.

Figure C9.4 The completed armour revetment

C9.6 Post construction issues


The scheme is included in an annual asset survey. This entails a qualitative inspection
of the general condition of the rock mound including observations on settlement, wear
and tear.
Since its construction in 1990, the protection measures appear to have performed well
in stalling scour at the toe.
The main problem has been attrition of the limestone rocks through abrasion by the
hard flint fraction of the beach sediments. As a result, the exposed edges of the rock
armour units are becoming rounded. There also appears to have been some
settlement in some sections. The rock interlocks have, however, remained reasonably
sound.

300
In places, there appears has been some localised increased wear of the steps and
apron where overtopping water jets have penetrated through gaps between the rocks
and apron. However, generally the rock revetments have protected the steps and
apron from excessive wear.

C9.7 Lessons learnt


The simple design comprising largely single size rocks with no filter layers has worked
very well.
The durability of the limestone may eventually limit the lifespan of the revetment.
However, whether the use of more expensive granite would have represented better
value in the long term is open to question.

C9.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Simon
Howard, Royal Haskoning, and David Lowsley, Chichester District Council, in the
preparation of this case study.

C9.9 References
Lewis and Duvivier, 1950. Report on coastal protection. Report to Chichester Rural
District Council.
C10 CASE STUDY: Fort Wall, Canterbury
Courtesy of Canterbury City Council

C10.1 Identification of the problem


The site is located just to the east of Herne Bay on the north Kent coast (Figure C10.1)
and is landmarked by St Mary’s Church immediately to the east. The subject of the
case study is known as Fort Wall; a short defence of 80 m in length.

Figure C10.1 Location of Fort Wall in north Kent


Fort Wall, which is owned by English Heritage, protects the archaeological remains of a
Roman fort. The historic site is located on a slight hill but the land behind falls away to
about high tide level. Hence, in addition to securing the archaeological remains, the
Fort Wall defence together with the immediate land strip provide flood protection to the
lower area behind.
The site surface geology comprises a mixed sand and shingle beach overlying Thanet
sandstone (at about 1 m below MHWS). The sandstone is easily eroded once exposed,
though there is no quantitative evidence of sandstone downcutting. However, Roman
ruins have been discovered 2 km out to sea implying, in very approximate terms, a
horizontal rate of coastal retreat of about 1 m per year.
The history of the seawall that preceded the present structure is relevant to the case
study and is outlined below. Built in 1965 by the then Ministry of Works, the earlier wall
(Figure C10.2) comprised:
• a mass concrete block (formed in 6 m lengths ) founded on the sandstone;
• bearing onto this from above was a revetment (variable slope) which
contained the mixed fill embankment.
The revetment consisted of a concrete bed over which were laid Ragstone blocks.

302
Figure C10.2 Sketch of former wall section (1965 construction)
Initially, beach levels at the site had been healthy. In the early 1990s, however, a major
scheme (not by Canterbury City Council) including a large terminal groyne was put into
place coming to within 100 m east of the old defence. Though not well recorded, beach
lowering at the Fort Wall section ensued.
A storm with an estimated return period of 10 years occurred in February 1996,
resulting in substantial destruction of the old wall. About half the Ragstone blocks came
off the revetment. It appeared that the large toe blocks had just turned over rather than
being undermined. An estimated 4 m of erosion occurred within four days as the mixed
fill became exposed to subsequent wave action (Figure C10.3).

Figure C10.3 Failure damage, 1996


An immediate fix was needed to prevent further retreat, which might have put the
archaeology at risk. This was achieved by backfilling the remaining toe blocks using
pumped concrete and then carefully shoring up the exposed face (Figure C10.4).
These emergency works were carried out in February 1996 and consisted of about 100
m3 of mass concrete to the failed toe and 90 m2 of sprayed concrete with mesh
reinforcement to the exposed slope face. The cost of emergency repairs was about
£27,000 in 1996. Further works to the slopes were again necessary in May 1996 with
further concrete infilling (costing about £10,000). These emergency works then
remained intact until the permanent works commenced in August 1998.
Figure C10.4 Emergency repair works, 1996

C10.2 Appraisal
The discussion below relates to the long-term repair works carried out from 1998
onwards.
Application was made to Defra in March 1996 to undertake a coast protection study for
this and the adjacent Reculver frontage. However, Defra required a full coastal defence
strategy plan to be carried out to also include the Environment Agency frontage to the
east. This Strategy, covering a total frontage of 5 km was carried out in-house by
Canterbury City Council (CCC); it began in June 1996 and was completed in October
1997.
The Strategy covered a full range of options for the full frontage. Options for the Fort
Wall section included the use of timber groynes with nourishment, this being required
due to lack of any natural feed. This would have to be a pocket beach including a
terminal groyne which would, therefore, have had its own downdrift effect. Difficult
access would have made the scheme very expensive; moreover, the fact that it was a
small scheme would have made the necessary use of large marine plant and
operations uneconomic. As beach amenity was not an important issue at this site, a
hard defence was the preferred option.
Other significant issues to consider included:
• the scheme was required to protect archaeology and so this introduced
important site access considerations;
• construction had to be during summer to avoid disruption to overwintering
birds.
Application was made to Defra in November 1997 for implementation of a coast
protection scheme, as recommended in the Strategy, apportioned between CCC and
English Heritage (385 m) and the Environment Agency (95 m). Defra scheme approval
for both the CCC and Environment Agency lengths was received in March 1998.

304
C10.3 Outline design
The design took into account the (then) lowered beach profile and wave attack which
was depth limited at the new structure. The design of the Fort Wall section was to
comprise:
• steel piling driven into the sandstone;
• concrete capping that would encompass the old concrete wall;
• a rock armoured revetment in front of the piled wall;
• a replacement concrete and Ragstone composite revetment.

C10.4 Detailed design


CCC undertook the site investigation to determine sandstone levels and the strength of
the material, which turned out to be very varied. Over the length of the works the
sandstone base varied by only about 0.6 m.
Detail design and contract preparation for the whole project (including the Environment
Agency and English Heritage lengths) was carried out in-house by CCC. Figure C10.5
shows details of the new Fort Wall.

C10.5 Construction issues


The tender for construction was issued in May 1998 and works commenced on-site in
August 1998. Harbour & General were the appointed contractor. The contract period
was 32 weeks but the works were actually completed within 20 weeks.
As access overland was difficult, the armour rock had to be delivered be sea
(Figures C10.6 and C10.7). Limestone rock armour was sourced from Boulogne.
Pile driving into the sandstone went well (Figure C10.7). Where possible the original
Ragstone blocks were salvaged and reused.
Figure C10.8 shows the final product.
The contract value for the whole project was £1.18 million, of which the value of Fort
Wall works was £220,000. The main quantities included:
• 1,510 m3 of 3–6 tonne armour rock;
• 190 m3 of reinforced concrete;
• 430 m2 of sheet piles driven to a depth of 3 m.
Figure C10.5 New design (extract)

306
Figure C10.6 Rock delivery

Figure C10.7 Pile driving


Figure C10.8 Final construction

C10.6 Post construction issues


A better beach has built up since installation of the rock revetment, most probably due
to the reduced reflection (compared with that of the earlier vertical structure). Overall,
the scheme has been very successful.
The site is now monitored regularly (three times per year) as part of the south-east
regional strategic monitoring programme. Little maintenance has been needed other
than repointing of the Ragstone blocks from time to time.

C10.7 Lessons learnt


The 1965 design could not withstand the combined effects of a lowered beach and a
10-year return period storm wave attack. Alongside this, better beach monitoring
should have been in place to identify the possibility and imminence of the 1996
collapse but this was not a major consideration at the time.
The new design learnt from the earlier failure and accounted properly for appropriate
design conditions.
Because of the very small extent of the works, it was packaged with other schemes to
reduce costs due to mobilisation. This made for an economically attractive project and,
indeed, a positive lesson for future schemes.

C10.8 Acknowledgements
We would like to acknowledge the invaluable advice and assistance provided by Ted
Edwards, Canterbury City Council, in the preparation of this case study.

308
C11 CASE STUDY: Prestatyn, North Wales

C11.1 Background history


The coastline between Rhyl and Prestatyn has been eroding for several decades
through a combination of reduced sediment supply and coastal squeeze. At the end of
the 19th century, some two million tonnes of gravel were removed from the beaches in
the Rhyl area (updrift and to the west of Prestatyn) to provide ballast for building
Liverpool Docks. As a result, the pebble storm beaches that once extended along
much of this frontage have largely disappeared. Tourist pressures have also led to the
reclamation of large stretches of marshland for the construction of holiday camps, golf
courses and so on. The holiday camp and housing at Prestatyn are situated close to
the shoreline and are protected by a seawall. To the east of this wall, there is a line of
dunes that have been eroding.
Prestatyn was first protected by a stepped concrete seawall in 1960. At the same time
the foreshore was protected by a series of long timber groynes. Already by the 1970s,
the beach in front of this wall had fallen significantly. The flatter gradient allowed the
ridge and runnel systems, common on this wide foreshore, to migrate shoreward. The
increased water depth at the toe of the wall then caused strong overtopping. The wall
itself was at risk of foundation failure. The sand transport was then concentrated at
some distance seawards of the wall, effectively starving the dunes immediately
downdrift of the sand supply.
The photograph shown in Figure C11.1, which was taken in early 1990, shows the tidal
runnel very close to the wall. Note that the runnel extends a considerable distance
alongshore, cutting across several timber groynes. All that remains at the toe of the
wall is a narrow strip of pebbles. Some emergency works in the form of a fillet of rock
armour-stone can be seen at the toe of the wall. The growth of algae on the concrete
steps indicates frequent wave overtopping.

Figure C11.1 Toe scour in front of stepped concrete seawall, 1990


C11.2 Mitigation measures
This frontage has required regular maintenance. In the 1980s, field investigations were
carried out into the problems of wave and tidal induced scour and the strength of tidal
currents over the foreshore. Rock groynes were constructed to reduce inshore tidal
currents. This improved beach levels over the foreshore, but the wall toe remained
vulnerable to wave attack.
Following substantial damage in 1990–1991, a major scheme was implemented,
including massive sand recharge coupled with the construction/upgrading of the rock
groynes. In addition, the vertical face of the seawall below the concrete stepped face
was replaced by an asphaltic sloping apron.

C11.3 Performance of mitigation measures


The increased height and width of the upper foreshore has resulted in the
disappearance of the ridge and runnel features from in front of the seawall. The high
foreshore levels have also removed the problems of wave overtopping. Figure C11.2
shows the swash limit along the line of the new sloping revetment. There is some sand
build up above the revetment and on the seawall steps. The amenity value of the
promenade, at the crest of the wall, has also been greatly improved.

Figure C11.2 Sloping asphalt apron and rock groynes in front of steeped
concrete seawall

C11.4 Other comments


A scheme of this type needs to be monitored carefully. Not only do beach levels in front
of the seawall need to be checked regularly, but the evolution of the downdrift beaches
must also be assessed.

310
C12 CASE STUDY: Colwyn Bay, North Wales

C12.1 Background history

Figure C12.1 Site location


Colwyn Bay is a popular tourist resort situated on the North Wales coast
(Figure C12.1). The sand beaches in this area have been eroding as a result of coastal
squeeze and a lack of sediment supply from the west (updrift). The reasons for this are
also described in the case studies for Penrhyn Bay (C14) and Rhos-on-Sea (C13). Due
to the fallen beach levels, the promenade and the road immediately to the landward
have been affected by heavy wave overtopping.
The construction of sea defences at Colwyn Bay dates back to the late 19th century.
The masonry seawall has, over the years, suffered considerable damage, requiring
extensive repairs and reconstruction. In the 1970s there was a groyned upper beach of
shingle, with an almost continuous sand cover over the flatter, lower part of the beach.
Subsequently the groyne system fell into disrepair, allowing beach levels at the wall toe
to fall. Thus, by the 1980s the shingle beach had largely disappeared from the wall toe,
causing foundation problems (Figure C12.2). In addition, the falling sand levels had
exposed the underlying pebbles over much of the lower foreshore.
Figure C12.2 Beach lowering in front of vertical seawall, Colwyn Bay

C12.2 Mitigation measures


In 1987 a rock berm was constructed along the most severely affected stretch of wall.
This encouraged some beach build up in the immediate vicinity of the wall toe,
assisting seawall stability. In the early 1990s, a number of low rock groynes were
constructed from the wall out to the low water line. These project no more than 1 m
above the beach surface and hence are not visually intrusive. Since then, other
sections of wall have required additional toe protection, usually comprising a rock toe
or sheet piling and concrete infills.

C12.3 Performance of mitigation measures


The Rhos-on-Sea breakwater may be responsible for trapping in its lee what little
shingle drift there is. Therefore, the construction of the rock berm has encouraged sand
rather than shingle accretion. Furthermore, the sand cover has increased significantly
seaward, so that only small areas have the underlying pebbles exposed.
This scheme is a good example of how relatively modest defences can be used
effectively to improve beach levels. The photograph in Figure C12.3, which was taken
in 2002, shows the sand build up at the foot of the seawall. The long rock groynes can
(just) be seen in the background, while a redundant timber groyne can be seen in the
middle distance.

312
Figure C12.3 Rock berm in front of vertical seawall, 2002
C13 CASE STUDY: Rhos-on-Sea, North Wales

C13.1 Background history

Figure C13.1 Site location


Rhos-on-Sea is situated on a small headland at the eastern end of Penrhyn Bay
(Figure C13.1). The headland is a focal point for wave action and the residential
development on low-lying land to the landward was at risk from heavy wave
overtopping in the recent past.
As described in the Penrhyn Bay case study (C14), there is a shortfall in the supply of
beach sediments in Penrhyn Bay and to the east. The headland rock promontory of the
Little Orme acts as a groyne, cutting off the majority of supply of shingle from the west,
as well as reducing the amount of sand supply. By contrast, the smaller promontory of
Rhos Point at the eastern end of Penrhyn Bay had not prevented beach material from
being transported eastwards (downdrift) into Colwyn Bay.
The construction of sea defences within Penrhyn Bay in the late 19th and 20th
centuries effectively cut off the supply of sediments from the erosion of boulder clay
cliffs at the western end of the bay. This left only a small supply of sand from the
nearshore zone, feeding around the Little Orme headland in suspension. Beach levels
had therefore gradually deteriorated both within Penrhyn Bay and around Rhos Point
itself.
The seawall around Rhos Point was built in the 1860s and, prior to the breakwater
protection scheme described here, had been breached and repaired in the recent past.
Further falls in beach levels were anticipated, similar to the progressive deterioration of
the beaches that had taken place earlier in Penrhyn Bay. In view of the falling beach
levels, upgrading the existing sea defences, for example, was considered to be
insufficient as a long-term solution to the problems that had developed around Rhos
Point.

314
C13.2 Mitigation measures
Following wave overtopping studies by HR Wallingford, a rock armour breakwater was
constructed off Rhos Point in 1983. This was located opposite low-lying land to the
south of the Point. Rock left over from the breakwater construction was used to
construct a short groyne on the coast immediately to the north, to encourage material
to collect around the Point itself.

C13.3 Performance of mitigation measures


The breakwater has eliminated the problems of wave overtopping that were becoming
increasingly more serious to the south of Rhos Point. The sheltered conditions in the
lee of the breakwater have allowed small boats to anchor there (Figure C13.2). This
has been possible because the breakwater is sited some distance away from the wall,
but not so far offshore that its sheltering effect would be significantly reduced. The low
groyne has been overtopped by beach material and shingle and sand have tended to
collect in the lee of the breakwater, further reducing any potential risk of wave
overtopping (Figure C13.3).
In view of the 7 m tidal range on spring tides, the offshore breakwater is a large
structure. As well as trapping the small volume of littoral drift from Penrhyn Bay, the
breakwater has also attracted a reverse westerly drift of material from Colwyn Bay to
the east. In addition, the high degree of shelter has attracted a small amount of mud
from offshore. The accumulation has not affected the development of sailing leisure
facilities in the lee of the structure.

Figure C13.2 View to the east showing usage by small boats


Figure C13.3 View to the west showing beach recharge

316
C14 CASE STUDY: Penrhyn Bay, North Wales

C14.1 Background history

Figure C14.1 Site location


Penrhyn Bay is situated to the east (downdrift) side of the headland of the Little Orme
(Figure C14.1). The problems of beach erosion in Penrhyn Bay are primarily due to a
lack of contemporary sediment supply. Historically, the main source of beach material
for the bay was the erosion of boulder clay cliffs on the east side of the Little Orme and
a (potential) feed from the nearshore seabed.
Since the beginning of the 20th century, the construction of sea defences has
progressively cut off the supply of beach material derived from the cliff erosion. The
seawalls themselves, the earliest of which dates back to the 1900s, have contributed to
coastal squeeze, causing further deterioration of beach levels. The present seawall
structures date from the period 1945–1960.
By the 1970s, however, the beach conditions had deteriorated to such an extent that
the beach material had largely disappeared from the eastern (downdrift) part of
Penrhyn Bay. Even in the more sheltered western part of the bay there was little beach
material remaining. As a result, groynes that had been introduced to try and control
beach movement were no longer effective and, having not been maintained, had
deteriorated (Figure C14.2). The lack of beach cover and the serious overtopping in the
exposed central frontage resulted in the construction of a short length of timber
breastwork to the seaward of the existing wall in an attempt to reduce wave
overtopping. These measures were only partially successful and did not deal with the
root cause of the problem, which was the lack of sediment supply.
Although partly sheltered from the west, the refraction/diffraction around the headland
of the Little Orme and the edge-wave effects, as waves propagate alongshore along
the line of the seawalls, produces a significant eastward littoral drift within Penrhyn
Bay. Since Penrhyn Bay is not fully enclosed at its eastern end, this has produced a
gradual emptying of sediments out of the bay.
As a consequence the walls were in danger of being undermined and significant
overtopping of the defences occurred during storm events.
In the late 1980s, investigations were carried out to identify a scheme that would
remedy the situation.

Figure C14.2 Lowered beach exposing shore platform

C14.2 Mitigation measures


The scheme introduced at Penrhyn Bay in 1989–1990 consisted of the construction of
two fishtail-type rock groynes, with the beach between the groynes recharged with a
mixture of sand, shingle and cobble sized material. The finer material was used in the
more sheltered parts of the bay and the coarser material towards the eastern end of
the bay which was more exposed. The groynes provided terminal structures to keep
the recharged material within the artificial embayment.
Along the downdrift frontage to the east, where beaches remained low, an alternative
form of construction was adopted and a new full height armour stone revetment was
constructed in the mid-1990s. Both schemes made use of locally available quarried
rock in the structures and locally available sand and quarried rock in the artificial
nourishment.

C14,3 Performance of mitigation measures


This innovative scheme has been very successful. The combination of the artificial
beach recharge and the fishtail groynes has effectively reduced the erosion of the
upper beach and contained the beach sediments within Penrhyn Bay. The artificially
formed beach has generally a sufficiently high crest to prevent waves reaching the
seawall.
The beach was graded from sand in the west to an (artificial) cobble beach formed of
quarried rock in the east. Movement of the material has caused a mixing of sediments
to occur with mixed sand and shingle tending to stay in the (sheltered) western corner
of the bay (Figure C14.3).

318
Figure C14.3 View of beach recharge in western part of Penrhyn Bay
Further east, material from the centre of the embayment tends to be moved towards
the easterly groyne where it can overtop the root of the groyne structure
(Figure C14.4). Regular recycling of material back towards the centre of the bay (once
or twice a year) is carried out in order to maintain Standards of Protection.

Figure C14.4 View from easterly groyne (CEUK)

C14.4 Other comments


A scheme of this type alters the character of the beach significantly. Initially the use of
angular quarried limestone was alien but overtime the material has rounded to a more
natural appearance. Conversely this has reduced the size of the material and
consequently increased its mobility.
Regular (bi-annual and post storm) monitoring is carried out to inform defence
performance assessment and the recycling regime. Losses of recharge material are
minimal and the structure layout attracts some material from offshore into the bay.
Further to the east beach levels remain low but reasonably stable and devoid of fines
(Figure C14.5).
Figure C14.5 View along rock revetment protected frontage to east
(CEUK)

320
C15 CASE STUDY: Sandbanks Peninsula, Poole,
Dorset

C15.1 Background history


The Sandbanks peninsula is a long, heavily developed sandy peninsula situated
immediately to the east of the entrance to Poole Harbour. The flow patterns here are
complex. There are rapid tidal currents in and out of the harbour. There is also a
subsidiary inshore channel, called the East Looe Channel, which allows fast tidal
currents to flow parallel to the peninsula and close inshore. To the seaward of this
channel there is a sandbank whose form changes in response to the wave climate, as
well as these complex tidal flows. The resulting sediment transport in this area is thus
extremely complex. Sand can be transported alongshore by breaking wave action as
well as by the tidal currents. As a result of these processes, the direction and
magnitude of longshore sediment transport is temporally and spatially very variable. In
addition, there is intermittent onshore movement of sand from Hook Sand, by swell
wave action.
The sand beaches along the Sandbanks peninsula were originally protected by a
system of crib type rock groynes. Historic charts show that these maintained high
beach levels over much of the frontage. However, these groynes fell into disrepair,
being removed in 1991 for health and safety reasons. Beach lowering was noted
subsequently, becoming most serious at the western end of the frontage, where a
seawall surrounds the head of the peninsula (Figure C15.1). The offshore transport of
sand, due to waves being reflected from the wall and its subsequent removal by the
fast tidal currents, caused concerns that the wall would become undermined.

Figure C15.1 Vertical wall and low beach near Haven Hotel, Sandbanks,
c.1998
C15.2 Mitigation measures
In 1991 a rock groyne was constructed near the western end of Sandbanks. In 1994 a
rock fillet was constructed at the base of the wall. Following the completion of a beach
management strategy for the entire Poole frontage in 1994–1995, HR Wallingford was
commissioned to produce an outline design for a coast protection scheme for the
western end of Sandbanks. This included the modelling of a groyne scheme to reduce
flows over the beach in front of the seawall. The optimum plan shape that was derived
ensured that the flows of the East Loe Channel were deflected away from the seawall.
The scheme was constructed in 1995–1996. Overall, this was a great success with the
earlier beach erosion being reversed and sand dunes forming where there were
previously low beach levels (Figure C15.2).
The rock groynes have subsequently been extended southwards in the second phase
of the works. These have also been very successful so that virtually the whole of the
Sandbanks frontage now has a high level of protection.

Figure C15.2 Widened beach after installation of sill and rock groynes,
2002

C15.3 Performance of mitigation measure


Before the scheme was implemented, there had been an increase in water depths
within the East Looe Channel and an onshore movement towards the line of the
seawall. When the scheme was built, the tidal currents were deflected away from the
wall, enabling sand to settle out of suspension.
This scheme demonstrates how the role of tidal currents on beach lowering should not
be overlooked, especially near estuary and inlet mouths.

322
C16 CASE STUDY: Seaford, East Sussex

C16.1 Background history


Now a resort, Seaford was a port before a great storm in 1579 caused the build-up of
shingle to block off the entrance channel and diverted the course of the river Ouse
westwards in the direction of Newhaven.
The subsequent development of the port of Newhaven has included the construction of
training walls to prevent the littoral from blocking up the new entrance. Successive
extensions of the western training wall were necessary as the beach to the west of the
entrance accreted. This meant that the shingle beach at Seaford has received a
dwindling supply of shingle. This has caused beach levels in front of the seawall to fall
at an increasing rate during the last century. By 1980, the beach in the eastern half of
the frontage had fallen to such an extent that it was providing very little support to the
old mass concrete seawall; this wall was originally a secondary defence behind a then
substantial shingle ridge. In places, erosion had exposed the underlying chalk platform,
allowing waves up to 6 m high to reach the wall without breaking. In 1981 parts of the
wall had become badly damaged and in 1985 undermining had caused local collapse
of the promenade. Figure C16.1 shows the situation around 1982.

Figure C16.1 Low beach levels and damaged concrete seawall, c.1982
C16.2 Mitigation measures
The local water authority was aware that reconstructing the seawall would be difficult to
justify economically and that such a course of action would be unsustainable given the
likely continued fall in beach levels. Following hydraulic and numerical model testing at
HR Wallingford, an open beach (ungroyned) nourishment scheme was adopted as the
most economical solution to the problem of overtopping and continuing deterioration of
the seawall. Due to its south-westerly aspect the shingle beach at Seaford is aligned
almost perpendicularly to the predominant direction of approaching south-westerly
waves. Because of the relatively low rate of littoral drift generated by these obliquely
incident waves at the western end of the nourished frontage, it was determined that a
terminal groyne was not necessary there. The large concrete groyne already in place at
the eastern end of the frontage was reconstructed to a greater height and length, so as
to prevent loss of material to the natural cliffed (and undeveloped) coastline to the east.
In 1987 the central and eastern end of Seaford was nourished with 1.5 million m3 of
shingle won from offshore (Figure C16.2). The material was won by using a trailer-
suction dredger, extracting the material from an existing licence area on the Owers
Bank, south of Selsey Bill. The material was spread over a 2.5 km frontage. The
western frontage was left untouched, as the beach there was already wide.

Figure C16.2 Beach after renourishment

C16.3 Performance of mitigation measure


Following the initial period of adjustment, the beach actually increased in volume within
the active beach profile (taken as above –4 mAOD) This is because the wave
reflectivity was significantly reduced, causing pebbles to migrate landward from the
hard chalk seabed on which material was very mobile.
Monitoring has been critical to the long-term viability of the scheme. In order to
maintain sufficient beach width at all points on the frontage, recycling needs to be
carried out. Modelling indicated that the average annual recycling volume was likely to
between 20,000 and 25,000 m3 per year. The volume that has had to be recycled has,
in fact, varied considerably from year to year and the average value has been

324
considerably higher than anticipated. Nevertheless, the scheme has successfully
protected the ageing seawall from wave attack and stopped the heavy wave
overtopping that used to take place.

C16.4 Possible improvement measures


It would appear that a nourished fill is considerably more mobile than the native beach
material. The reasons for this are not particularly well understood, but it is considered
likely that shingle transport on what is a relatively steep beach may be enhanced by
tidal current action. Certainly there is evidence in the form of shingle waves, showing
enhanced mobility. This type of response has been observed in a number of other
beach nourishment schemes involving shingle.
The disadvantages of beach nourishment are that it is difficult to predict the expected
life-span of beach nourishment material in view of the unpredictability of the UK wave
climate. There may also be difficulty in obtaining the right grade of material, as offshore
dredging operations are dependent upon a licence being available. Massive
nourishment schemes, particularly where they involve recycling, also have an adverse
impact on the usage of the beach. This can be minimised by targeting the recycling
operations so as to avoid holiday periods, etc.
C17 CASE STUDY: Selsey Bill, West Sussex

C17.1 Background history


Selsey Bill is situated at the southern tip of a low-lying headland that juts out into the
English Channel. The coastline is formed of Bracklesham Clays, which are overlain by
gravel deposits at Selsey Bill. These gravels form low cliffs that are easily eroded.
Because of its open aspect, Selsey Bill is a focus for wave energy. It is also a drift
divide, the cliff erosion having provided material for the development of the beaches
east and west of the Bill. In addition marine sediments – principally coarse sands and
gravel – are driven onshore during storms from barely submerged banks lying off the
Bill. This movement takes place in pulse fashion, so that the thickness of the shingle
cover at any location varies greatly with time. The geomorphology of the area is thus
very complex and ever changing.
During the early part of the 1900s the coastline west of the Bill had undergone
continued long-term recession, which reached an annual rate of the order of 6 m per
year. The erosion of the sandy clays and gravel provided large drift along the frontage
to the west, providing sediments for the East Head spit at the western side of the
entrance to Chichester Harbour, as well as for the ebb bar across the entrance.
However, had erosion continued much of the shorefront development would have been
lost (some had already been lost before the scheme was implemented)
Sea defences were begun in the 1950s. With an onshore supply near the Bill, the
beaches did not deteriorate as rapidly as might have been expected in view of the
earlier very rapid rates of retreat. By the late 1980s, however, the walls had
deteriorated through continuous wave action and falling beach levels. Figure C17.1
shows the situation in 1988.

Figure C17.1 Low beach levels in front of stepped concrete wall, 1988

326
C17.2 Mitigation measures
By the early 1990s, the stepped concrete wall at Selsey West Beach was in danger of
being undermined. In addition, there was heavy wave overtopping on virtually every
high tide, resulting in damage to developments on the immediate backshore.
In 1992 a major scheme was initiated along much of the Selsey frontage. At Selsey
West Beach, the seawall was reconstructed at strategic locations where heavy
overtopping could not be tolerated. In other areas the wall was strengthened. In places
the wall was extended by the addition of concrete armour units and a rock berm, as
shown in Figure C17.2. In addition, some shingle was added to the upper beach, to
help fill the groyne compartments.

Figure C17.2 Toe berm of rock and concrete armour units, 1994

C17.3 Performance of mitigation measures


This is a very exposed location and it is not possible to maintain a shingle beach in
front of the seawall permanently; Figure C17.2 shows the face of a groyne, against
which the shingle beach has recently been drawn down. Shingle beach levels continue
to fluctuate strongly from season to season. The level of the lower sandy foreshore
appears to have been maintained.
While the problems of overtopping along this frontage have not been eliminated, the
volume and frequency of overtopping has been significantly reduced. In addition, the
stability of the toe of the wall has been secured.
C18 CASE STUDY: Sidmouth, Devon

C18.1 Background history


Sidmouth is situated at the mouth of a river valley that is flanked by cliffs of red marl.
The net littoral drift is from west to east, pinching in the river Sid against the cliffs on the
eastern end of the valley. (Sidmouth was a port, which became unusable when the
river entrance was infilled with shingle.) The shingle beach is formed of material
derived from the erosion of the marl cliffs. These cliffs are largely unprotected, so that
the supply has been maintained. Over the town frontage, the shingle beach is backed
by a seawall, and is therefore vulnerable when waves are able to reach the wall.
Apart from changes due to fluctuations in the rate of west to east littoral drift, the
shingle beach in front of the seawall is affected by drawdown during severe storms,
leaving the wall exposed to wave attack. In the early 1990s a severe storm caused
serious beach lowering and major wave overtopping over the town frontage. After this
storm the beach did not recover its former levels.

C18.2 Mitigation measures


Following model testing at HR Wallingford, a beach nourishment scheme was
implemented using local beach material, together with the construction of two offshore
breakwaters and a groyne at the eastern (downdrift) end (Figure C18.1). The purpose
of the two breakwaters at the western end of the frontage is to protect the town
frontage against the predominant westerly storms. The groyne is there to prevent
material from being transported out of the area by the net west to east drift.

Figure C18.1 Beach recharge, groyne and offshore breakwaters

328
C18.3 Performance of mitigation measures
The scheme has eliminated beach drawdown and overtopping during westerly storms.
A succession of storms from the east caused shingle to migrate into the lee of the
breakwaters, reducing the beach width at the eastern end of the frontage. This was
remedied by constructing an additional rock groyne to reduce littoral drift from the east.
Since the second groyne was added the beach has maintained an adequate width over
the whole frontage.
Figure C18.2 shows the eastern end of the nourished frontage. The improvement in
beach width has not only effectively reduced the former problems of beach lowering but
has also provided a more attractive beach. In addition, the offset breakwaters are also
not visually intrusive, being below the horizon at promenade level.

Figure C18.2 Rock groynes and beach recharge, eastern end of


promenade
C19 CASE STUDY: Portobello Beach, Edinburgh

C19.1 Background history


The justification for many beach nourishment schemes in the UK has been that the
cost of nourishment is considerably less than the cost of reinstating existing hard
defences. One of the first schemes to be justified on this basis was carried out at
Portobello Beach, Edinburgh. This beach is situated on the western side of Edinburgh
and faces directly into the mouth of the Firth of Forth. It became denuded as a result of
sand abstraction for the glass industry, which began in the 19th century and continued
up to the 1930s. The promenade seawall at Portobello dates back to 1860. By the late
1950s the lowered beach meant that the wall was under continuous wave attack, with
resulting frequent serious overtopping (Figure C19.1). By this time the beach had
flattened and the median sand grain size was 0.2 mm.

Figure C19.1 Portobello Beach, c.1970

C19.2 Mitigation measures


Following studies by HR Wallingford, the beach was nourished in 1972. The sand
nourishment material had a median size of 0.27 mm, considerably coarser than the
beach material. The sand was obtained from a sub-tidal borrow area some 3 km east
of Portobello in a sheltered part of the Firth of Forth. This material is so close to
Portobello that it may well be the natural sand size for the area. The beach at
Portobello had become so eroded that the sand was no longer representative of the
beach material under healthier conditions.
Some 180,000 m3 of coarse sand was extracted from a borrow area by bucket dredger,
transported by barge and pumped over a 1.6 km frontage to a foreshore gradient of 1
in 20. The material was placed over a depleted beach whose gradient had fallen as a
result of beach lowering to about 1 in 42. The nourished beach was held in place by a

330
number of timber groynes, with an (easily adjustable) gabion-type groyne at the
eastern (updrift) end of the frontage (Figure C19.2).

Figure C19.2 Post-nourishment view at Portobello

C19.3 Performance of mitigation measure


From the start, it was recognised that careful monitoring was crucial to the long-term
success of the scheme. The beach was monitored in the early year of the scheme by
HR Wallingford and then by the local Coast Protection Authority (first Lothian Regional
Council, then City of Edinburgh Council)
The beach profile surveys show that, after 18 months, the beach slope had adjusted to
1 in 23, but other than the seaward movement of the toe of the nourished beach, there
were no significant offshore losses of beach material. Littoral drift in this area is low so
that end losses are small.
Beach volumes remained relatively unchanged until 1978. By 1981 the losses had
increased to about 50 per cent of the original nourishment volume as a result of severe
storms. Following calmer weather the beach recovered, so that in 1984 there was still
some 75 per cent of the renourishment volume remaining above the low water mark!
By 1988 the nourishment volume had reduced to 70 per cent of the placed volume. In
late 1988 a further 102,000 m3 of sand were added as a topping up and improvement
operation. The trend of beach erosion after storms and subsequent recovery has
continued since. Despite a trend of gradually declining beach volume, the beach
remains above its pre-1988 renourishment level. 12 (HR Wallingford 2002).

C19.4 Possible improvement measures


This scheme has been so successful that no significant improvements to the mitigation
technique employed can be envisaged. The beach has a low littoral drift and the swell
waves penetrating through the mouth of the Firth of Forth almost balance the
destructive action of locally generated, hence short period and destructive waves. Had

12
Southern North Sea Sediment Transport Study, Phase 2. HR Wallingford Report EX 4526.
finances been available the scheme might have been extended westwards over the
partly industrial frontage to Leith Docks.

332

You might also like