You are on page 1of 1009

)('(5$/'(02&5$7,&5(38%/,&2)

('(5$/'(02&5$7,&5(38%/,&2)

(7+,23,$152$'6$87+25,7<

,7 <
( 7+

25
,2

,$
15 87
+
3

2$'6 $

'(6,*10$18$/)25/2:92/80(52$'6
3$57$3$57%$1'3$57&
),1$/'5$)7$35,/
Foreword - i

FOREWORD

Low volume roads in Ethiopia typically carry less than 300 vehicles per day and provide important links
from homes, villages and farms to markets and offer the public access to health, education and other
essential services. These roads also provide important links between Wereda Centres and the Federal
road network.

Many aspects of the design and construction of roads in Ethiopia, has stemmed from technologies
and practices emanating from Europe and the USA some 40 years ago. These practices have to some
extent been modified in the intervening years, but the basic philosophy of road provision has remained
the same. While these “standard” approaches might still be appropriate for much of the main trunk
and link road network, they remain overly conservative, inappropriate and far too costly for application
on much of the country’s rural road network. In facing the challenges of improving and expanding
Ethiopia’s low volume rural road network, application of the traditional planning, design, construction
and maintenance approaches cannot provide the solution.

Many innovative practices and unconventional techniques, often developed and proved through
years of research, have not found the degree of application and implementation that they should.
Opportunities are missed that would provide better and lower cost engineering solutions and more
sustainable low volume roads.

There is a wealth of local and international information, experience and research that when utilised,
can change current practices and thinking and provide Ethiopia with an appropriate and affordable
low volume road network. To benefit fully from these advances and to see necessary improvements
implemented on the ground, Ethiopian Roads Authority (ERA) has developed its first comprehensive
national road design manual, technical specifications and bidding documents specifically for low volume
roads. The task was completed with the assistance of a team of international experts commissioned
through DFID’s Africa Community Access Programme (AFCAP).

Compilation of the documents was undertaken in close consultation with the local industry and regional
authorities. The Federal and Regional Roads Authorities, the contracting and consulting industry, the
universities, training schools, the road fund and other industry stakeholders all participated in the
formulation of the documentation. Local issues and experience on the geometric, earthwork, drainage,
pavement and surfacing design for low volume roads were discussed and debated at length. Of
particular interest were aspects of better use of local materials; materials selection and specification;
testing and improvement of materials; construction methods and the utilisation of approaches such
as labour-based and intermediate equipment technology; route selection; works specifications; and
contracting of small works. Much of this debate and other resource materials developed during the
compilation of the manual have been captured and are available from the ERA website: www.era.gov.et

Importantly, in supporting the preparation of the documents, a series of thematic peer review panels
were established that comprised local experts from the public and private sector who provided guidance
and review for the project team. Mainstreaming of social and cross-cutting aspects received special
attention and a peer panel of local experts developed a welcome addition to the design process in
terms of complementary interventions.

On behalf of the Ethiopian Roads Authority I would like to take this opportunity to thank DFID, Crown
Agents and the AFCAP team for their cooperation, contribution and support in the development of
the low volume roads manual and supporting documents for Ethiopia. I would also like to extend my
gratitude and appreciation to all of the industry stakeholders and participants who contributed their
time, knowledge and effort during the development of the documents. Special thanks are extended to
the members of the various Peer Panels whose active support and involvement guided the lead authors
and the process.

FOREWORD
Foreword - ii

I trust that the low volume roads manual will provide the essential information needed to guide our
design engineers in the provision of appropriate and sustainable low volume roads.

Zaid Wolde Gebriel

Director General of the Ethiopian Roads Authority

FOREWORD
Preface - iii

PREFACE

The Ethiopian Roads Authority is the custodian of the series of technical manuals, standard specifications
and bidding documents that are written for the practicing engineer in Ethiopia. The series describe
current and recommended practice and set out the national standards for roads and bridges. They are
based on national experience and international practice and are approved by the Director General of the
Ethiopian Roads Authority.

This Design Manual for Low Volume Roads (2011) forms part of the Ethiopian Roads Authority series of
Road and Bridge Design documents.

Companion documents and manuals include the Standard Technical Specifications, Standard Detailed
Drawings and Standard Bidding Documents.

The complete series of documents, covering all roads and bridges in Ethiopia, are contained within the
series:
ƒ Geometric Design Manual – 2011
ƒ Site Investigation Manual – 2011
ƒ Geotechnical Design Manual for Slopes - 2011
ƒ Pavement Design Manual Volume I – 2011 (Flexible and Unpaved Pavements)
ƒ Pavement Design Manual Volume II – 2011 (Rigid Pavements)
ƒ Drainage Design Manual – 2011
ƒ Bridge Design Manual – 2011
ƒ Low Volume Roads Design Manual – 2010
ƒ Standard Environmental Procedures Manual – 2002
ƒ Standard Technical Specifications – 2011
ƒ Standard Detailed Drawings – 2011
ƒ Standard Bidding Documents for Road Work Contracts – A series of Bidding Documents covering a
full range from large scale projects unlimited in value to works with an upper threshold of $300,000.
The higher level documents have both Local Competitive Bidding and International Competitive
Bidding versions – 2011

These documents are available to registered users through the ERA website: www.era.gov.et

1.1 Manual Updates


Significant changes to criteria, procedures or any other relevant issues related to new policies or revised
laws of the land or that is mandated by the relevant Federal Government Ministry or Agency should be
incorporated into the manual from their date of effectiveness.

Other minor changes that will not significantly affect the whole nature of the manual may be accumulated
and made periodically. When changes are made and approved, new page(s) incorporating the revision,
together with the revision date, will be issued and inserted into the relevant chapter.

This version of the Low Volume Roads Manual is released in draft for a limited period during which the
industry at large is encouraged to put it into practice and to feed back to the ERA Director General any
suggestions for inclusion or improvement.

All suggestions to improve the draft Low Volume Roads Design Manual should be made in accordance
with the following procedures:

ƒ Users of the manual must register on the ERA website: www.era.gov.et


ƒ Proposed changes should be outlined on the Manual Change Form and forwarded with a covering
letter of its need and purpose to the Director General of the Ethiopian Roads Authority.

PREFACE
Preface - iv

ƒ After completion of the draft review period, proposed modifications will be assessed by the
requisite authorities in ERA, in close consultation with the Peer Panels and the AFCAP project
team.
ƒ Agreed changes will be approved by the Director General of the Ethiopian Roads Authority on
recommendation from the Deputy Director General (Engineering Operations).
ƒ All changes to the manual will be made prior to release of the final version of the manual.
ƒ The release date of the final version will be notified to all registered users and authorities.

PREFACE
Preface - v

ETHIOPIAN ROADS AUTHORITY


CHANGE CONTROL DESIGN MANUAL

This area to be completed by the ERA


MANUAL CHANGE Director of Quality Assurance

Manual Title CHANGE NO.


(SECTION NO. CHANGE NO

Section
Table
Explanation Suggested Modification
Figure
Page

Submitted by: Name/Designation:


Company/Organisation Address
email: Date:
Manual Change Action

Authority Date Signature Recommended Action Approval


Registration

Director Quality Assurance

Directors Region

Deputy Director General Eng.Ops

Approval / Provisional Approval / Rejection of Change:

Director General ERA Date:

PREFACE
Acknowledgements - vi

ACKNOWLEDGEMENTS

The Ethiopian Roads Authority (ERA) wishes to thank the UK Government’s Department for International
Development (DFID) through their Africa Community Access Programme (AFCAP) for their support
in developing this Low Volume Roads design manual. The manual will be used by all authorities and
organisations responsible for the provision of low volume roads in Ethiopia.

From the outset, the approach to the development of the manual was to include all sectors and
stakeholders in Ethiopia. The input from the international team of experts was supplemented by our own
extensive local experience and expertise. Local knowledge and experience was shared through a series
of four “information gathering” workshops followed by a review workshop to discuss and debate the
contents of the draft manual. ERA wishes to thank all the individuals who gave their time to attend the
five workshops and provide valuable inputs to the compilation of the manual.

In addition to the workshops, Peer Groups comprising specialists drawn from within the local industry,
were established to provide advice and comments in their respective areas of expertise. The contribution
of the Peer Group participants is gratefully acknowledged.

The final review and acceptance of the document was undertaken by an Executive Review Group. Special
thanks are given to this group for their assistance in reviewing the final draft of the document.

Finally, ERA would like to thank Crown Agents for their overall management of the project

Executive Review Group


Individual Organisation
Bizuneh Kebebe Civil Works Consulting Engineers
Efrem Degefu Beacon Consulting Engineers plc
Bekele Jebessa Ethioinfra Consultants
Dawit Dejene Civil Works Consulting Engineers
Shimeles Tesfaye SPICE
Teshome Worku CORE Consulting Engineers
Assefa Addisu Assefa Social Consult
Frew Bekele Ethiopian Roads Authority
Biniam Girma Ethiopian Roads Authority
Alemayehu Ayele Ethiopian Roads Authority

ACKNOWLEDGEMENTS
Acknowledgements - vii

List of Persons Contributing to Peer Group Reviews


Peer Group

Materials & Pavement Design and


Geometric Design and Standard

Complementary Interventions
Standard Bidding Document
Technical Specifications
Site Investigation

Route Selection
Drawings

Drainage

Bridges
No Name Organisation
1 Abdissa Megerssa ERA 9
Abdo
2 ERA 9
Mohammed
Abdurkerim Amhara Road
3 9 9
Mohamed Authority
4 Abeba Berhanu ERA 9
ERA: Design &
5 Abebe Asefa, Ato Technical Support 9 9
Branch
Oromia Roads
6 Addisu Mosissa 9
Authority
7 Alemayehu Mamo CWCE 9
Amanuel
8 ERA 9
Haddush
9 Amarech Fikera ERA 9
Andualem H/ Beza Consulting
10 9
Georgis Engineers
Omega Consulting
11 Asnake Adamu 9
Engineers
Omega Consulting
12 Asnake Haile 9
Engineers
13 Asrat Sewit Saba Engineering 9
Assefa Social
14 Assefa Addisu 9
Consult
Ethioinfra
15 Bekele Jebessa 9 9
Consultants
Beneyam Dire Dawa Roads
16 9 9
Wubshet Authority
17 Berhanu Tadese AEC 9
18 Biniam Girma ERA 9 9
19 Bizuneh Kebebe CWCE 9 9
Laber Base
20 Brook Shamsu 9 9 9
Association

ACKNOWLEDGEMENTS
Acknowledgements - viii

List of Persons Contributing to Peer Group Reviews


Peer Group

Materials & Pavement Design and


Geometric Design and Standard

Complementary Interventions
Standard Bidding Document
Technical Specifications
Site Investigation

Route Selection
Drawings

Drainage

Bridges
No Name Organisation
21 Daniel Nebro ERA 9 9 9 9
22 Dawit Degenie CWCE 9 9
Dessalegn Ministry of
23 9 9
Bezabih Agriculture
Beacon Consulting
24 Efrem Degefu 9
Engineers
Efrem Gebre Core Consulting
25 9
Egziabher Engineers
Engda
26 Private 9
Zemedagegnehu
27 Ermias Abate CWCE 9
Transnational
28 Ermias Ketema 9 9
Engineers
29 Ezra Mersha AEC 9
Harar Road
30 Fethi Tewfik 9
Authority
31 Fikert Arega ERA 9 9
32 Frew Bekele ERA 9 9
33 Gashawe Girma ERA 9
34 Genet Woldie Bridge Engineer 9
35 Girma Worku ERA 9
36 Haddis Tesfaye ERA 9 9
Tigray Road
37 Hailay Techlu 9
Authority
Oromia Roads
38 Iffa Hussein 9
Authority
Oromia Roads
39 Jamariyo Jarso 9
Authority
40 Jemal Ali WT Consult 9 9 9
Dire Dawa Roads
41 Marshet Biru 9 9
Authority

ACKNOWLEDGEMENTS
Acknowledgements - ix

List of Persons Contributing to Peer Group Reviews


Peer Group

Materials & Pavement Design and


Geometric Design and Standard

Complementary Interventions
Standard Bidding Document
Technical Specifications
Site Investigation

Route Selection
Drawings

Drainage

Bridges
No Name Organisation
Metaferia
Mekari
42 Consulting 9
Zemedagegnehu
Engineers
Construction
Contractors
43 Melaku Tadesse 9 9 9
Association of
Ethiopia
Tigray Road
44 Mewael G/Kidan 9
Authority
Somali Rural Roads
45 Mohamed Muktar 9 9
Authority
Mulu Haile Dire Dawa Roads
46 9
Michael Authority
47 Muse Belew ERA 9 9
Core Consulting
48 Saba Million 9
Engineers
49 Shimeles Tesfaye SPICE 9
Dire Dawa Roads
50 Sintayehu Melaku 9
Authority
SNNP Roads
51 Sisay Belachew 9
Authority
52 Solomon Kulich Scott Wilson ARE 9
53 Taddele Debela Private 9
54 Tadesse, Ato Private 9
55 Tamrat Fikru Adigerate DRMC 9
56 Teferra Mengesha WT Consult 9
Core Consulting
57 Teshome Worku 9 9
Engineers
58 Tewodros Alene ERA 9 9
Ethioinfra
59 Tibebu Gebre 9
Consultants
60 Yoseph Kidane ERA 9 9

ACKNOWLEDGEMENTS
Introduction
Acknowledgements - x

Organisations represented in the LVR Project Workshops


1 Adigerate DRMC 28 Helvetase
2 AEC 29 Laber Base Association
3 Afar Rural Roads Authority 30 Lea Consult
4 AFCAP/Crown Agents 31 Mekele University
5 Amhara Region TVET Centre 32 METAFERIA
6 Amhara Rural Roads Authority 33 Ministry of Agriculture
7 ARRB, Australia 34 Mycube, South Africa
Beacon Consulting Engineers Nazerete Tractor Assembly Factory
8 35
plc Plant
Benishanguel-Gumuz Rural
9 36 Omega Consulting Engineers
Roads Authority
10 Beza Consult 37 Oromia Roads Authority
Construction Contractors
11 38 Roughton
Association of Ethiopia
12 Core Consulting Engineers plc. 39 Saba Engineering
13 C-Tech Engineering 40 Scott Wilson ARE
14 CWCE 41 Shashemene
15 DFID, UK 42 SNNP Road Authority
Somali Region Rural Roads
16 Dire Dawa Roads Authority 43
Authority
17 East Hararge 44 Sunshine Construction
18 Ethiopia Road Fund 45 Tigray Road Authority
19 Ethiopian Roads Authority 46 Transnational Engineers
20 Ethioinfra Consultants 47 TRL
Ethiopian Institue of Agricultural
21 48 TRRA
Research
22 Finn Road 49 Unicone
23 Gambella Rural Roads Authority 50 West Hararge
Gimbichu Wereda
24 51 World Bank
Administration
Gonder District Engineer
25 52 WSP
Division
26 Hamda Engineering Consult plc 53 WT Consult
27 Harar Road Authority 54 Yencomad

ACKNOWLEDGEMENTS
Acknowledgements - xi

Project Team
Name Organisation Role
Bekele Negussie ERA AFCAP Coordinator for Ethiopia
Abdo Mohammed ERA Project Coordinator
Daniel Nebro ERA Project Coordinator
Abeba Berhanu ERA Project Coordinator
Frew Bekele ERA Project Coordinator
Colin Gourley ERA and DFID Specialist Adviser
Rob Geddes AFCAP/Crown Agents Technical Manager
Les Sampson AFCAP/Crown Agents Project Director
John Rolt AFCAP/Crown Agents Lead Author
Mike Pinard AFCAP/Crown Agents Lead Author
Lulsaged Ayalew AFCAP/Crown Agents Lead Author
Bob Carson AFCAP/Crown Agents Contracts Expert
Rob Petts AFCAP/Crown Agents Specialist Contributor
Jasper Cook AFCAP/Crown Agents Specialist Contributor
Gareth Hearns AFCAP/Crown Agents Specialist Contributor
Gerrie van Zyl AFCAP/Crown Agents Specialist Contributor
Pat Loots AFCAP/Crown Agents Specialist Contributor

ACKNOWLEDGEMENTS
Acronyms - xii

ACRONYMS

> : Greater than


< : Less than
% : Percentage

A
AADT : Annual Average Daily Traffic
AASHTO : American Association of State Highway and Transportation Officials
AFCAP : Africa Community Access Programme
AIDS : Acquired Immune Deficiency Syndrome or
Acquired Immunodeficiency Syndrome
ALD : Average Least Dimension
ARRB : ARRB Group, formerly the Australian Road Research Board
ARVs : Antiretroviral
ASTM : American Standard Test Methods

B
BDS : Bid Data Sheet

C
CB : Clay Brick (fired)
CBO : Community Based Organisation
CBR : California Bearing Ratio
CI : Complementary Interventions
CMG : Crown Agents Core Management Group
COLTO : Commission of Land Transport Officials (South Africa)
CPT : Cone Penetrometer Test
CS : Cobblestone

D
DBM : Drybound macadam
DC : Design Class
DCP : Dynamic Cone Penetrometer
DF : Drainage Factor
DFID : UK Government’s Department of International Development
DMT : Dilatometer Test
DBBM : Drybound Macadam
DS : Dressed Stone
DV : Design Vehicles

E
EF : Equivalency Factor
eg : For example (abbreviation for the Latin phrase exempli gratia)
EIA : Environmental Impact Assessment
EMP : Environmental Management Plan
ENS : Engineered Natural Surfaces
EOD : Environmentally Optimised Design
ERA : Ethiopian Roads Authority
ERTTP : Ethiopian Rural Travel and Transport Programme
esa : Equivalent standard axles
EVT : Equiviscous Temperature

ACRONYMS
Acronyms - xiii

F
FACT : Fine Aggregate Crushing Test
FED : Final Engineering Design

G
g/m² : Grams per Square Metre
GDP : Gross Domestic Product
GM : Grading Modulus
gTKP : Global Transport Knowledge Partnership
GVW : Gross Vehicle Weight

H
ha : Hectare
HDM 4 : World Bank’s Highway Design and Maintenance Standards model
HIV : Human immunodeficiency virus
HPS : Hand Packed Stone
HVR : High Volume Road

I
ICB : International Competitive Bidding
ICT : Information Communication Technology
IDA : International Development Agency
ie : That is (abbreviation for the Latin phrase id est)
ILO : International Labour Organisation
IMT : Intermediate Means of Transport
IRR : Internal Rate of Return
ITB : Instructions to Bidders

K
km : Kilometre
km² : Square Kilometre
km/h : Kilometres per hour
km/hr : Kilometres per hour

L
LIC : Labour Intensive Construction
LVR : Low Volume Road

M
m : Metre
m² : Square Metre
m³ : Metres Cubed
MCB : Mortared Clay Brick (fired)
MCS : Mortared Cobblestones
MDS : Mortared Dress Stone
Mesa : Million equivalent standard axles
mg/m³ : Milligram per metre cubed
mm : Millimetre
mm² : Square Millimetre
mm³ : Millimetres Cubed
m/s : Metres per second
MC : Medium Curing
MoFED : Ministry of Finance and Economic Development
MPa : megapascal (a unit of pressure equal to 1000 kilopascals (kPa),
commonly used in the building industry to measure crushing pressure
of bricks)
MS : Mortared Stone

ACRONYMS
Acronyms - xiv

MSSP : Mortared Stone Setts or Pavé


MWUD : Ministry of Works and Urban Design

N
NBP : Non-Bituminous Pavement
NCB : National Competitive Bidding
NCT : National Competitive Tendering
NGO : Non-Government Organisation
nm : Nanometre
NMT : Non-Motorised Transport
NRC : Non-reinforced concrete
NRCP : Non-reinforced concrete pavement

O
OMC : Optimal Moisture Content
ORN : Overseas Road Note
ORRA : Oromiya Regional Roads Authority

P
PCU : Passenger Car Unit
PDM : Pavement Design Manual
Pen. : Penetration
PI : Plasticity Index
PM : Pressure Meter
PPA : Public Procurement Agency
PPP : Public Private Partnership
PSD : Particle Size Distribution
PSNP : Productive Safety Net Programme

R
R : Radius
RC : Reinforced concrete
Ref : Reference (Part C, Page 13)
RFP : Request for Proposals
RRA : Regional Road Authority
RS : Road Safety
RSC : Research Steering Committee
RTS : Road Transport Service

S
SADC : Southern Africa Development Community
SBL : Sand Bedding Layer
SDMS : Surfacing Decision Management System
SE : Super Elevation
SMEs : Small and Medium Enterprises
SSP : Stone Setts or Pavé

T
TBA : To Be Advised
Tc : Time of Concentration
ToR : Terms of Reference
TRL : Transport Research Laboratory

U
UK : United Kingdom
USA : United States of America
USCS : Unified Soil Classification System

ACRONYMS
Acronyms - xv

USD : United States Dollar


UTRCP : Ultra Thin Reinforced Concrete Pavement

V
VI : Impinging Velocity
VAVE : Average Velocity
VP : Parallel Velocity
vpd : vehicles per day
VOCs : Vehicle Operating Costs
VST : Vane Shear Test

W
WBM : Waterbound Macadam
WC : Wearing Course

ACRONYMS
Glossary of technical terms - xvi

GLOSSARY OF TECHNICAL TERMS

Aggregate (for construction)


A broad category of coarse particulate material including sand, gravel, crushed stone, slag and recycled
material that forms a component of composite materials such as concrete and asphalt.

Asphalt
A mixture of inert mineral matter, such as aggregate, mineral filler (if required) and bituminous binder in
predetermined proportions.

Atterberg limits
Basic measures of the nature of fine-grained soils which identify the boundaries between the solid, semi-
solid, plastic and liquid states.

Binder, Bituminous
Any bitumen based material used in road construction to bind together or to seal aggregate or soil
particles.

Binder, Modified
Bitumen based material modified by the addition of compounds to enhance performance. Examples of
modifiers are polymers, such as PVC, and natural or synthetic rubbers.

Bitumen
A non-crystalline solid or viscous mixture of complex hydrocarbons that possesses characteristic
agglomerating properties, softens gradually when heated, is substantially soluble in trichlorethylene and
is obtained from crude petroleum by refining processes.

Bitumen, Cutback
A liquid bitumen product obtained by blending penetration grade bitumen with a volatile solvent to
produce rapid curing (RC) or medium curing (MC) cutbacks, depending on the volatility of the solvent
used. After evaporation of the solvent, the properties of the original penetration grade bitumen become
operative.

Bitumen, Penetration Grade


That fraction of the crude petroleum remaining after the refining processes which is solid or near solid at
normal air temperature and which has been blended or further processed to products of varying hardness
or viscosity.

Bitumen emulsion
An emulsion of bitumen and water with the addition of an emulsifier or emulsifying agent to ensure
stability. Conventional bitumen emulsion most commonly used in road works has the bitumen dispersed
in the water. An invert bitumen emulsion has the water dispersed in the bitumen. In the former, the
bitumen is the dispersed phase and the water is the continuous phase. In the latter, the water is the
dispersed phase and the bitumen is the continuous phase. The bitumen is sometimes fluxed to lower its
viscosity by the addition of a suitable solvent.

Bitumen Emulsion, Anionic


An emulsion where the emulsifier is an alkaline organic salt. The bitumen globules carry a negative
electrostatic charge.

Bitumen Emulsion, Cationic


An emulsion where the emulsifier is an acidic organic salt. The bitumen globules carry a positive
electrostatic charge.

GLOSSARY OF TECHNICAL TERMS


Glossary of technical terms - xvii

Bitumen Emulsion Grades


Premix grade: An emulsion formulated to be more stable than spray grade emulsion and suitable
for mixing with medium or coarse graded aggregate with the amount smaller than
0.075mm not exceeding 2%.
Quick setting grade: An emulsion specially formulated for use with fine slurry seal type aggregates,
where quick setting of the mixture is desired.
Spray grade: An emulsion formulated for application by mechanical spray equipment in chip
seal construction where no mixing with aggregate is required.
Stable mix grade: An emulsion formulated for mixing with very fine aggregates, sand and crusher
dust. Mainly used for slow-setting slurry seals and tack coats.

Cape Seal
A single application of binder and stone followed by one or two applications of slurry.

Cement (for construction)


A dry powder which on the addition of water and other additives, hardens and sets independently to bind
aggregates together to produce concrete.

Chip Seal, Single


An application of bituminous binder followed by a layer of stone or clean sand. The stone is sometimes
covered with a fog spray.

Chip Seal, Double


An application of bituminous binder and stone followed by a second application of binder and stone or
sand. A fog spray is sometimes applied on the second layer of aggregate.

Collapsible soil
Soil that undergoes a significant, sudden and irreversible decrease in volume upon wetting.

Complimentary Interventions
Actions that are implemented through a roads project which are targeted toward the communities that lie
within the influence corridor of the road and are intended to optimise the benefits brought by the road
and to extend the positive, and mitigate the negative, impacts of the project.

Concrete
A construction material composed of cement (commonly Portland cement) as well as other cementitious
materials such as fly ash and slag cement, aggregate (generally a coarse aggregate such as gravel or
crushed stone plus a fine aggregate such as sand), water, and chemical admixtures.

Concrete Block Paving


A course of interlocking or rectangular concrete blocks placed on a suitable base course and bedded and
jointed with sand.

Crushed Stone
A form of construction aggregate, typically produced by mining a suitable rock deposit and breaking the
removed rock down to the desired size using crushers.

Design speed
The maximum safe speed that can be maintained over a specified section of road when conditions are so
favourable that the design features of the road govern the speed.

Dispersive soil
Soil in which the clay particles detach from each other and from the soil structure in the presence of water
and go into suspension.

GLOSSARY OF TECHNICAL TERMS


Glossary of technical terms - xviii

Distributor
A vehicle comprising an insulated tank with heating and circulating facilities and a spray bar capable of
applying a thin, uniform and predetermined layer of binder.

Expansive soil
Typically clayey soil that undergoes large volume changes in direct response to moisture changes.

Filler
Mineral matter composed of particles smaller than 0.075mm.

Fog Spray
A light application of diluted bitumen emulsion to the final layer of stone of a reseal or chip seal or to an
existing bituminous surfacing as a maintenance treatment.

Gravel
A naturally-occurring, weathered rock within a specific particle size range. In geology, gravel is any loose
rock that is larger than 2mm in its largest dimension and not more than 63mm.

Kabele
Administrative division in Ethiopia equivalent to sub-district or ward. Smallest administrative unit in
Ethiopia.

Labour Intensive Construction


Economically efficient employment of as great a proportion of labour as is technically feasible throughout
the construction process to produce the standard of construction as demanded by the specification
and allowed by the available funding. Substitution of equipment with labour as the principal means of
production.

Low Volume Road


Roads carrying up to about 300 vehicles per day and less than about 1 million equivalent standard axles
over their design life.

Otta Seal
A carpet of graded aggregate spread over a freshly sprayed hot bituminous binder.

Prime Coat
A coat of suitable bituminous binder applied to a non-bituminous granular pavement layer as a preliminary
treatment before the application of a bituminous base or surfacing. While adhesion between this layer
and the bituminous base or surfacing may be promoted, the primary function of the prime coat is to assist
in sealing the surface voids and bind the aggregate near the surface of the layer.

Reseal
A surface treatment applied to an existing bituminous surface.

Rejuvenator
A material (which may range from a soft bitumen to petroleum) which, when applied to reclaimed asphalt
or to existing bituminous surfacing, has the ability to soften aged, hard, brittle binders.

Seal
A term frequently used instead of “reseal” or “surface treatment”. Also used in the context of “double
seal” and “sand seal” where sand is used instead of stone.

Selected layers
Pavement layers of selected gravel materials used to bring the subgrade support up to the required
structural standard for placing the subbase or base course.

GLOSSARY OF TECHNICAL TERMS


Glossary of technical terms - xix

Site Investigation
Collection of essential information on the soil and rock characteristics, topography, land use, natural
environment, and socio-political environment necessary for the location, design and construction of a
road.

Slurry
A mix of suitably graded fine aggregate, cement or hydrated lime, bitumen emulsion and water, used for
filling the voids in the final layer of stone of a new surface treatment or as a maintenance treatment (also
referred to as a slurry seal).

Slurrybound Macadam
A surfacing layer constructed where the voids in single-sized stone skeleton are filled using bituminous
slurry.

Subgrade
The native material underneath a constructed road pavement.

Surface Treatment
A general term incorporating chip seals, micro surfacing, fog sprays or tack coats.

Surfacing
The layer with which traffic makes direct contact.

Tack Coat
A coat of bituminous binder applied to a primed layer or to an existing bituminous surface as a preliminary
treatment to promote adhesion between the existing surface and a subsequently applied bituminous
layer.

Ultra-thin Reinforced Concrete Pavement (UTRCP)


A layer of concrete, 50 mm thick, continuously reinforced with welded wire mesh.

Wearing Course
The upper layer of a road pavement on which the traffic runs and is expected to wear under the action
of traffic.

Waterbound Macadam
A pavement layer constructed where the voids in a large single-sized stone skeleton are filled with a fine
sand.

Wereda
Administrative division in Ethiopia equivalent to district.

GLOSSARY OF TECHNICAL TERMS


Table of contents - xx

TABLE OF CONTENTS

PART A
A. TABLE OF CONTENTS ......................................................................................................A.III
A. LIST OF FIGURES ............................................................................................................. A.IV
A. LIST OF PLATES ................................................................................................................ A.V
1. CONTEXT AND SCOPE OF THE LOW VOLUME ROADS MANUAL .................................. A.1
2. STRUCTURE AND LAYOUT ............................................................................................. A.10
3. ROAD PROJECT IMPLEMENTATION ............................................................................. A.14

PART B
B. TABLE OF CONTENTS ........................................................................................................B.I
B. LIST OF TABLES ................................................................................................................ B.III
B. LIST OF FIGURES .............................................................................................................. B.V
1. INTRODUCTION ................................................................................................................B.1
2. POLICY AND LEGISLATIVE CONTROLS ............................................................................B.2
3. DESIGN PARAMETERS ......................................................................................................B.4
4. GEOMETRIC DESIGN STANDARDS ...................................................................................B.9
5. MATERIALS ......................................................................................................................B.27
6. PAVEMENT DESIGN ........................................................................................................B.38
7. DRAINAGE AND EROSION CONTROL ...........................................................................B.50
8. WATER CROSSINGS AND ASSOCIATED STRUCTURES ..................................................B.70
9. ROAD FURNITURE AND SIGNAGE .................................................................................B.71
B. APPENDIX B.1 - RAINFALL INTENSITY-DURATION-FREQUENCY CHARTS .....................B.73

PART C
C. TABLE OF CONTENTS ......................................................................................................C.III
C. LIST OF TABLES ............................................................................................................... C.IV
C. LIST OF PLATES ................................................................................................................ C.V
1. CONTEXT AND APPLICATION OF COMPLEMENTARY INTERVENTIONS .........................C.1
2. PLANNING, IDENTIFICATION AND IMPLEMENTATION OF COMPLEMENTARY
INTERVENTIONS ...............................................................................................................C.3
3. EMPLOYMENT AND HUMAN RESOURCE ISSUES ............................................................C.7
4. CONTRACT PROVISIONS TO SUPPORT COMPLEMENTARY INTERVENTIONS ...............C.10
5. SUPPORTING SMALL SCALE CONTRACTORS.................................................................C.16
6. SUPERVISION CONSULTANT’S CONTRACT ....................................................................C.17
7. MANAGEMENT, MONITORING AND ENFORCEMENT ...................................................C.18
8. REFERENCES ...................................................................................................................C.19
C. APPENDIX C.1 - COMPLEMENTARY INTERVENTIONS – PLANNING,
FEASIBILITY AND PRELIMINARY DESIGN STAGES..........................................................C.20
C. APPENDIX C.2 - INDICATIVE EXAMPLES OF CIS BY CATEGORY AND THEME ..............C.23

TABLE OF CONTENTS
Table of contents
Introduction
- xxi

PART D
D. TABLE OF CONTENTS ....................................................................................................... D.I
D. LIST OF TABLES ............................................................................................................... D.VI
D. LIST OF FIGURES ............................................................................................................. D.IX
D. LIST OF PLATES ................................................................................................................ D.X
1. GENERAL INTRODUCTION .............................................................................................. D.1
2. SITE INVESTIGATION FOR ROUTE SELECTION AND DESIGN ......................................... D.2
3. ROADSIDE SLOPE STABILISATION ................................................................................ D.26
4. GEOMETRIC DESIGN ..................................................................................................... D.43
5. DRAINAGE ..................................................................................................................... D.75
6. MATERIALS AND PAVEMENT DESIGN ........................................................................ D.110
7. SURFACING .................................................................................................................. D.179
D. APPENDIX D.1 ............................................................................................................. D.208
D. APPENDIX D.2 ............................................................................................................. D.257
D. APPENDIX D.3 ............................................................................................................. D.261

PART E
E. TABLE OF CONTENTS ...................................................................................................... E.III
E. LIST OF TABLES ................................................................................................................E.IX
E. LIST OF FIGURES ..............................................................................................................E.XI
E. LIST OF PLATES ............................................................................................................. E.XIV
1. INTRODUCTION ................................................................................................................E.1
2. PROJECT PLANNING ........................................................................................................E.7
3. DESIGN CRITERIA ...........................................................................................................E.14
4. STRUCTURAL OPTIONS ..................................................................................................E.21
5. SITE SELECTION AND APPRAISAL ..................................................................................E.35
6. WATERCOURSE CHARACTERISTICS ................................................................................E.47
7. MATERIALS ......................................................................................................................E.56
8. STRUCTURE DESIGN .....................................................................................................E.100
9. CONSTRUCTION ...........................................................................................................E.168
10. MAINTENANCE .............................................................................................................E.192
11. REFERENCES .................................................................................................................E.199

TABLE OF CONTENTS
PART A
INTRODUCTION TO LOW VOLUME ROAD DESIGN
Road network classification

Low volume design principles

Low volume earth and


gravel roads

Context and scope Low volume paved roads

Surface improvement technology

Context sensitivity
Part A
Introduction to low
volume road design Protection of the environment

Structure and layout

Contacts of works

Project implementation
Supervision services
iii

A TABLE OF CONTENTS

A. TABLE OF CONTENTS ......................................................................................................A.III


A. LIST OF FIGURES ............................................................................................................. A.IV
A. LIST OF PLATES ................................................................................................................ A.V
1. CONTEXT AND SCOPE OF THE LOW VOLUME ROADS MANUAL .................................. A.1
1.1 Road network classification .............................................................................................A.2
1.2 Low volume road design principles.................................................................................A.4
1.3 Low volume earth and gravel roads ................................................................................A.5
1.4 Low volume paved roads ................................................................................................A.6
1.5 Surface improvement technology ...................................................................................A.7
1.6 Context sensitivity ..........................................................................................................A.7
1.6.1 Political support ................................................................................................A.7
1.6.2 Social acceptance .............................................................................................A.8
1.6.3 Institutional capacity .........................................................................................A.8
1.6.4 Technology choice ............................................................................................A.8
1.6.5 Economic viability ............................................................................................A.9
1.6.6 Financially sound .............................................................................................A.9
1.7 Protection of the environment.........................................................................................A.9
2. STRUCTURE AND LAYOUT ............................................................................................. A.10
3. ROAD PROJECT IMPLEMENTATION ............................................................................. A.14
3.1 Contract of works ..........................................................................................................A.14
3.2 Supervision services .....................................................................................................A.16

PART A: DESIGN STANDARD APPROACHES


iv

A LIST OF FIGURES

Figure A.1.1: Road classes in Ethiopia ................................................................................................. A.3


Figure A.1.2: Framework for sustainable provision of low volume roads ............................................ A.5
Figure A.1.3: Framework for sustainable provision of low volume roads ............................................ A.8
Figure A.2.1: The structure of the Manual ......................................................................................... A.11

PART A: DESIGN STANDARD APPROACHES


v

A LIST OF PLATES

Plate A.1.1: Typical access problems ..................................................................................................... A.5

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 1

1. CONTEXT AND SCOPE OF THE LOW


VOLUME ROADS MANUAL
This Low Volume Roads Manual promotes rational, appropriate and affordable designs for low volume
roads in Ethiopia. In doing so it aims at making cost effective and sustainable use of local resources. The
Manual reflects local experience and advances in low volume road technology gained in Ethiopia and
elsewhere.

Application of appropriate design standards for low volume roads in Ethiopia aims to optimise construction
and maintenance costs and meet requirements to:
ƒ Improve the economic and social well being of rural communities and access to social and other
services;
ƒ Facilitate inclusion of different ethnic and other groups in society;
ƒ Lower road user costs and promote socio-economic development, poverty reduction, trade growth
and wealth creation in rural areas;
ƒ Protect and manage non-renewable natural resources and reduce import dependency.

This manual is intended for use by roads practitioners responsible for the design of low traffic earth, gravel
or paved roads in Ethiopia. It is appropriate for roads which, over their design life, are required to carry
an average of up to about 300 vehicles per day, and less than about 1.0 million equivalent standard axles
(Mesa) in one direction. The Manual complements and links to the “Pavement Design Manual – Volume
1 - Flexible Pavements” and is accompanied by separate volumes dealing with Technical Specifications,
Drawings and appropriate level Bidding Documents.

The client for the low volume road works could be a Wereda Administration, a Regional Roads Authority
or the Ethiopian Roads Authority. The client could also be a local level administration such as a Kebele
Administration, community organisation or cooperative. Road works, whether undertaken by a contractor,
through an in-house capability or by community will require a design. This design will work towards
satisfying a national standard set for a particular type of road. The degree of sophistication of the design
will in general increase as the standard of the road increases. However, this does not mean that unpaved
earth or gravel roads are any easier to design than a first generation low volume sealed road. Often it is
quite the contrary.

The road design engineer is normally supported by a team of individuals, with varying specialities, and
equipped to deal with all aspects of the road design. The job of the design team is to provide a robust
technical design (geometric, drainage and pavement) and to reflect this in the instructions to bidders,
conditions of particular application, the specifications, the bills of quantities and the detailed drawings.
The design team should also include (or consult) environmentalists and sociologists for additional
specialist inputs.

The general approach to the design will be guided by the client and will build on information and data
collected during the project pre-feasibility and feasibility stages. The client will have a budget in mind for
the works, the location and route will be known in outline, and the preferred approach to the works will
also be known, for example labour or equipment based. The client may also have views and guidance on
apportioning works and contract size, technical issues, social, environmental and time constraints. The
job of the road design engineer will then be to develop the project within and around these boundaries
and limitations, whilst at the same time alerting the client to issues and problems that may limit or require
adjustment of expectations.

The approach to the design of low volume roads follows the general principles of any good road design
practice. There are, however, subtle differences from the traditional road design practice. This manual sets
out to provide the design engineer with the requisite tools that will provide the client with an optimised
design based on the financial, technical and other constraints that define the project.

Optimising a design requires a multi-dimensional understanding of all of the project elements and in this
respect all design elements become context specific. The design team therefore needs to be able to

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 2

work outside their normal areas of expertise and to understand implications of their recommendations or
decisions on all other elements of the design.

The successful design of low volume roads relies on:


ƒ A full understanding by the design engineer of the local environment (natural and social);
ƒ An ability to work within the demands of the local environment and to turn these to a design
advantage;
ƒ Recognition and management of risk;
ƒ Innovative and flexible thinking through the application of appropriate engineering solutions
rather than following traditional thinking related to road design;
ƒ A client who is open and responsive to innovation;
ƒ Guaranteed routine and periodic maintenance.

There is an onus on the design engineer to provide a road that meets the expected level of service. Design
engineers are traditionally conservative and build in factors of safety that cater for their perceptions
of risk and extremes of caution. This approach prevents the application of innovation, uses scarce or
inappropriate resources and results in high financial costs for the client and the country. There is also
often a temptation to provide or upgrade roads to a future level of service not justified by the economic
or other project projections; or road user requirements. This type of approach absorbs available resources
and prevents extension of access. It is the role of the design engineer to properly represent the clients
and country’s interests.

The level of attention and engineering judgement required for optimal provision of low volume roads is
no different and in most cases is higher than that required for the provision of other roads. The design
engineer needs to draw on all of his engineering skills, judgement and local experience if appropriate
designs are to be developed without incurring unacceptable levels of risk. This manual will assist the
engineer in that task.

The Manual is fully compatible with the ERA Standard Specifications for Road Works, the ERA 2011 series
of design manuals for higher volume road, and ERA’s Standard Bidding Documents.

1.1 Road network classification


The functional classification of roads in Ethiopia is based on five classes:
ƒ Trunk roads: roads linking Addis Ababa to centres of international importance and to international
boundaries;
ƒ Link roads: connecting centres of national and international importance such as principal towns
and urban centres;
ƒ Main access: connecting centres of provincial importance;
ƒ Collectors: connecting locally important centres to each other or to a more important centre or to
a higher class road; and
ƒ Feeder roads: connecting minor centres such as a market to other parts of the network.

Low volume roads can be represented in all five of these functional classes.

Roads in Ethiopia can be further divided into three categories depending on ownership and the authority
responsible for them. These are:
ƒ Federal (the responsibility of the Ethiopian Roads Authority);
ƒ Regional (the responsibility of the Regional or Rural Roads Authorities); and
ƒ Other rural roads (the responsibility of local authorities at Wereda or Kebele level or communities).

ERA is responsible for major roads falling into the higher design classes, predominantly DC5 and above,
but also has a substantial stock of roads below DC5. Regional and local authorities are responsible for
roads in classes DC4 – DC1.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 3

Figure A.1.1 shows the definitive classification of roads in Ethiopia based on geometric standards with the
appropriate level of service as defined below Figure A.1.1.

Geometric Level of
Road Functional Classification AADT
Standards Service

DC8 >10,000

HIGH VOLUME
A
DC7 3,000 - 10,000

TRUNK
DC6 1,000 - 3,000

LINK
B
DC5 300 - 1,000
MAIN ACCESS

DC4 150 – 300


COLLECTOR

DC3 75 – 150
LOW VOLUME
C
FEEDER

DC2 25 – 75

DC1 <25
D
Track

Figure A.1.1: Road classes in Ethiopia

Level A: The highest level of service. Traffic is free flowing, with the volumes and types of traffic easily
accommodated. Safety is a high priority. Design speed is very important and takes precedence over
topographic constraints.

Level B: Traffic may not flow smoothly in all situations. Safety is a high priority, but some safety controls
may need to be enforced. Design speed is important, but topography may dictate some design changes
and controls.

Level C: The efficiency of traffic movement and flow is not a limiting factor. Traffic will be accommodated,
but some design controls may need to be applied. Safety provisions are adapted to lower and variable
speed scenarios. The topography will dictate alignment and the design speed.

Level D: Service level is geared to provision of access rather than efficiency. Design standards for water-
crossings may allow service interruption and some roads may even be closed to protect these assets.
Other design standards for geometrics, surfacing and safety will reflect lower speed environments and
access requirement.

The density of roads in most areas of Ethiopia is relatively low and many existing low volume roads
are relatively long (>25km). Alternative routes are often long or non-existent and the consequences of
disruption are high. It is prudent therefore to adopt design standards that provide an appropriate level
of reliability and service commensurate with the functional characteristics of the road.

While there are exceptions to every rule, low volume roads in Ethiopia can be considered as roads
carrying less than 300vpd and generally of DC4 standard or below and meeting C or D service level
criteria. The majority of roads in Ethiopia carry relatively low levels of traffic, and most carry less than

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 4

about 300 vehicles per day. Such roads are referred to as “low-volume” roads in this manual and all are
an essential component of the road system. Their importance and reach extends to all aspects of the
economic and social development of rural communities and the country at large.

1.2 Low volume road design principles


Ethiopia is a country of great geographical, geological and climatic diversity. Altitudes range from the
highest peak at Ras Dejen, 4,620 meters above sea level, down to the Afar Depression at about 110
meters below sea level. Ethiopia’s high plateaus and mountain ranges, usually above 1500 meters are
characterised by precipitous edges and dissection by numerous rivers and streams. These areas constitute
about 45 percent of the total area and are inhabited by close to 80 percent of the population.

Below 1500 meters are the lowland areas, located in the north-west, east and south. The vast majority
of these areas support nomadic and semi-nomadic pastoralism. The descent to the southwest and west
leads to the semi–humid lowlands. Climatic regions range from arid, tropical rainy to temperate rainy
areas.

Soils are highly variable and often problematic, available materials for construction can be very variable
or scarce and involve long haulage.

Hence, traditional engineering, and traditional road engineering in particular, is challenging in the face
of such diversity. For low volume road provision the challenges can be even greater. Low volume roads
provide important links from homes, villages and farms to markets and offer communities access to
health, education and other services. These roads also provide important links between Kebele and
Wereda centres and the Federal road network. Given their importance, design engineers need to work
with and around such challenges. Clients also need to be flexible and adaptable, if low volume roads are
to be provided at reasonable cost.

Typically in Ethiopia, low volume roads are unpaved with a gravel or earth wearing surface. Very few
are in good condition and can provide the level of all-weather access that is required. Budgets for the
improvement and maintenance of these roads are constrained. In facing the challenge of improving
the low volume road network, the application of the conventional planning, design, construction and
maintenance philosophies used for higher traffic roads is unlikely to provide an optimal solution.

In determining cost-effective solutions for the provision of low volume roads it is important to understand
the mechanics of how the road deteriorates in the first place. Deterioration of the existing unpaved low
volume roads in Ethiopia is governed by the type of material used on the surface (gravel to soil); the
strength of the underlying soil (soft, erodible and/or expansive), the type and action of traffic (heavy
vehicle to pedestrian) and probably most importantly, the influence of the “road environment”. The term
“road environment” includes both the natural or bio-physical environment and the human environment.
It includes the interaction between the different environmental factors and the road structure. Some of
these factors are uncontrollable, such as those attributable to the natural environment, including the
interacting influence of climate (eg wind, rainfall and intensity), local hydrology and drainage, terrain and
gradient. Collectively, these will influence the performance of the road and the design approach needs
to recognise such influence by providing options that minimise the negative effects. Others factors, such
as the construction and maintenance regime; safety and environmental demands; and the extent and
type of traffic are largely controllable and can be more readily built into the design approach.

Typical road environment factors are presented in Figure A.1.2 and covered in more detail in Part D,
Chapter 6 of the Manual.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 5

Road Safety
Climate
Regime

Maintenance Surface/sub-
Regime Surface Hydrology

Construction The Road


Regime Environment Subgrade

“Green” Terrain
Environment

Controllable Construction Uncontrollable


Traffic Materials Factors
Factors

Figure A.1.2: Framework for sustainable provision of low volume roads

1.3 Low volume earth and gravel roads


Surface materials, where these are present, need to resist wear and abrasion in dry weather and promote
surface drainage and run-off in wet weather. Under traffic they need to resist whip off, dust generation and
be stable enough when compacted to resist deformation. The compacted material needs to resist erosion
and scour.

The nature and strength of the underlying soil is critical in determining the performance of low volume roads,
particularly in periods of wet weather. Many rural roads are characterised by deep rutting, where the road
formation is not strong enough to support the traffic loads. Some roads have loose and/or stony materials on
the surface, leading to dusty, rough and/or slippery conditions. Potholes create difficult and unsafe surfaces.
Severe erosion and scouring may prevent access by any form of motorised, and many types of intermediate,
transport. Transverse scouring can start at the edge or on the side slope of the road and work its way to the
centre of the carriageway. Longitudinal scour occurs where water flows against the direction of road crossfall.
Inadequate scour protection in drainage ditches may lead to serious erosion, dangerous conditions for road
users, local access restrictions, and loss of valuable agricultural land along the road.

Rural access may be prevented for long periods during the rains when streams and rivers start to flow. In
some situations wash-aways may occur. When the rains have eased or stopped the same points may be
subject to saturation and ponding. This weakens the underlying soils and any movement on the surface can
churn up the surface causing deep rutting and the bogging down of vehicles.

Plate A.1.1: Typical access problems

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 6

This problem is worsened in areas where there is a prevalence of expansive, black cotton soils. These soils
have high agricultural potential, but become weak and slippery when wet. They often cannot support
even the lightest vehicles. Where gravel is placed directly over this material it may rut under the influence
of traffic and mix into the weak soil below.

Vehicle operating costs (VOCs) are high on roads with high roughness and restricted access. VOCs
include repairs, maintenance, fuel and tyre replacement. The consequence is that transport operators
tend to avoid roads with high roughness and other defects forcing people to walk long distances to reach
the nearest point where transport services are prepared to operate.

Dust is often overlooked as a problem on unpaved roads. It is caused by the action of traffic and wind.
Unpaved roads lose fine material which can travel over 100 metres from the road. The dust affects other
road users, pedestrians and school children, houses, shops and crops near the road. Roads in dry areas
can lose up to 33 tonnes of surface fines per kilometre per year. Dust has significant and costly social
(cleanliness), health (eye and respiratory hazards), environmental (crop and natural habitat damage) and
economic (vehicle and equipment damage, pedestrian and vehicle safety) consequences. Approaches to
alleviate dust problems, particularly in populated areas are offered in the Manual.

Gravel for road works is a non-renewable natural resource. On unpaved roads it is used as a sacrificial
layer and must be periodically replaced. Optimal materials for gravel surfaced roads are not commonly
found in Ethiopia, and it is possible to lose up to 150mm of gravel per year depending on conditions.
Gravel roads require a continuous cycle of reshaping and regravelling to maintain the required running
surface and the desired level of service. The type of materials prevalent in Ethiopia, the nature of the
climate and the terrain presents significant challenges to achieving this type of maintenance. Screening
and blending techniques are available to improve the properties and such techniques are described in
the Manual.

The major technical challenges for unpaved roads are to provide durable and functional water crossings,
surfacing with materials that provide the desired and necessary level of service and to provide effective
maintenance management. These challenges are recognised in the Manual and in many cases options
and solutions are offered to mitigate and manage problems.

1.4 Low volume paved roads


The design manual draws on international research carried out over several decades. This research was
carried out by a number of research organisations in collaboration with national road authorities, including
the Ethiopian Roads Authority. Much of the research was aimed at deriving local specifications, designs
and techniques for improving the cost-effective provision of low volume roads sealed with a bituminous
or alternative, non-bituminous surfacing and advanced thinking on provision of more appropriate
geometric, drainage and pavement design standards. Innovative construction techniques and methods
were identified that optimised the use of local labour, introduced intermediate equipment techniques and
improved opportunities for the local private sector to participate in road construction and maintenance.

The research questioned existing paradigms associated with paved surfacing design for low volume
roads. This research, combined with local experience, provided a basis for understanding how such roads
deteriorate leading to the development of revised standards, specifications and design methods that
make better use of locally available materials and demonstrate an effective range of viable bituminous
and non-bituminous surfacing technologies.

The approaches adopted for the design of low volume paved roads differs in a number of fundamental
respects from roads carrying higher traffic volumes. In particular, the relative influences of road
deterioration factors are significantly different for low volume roads compared with higher volume roads.
A critical observation is that for sealed roads, carrying below about 1.0 Mesa pavement deterioration
was controlled mainly by how the road responded to environmental factors, such as moisture changes
in the pavement layers, fill and subgrade and to the effects of age hardening of bituminous surfacings.
The appropriate design options for low volume roads therefore need to be responsive to a wider range
of factors captured in the road environment, the most critical being the internal and external drainage.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 7

The role of the design engineer is to recognise and design to these parameters and optimise the design
to the expected performance. This is known as an environmentally optimised design (EOD) approach.
EOD takes account of road environment changes along the alignment and the design responds to these
changes (for more details see Part D, Section 6.13).

1.5 Surface improvement technology


Gravel and earth roads are particularly vulnerable to the effects of the road environment. A range of more
durable surfacing options, other than gravel or earth are available for low volume roads. These include
thin bituminous surfacings, and non-bituminous surfacings such as cobbles, hand packed stone and even
thin concrete. The selection, design and use of the various surfacing options are described in detail in
this manual in Part D, Chapter 7 with design standards presented in Part B, Chapter 3.4.

Improved surfacings may be provided for the entire length of a road, or only on the most vulnerable
sections. The approach may include dealing only with individual critical sections (weak or vulnerable
sections; roads through villages or settlements) on a road link (spot improvements), or providing a total
whole rural link design, which could comprise different design options along its length.

The choice of surfacing type, and when to use it, involves a trade-off between initial cost, level of service
and maintenance requirements. Cobblestone may use locally available resources and require very little
maintenance, but it gives a relatively rough riding surface. Surface dressings provide smoother riding
surfaces but may require more expensive earthworks and pavement layers, as well as imported bitumen,
specialised equipment and skilled operators. Appropriate selection will be driven to some extent by the
service level.

Surface dressings should not be constructed where there is no capacity for routine maintenance, including
pothole repairs and crack sealing, as well as periodic re-sealing. Edge break is a common problem on
sealed roads due to vehicles and pedestrians moving on and off the road, and needs to be controlled
through appropriate road width, provision of stopping places, and kerbing.

The challenge for the design engineer for a low volume road is to achieve the required level of service,
using appropriate engineering approaches and to minimise costs over the whole life of the road. This
should be done in a context sensitive way that recognises the needs of the client, the road environment
and the prevailing maintenance management regime.

1.6 Context sensitivity


In addition to ensuring that the design developed is technically appropriate and is within the financial
envelope, the design engineer needs to bear in mind other factors that could influence the success of the
low volume road design approach, its implementation and its long term sustainability.

This requires a broadly focused, multi-dimensional and context sensitive approach in which a number of
other influential factors are considered, illustrated in Figure A.1.3.

1.6.1 Political support

Demand for low volume road provision needs to be framed under a national policy driven by government
and should be supported at the highest level. The cross-sectoral influence of low volume road provision
and its role in under-pinning other sectoral development strategies and poverty alleviation programmes
should be highlighted, quantified and understood.

The approach adopted for low volume road provision should complement national plans, policies and
strategies and should be responsive to wider needs and demands, including:
ƒ The social and economic goals of poverty alleviation and development;
ƒ Increasing rural accessibility;
ƒ The use of appropriate technology, promotion of the domestic construction industry and
employment creation;
ƒ Protection of the environment;
ƒ Cost minimisation and improved efficiency.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 8

Mostly design
Mostly client
engineer
responsibility
responsibility
Politically supported

Environmentally
Socially acceptable
sustainable

Context sensitivity &


sustainability for low volume
road provision Institutionally
Financially sound
possible

Technologically
Economically viable
appropriate

Figure A.1.3: Framework for sustainable provision of low volume roads

There is a need to maintain dialogue with political and public stakeholders in order to highlight the
advantages of design approaches and alternative, often unfamiliar, solutions selected for low volume
road provision. The language used for advocacy should be carefully chosen and should avoid negative
connotations such as “low standard”; “low cost” and “marginal”.

1.6.2 Social acceptance

Provision of low volume rural road networks should be managed in a way that:
ƒ Ensures community participation in planning and decision making;
ƒ Eliminates gender bias and promotes participation by women in the road sector;
ƒ Promotes activities and investment for sustainable livelihoods (including Complementary
Interventions shown in Part C);
ƒ Promotes road safety in all aspects of low volume road provision.
ƒ Supports cost-effective labour-based and intermediate equipment methods of construction and
maintenance; and
ƒ Minimises resettlement and mitigates unavoidable resettlement through appropriate compensation.

1.6.3 Institutional capacity

Road authorities and clients should:


ƒ Promote institutional, economic and technical understanding in the provision and management of
low volume roads;
ƒ Promote commercial management practices;
ƒ Develop a conducive environment for the development of national contractors;
ƒ Ensure that design, construction and maintenance approaches for low volume roads are represented
on all tertiary civil engineering training curricula.

1.6.4 Technology choice

Technologies for designing, constructing and maintaining low volume roads should:
ƒ Employ appropriate design standards and specifications;

PART A: DESIGN STANDARD APPROACHES


A - Chapter 1 - 9

ƒ Utilise intermediate equipment technology options and reduce reliance on heavy equipment
imports;
ƒ Promote road construction and maintenance technologies that create employment opportunity;
ƒ Use types of contract that support the development of domestic contractors and consultants;
ƒ Be robust to the vagaries of climate and recognise potential impacts of a changing climate.

1.6.5 Economic viability

Economic appraisal for low volume roads should:


ƒ Employ tools for low volume roads that should be capable of quantifying social, economic and
environmental costs and benefits;
ƒ Ensure investment decisions for low volume roads are based on an assessment of whole life costs.

1.6.6 Financially sound

Sustainable provision of low volume roads depends on the sustainable provision of funding to the sector
in that:
ƒ Roads should not be upgraded to engineered standards if funding is not in place for routine and
periodic maintenance requirements.
ƒ Designs should not be forwarded that require excessive allocation of maintenance resources.

1.6.7 Protection of the environment

The design and management of low volume roads should:


ƒ Minimise the physical impacts of construction and maintenance activities on the natural environment;
ƒ Take account of socio-cultural impacts (community cohesion);
ƒ Minimise the carbon footprint;
ƒ Optimise resource management and allow for recycling of non-renewable materials; and
ƒ Minimise impacts and emissions that might contribute to climate change.

1.7 Protection of the environment


The design and management of low volume roads should:
ƒ Minimise the physical impacts of construction and maintenance activities on the natural environment;
ƒ Take account of socio-cultural impacts (community cohesion);
ƒ Minimise the carbon footprint;
ƒ Optimise resource management and allow for recycling of non-renewable materials; and
ƒ Minimise impacts and emissions that might contribute to climate change.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 2 - 10

2. STRUCTURE AND LAYOUT

The low volume road manual has been drafted with a view to it being fully adaptable for different clients.
The Manual has application for roads at a federal level and administered by the Ethiopian Roads Authority,
regional level for roads administered by the Regional and Rural Roads Authorities and district level for
roads administered by district (Wereda /Kebele) administrations or communities. The document can cater
for interventions that deal with individual critical areas on a road link (spot improvements) through to
providing total rural road link designs. In this latter case, this could comprise different design options
along the total length.

The Manual is divided into five parts:


ƒ Part A provides an overview of the Manual, its application, context, use and introduces the
philosophy of low volume road design.
ƒ Part B sets out typical design controls that should be considered during the design process and
the national design standards for low volume roads in Ethiopia. These are mandatory standards
that must be adhered to by the engineer. Departures from these standards are only permitted in
exceptional circumstances and with the prior approval of the relevant authorities.
ƒ Part C describes how complementary interventions and activities can be introduced into the road
works contract and how these can add value and impact to the project for the client and beneficiary
communities, and users.
ƒ Part D provides the engineering details and guidance on the application of the national standards
for the design of low volume roads given in Part B.
ƒ Part E provides the engineering details and guidance on the design of low level structures and
water crossings for low volume roads.

The structure of the Manual is shown in Figure A.2.1.

The low volume road manual is comprehensive, in that it deals with the design standards and technicalities
of all types of road (paved and unpaved), low level water crossing structures and socio-environmental
aspects. The Manual deals with design standards for road geometry, earthworks, drainage, pavement
structures and surfacing, with relevant approaches to route selection, site investigation, materials selection,
testing and treatment, and the design of low level water crossings and structures for low volume roads.

The structure of the Manual is such that it allows client authorities (Federal, Regional and Wereda) to
extract and use those parts of the Manual most appropriate to their individual set of circumstances. The
complete document can be downsized to capture specific and appropriate aspects.

In terms of “road type”, Wereda Administrations, Kebele committees, community organisations or


cooperatives would generally be concerned with the design of low volume unpaved, earth and gravel
roads. Regional Road Authorities and the ERA would be concerned with larger gravel roads and bituminous
sealed roads. Part D of the Manual captures all of the design elements, approaches and aspects for each
of these types of roads. The Manual deals with each of these road types and the national standards for
particular road types are set out and explained in Part B. Part B can be used by the engineer or others
who simply needs to look-up the values of the key parameters.

All authorities should find application for alternative surfacing technologies and spot improvement
approaches (Part D) and for utilisation of low level water crossings (Part E). All authorities should also be
interested to find application for complementary interventions and activities (Part C).

A clear understanding of the end users and their requirement for the Manual document is essential. The
low volume road manual caters for a range of road type, from basic earth tracks to bituminous sealed
roads. It is unlikely that one institution will cater for all standards of road and more likely that the document
will have application across a number of different authorities, agencies and ministries. Bituminised and
major gravel roads in Ethiopia are constructed and maintained by the Federal and Regional Roads
Authorities. The Districts (Wereda) and sub-districts (Kebele) have substantial networks of unpaved

PART A: DESIGN STANDARD APPROACHES


A - Chapter 2 - 11

gravel and earth roads. Substantial kilometres of unpaved road are also provided through community
programmes, cooperative ventures and by programmes operating through other line ministries, such as
the Ministry of Agriculture. The Manual contains technical explanations of all the steps in deriving the
standards for low volume roads related to specific environmental conditions. The Manual is structured
to provide the building blocks for authorities to compile their own tailored document that is applicable
to their prevailing circumstances.

Figure A.2.1: The structure of the Manual

Part A
Introduction to low
volume road design

Project implementation
Structure and layout
Context and scope

Part B
Design standards for
low volume roads
Geometric design standards

Road furniture and signage


associated structures
Policy and legislative

Water crossings and


Design parameters

Pavement design

erosion control
Drainage and
Introduction

Materials
controls

PART A: DESIGN STANDARD APPROACHES


A - Chapter 2 - 12

Context and application of complementary


interventions
Explanatory notes for low volume road
design general introduction
Planning, identification and implementation of
complementary interventions
Site investigation for
route selection and design
Employment and human resource issues

Roadside slope stabilisation


Contract provisions to support complementary
interventions

Geometric design
Part C

Part D
Supporting small scale contractors
Interventions
Complementary

volume road design


Drainage

Explanatory notes for low


Supervision consultant’s contract

Materials and pavement design

PART A: DESIGN STANDARD APPROACHES


Supervision consultant’s contract
Figure A.2.1: The structure of the Manual (continued)

Surfacing

Management, monitoring and enforcement


A - Chapter 2 - 13

Figure A.2.1: The structure of the Manual (continued)

Part E
Explanatory notes and
design standards for
small structures

Site selection and appraisal

Watercourse characteristics
Structural options

Structure design
Project planning

Design criteria

Maintenance
Construction
Materials

PART A: DESIGN STANDARD APPROACHES


A - Chapter 3 - 14

3. ROAD PROJECT IMPLEMENTATION

3.1 Contract of works


The key to successful execution of the low volume roads project will be to ensure clearly defined
requirements and adequate provisions are included in all bidding and contract documents. A clear
understanding of the requirements from both the works contractor and the supervising engineer is
needed.

Bidding documents need to contain all of the information and provisions for the interested companies
or organisations on all that is relevant to obtain the contract. The bidding document informs interested
bidders on all of the procedures to be followed, documents to be submitted, the bid evaluation procedure
and the award of contract.

Bidding documents need to provide the bidder with all essential information for successful completion
of the project. The approach used for execution of low volume roads can differ in many respects from
the traditional road provision approaches. For example, the client may favour labour-based approaches;
the use of intermediate equipment; sub-contracting to empower small enterprises; and/or additional
enhancements through complementary interventions (Part C). It is therefore important that the provisions
within the bidding documents clearly reflect these preferences and adjust the provisions of the contract
accordingly.

Failure to properly differentiate low volume road approaches can lead to complications during the
bidding procedure or execution of the contract. Moreover, clear and well prepared bidding documents
are essential to ensure that sufficient companies or organisations are confident to bid for works.

In order to facilitate preparation of documents a series of model bidding documents have been prepared
for use with works of differing complexity. These model documents which will vary in complexity depending
on the maximum contract value include:

ƒ Standard Bidding Documents For Major Work, International Competitive Bidding (ICB) and
National Competitive Bidding (NCB) versions – Unlimited Contract Value;
ƒ Standard Bidding Document for Intermediate Works (based on PPA 2006) - – both ICB and NCB
versions - Maximum contract value USD10 Million;
ƒ Standard Bidding Document for the Procurement of Minor Works – Maximum contract value USD3
Million;
ƒ Standard Bidding Document for the Procurement of Micro Works – Maximum contract value
USD300,000.

For each of the above documents a User Guide has been developed to provide guidance to the Client or
Employer organisation and to the design consultant who may be employed in preparing the documents.

In nearly all cases local authorities would be concerned with contract values below USD10M, whilst ERA
and Regional or Rural Road Authorities could utilise all available bidding documents.

The main issues for a bidder/contractor is to fully understand the scope of all the works, including
any complementary interventions, and the fundamental issues of measurement and payment. For the
preparation of Works Bidding Documents the key documents requiring attention are:

Instructions to Bidders (ITB) and the Bid Data Sheet (BDS): The ITB is generally a standard document,
slightly varying for the different clients. For a LVR project the client should include an additional item
that will draw the attention of the bidder to the low volume road approach and any requirements for
complementary interventions. The BDS is linked to the ITB and provides specific project information.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 3 - 15

Standard Technical Specifications: These will define the scope of the technical requirements of the
contract, including the type and quality of materials and equipment, the standards of workmanship. The
Standard Technical Specifications form a separate volume in the ERA 2011 series and these capture most
of the proposed interventions mentioned in this low volume road manual. The ERA Standard Technical
Specifications includes information on the format of Bill Items for the Bill of Quantities, on item coverage
and the method of payment.

Particular Specifications: This is where any detailed technical requirements and specifications, and
implementation mechanisms specific for the project should be clearly defined. The particular technical
specification should also include any specifications and limitations on the freedom of choice for the
contracting company related to the execution of works. Particular technical specifications add further
detail to complement or replace those stated in the Standard Technical Specifications.

Bills of Quantities or Schedules of Rates: This should be linked by item number to the Standard
Technical Specifications and to the Particular Specifications; and is where the schedule of activities and
estimated quantities are set out for the bidder to price.

Drawings: Some standard detailed drawings may be applied directly for low volume roads works (eg
cross-sections, standard culvert design and signage). Supplementary drawings, linked with the Particular
Specifications, may also be required where new, innovative or special approaches are included. Where
trail bridges are included within the contract, as a complementary intervention, a separate volume of
drawings is provided.

Conditions of Contract: This includes standard provisions for execution of the contract and unless
amended in the Conditions of Particular Application, these will apply. It is anticipated for Major and
Intermediate works projects that amendments may be required to reflect the desired approach for low
volume road works. The simplified Conditions of contract with the standard Bidding Documents for
Minor and Micro works should not need amending.

Conditions of Particular Application: This is where any Provisions in the General Conditions of Contract
may be amended, as required, to make them more appropriate to the requirements of the low volume
road approach, including complementary interventions.

In promoting small and medium scale enterprises; emergent contractors; employment intensive or
labour-based works; and utilisation of intermediate equipment options, due consideration should be
given to the clauses referring to Performance Security, Performance Program, Insurances, Cash Flow,
Plant, Equipment & Workmanship, Payments, Retention and Advances, Price Adjustment and currency
restrictions. There will be cases where the general conditions of contract may prevent, or work against,
the small scale industry and these should be adjusted accordingly to promote competition and fairness
to the emergent industry.

Due consideration should also be given to strengthening clauses aimed at promoting sub-contracting/
assignment; local employment and conditions (particularly for women); rights and insurances; and for
strengthening complementary interventions.

User Guide’s: These documents will provide detailed guidelines to the design consultant on the
preparation of the Works Bidding documents.

Works Contract Evaluation: Recognition that the context of the works is using low volume road
approaches must be included within the evaluation of bid process. The aim should be to ensure that
the bidder confirms an understanding of the Client’s perception for implementation. If complementary
intervention requirements are built into the prospective works, the evaluation should also ensure that the
complementary interventions are fully understood.

PART A: DESIGN STANDARD APPROACHES


A - Chapter 3 - 16

3.2 Supervision services


In addition to adequately defining the scope and understanding of the project and its approach for the
works bid, it is essential that the same level of understanding is reflected in the Request for Proposals
(RFP) for the supervising consultant. The RFP should specifically include appropriate inputs of key
personnel with the requisite skills to meet the requirements of the low volume roads project approach.
The supervising consultant should be fully familiar with the techniques and approaches to be employed
on the works, including any complementary interventions.

The RFP should include:


ƒ Clear definition of the role of the supervisor in the context of the project. If small scale or emergent
contractors are employed, the client may require the consultant to act both as supervisor and
mentor or to provide training, for example to client or authority staff;
ƒ Clearly defined and appropriate inputs for key personnel with requisite experience on low volume
road implementation;
ƒ Requisite skills to cater for socio-environmental supervision and oversight of any complementary
interventions;
ƒ Reference to this Manual and supporting documents.

PART A: DESIGN STANDARD APPROACHES


)('(5$/'(02&5$7,&5(38%/,&2)
('(5$/'(02&5$7,&5(38%/,&2)

(7+,23,$152$'6$87+25,7<

,7 <
( 7+

25
,2

,$
15 87
+
3

2$'6 $

'(6,*10$18$/)25/2:92/80(52$'6
3$57$3$57%$1'3$57&
),1$/'5$)7$35,/
Part B
DESIGN STANDARDS FOR LOW VOLUME ROADS

Introduction

Policy and legislative


controls

Design parameters

Geometric design standards


Part B
Design standards for
low volume roads
Materials

Pavement design

Drainage and
erosion control

Water crossings and


associated structures

Road furniture and signage


B - iii

B TABLE OF CONTENTS
B. TABLE OF CONTENTS ........................................................................................................B.I
B. LIST OF TABLES ................................................................................................................ B.III
B. LIST OF FIGURES .............................................................................................................. B.V
1. INTRODUCTION ................................................................................................................B.1
2. POLICY AND LEGISLATIVE CONTROLS ............................................................................B.2
2.1 Legal Framework ............................................................................................................B.2
2.2 Road Safety .....................................................................................................................B.3
3. DESIGN PARAMETERS ......................................................................................................B.4
3.1 Climate ............................................................................................................................B.4
3.2 Terrain ..............................................................................................................................B.5
3.3 Demographics .................................................................................................................B.5
3.4 Traffic ...............................................................................................................................B.5
3.4.1 Vehicle classification .........................................................................................B.5
3.4.2 Traffic volumes ..................................................................................................B.6
3.4.3 Traffic growth ....................................................................................................B.6
3.4.4 Geometric design .............................................................................................B.6
3.4.5 Structural design ...............................................................................................B.6
3.4.6 Equivalent standard axles per vehicle class ......................................................B.7
4. GEOMETRIC DESIGN STANDARDS ...................................................................................B.9
4.1 Traffic composition .......................................................................................................B.10
4.2 Roadside Population and non-motorised vehicles ........................................................B.10
4.3 Geometric Design Standards for LVRs...........................................................................B.11
4.4 Design-by-eye ...............................................................................................................B.17
4.5 Typical Cross Sections ..................................................................................................B.17
5. MATERIALS ......................................................................................................................B.28
5.1 Subgrades ....................................................................................................................B.28
5.1.1 Specifying the design subgrade class ............................................................B.28
5.1.2 Material depth ...............................................................................................B.29
5.1.3 Improved subgrade layers ..............................................................................B.30
5.1.4 Dealing with poor subgrade soils ...................................................................B.30
5.2 Pavement Materials .......................................................................................................B.30
5.2.1 Materials requirements for roadbase ..............................................................B.32
5.2.2 Material requirements for sub-base ................................................................B.33
5.2.3 Material requirements for gravel wearing course ...........................................B.35
5.2.4 Material Improvement ....................................................................................B.36
6. PAVEMENT DESIGN ........................................................................................................B.39
6.1 Design traffic classes ....................................................................................................B.39
6.2 Engineered natural surfaces ..........................................................................................B.39
6.3 Natural gravel roads .....................................................................................................B.40
6.3.1 Major gravel roads .........................................................................................B.40
6.3.2 Minor gravel roads ..........................................................................................B.42
6.4 Surfacing options and design standards for paved roads ............................................B.43
6.4.1 Bituminous surfaced roads ............................................................................B.43
6.4.2 Non bituminous surfaced roads .....................................................................B.44
7. DRAINAGE AND EROSION CONTROL ...........................................................................B.51
7.1 Size of watercourse .......................................................................................................B.51

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - iv

7.2 The Rational Method .....................................................................................................B.51


7.2.1 Catchment runoff coefficient, C......................................................................B.51
7.2.2 Rainfall intensity, I (mm/hour)..........................................................................B.52
7.2.3 Catchment area, A (km2) .................................................................................B.53
7.3 The SCS method ...........................................................................................................B.54
7.3.1 Catchment area ..............................................................................................B.54
7.3.2 Rainfall ............................................................................................................B.54
7.3.3 Runoff and Curve Numbers ............................................................................B.55
7.3.4 Time of concentration.....................................................................................B.57
7.3.5 Steps in the SCS procedure ............................................................................B.59
7.4 Design of culverts ..........................................................................................................B.61
7.4.1 Nomograph method for culvert sizing............................................................B.61
7.4.2 Correlation with successful practice ...............................................................B.61
7.4.3 Design of drifts and fords ...............................................................................B.61
7.5 Components of External Drainage ................................................................................B.62
7.5.1 General principles..........................................................................................B.62
7.5.2 Side drains .....................................................................................................B.62
7.5.3 Erosion control in the side drain .....................................................................B.66
7.5.4 Mitre drains or turnouts ..................................................................................B.68
7.5.5 Wet lands ........................................................................................................B.69
7.5.6 Subsoil Drains .................................................................................................B.69
7.5.7 Filters ..............................................................................................................B.69
7.5.8 Interceptor, cut-off or catch-water drains. .....................................................B.69
7.5.9 Chutes.............................................................................................................B.69
7.5.10 Slope protection .............................................................................................B.70
8. WATER CROSSINGS AND ASSOCIATED STRUCTURES ..................................................B.71
9. ROAD FURNITURE AND SIGNAGE .................................................................................B.72
9.1 Traffic Signs ...................................................................................................................B.72
9.2 Road Markings...............................................................................................................B.72
9.3 Marker Posts ..................................................................................................................B.72
9.4 Safety barriers................................................................................................................B.72
APPENDIX B.1 - RAINFALL INTENSITY-DURATION-FREQUENCY CHARTS .....................B.74

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B-v

B LIST OF TABLES

Table B.2.1: Summary of Existing Requirements ...............................................................................B.3


Table B.3.1: Vehicle Classification ......................................................................................................B.5
Table B.3.2: Average equivalency factors for different vehicle types ...............................................B.7
Table B.3.3: Factors for design traffic loading ...................................................................................B.8
Table B.4.1: PCU values ..................................................................................................................B.10
Table B.4.2: Increased ‘shoulder’ widths (each side) for unpaved LVRs ...........................................B.11
Table B.4.3: Shoulder widths (each side) for paved LVRs .................................................................B.11
Table B.4.4: Geometric design standards for paved DC4(1) (AADT 150-300) .................................B.12
Table B.4.5: Geometric design standards for unpaved DC4(1) (AADT 150-300) .............................B.13
Table B.4.6: Geometric design standards for paved DC3(1) (AADT 75-150) ...................................B.13
Table B.4.7: Geometric design standards for unpaved DC3(1) (AADT 75-150) ...............................B.14
Table B.4.8: Geometric design standards for DC2 paved(1) (AADT 25-75) .....................................B.14
Table B.4.9: Geometric design standards for DC2 (1, 2) unpaved (AADT 25-75) ............................B.15
Table B.4.10: Geometric design standards for DC1 (AADT 1-25)......................................................B.15
Table B.4.11: Minimum standards for basic access ............................................................................B.16
Table B.4.12: Adverse cross-fall to be removed if radii are less than shown .....................................B.16
Table B.4.13: Super-elevation development lengths .........................................................................B.16
Table B.4.14: Widening recommendations (m) ..................................................................................B.16
Table B.4.15: Slope dimensions for cross-sections (ratios are vertical:horizontal) .............................B.17
Table B.5.1: Subgrade classes .........................................................................................................B.28
Table B.5.2: Dependence of design subgrade on design traffic class .............................................B.29
Table B.5.3: Material depth by road category .................................................................................B.29
Table B.5.5: Particle size distribution for natural gravel base ..........................................................B.32
Table B.5.6: Plasticity requirements for natural gravel road base materials.....................................B.33
Table B.5.7: Guidelines for the selection of lateritic gravel road base materials .............................B.34
Table B.5.8: Typical particle size distribution for sub-bases .............................................................B.35
Table B.5.9: Plasticity characteristics for granular sub-bases ...........................................................B.35
Table B.5.10: Typical standardised gravel loss ...................................................................................B.37
Table B.5.11: Recommended material specifications(1,3) for unsealed rural roads ..........................B.37
Table B.5.12: Recommended material specifications for unsealed ‘urban’ roads .............................B.38
Table B.6.1: Traffic classes for flexible pavement design .................................................................B.39
Table B.6.2: Required minimum height (hmin) between road crown and invert level of drain in
relation to climate ........................................................................................................B.40
Table B.6.3 (a): Gravel base thickness for major gravel roads – strong gravel (G45) ............................B.41
Table B.6.3 (b): Gravel base thickness for major gravel roads – medium gravel (G30) .........................B.41
Table B.6.3 (c): Gravel base thickness for major gravel roads – weak gravel (G15) ..............................B.41

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - vi

Table B.6.4: Typical gravel loss ........................................................................................................B.42


Table B.6.5: Design Chart for minor gravel roads ............................................................................B.43
Table B.6.6: Bituminous Pavement Design Chart 1 ........................................................................B.44
Table B.6.7: Bituminous Pavement Design Chart 2 ........................................................................B.44
Table B.6.8: Non-bituminous pavement surfacing options ..............................................................B.45
Table B.6.9: Substitution of pavement layer material ......................................................................B.45
Table B.6.10: Thickness designs for WBM pavements.......................................................................B.46
Table B.6.11: Thicknesses designs for Hand Packed Stone (HPS) pavement (mm) ...........................B.47
Table B.6.12: Thicknesses designs for various discrete element surfacings (mm) .............................B.48
Table B.6.13: Thicknesses (mm) - Non-Reinforced Concrete Pavement (NRC) ..................................B.49
Table B.6.14: Ultra-Thin Reinforced Concrete Pavement (UTRCP) Design .......................................B.50
Table B.7.1: Runoff coefficient: Humid catchment ...........................................................................B.51
Table B.7.2: Runoff coefficient: semi-arid catchment .......................................................................B.52
Table B.7.3: Storm design return period (years) ..............................................................................B.53
Table B.7.4: Storm design return period (years) for severe risk situations .......................................B.53
Table B.7.5: Hydrological characteristic soil groups ........................................................................B.55
Table B.7.6: Runoff Curve Numbers (CN) ........................................................................................B.56
Table B.7.7: Conversion from average to wet and dry antecedent moisture conditions .................B.57
Table B.7.8: Antecedent moisture conditions ..................................................................................B.57
Table B.7.9: Manning’s roughness coefficients for sheet flow..........................................................B.58
Table B.7.10: Roughness coefficient (n) for drains .............................................................................B.66
Table B.7.11: Permissible flow velocities (m/sec) in excavated ditch drains ......................................B.67
Table B.7.12: Spacing between scour checks ....................................................................................B.68
Table B.7.13: Maximum spacing of mitre drains ................................................................................B.68
Table B.7.14: Common cut-slope ratios for LVRs. ..............................................................................B.70
Table B.7.15: Common fill slope batters for LVRs. .............................................................................B.70
Table B.9.1: Spacing of Guide Posts at Curves ................................................................................B.73

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - vii

B LIST OF FIGURES

Figure B.3.1: Climatic N-value map for Ethiopia .................................................................................B.4


Figure B.4.1: Selection of appropriate geometric design standards ...................................................B.9
Figure B.4.2: Typical cross section, DC1 – 4, Flat Terrain, Unpaved..................................................B.17
Figure B.4.3: Typical cross section, DC1 – 4, Flat Terrain, Paved ......................................................B.18
Figure B.4.4: Typical cross section, DC1 – 4, Rolling Terrain, Unpaved ............................................B.19
Figure B.4.5: Typical cross section, DC1 – 4, Rolling Terrain, Paved .................................................B.20
Figure B.4.6: Typical cross section, DC1 – 4, Mountainous Terrain, Unpaved...................................B.21
Figure B.4.7: Typical cross section, DC1 – 4, Mountainous Terrain, Paved .......................................B.22
Figure B.4.8: Typical cross section: DC1 – 4, Escarpment Terrain, Unpaved.....................................B.23
Figure B.4.9: Typical cross section: DC1 – 4, Escarpment Terrain, Paved .........................................B.24
Figure B.4.10: Typical cross section, DC1 – 4, Populated areas, Unpaved .........................................B.25
Figure B.4.13: Typical cross section, DC1 – 4, Flat Terrain, Expansive soils, Paved ............................B.26
Figure B.5.1: Illustration of CBR strength cumulative distribution.....................................................B.28
Figure B.5.2: Figure B.5.2: Material quality zones .............................................................................B.35
Figure B.6.1: Cross-section details ENS ............................................................................................B.38
Figure B.6.2: Typical gravel road cross section in flat terrain. ...........................................................B.39
Figure B.7.1: Velocity of flow .............................................................................................................B.51
Figure B.7.2: Relationship between Precipitation, Direct Runoff and CN .........................................B.54
Figure B.7.3: B.24 hour depth-frequency curves ...............................................................................B.59
Figure B.7.4: Unit peak discharge (Type II rainfall) ............................................................................B.60
Figure B.7.5: Manning’s Formula .......................................................................................................B.61
Figure B.7.6: Headwater depth and capacity for corrugated metal pipe culverts with inlet control
(Adapted from FHWA, 1998) ......................................................................................B.62
Figure B.7.7: Headwater depth and capacity for concrete pipe culverts with inlet control
(Adapted from FHWA, 1998) ......................................................................................B.63
Figure B.7.8: Headwater depth and capacity for concrete box culverts with inlet control
(Adapted from FHWA, 1998) ......................................................................................B.64

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 1 - 1

1. INTRODUCTION

Low Volume Roads are defined as those roads carrying:


ƒ Up to about 300 vehicles per day; and
ƒ Less than about 1 million equivalent standard axles.

For geometric designs, roads carrying an excess of 300 vpd should be designed in accordance with
the Geometric Design Manual-2011. For roads carrying in excess of 300 vpd, but with a total traffic
loading of less than 1 Mesa, the structural pavement design should be carried out in accordance with the
standards in this document.

For structural pavement design, roads carrying in excess of 1 Mesa should be designed in accordance
with the 2011 Pavement Design Manual. For roads carrying in excess of 1 Mesa, but with a traffic volume
of less than 300 vpd, the geometric design should be carried out in accordance with the standards in this
document.

Low Volume Roads fall under the responsibility of several authorities including community/cooperative
structures, kebele and wereda administrations, and the regional and federal road authorities. This
manual provides the requirements for the design of low volume roads under the responsibility of these
authorities. The standards provide an appropriate level of service for each class of road.

The custodian of design standards for all roads, including the associated specifications and standard
drawings, is the Ethiopian Roads Authority (ERA).

There are four classes of LVR known as DC1 to DC4 based on traffic levels, with each class being defined
by appropriate geometric design standards (Part A, Figure A.1.1). Once the geometric standards are
fixed, the design approach for LVRs requires the selection of a surfacing technology, pavement design,
and drainage appropriate to the road environment.

The Environmentally Optimised Design approach (outlined in Part A) allows different solutions to be
adopted along the road. The manual also provides comprehensive guidance on the design of water
crossings and retaining structures that provide a level of service commensurate with the standard of the
road.

ERA’s General Technical Specifications contain the detailed engineering requirements supporting
the design. These may be modified and added to in exceptional circumstances depending on the
requirements of the specific project or road environment.
The design options for low volume roads, drainage and retaining structures, assume that adequate
maintenance is carried out on the road.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 2 - 2

2. POLICY AND LEGISLATIVE CONTROLS

Government policy, national legislation and development planning dictate the underlying principles
of low volume road design. This includes, for example, environmental controls, road safety legislation,
promotion of the use of labour or application of intermediate equipment based technologies to encourage
local participation and SME development. Authorities may choose to put emphasis on Complementary
Interventions, as set out in Part C.

2.1 Legal Framework


Environmental and Social Impact Assessments promote maintenance of the road corridor environment in
at least the same condition as it was before the road construction project started. Engineering designs
must make provision for protective and mitigation measures. Key documents that must be referred to
in the design of low volume road projects to assess and address environmental and social safeguards
during project planning, design and construction are.
ƒ Legal Framework - in particular Articles 35, 40, 41, 43, 44, 91, 92 – Constitution of the Federal
Democratic Republic of Ethiopia;
ƒ Proclamation No. 299/2002 – Environmental Impact Assessment;
ƒ Proclamation No. 300/2002 – Environmental Pollution Control;
ƒ Proclamation No. 295/2002 – Environmental Protection Organs;
ƒ Proclamation No. 209/2002 – Cultural Heritage;
ƒ Proclamation No 455/2005 – Expropriation of land holdings for public purposes and payment of
compensation;
ƒ Proclamation No 135/2007 – Council of Ministers regulation on the payment of compensation for
property situated on land holdings expropriated for public purposes;
ƒ Proclamation No 456/2005 – Rural land administration and land use;
ƒ Proclamation No. 94/1994 – Conservation of Forests;
ƒ Proclamation No. 192/1980 and 416/1972 – Wildlife Conservation and Conservation Areas;
ƒ Proclamation No. 197/2000 – Water Resources;
ƒ Proclamation No. 200/2000 – Healthy Environment;
ƒ Development plans of the Federal Government, Regional Governments and Weredas;
ƒ Regional States’ environmental legislation;
ƒ Environmental Protection Authority (EPA) requirements for preparation of EIAs and EMPs;
ƒ Funding agency policies, regulations and guidance notes.

The requirements of existing environmental legislation and related government proclamations are
summarised in Table B.2.1.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 2 - 3

Table B.2.1 Summary of Existing Requirements

No. Technical Area Existing Documentation Requirements


1 Environmental Proclamation on EIAs, EPA Full ESIA on all major urban roads, all
& Social Impact guidance note, GTP rural road programmes, upgrading
Assessment (ESIA) or rehabilitation of major rural roads.
Preliminary ESIA on other rural road
works.
2 Land Acquisition Proclamations on Compensation to be made in
(LA) and Land Acquisition and accordance with relevant proclamations
Compensation Compensation, ERA QMS and regulations.
3 Resettlement Action Funder policies and Only if project is funded by international
Plans (RAP) guidelines; ERA guideline on agency and more than 200 people are to
Resettlement Rehabilitation be relocated
Policy Framework; and
FDRE proclamation on
land expropriation and
compensation payment.
4 Environmental & EPA guidance note Full ESMP required for all road projects.
Social Management
Plans (ESMP)

2.2 Road Safety


For the prevention of accidents, adherence to the following points and principles are of particular
importance for the design of safer roads. These principles underpin the geometric design standards
contained in this manual.
ƒ Design for all road users.
• Consider pedestrians and intermediate means of transport in road design, including
carriageway width, shoulder design, side slopes and side drains, and road surface drainage;
• Provide for traffic calming, improved surfacing and segregation of vulnerable road users in
populous areas.
ƒ Provide a clear and consistent message to the driver.
• Roads should be easily “read” and understood by drivers and should not present them with
any sudden surprises;
• Ensure adequate sight distances and harmonisation of horizontal and vertical alignments.
ƒ Encourage appropriate speeds and driver behaviour.
• Ensure prominent and proper location of signage, road markings, traffic calming and other
traffic control devices.
ƒ Reduce conflicts.
• Reduce conflicts through appropriate design, eg by staggering junctions or by using guard
rails to channel pedestrians to safer crossing points.
ƒ Create a forgiving road environment.
• Allow for driver mistakes or vehicle failure, to the extent that this is possible without
significantly increasing costs;
• Ensure that demands are not placed upon the driver which are beyond his or her ability to
manage;
• Avoid situations where drivers must make more than one decision at a time.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 3 - 4

3. DESIGN PARAMETERS

The principal factors affecting the design of low volume roads are climate, terrain, demographics and
traffic.

3.1 Climate
The climatic descriptor which is used for the pavement design catalogues is the Weinert ‘N’ value
(Weinert, 1974). This index is calculated as follows:

N = 12.Ej/Pa Equation 3.1

where:
Ej = evaporation for the warmest month
Pa = total annual precipitation

N-values less than 4 apply to a climate that is seasonally tropical and wet (the Kolla, Woina Dega, Dega
and Wurch regions of Ethiopia), whereas N-values greater than 4 apply to a climate that is arid, semiarid
or dry (the Bereha region of Ethiopia). A map of equivalent N-values for Ethiopia is shown in Figure B.3.1
and provides the means of placing a road in the appropriate climatic zone for design purposes.

Figure B.3.1: Climatic N-value map for Ethiopia

The climatic zones demarcated by the N-values are macro-climates and it should be kept in mind that
different micro-climates may occur within these regions. This is particularly important where such local
micro-climates can play a significant role in determining the in-situ moisture content of the various
pavement layers; a factor which needs to be considered in the choice of N-Value or the subgrade class
used for design purposes.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 3 - 5

3.2 Terrain
Terrain class is determined by the number of 5-metre contours crossed by a straight line connecting the
two ends of the road section in question according to the following definitions:

0 to 10 five-metre contours per km. The natural ground slopes perpendicular to


Flat
the ground contours are generally below 3%.
11 to 25 five-metre contours per km. The natural ground slopes perpendicular to
Rolling
the ground contours are generally between 3 and 25%.
26 to 50 five-metre contours per km. The natural ground slopes perpendicular to
Mountainous
the ground contours are generally above 25%.
Escarpments are geological features that require special geometric standards
Escarpment because of the engineering risks involved. Typical gradients are greater than
those encountered in mountainous terrain.

It should be noted that it is not dependent on the alignment chosen for the road.

3.3 Demographics
Appropriate design approaches must be introduced in populous areas to mitigate the effects of dust and
improve the safety of road users (additional road widths, parking, bus lay-bys) and appropriate drainage
systems.

3.4 Traffic
The use of traffic data varies depending on whether it is being used for geometric design or pavement
structural design. Traffic growth needs to be taken into account in the design process.

3.4.1 Vehicle classification

Table B.3.1 shows the vehicle classification.

Table B.3.1: Vehicle Classification

Class Type Axles Description


1 Car 2 Passenger cars and taxis
2 Pick-up/4-wheel drive 2 Pick-up, minibus, Land Rovers, Land Cruisers
3 Small bus 2 ≤ 27 seats
4 Bus/coach 2 > 27 seats
5 Small truck 2 ≤ 3.5 tonnes
6 Medium truck 2 or 3 3.5 – 7.5 tonnes
7 Large 2-axled truck 2 > 7.5 tonnes
8 3-axled truck 3 >7.5 tonnes
9 4-axled truck 4 *
10 5-axled truck 5 *
11 6-axled truck 6 *
12 3-axled trailer 3 *
13 4-axled trailer 4 *
* Not needed for definition

For geometric design purposes it is also necessary to count non-motorised and intermediate means of
transport including pedestrians, bicycles, animal transport, motorcycles, tractors and trailers.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 3 - 6

3.4.2 Traffic volumes

The (Annual) Average Daily Traffic (AADT) is defined as the total annual traffic summed for both directions
and divided by 365. It is usually obtained by recording actual traffic volumes over a much shorter period
from which the AADT is then estimated.

Where there is no existing road of any sort, the existing pedestrian traffic can be used to estimate the
likely vehicular traffic after the road is constructed. Alternatively, traffic information might be available
from an economic evaluation carried out to justify the road in the first place. In the unlikely event that
there is no information available, the lowest class of engineered road (DC1) should be provided.

3.4.3 Traffic growth

Future traffic falls into the following three categories:


ƒ Normal traffic. Traffic which would pass along the existing road or track even if no new pavement
were provided.
ƒ Generated traffic. Additional traffic which occurs in response to the provision or improvement of
the road.
ƒ Diverted traffic. Traffic that changes from another route (or mode of transport) to the project road
because of the improved pavement, but still travels between the same origin and destination.

The AADT in both directions in the first year of analysis consists of the current traffic plus an estimate of
the diverted traffic. If the total traffic is denoted by AADT0 and the general growth rate is i per cent per
annum, then the traffic in any subsequent year, x, is given by the following equation:

AADTx = AADT0 (1+i/100)x Equation B.3.1

3.4.4 Geometric design

Four different basic geometric standards (DC1-DC4) are defined for LVRs based on the number of
4-wheeled (and more) vehicles defined in Table B.3.1. The traffic level is the sum for both directions
and is estimated at the middle of the design life period. A design life of 10 years is recommended for
unpaved roads and 15 years for paved roads hence Equation B.3.1 is used to calculate the traffic after 5
or 7 years respectively. Where the expected traffic is near to a traffic boundary, the higher classification
should be adopted.

Geometric design also requires the traffic level of pedestrians, non-motorized and intermediate forms of
traffic and this is calculated in the same way using Equation B.3.1.

3.4.5 Structural design

For structural pavement design the cumulative traffic loading of each of the motorised vehicle classes
over the design life of the road in one direction is required. For a given class, m, this is given by the
following equation:

T(m) = 0.5 x 365 x AADT(m)0 [(1+i/100)N – 1]/(i/100) Equation B.3.2

Where
T(m) = the cumulative traffic of traffic class m
AADT(m)0 = The AADT of traffic class m in the first year
N = the design period in years
i = the annual growth rate of traffic in per cent

The cumulative traffic for each class of vehicle is multiplied by the average number of equivalent standard
axles of vehicles in that class to calculate the cumulative total number of equivalent standard axles over
the life of the road.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 3 - 7

3.4.6 Equivalent standard axles per vehicle class

The number of equivalent standard axles (ef) of an axle is related to the axle load as follows:
ef = (P/8160)n (for loads in kg) Equation B.3.3
or ef = (P/80)n (for loads in kN) Equation B.3.4
Where:
ef = number of equivalent standard axles (esas)
P = axle load (in kg or kN)
n = damage exponent (n = 4 for LVRs).

The sum of the individual ef values for each axle of the vehicle gives the equivalence factor for the
vehicle as a whole, EF(m). Guidance on the likely average EF(m) for different vehicle classes derived from
historical data is given in Table B.3.2. However, data from any recent axle load survey on the road in
question or a similar road in the vicinity is better than using countrywide averages.

The cumulative esas over the design period for each vehicle class is obtained by multiplying EF(m) by
the cumulative traffic, T(m). The total number of cumulative standard axles for all vehicle classes is then
obtained by adding together the values of EF(m) x T(m) for all the classes.

In some cases there will be distinct differences in each direction and separate vehicle damage factors for
each direction should be derived. The higher of the two directional values should be used for design.

Table B.3.2: Average equivalency factors for different vehicle types

No of Average esa per vehicle Average esa per vehicle


Class Type
axles - all loaded - half loaded(1)
1 Car 2 - -
2 4-wheel drive 2 - -
3 Minibus 2 0.3 0.15
4 Bus/coach 2 2.0 1.0
5 Small truck/PU 2 1.5 0.7
6 Medium truck 2 5 2.5
7 Large 2-axled truck 2 10 5
8 3-axled truck 3 12 3.5
9 4-axled truck 4 15 7.5
10 5-axled truck 5 17 8.5
11 6-axled truck 6 17 8.5
12 2-axled trailer 2 10 5
13 4-axled trailer 4 12 6
Note:
1. It is common to find that vehicles have no back load hence half the vehicles are likely to be empty, or nearly so.

On narrow roads the traffic tends to be more channelised than on wider two lane roads. In such cases
the effective traffic loading is greater than that for a wider road and the design traffic loading (esas) is
calculated using the relationships given in Table B.3.3.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 3 - 8

Table B.3.3: Factors for design traffic loading

Cross Paved Corrected design


Explanatory notes
Section width traffic loading (esa)
The driving pattern on
Double the sum of
< 3.5m this cross-section is very
esas in both directions
channelized.
Min. 3.5m but less than The sum of esas in Traffic in both directions
4.5.m both directions uses the same lane
Single
carriageway To allow for overlap in
80% of the esas in
Min. 4.5m but less than 6 m the centre section of
both directions
the road
Total esas in the Minimal traffic overlap
6m or wider heaviest loaded in the centre section of
direction the road.
90% of the total The majority of vehicles
More than one lane
esas in the studied use one lane in each
in each direction
direction direction.

Construction traffic can also be a significant proportion of total traffic on LVRs (sometimes 20 – 40 % of
total traffic) and should be taken into account in the design of the pavement.

For very low volume roads (traffic <25 vpd), a detailed traffic analysis is seldom warranted because
environmental rather than traffic loading factors generally determine the performance of roads.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 9

4. GEOMETRIC DESIGN STANDARDS

The flow diagram in Figure B.4.1 shows the process for the geometric design of low volume roads. This
is followed by Tables of key data. Further details of the geometric design process is provided in Chapter
4 of Part D.

Step 1 Step 1
Determine AADT of motorised Determine AADT of heavy
traffic trucks (3-axles or more)

Select road class

Step 2 Step 4
Step 3
Determine daily PCUs of non Determine nature of roadside
Determine terrain class
-motorised traffic population

Step 5
Select Road Type or Types

Step 6
Select widths of
carriageway & shoulders

Determine a trial alignment


using the parameters
selected.

Figure B.4.1: Selection of appropriate geometric design standards

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 10

4.1 Traffic composition


The proportion of heavy vehicles in the traffic stream on LVRs is often quite high. The Geometric Design
standards for DC2, DC3 and DC4 include a modification to cater for this.

In order to quantify traffic for normal capacity design the concept of equivalent PCUs is used. The PCU
values are shown Table B.4.1.

Table B.4.1: PCU values

Vehicle PCU value


Pedestrian 0.15
Bicycle 0.2
Motor cycle 0.25
Bicycle with trailer 0.35
Motor cycle taxi (bajaj) 0.4
Motor cycle with trailer 0.45
Small animal-drawn cart 0.7
Bullock cart 2.0
All based on a passenger car = 1.0

4.2 Roadside population and non-motorised vehicles


If the road is passing through a Wereda seat or a larger populated area, an extra carriageway of 3.5m
width is provided in each direction for parking and for passenger pick-up and a 2.5m pedestrian footpath
is also specified. The latter is essentially the road shoulder (Tables B.4.2 and B.4.3). In addition, the main
running surface is paved and is 7.0m wide.

When passing through a Kebele seat a 2.5m paved shoulder is specified but no additional footpath,
though one could be provided if required. The carriageway is also increased to 7.0m and therefore the
standard is very similar to DC4 but with wider shoulders.

These standards are not justified for the lower traffic levels of DC2, which is a single carriageway, unless
the road is passing through a particularly well populated area that is not classified as a Kebele or Wereda
seat but where additional traffic may be expected. In such circumstances the shoulders should be widened
to 2.5 metres for the extent of the populated area.

If the road is passing through a Wereda seat or a larger populated area, an extra carriageway of 3.5m
width is provided in each direction for parking and for passenger pick-up and a 2.5m pedestrian footpath
is also specified. The latter is essentially the road shoulder (Tables B.4.2 and B4.3). In addition, the main
running surface is paved and is 7.0m wide.

When passing through a Kebele seat a 2.5m paved shoulder is specified but no additional footpath,
though one could be provided if required. The carriageway is also increased to 7.0m and therefore the
standard is very similar to DC4 but with wider shoulders.

These standards are not justified for the lower traffic levels of DC2, which is a single carriageway, unless
the road is passing through a particularly well populated area that is not classified as a Kebele or Wereda
seat but where additional traffic may be expected. In such circumstances the shoulders should be widened
to 2.5 metres for the extent of the populated area.

Additional shoulder widths are also provided if there is a high number of PCUs of non-motorised
vehicles,(defined as more than 300 PCUs per day on average (Tables B.4.2 and B4.3)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 11

Table B.4.2: Increased ‘shoulder’ widths (each side) for unpaved LVRs

Basic shoulder widths (m) High


Design standard ‘Populated’areas
Flat Rolling Mount’n Escarp’t PCUs
DC4 unpaved +3.5 +2.0
DC3 unpaved Shoulders not defined for unpaved roads. +3.5 +1.5
The figure is for increased width each side
DC2 unpaved but DC1 is unlikely in populated areas +2.5 +1.25(1)
DC1 +2.0
Notes
1. DC2 is effectively a single carriageway hence less extra width is needed for PCUs.

Table B.4.3 Shoulder widths (each side) for paved LVRs

Basic shoulder widths (m) Populated


High
Design standard section for
Flat Rolling Mount’n Escarp’t PCUs
‘parking’
DC4 paved 1.25 1.25 0.5 0.5 3.5 +2.0
DC3 paved 1.0 1.0 0.5 0.5 3.5 +1.5
(1)
DC2 paved 1.5 1.5 1.0 1.0 2.5 +1.25(2)
DC1(1) 2.0
Notes
1. Paved sections are single carriageway resulting in a wider shoulder.
2. DC2 is effectively a single carriageway hence less extra width is needed for high PCUs.

4.3 Geometric design standards for LVRs


The design standards are shown in Tables B.4.4 to B.4.10. In these Tables ‘large vehicles’ are defined as
trucks with three or more axles and gross vehicle weights greater than 10 tonnes. Two sets of tables are
shown, one for paved roads and one for unpaved roads.

Sometime there will be cases where it is impossible to meet any of the standards mainly due to severe
terrain conditions. Under such circumstances the standards must be relaxed and suitable permanent
signage used to warn road users.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 12

Table B.4.4: Geometric design standards for paved DC4(1) (AADT 150-300)

Populated
Design Element Unit Flat Rolling Mountain Escarpment
areas
Design speed km/hr 70 60 50 25 50
Width of running
m 6.5(2) 6.5(2) 6.5 6.5 6.5(1)
surface
Width of shoulders m 1.25(2) 1.25(2) 0.5 0.5 1.25(3)
Total width m 9.0 9.0 7.5 7.5 9.0
Min stopping sight
m 110 90 70 25 65
distance
Min horizontal
m 195 135 85 15(4) 85
radius for SE=4%
Min horizontal
m 170 120 75 17(4) NA
radius for SE=7%
Min horizontal
m 150 105 70 22(4) NA
radius for SE=10%
Max desirable
% 4 7 10 12 4
gradient
Maximum gradient % 7 10 12(5) 12(5) 6
Min crest vertical
K 21 12 7 4 7
curve
Min sag vertical
K 4.8 3.5 2.2 1.3 2.2
curve
Normal cross-fall % 3 3 3 3 3
Shoulder cross-fall % 6 6 3 3 6
Notes:
1. If there are more than 80 large vehicles then DC5 should be used.
2. If the number of large vehicles is >40 then this should be increased to 7.0m and shoulders reduced to 1.0m.
3. Parking lanes and footpaths may be required.
4. On hairpin stacks the minimum radius may be reduced to a minimum of 15m.
5. Length not to exceed 200m and relief gradients required (<6% for minimum of 200m).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 13

Table B.4.5: Geometric design standards for unpaved DC4(1) (AADT 150-300)

Design Element Unit Flat Rolling Mountain Escarpment Populated areas


Design speed km/hr 70 60 50 25 50
(3) (3)
Road width m 7.0 7.0 7.0 7.0 7.0(2,3)
Min stopping sight
m 125 105 75 28 70
distance
Min horizontal radius m 245 175 110 23(4) 110
Max desirable gradient % 4 6 6 6 4
Max gradient % 6 9 9 9 6
Max. super-elevation % 6 6 6 6 6
Min crest vertical curve K 34 19 11 6 11
Min sag vertical curve m 4.8 3.5 2.2 1.3 2.2
Normal cross-fall(5) % 6 6 6 6 6
Notes:
1. If there are more than 80 large vehicles then DC5 should be used.
2. Parking lanes and footpaths may be required.
3. If the number of large vehicles >40 then this should be increased to 7.5m.
4. On hairpin stacks the minimum radius may be reduced to a minimum of 15m.
5. Cross fall can be reduced to 4% where warranted (eg poor gravel (for safety), low rainfall).

Table B.4.6: Geometric design standards for paved DC3(1) (AADT 75-150)

Design Element Unit Flat Rolling Mountain Escarpment Populated areas


Design speed km/hr 70 60 50 25 50
Width of running
m 6.0 6.0 6.0 6.0 6.0
surface
Width of shoulders m 1.0 1.0 0.5 0.5 1.0(2)
Total width m 8.0 8.0 7.0 7.0 8.0
Min stopping sight
m 110 90 70 25 65
distance
Min horizontal radius
m 195 135 85 20(3) 85
for SE=4%
Min horizontal radius
m 170 120 75 18(3) NA
for SE=7%
Min horizontal radius
m 150 105 70 16(3) NA
for SE=10%
Max desirable gradient % 4 7 10 12 4
(4,5) (4,5)
Maximum gradient % 7 10 12 12 6
Min crest vertical curve K 21 12 7 2 7
Normal cross-fall % 4 4 4 4 4
Minimum sag vertical
m 4.8 3.5 2.2 1.3 2.2
curve
Normal cross-fall % 3 3 3 3 3
Shoulder cross-fall % 6 6 3 3 6
Notes:
1. If there are more than 30 large vehicles then DC4 should be used.
2. Parking lanes and footpaths may be required.
3. On hairpin stacks the minimum radius may be reduced to a minimum of 15m.
4. Length not to exceed 200m and relief gradients required (<6% for minimum of 200m).
5. 5 If the number of large vehicles <20 this can be increased to 15%.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 14

Table B.4.7: Geometric design standards for unpaved DC3(1) (AADT 75-150)

Design Element Unit Flat Rolling Mountain Escarpment Populated areas


Design speed km/hr 70 60 50 25 50
Road width m 7.0 7.0 6.5 6.5 7.0(2)
Min stopping sight
m 125 105 75 28 70
distance
Min horizontal radius m 245 175 110 23(4) 110
Max desirable
% 4 6 6 6 4
gradient
Max gradient % 6 9 9 9 6
Max. super-elevation % 6 6 6 6 6
Min crest vertical
K 34 19 11 3 11
curve
Minimum sag vertical
K 4.8 3.5 2.2 1.3 2.2
curve
Normal cross-fall(3) % 6 6 6 6 6
Notes:
1. If the number of large vehicles is >30, then DC4 should be used.
2. Parking lanes and footpaths may be required.
3. Cross fall can be reduced to 4% where warranted (eg poor gravel (for safety), low rainfall).
4. On hairpin stacks the minimum radius may be reduced to a minimum of 15m.

Table B.4.8: Geometric design standards for DC2 paved(1) (AADT 25-75)

Populated
Design Element Unit Flat Rolling Mountain Escarpment
areas
Design speed km/hr 60 50 40 20 50
Width of running surface m 3.3 3.3 3.3 3.3 3.3
Width of shoulders m 1.5 1.5 1.0 1.0 1.5(2)
Total width m 6.3 6.3 5.3 5.3 6.3
Min stopping sight distance m 85 70 50 17 65
Min horizontal radius for
m 135 85 50 15(3) 85
SE=4%
Min horizontal radius for
m 120 75 45 15(3) NA
SE=7%
Min horizontal radius for
m 105 70 40 15(3) NA
SE=10%
Max desirable gradient % 4 7 10 12 4
(4) (4)
Max gradient % 7 10 12 15 6
Max. super-elevation % 6 6 6 6 6
Min crest vertical curve K 12 7 4 2 7
Minimum sag vertical curve K 3.5 2.2 1.3 0.7 2.2
Normal cross-fall % 3 3 3 3 3
Shoulder cross-fall % 6 6 3 3 6
Notes:
1. If the number of large vehicles >20 then DC3 should be used.
2. Parking lanes and footpaths may be required.
3. On hairpin stacks the minimum radius may be reduced to a minimum of 13m.
4. Length not to exceed 200m and relief gradients required (<6% for minimum of 200m).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 15

Table B.4.9: Geometric design standards for DC2 (1, 2) unpaved (AADT 25-75)

Populated
Design Element Unit Flat Rolling Mountain Escarpment
areas
Design speed km/hr 60 50 40 20 50
(5)
Road width m 6.0 6.0 6.0 6.0 6.0(3)
Min stopping sight distance m 95 75 55 20 70
(4)
Min horizontal radius m 175 110 70 15 110
Max desirable gradient % 4 6 6 6 4
Max gradient % 6 9 9 9 6
Max. super-elevation % 6 6 6 6 6
Min crest vertical curve K 19 11 6 3 11
Minimum sag vertical curve K 3.5 2.2 1.3 0.7 2.2
Normal cross-fall % 6 6 6 6 6
Notes:
1. If the number of large vehicles is >20 then DC3 should be used.
2. If the number of large vehicles is <10 then DC1 may be used
3. Parking lanes and footpaths may be required.
4. On hairpin stacks the minimum radius may be reduced to a minimum of 13m.
5. Road widths may be reduced at the discretion of the engineer and approval of the client to address specific local
conditions, especially in mountainous areas

Table B.4.10: Geometric design standards for DC1 (AADT 1-25)

Populated
Design Element Unit Flat Rolling Mountain Escarpment
areas
Desirable speed km/hr 50 40 30 20 40
Road width m 4.5 4.5 4.5 4.5 4.5
Min stopping sight distance m 70 55 35 18 50
(1)
Min horizontal radius m 110 70 35 15 70
Max desirable gradient % 4 6 6 6 4
(2) (2) (2) (2)
Max gradient % 12 12 12 12 6
Min crest vertical curve K 11 6 3 2 6
Minimum sag vertical curve K 2.2 1.3 0.7 0.5 1.3
Normal cross-fall % 6 6 6 6 6
Notes:
1. On hairpin stacks the minimum radius may be reduced to 13m.
2. Length not to exceed 200m and relief gradient required (<6% for minimum of 200m)

For the lowest category of road it may sometimes be necessary to adopt a basic access only approach.
For such roads it may be too expensive to provide a design speed but minimum absolute standards must
be applied. These are summarised in Table B.4.10.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 16

Table B.4.11: Minimum standards for basic access

Characteristic Minimum requirements


Radius of horizontal curvature 12m absolute but up to 20m depending on expected vehicles
Vertical curvature
K value for crests 2.5
K value for sags 0.6
Maximum gradients
Open to all vehicles 14%
Open only to cars and pick-ups 16%
Minimum stopping sight distance Flat and Rolling terrain 50m
Mountainous 35m
Escarpments 20m

For classes of road with the higher design speeds, adverse cross fall should be removed for curves with
low radii as indicated in Table B.4.12.

Table B.4.12: Adverse cross-fall to be removed if radii are less than shown

Minimum radii (m)


Design speed (km/h)
Paved Unpaved
<50 500 700
60 700 1000
70 1000 1300
85 1400 _
100 2000 _

Where super-elevation is required it should be developed gradually as indicated in Table B.4.13.

Table B.4.13: Super-elevation development lengths

Design speed (km/h) Development length (m)


30 25
40 30
50 40
60 55
70 65
80 80

In situations where low radii of curvature are necessary, the curves must be widened on the inside as
indicated in Table B.4.14.

Table B.4.14: Widening recommendations (m)

Single lane roads Two lane roads


Curve radius 20 30 40 60 <50 51-150 151-300 301-400
(1)
Increase in width 1.5 1.0 0.75 0.5 1.5 1.0 0.75 0.5
Notes:
1. See Section D.4.6.4 dealing with hairpin stacks.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 4 - 17

4.4 Design-by-eye
The design-by-eye method is best suited to rehabilitation or upgrading projects where a road alignment
already exists and is the preferred method for developing a design for a track or undesignated road under
a community roads programme where a walking track is being improved to enable it to carry occasional
vehicles. Nevertheless, considerable experience and skill is needed to carry out the design-by-eye method
and the approach should only be used under the guidance and supervision of an experienced engineer.

4.5 Typical Cross Sections


Typical cross sections for a range of conditions are shown in Figures B.4.2 to B.4.13. They include:
ƒ Roads on flat terrain;
ƒ Roads on rolling terrain;
ƒ Roads on mountainous terrain;
ƒ Roads on escarpments;
ƒ Roads through populated areas;
ƒ Roads on expansive soils.

Slope dimensions for the various conditions are summarised in Table B.4.15.

Table B.4.15: Slope dimensions for cross-sections (ratios are vertical:horizontal)

Side slope Safety


Material Height of slope (m) Back slope
Cut Fill classification
0.0-1.0 1:4 1:4 1:3 Recoverable
Not
Earth(1) 1.0-2.0 1:3 1:3 1:2
recoverable
>2.0 1:2 1:2 1:1.5 Critical
Rock Any height Dependant on costs Critical
Expansive 0-2.0 n/a 1:6 Recoverable
clays(2) >2.0 n/a 1:4
Notes:
1. See Cross Section
2. Certain soils may be unstable at slopes of 1:2. Geotechnical advice required.
3. The drainage ditch should be moved away from the embankment

The detailed cross-sections to scale are given in the Standard Detail Drawings (2011).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS



 
B - Chapter 4 - 18



 
 




 

Figure B.4.2: Typical cross section, DC1 – 4, Flat Terrain, Unpaved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 5.0 5.5 6.0
B Shoulder width (m) 0.6 0.5 0.75 0.75
C Min Crossfall/Camber (%) 4 4 4 4
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) Varies
H Crown height (m) 0.35 0.35 0.5 0.5
J Cleared width (m) 15 20 20 20
Notes:

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


1. Section not drawn to scale;
2. V-shape is the standard shape of the drainage ditch constructed by motor or towed grader;
3. Trapezoidal drains are commonly used and are much easier to dig and clean using labour-intensive methods. The minimum recommended width is 400mm and the typical cross-
section is shown below
!
6400

1 1-3,1'
2
4. Rectangular drains need to be lined with rock, brick stone masonry or concrete to maintain their shape;
5. More detail on side drains is provided in Part D, Section 5.4.4.

 


 
 


0 

 

Figure B.4.3: Typical cross section, DC1 – 4, Flat Terrain, Paved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m), minimum 3.3 6.0 6.5
B Shoulder width (m) 1.5 1.0 1.25
B1 Shoulder Crossfall (%) 6 6 6
C Crossfall/Camber (%) 3 3 3
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of side ditch (m) Varies

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


H Crown height (m) 0.75 0.75 0.75
J Cleared width (m) 20 20 20
Notes:
1. Section not drawn to scale;
2. V-shape is the standard shape of the drainage ditch constructed by motor or towed grader;
3. Trapezoidal drains are commonly used and are much easier to dig and clean using labour-intensive methods. The minimum recommended width is 500mm;
4. Rectangular drains need to be lined with rock, brick stone masonry or concrete to maintain their shape;
5. More detail on side drains is provided in Part D, Section 5.4.4.
B - Chapter 4 - 19
B - Chapter 4 - 20


 


 

 



 
Figure B.4.4: Typical cross section, DC1 – 4, Rolling Terrain, Unpaved
Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 5.0 5.5 6.0
B Shoulder width (m) 0.6 0.5 0.75 0.75
C Min Crossfall/Camber (%) 4 4 4 4
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) Varies
G Side slope (v:h ratio) See Table B.4.15

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


H Crown height (m) 0.35 0.35 0.5 0.5
J Cleared width (m) 15 20 20 20
K Embankment toe (m) Varies
Notes:
1. Section not drawn to scale.

 


 

  0 


 

Figure B.4.5: Typical cross section, DC1 – 4, Rolling Terrain, Paved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 6.0 6.5
B Shoulder width (m) 1.5 1.0 1.25
B1 Shoulder Crossfall (%) 6 6 6
C Crossfall/Camber (%) 3 3 3
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


F Depth of side ditch (m) Varies
G Side slope See Table B.4.15
H Crown height (m) 0.75 0.75 0.75
J Cleared width (m) 20 20 20
K Embankment toe (m) Varies
Notes:
1. Section not drawn to scale.
B - Chapter 4 - 21


"
B - Chapter 4 - 22

 #
 

/+3

 

/+5

 0+/
 

 5; 







 

Figure B.4.6: Typical cross section, DC1 – 4, Mountainous Terrain, Unpaved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 5.0 5.5 6.0
B Shoulder width (m) 0.6 0.5 0.5 0.5
C Min Crossfall/Camber (%) 4 4 4 4

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) 0.35
G Side slope (v:h ratio) See Table B.4.15
J Cleared width (m) 15 20 20 20
L Ditch width (m) Varies
Notes:
1. Section not drawn to scale.


"
 #
 

/+3

 

/+5

 0+/
 

 5; 
 0








Figure B.4.7: Typical cross section, DC1 – 4, Mountainous Terrain, Paved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 6.0 6.5
B Shoulder width (m) 1.0 0.5 0.5
B1 Shoulder crossfall (%) 3 3 3

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


C Crossfall/Camber (%) 3 3 3
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) 0.5
G Side slope (v:h ratio) See Table B.4.15
J Cleared width (m) 20 20 20
L Ditch width (m) Varies
Notes:
B - Chapter 4 - 23

1. Section not drawn to scale;


1.
B - Chapter 4 - 24




 /+3

 "#
/+5

 
 

 $
 
%$ !

$"%$%"
0+/

 









 "

Figure B.4.8: Typical cross section: DC1 – 4, Escarpment Terrain, Unpaved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 5.0 5.5 6.0
B Shoulder width (m) 0.6 0.5 0.5 0.5
C Min Crossfall/Camber (%) 4 4 4 4

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) Min 0.35
J Cleared width (m) 15 20 20 20
L Ditch width (m) Varies
M Slope of retaining structure Varies
Notes:
1. Section not drawn to scale;




 /+3

 "#
/+5

 
 

 $
 
%$ !

$"%$%"
0+/

 
0 
 







 "



Figure B.4.9: Typical cross section: DC1 – 4, Escarpment Terrain, Paved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m) 3.3 5.5 6.5
B Shoulder width (m) 1.0 0.5 0.5
B1 Shoulder crossfall (%) 3 3 3

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


C Crossfall/Camber (%) 3 6 3
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of Side ditch (m) Min 0.5
J Cleared width (m) 20 20 20
L Ditch width (m) Varies
M Slope of retaining structure Varies
Notes:
B - Chapter 4 - 25

1. Section not drawn to scale;



B - Chapter 4 - 26

 

 
 $1
 
0+/  $0
/+14 /+4
0 0
3 3

 $0

Figure B.4.10 Typical cross section, DC1 – 4, Populated areas, Unpaved


Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m), minimum 3.3 5.0 5.5 6.0
B Shoulder width (m) 0.6 0.5 0.75 0.75
C Min Crossfall/Camber (%) 4 4 4 4
J Cleared width (m) 15 20 20 20
Notes:
1. Open channel type A – 25 cm thick mortared stone pitching
Open channel type B – 25 cm thick mortared stone pitching

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


2. Wearing course
3. Choice of open channel dependent on local conditions
4. Provide lined channels only where maintenance of road surface and camber at original levels is guaranteed.


 
 $
#
  
 


Figure B.4.13: Typical cross section, DC1 – 4, Flat Terrain, Expansive soils, Paved
Design Classes
Label Design Criteria
DC1 DC2 DC3 DC4
A Carriage width (m), minimum 3.3 6.0 6.5
B Shoulder width (m) 1.5 1.0 1.25
B1 Shoulder Crossfall (%) 6 6 6
C Crossfall/Camber (%) 3 3 3
D Backslope of ditch (v:h ratio)
See Table B.4.15
E Side slope of ditch (v:h ratio)
F Depth of side ditch (m) Varies

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


G Side slope Varies
H Crown height (m) 0.75 0.75 0.75
J Cleared width (m) 20 20 20
Notes:
1. Section not drawn to scale.
B - Chapter 4 - 27
B - Chapter 5 - 28

5. MATERIALS

For materials specifications see Part D, Section 6.7.

5.1 Subgrades
Subgrades are classified on the basis of the laboratory soaked CBR tests on samples compacted to 97%
AASHTO T180 compaction. Samples are soaked for four days or until zero swell is recorded. The subgrade
strength for design is assigned to one of six strength classes reflecting the sensitivity of thickness design
to subgrade strength. The classes are defined in Table B.5.1.

For the design of earth and gravel roads, if no suitable laboratory is available, the existing subgrade can
be assessed using a DCP at the time of the year that the soil is at its wettest.

Table B.5.1: Subgrade classes

Subgrade Class
Design CBR S2 S3 S4 S5 S6
Range % 3-4 5-8 9 - 14 15 - 29 30+

No allowance for CBRs below 3% has been made because, from both a technical and economic
perspective, it would normally be inappropriate to lay a pavement on soils of such poor bearing capacity.
For such materials, special treatment is required (see Section D.6.19.7).

The use of Class S2 soils as direct support for the pavement should be avoided as much as possible.
Wherever practicable, such relatively poor soils should be excavated and replaced, or covered with an
improved subgrade.

Class S6 covers all subgrade materials having a soaked CBR greater than 30 and which comply with the
plasticity requirements for natural sub-base. In such cases, no sub-base is required.

5.1.1 Specifying the design subgrade class

The CBR results obtained from the subgrade soils testing are used to determine which subgrade class
should be specified for design purposes in accordance with Table B.5.1. The variation in results may make
selection unclear. In such cases it is recommended that, firstly, the laboratory test process is checked to
ensure uniformity (to minimise inherent variation arising from, for example, inconsistent drying out of
specimens). Secondly, more samples should be tested to build up a more reliable basis for selection.
Plotting these results as a cumulative distribution curve (S-curve) in which the y-axis is the percentage
of samples less than a given CBR value (x-axis) provides a method of determining a design CBR value
(Figure B.5.1).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 29

Figure B.5.1: Illustration of CBR strength cumulative distribution

The actual subgrade CBR values used for design depends on the traffic class as shown in Table B.5.2. For
example, as indicated in the Table, for a design traffic class of LV5 the design CBR value should be the
lower 10th percentile (ie the value exceeded by 90% of the CBR measurements).

Table B.5.2: Dependence of design subgrade on design traffic class

Traffic class Design CBR


LV5 (0.5-1.0 mesa) Lower 10-percentile
LV3 and LV4 (0.1-0.5 mesa) Lower 15-percentile
LV1 and LV2 (<0.1mesa) 30th percentile

5.1.2 Material depth

The concept of “material depth” is used to denote the depth below the finished level of the road to
which soil characteristics have a significant effect on pavement behaviour and throughout which the
nominal subgrade strength selected for design should be maintained.

Table B.5.3 specifies typical material depths used for determining the design CBR of the subgrade. Note
that this depth may be insufficient in certain special cases where “problem” soils occur (See Part D,
Section 6.19).

Table B.5.3: Material depth by road category

Road Category Material Depth (mm)


DC 7 and DC 8 1,000 – 1,200
High volume roads
DC 5 and DC 6 800 – 1,000
DC 3 and DC 4 800
Low volume roads
DC 1 and DC 2 700

The minimum depths indicated in the Table are not depths to which re-compaction and reworking is
necessarily required. Rather, they are the depths to which the Engineer should confirm that the nominal

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 30

subgrade strength is available. In general, unnecessary working of the subgrade should be avoided and
limited to rolling prior to constructing overlying layers

For the stronger subgrades, especially Class S4 and higher (CBR 9-14% and more) the depth check is to
ensure that there is no underlying weaker material which could lead to detrimental performance.

It is recommended that the Dynamic Cone Penetrometer (DCP) be used during construction to monitor
the uniformity of subgrade support to the recommended minimum depths given in Table B.5.3.

5.1.3 Improved subgrade layers

There are many advantages to improving the CBR strength of the in situ subgrade to a minimum of
15% (Subgrade Class S5) by constructing one or more improved layers where necessary. In principle,
where a sufficient thickness of improved subgrade is placed, the overall subgrade bearing strength is
increased to that of a higher class and the sub-base thickness may be reduced accordingly. This is often
an economic advantage as sub-base quality materials are generally more expensive than fill materials,
hence the decision whether or not to consider the use of an improved subgrade layer(s) will generally
depend on the respective costs of sub-base and improved subgrade materials.

5.1.4 Dealing with poor subgrade soils

Methods of design and treatment for problem soils are described in Part D Section 6.19.

5.2 Pavement Materials


The material code and characteristics of the material types for both paved and unpaved roads are
described in Table B.5.4.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 31

Table B.5.4: Pavement material types and abbreviated nominal specifications


used in the paved and unpaved catalogue of designs

Code Material Abbreviated Specifications

Min. CBR: 80% @ 98/100% AASHTO T180 and 4 days soaking


Max. Swell: 0.2%
G80 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified.
PI: < 6 or as otherwise specified (material specific).
Min. CBR: 65% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G65 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 55% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G55 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 45% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G45 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 30% @ 95/97% AASHTO T180 & highest anticipated moisture
content
G30 Natural gravel Max. Swell: 1.0% @ 100% AASHTO T180
Max. Size and grading: Max size 63mm or 2/3 layer thickness
PI: < 12 or as otherwise specified (material specific)
Min. CBR: 30% @ 95/97% AASHTO T180 & highest anticipated moisture
content
G25 Natural gravel Max. Swell: 1.0% @ 100% AASHTO T180
Max. Size and grading: Max sixe 63mm or 2/3 layer thickness.
PI: <12 or as otherwise specified (material specific)
Min. CBR: 15% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G15 Gravel/soil Max. Swell: 1.5% @ 100% AASHTO T180
Max. Size: 2/3 of layer thickness
PI: < 12 or 3GM + 10 or as otherwise specified (material specific)
Min. CBR: 7% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G7 Gravel/soil Max. Swell: 1.5% @ 100% AASHTO T180
Max. Size: 2/3 layer thickness
PI: < 12 or 3GM + 10 or as otherwise specified (material specific)
Min. CBR: 3% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G3 Gravel/soil
Max. Swell: N/A
Max. Size: 2/3 layer thickness
Note:
Two alternative minimum levels of compaction are specified. Where the higher densities can be realistically attained in
the field (from field measurements on similar materials or other established information) they should be specified by the
Engineer.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 32

5.2.1 Materials requirements for roadbase

A wide range of materials including lateritic, calcareous and quartzitic gravels, river gravels and other
transported and residual gravels, or granular materials resulting from weathering of rocks can be used
successfully as road base materials.

Particle size distribution: The grading envelopes to be used for road base are shown in Table B.5.6.
Envelope A varies depending whether the nominal maximum particle size is 37.5mm, 20mm or 10mm. A
requirement of five to ten per cent retained on successive sieves may be specified at higher traffic (>0.3
mesa) to prevent excessive loss in stability. Envelope C extends the upper limit of envelope B to allow
the use of sandy materials, but its use is not permitted in wet climates. Envelope D is similar to a gravel
wearing course specification, and is used for very low traffic volumes. The grading is specified only in
terms of the grading modulus (GM) and can be used in both wet and dry climates.

Table B.5.5: Particle size distribution for natural gravel base

Per cent by mass of total aggregate passing test sieve


Test Sieve Envelope A
size Nominal maximum particle size Envelope B Envelope C
37.5mm 20mm 10mm
50mm 100 100
37.5mm 80-100 100 80-100
20mm 55-95 80-100 100 55-100
10mm 40-80 55-85 60-100 40-100
5mm 30-65 30-65 45-80 30-80
2.36mm 20-50 20-50 35-75 20-70 20-100
1.18mm - - - - -
425µm 8-30 12-30 12-45 8-45 8-80
300µm - - - - -
75µm 5-20 5-20 5-20 5-20 5-30
Envelope D
1.65 < GM < 2.65

Strength and plasticity: The strength requirement varies depending on the traffic level and climate
as outlined in the Catalogue of Structures (Chapter B.6). The soaked CBR test is used to specify the
minimum road base material strength.

The plasticity requirement also varies depending on the traffic level and climate as shown in Tables B.5.7
and B.5.8. A maximum plasticity index of 6 has been retained for higher traffic levels and also on weaker
subgrades. For designs in dry environments the plasticity modulus for each traffic and subgrade class
can be increased depending on the crown height and whether unsealed or sealed shoulders are used as
described in Part D, Section 6.17.2 and Figure D.6.22.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 33

Table B.5.6: Plasticity requirements for natural gravel road base materials

Traffic class (Mesas)


LV1 LV2 LV3 LV4 LV5

Subgrade

Property
class4
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

S2 Ip <12 <9 <6 <6 <6


PM <400 <150 <120 <90 <90
Grading B B A A A
S3 Ip <15 <12 <9 <6 <6
PM <550 <250 <180 <90 <90
Grading C(1) B B A A
S4 Ip Note(2) <12 <12 <9 <9
PM <800 <320 <300 <200 <90
Grading D(3) B B B A
S5 Ip Note(2) <15 <12 <12 <9
PM - <400 <350 <250 <150
Grading D(3) B B B A
S6 Ip Note(2) <15 <15 <12 <9
PM - <550 <500 <300 <180
Grading D(3) C(1) B B A
Notes:
1. Grading ‘C’ is not permitted in wet environments or climates (N<4); grading ‘B’ is the minimum requirement
2. Maximum Ip = 8 x GM
3. Grading ‘D’ is based on the grading modulus 1.65 < GM < 2.65
4. All base materials are natural gravels; Subgrades are non-expansive

Lateritic road base gravels: The requirements for selection and use of lateritic gravels for bases are
slightly different to those given for other natural gravels. These are presented in Table B.5.8. A maximum
plasticity index of 9 has been specified for higher traffic levels and weak subgrades. For design traffic
levels greater than 0.3 mesa, a requirement is set that the liquid limit should be less than 30. Below this
traffic level, this requirement is relaxed to a liquid limit of less than 35. Where sealed shoulders over one
metre wide are specified in the design, the maximum plasticity modulus may be increased by 40 per cent.
A minimum field compacted dry density of 2.0 Mg/m3 is required for these materials.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 34

Table B.5.7: Guidelines for the selection of lateritic gravel road base materials

Traffic class (Mesas)


LV1 LV2 LV3 LV4 LV5

Subgrade

Property
class <0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

Ip <15 <12 <9 <9 <6


S2 PM <400 <150 <150 <120 <90
Grading B B A A A
Ip <18 <15 <12 <9 <6
S3 PM <550 <250 <180 <120 <90
Grading C(1) B B A A
Ip <20(1) <15 <15 <9 <9
S4 PM <800 <320 <300 <200 <90
Grading GM 1.6-2.6 B B B A
Ip <25(1) <18 <15 <12 <9
S5 PM - <400 <350 <250 <150
Grading GM 1.6-2.6 B B B B
Ip <25(1) <20 <18 <15 <12
S6 PM - <550 <400 <300 <180
Grading GM 1.6-2.6 B B B A

Notes:
1. Maximum Ip = 8 x GM
2. Unsealed shoulders are assumed. Further modification to the limits can be made if the shoulders are sealed.
3. The compaction requirement for the soaked CBR test to define the subgrade classes is 100% Mod. AASHTO with a
minimum soaking time of 4 days or until zero swell is recorded. This is a relaxation of the soaked CBR requirement for
natural gravel base materials given in the catalogues.

Basic igneous rock (including basaltic and doleritic gravels): These materials occur extensively in
Ethiopia and their more wide-spread use could result in significant savings provided the characteristics
of the material are good enough to serve as a road base material.. The following indicative limits can
contribute to successful use of the material in road bases:
ƒ Maximum secondary mineral content of 20 per cent (determined from petrographic analysis);
ƒ Maximum loss of 12 or 20 per cent after 5 cycles in the sodium or magnesium sulphate soundness
tests, respectively;
ƒ Clay index of less than 3 in the dye absorption test;
ƒ Increase in modified glycol-soaked AIV from the wet modified AIV should be < 4% units.
ƒ Durability mill index of less than 125.

In drier climatic areas (N>4), the materials can be used unmodified up to a maximum plasticity index
of 10. However, it is suggested that the materials should not be used in wet areas unless chemically
modified. The risk of using the material can be minimised if consideration is given to:
ƒ The variability of the material deposit, with good selection and control procedures in place for the
operation of the pit and on site;
ƒ The provision of good drainage conditions (these materials are particularly sensitive to moisture);
ƒ The adequacy of the pavement design (the use of Pavement Catalogue 2 with sealed shoulders is
suggested);
ƒ The use of double surface treatments or similar.

Engineers need to use considerable judgement, experience and information from other roads in the area
to utilise these materials successfully. Risks must be identified and controlled.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 35

Cinder gravels: Cinder gravels have been used successfully as road base on experimental sections
constructed in the late 1970s (TRL, 1987). From these trials it was concluded that with careful selection,
cinder gravels can be used for lightly trafficked paved roads in accordance with the requirements of the
pavement design chart 2 (Table B.6.7)

5.2.2 Material requirements for sub-base

Strength requirements: A minimum CBR of 30% is required at the highest anticipated moisture content
when compacted to the specified field density, usually a minimum of 95% (preferably 97% where
practicable) AASHTO T180 compaction.

Under conditions of good drainage and when the water table is not near the ground surface, the field
moisture content under a sealed pavement will be equal to or less than the optimum moisture content
in the AASHTO T180 compaction test. In such conditions, the sub-base material should be tested in the
laboratory in an unsaturated state.

If the road base allows water to drain into the lower layers, as may occur with unsealed shoulders and
under conditions of poor surface maintenance where the road base is pervious, saturation of the sub-
base is likely. In these circumstances the bearing capacity should be determined on samples soaked in
water for a period of four days. The test should be conducted on samples prepared at the density and
moisture content likely to be achieved in the field.

Particle size distribution and plasticity requirements: In order to achieve the required bearing capacity,
and for uniform support to be provided to the upper pavement, limits on soil plasticity and particle size
distribution may be required. Materials which meet the recommendations of Tables B.5.9 and B.5.10 will
usually be found to have adequate bearing capacity.

Table B.5.8: Typical particle size distribution for sub-bases

Per cent by mass of total


Sieve Size (mm)
aggregate passing test sieve
50 100
37.5 80 – 100
20 60 – 80
5 30 – 100
1.18 17 – 75
0.3 9 – 50
0.075 5 - 25

Table B.5.9: Plasticity characteristics for granular sub-bases

Climate Liquid Limit Plasticity Index Linear Shrinkage


Moist tropical and wet tropical (N<4) < 35 <6 <3

Seasonally wet tropical (N<4) < 45 < 12 <6

Arid and semi-arid (N>4) <55 < 20 <10

5.2.3 Material requirements for gravel wearing course

The specifications identify the most suitable materials in terms of two basic soil parameters – Shrinkage
Product and Grading Coefficient – which are determined from particle size distribution and linear
shrinkage as shown in Figure B.5.2.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 36

An alternative to using linear shrinkage and the shrinkage product is to use the plasticity index and the
associated plasticity product. For the range of materials likely to be used for gravel wearing course, the
plasticity index can be assumed to be 2 x linear shrinkage. The linear shrinkage (shrinkage product) is
recommended as it is based on one relatively simple test which has good precision limits in the shrinkage
ranges of acceptable gravel wearing course material.

Figure B.5.2: Material quality zones

The characteristics of materials in each zone are as follows:


A: Materials in this area generally perform satisfactorily but are finely graded and particularly prone
to erosion. They should be avoided if possible, especially on steep grades and sections with steep
cross-falls and super-elevations. Roads constructed from these materials require frequent periodic
labour intensive maintenance over short lengths and have high gravel losses due to erosion.
B: These materials generally lack cohesion and are highly susceptible to the formation of loose
material (ravelling) and corrugations. Regular maintenance is necessary if these materials are used
and the road roughness is to be restricted to reasonable levels.
C: Materials in this zone generally comprise fine, gap-graded gravels lacking adequate cohesion,
resulting in ravelling and the production of loose material.
D: Materials with a shrinkage product in excess of 365 tend to be slippery when wet.
E: Materials in this zone perform well in general, provided the oversize material is restricted to the
recommended limits.

Gravel loss: Gravel loss is the single most important reason why gravel roads are expensive in whole
life cost terms and often unsustainable, especially when traffic levels increase. Reducing gravel loss by
selecting better quality gravels or modifying the properties of poorer quality materials is one way of
reducing long term costs. Gravel losses (gravel loss in mm/year/100vpd) are determined in relation to
the quality of the gravel wearing course (Table B.5.11).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 37

Table B.5.10: Typical standardised gravel loss

Material Quality Zone1 Material Quality Typical gravel loss (mm/yr/100vpd)


Zone A Satisfactory 20
Zone B Poor 45
Zone C Poor 45
Zone D Marginal 30
Zone E Good 10
Notes:
1. See Figure B.5.2

The gravel losses shown in Table B.5.11 probably hold only for the first phase of the deterioration cycle
lasting possibly two or three years. Beyond that period, as the wearing course is reduced in thickness,
other developments, such as the formation of ruts, will also affect the loss of gravel material. However,
the rates of gravel loss given in the Table can be used as an aid to the planning for regravelling in the
future. A more accurate indication of gravel loss for a particular section of road can be obtained from
periodic measurement of the gravel layer thickness.

Material requirements for gravel roads in rural areas: Table B.5.12 shows the recommended specifications
for materials for unsealed rural roads

Table B.5.11: Recommended material specifications(1,3) for unsealed rural roads

Maximum size (mm) 37.5


Oversize index (Io)a ≤5%
Shrinkage product (Sp)b (2) 100 - 365 (max. of 240 preferable)
Grading coefficient (Gc)c (2) 16 - 34
Soaked CBR (at 95 per cent Mod AASHTO) ≥ 15 %
Treton impact value (%)(4) 20 – 65
a Io = Oversize index (percent retained on 37.5 mm sieve
b Sp = Linear shrinkage x percent passing 0.425 mm sieve
c Gc = (Percentage passing 26.5 mm - percentage passing 2.0 mm) x percentage passing
4.75 mm)/100
Notes:
1. Specifications should be applicable after placement and compaction
2. The Grading Coefficient and Shrinkage Product must be based on a conventional particle size distribution determination
which must be normalised for 100% passing the 37.5 mm screen.
3. Only representative material samples are to be tested.
4. The Treton Impact Value (TIV) limits exclude those materials that are too hard to be broken with a grid roller (TIV < 20%)
or too soft to resist excessive crushing under traffic (TIV > 65%).

Material requirements for gravel roads in ‘urban’ areas: The specifications in Table B.5.13 are
recommended for unsealed roads in areas where there is a significant number of dwellings and local
businesses. In comparison with the limits for rural roads, the limits for the oversize index have been reduced
to eliminate stones whilst the shrinkage product has been reduced to a maximum of 240 to reduce the
dust as far as practically possible. This lower limit reduces the probability of having unacceptable dust
from about 70% to 40%.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 5 - 38

Table B.5.12: Recommended material specifications for unsealed ‘urban’ roads

Maximum size (mm) 37.5


Oversize index (Io) 0
Shrinkage product (Sp) 100 - 240
Grading coefficient (Gc) 16 - 34
Soaked CBR (at 95 per cent Mod AASHTO) ≥ 15 %
Treton impact value (%) 20 – 65

5.2.4 Material Improvement

Obtaining materials that comply with the necessary grading and plasticity specifications for a gravel
wearing course can be difficult. Many of the natural gravels tend to be coarsely graded and relatively non
plastic and the use of such materials results in very high roughness levels and high rates of gravel loss in
service and, in the final analysis, very high life-cycle costs.

In order to achieve suitable wearing course properties a suitable Particle Size Distribution (PSD) can be
obtained by breaking down oversized material to a maximum size of 50 mm or smaller. Atterberg limits
may be modified by granular/mechanical stabilisation (blending) with other materials. These material
improvement measures are discussed in Part D, Section 6.7.6.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 39

6. PAVEMENT DESIGN

6.1 Design traffic classes


For structural pavement design, five traffic classes have been defined as shown in Table B.6.1. If the
estimate of cumulative traffic is close to the boundaries of a traffic class, then the basic traffic data and
forecasts should be re-evaluated and sensitivity analyses carried out to ensure that the choice of traffic
class is appropriate. If there is any doubt about the accuracy of the traffic estimates the next higher traffic
class should be selected for the design.

Table B.6.1: Traffic classes for flexible pavement design

Traffic range LV1 LV2 LV3 LV4 LV5/T2(1)


(mesas) < 0.01 0.01 – 0.1 0.1 – 0.3 0.3 – 0.5 0.5 – 1.0

LV5/T2 is the transition traffic zone between low-volume and high-volume roads with the former traffic
class (LV5) applying to the lower boundary of the traffic range and the latter traffic class (T2) applying to
the upper boundary.

6.2 Engineered natural surfaces


The following design standards are recommended for Engineered Natural Surfaces (ENS) in the DC1
design class carrying < 25 vpd. The details of the cross-section are given in Chapter B.4 but shown
schematically in Figure B.6.1 for convenience. Further supporting information is given in Section 6.15 of
Part D.

Figure B.6.1: Cross-section details ENS

ƒ The crown height of the earth road should be at least 35 cm above the bed of the drain.
ƒ Where the topography allows, wide, shallow longitudinal drainage for earth roads are preferred.
They minimise erosion, and will not block as easily as narrow ditches. The ditches grass over in
time, binding the soil surface and further slowing down the speed of water, both of which act to
prevent or reduce erosion.
ƒ The surface of earth roads should be graded and compacted to provide a durable and level
running surface for traffic and the road surface should have a minimum camber of 4% to ensure
water runs off the surface and into the side drains.
ƒ Areas where there are specific problems (usually due to water or to the poor condition of the
subgrade) may be treated in isolation by localised replacement of subgrade, gravelling, installation
of culverts, raising the roadway or by installing other drainage measures. This is the basis of a “spot
improvement” approach.
ƒ Water should be drained away from the carriageway side drains by installing lead off (mitre) drains,
to divert the flow into open space.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 40

6.3 Natural gravel roads


A gravel road consists of a wearing course and a structural layer (base) which covers the in situ material.
The minimum thickness of the structural layer is maintained in service by providing a wearing course
throughout the design life of the road.

To achieve adequate external drainage, the road must also be raised above the level of existing ground
such that the crown of the road is maintained at a minimum height (hmin) above the table drain inverts.
Cross sections are shown in detail in Chapter B.4 and shown here schematically for convenience (Figure
B.6.2).

Figure B.6.2 – Typical gravel road cross section in flat terrain.

The minimum height is dependent on the climate and road design class as shown in Table B.6.2.

Table B.6.2: Required minimum height (hmin) between road crown


and invert level of drain in relation to climate

Climate
Road Class Wet (N < 4) Dry (N > 4)
hmin (mm) hmin (mm)
DC-1 350 250
DC-2 400 450
DC-3 500 300
DC-4 350 400

Gravel roads are divided into two broad categories for design purposes namely ‘major’ and ‘minor’ gravel
roads. Gravel roads in classes DC3 and DC4 are defined as major gravel roads, minor gravel roads are
classes DC1 and DC2, except where the number of heavy vehicles exceeds about 10 per day. Major
gravel roads are engineered to a higher specification.

6.3.1 Major gravel roads

For major gravel roads the approach is as follows:


ƒ The sub-grade should be prepared in the same way as for a low volume sealed road.
ƒ It is assumed that the wearing course will be replaced at intervals related to the expected annual
gravel loss and before the structural layer is exposed to traffic and itself begins to wear away;
ƒ The geometry and drainage are upgraded to acceptable minimum levels during construction.
This may require the introduction of a fill layer between the compacted in situ sub-grade and the
wearing course.

Major gravel roads are likely to incur high maintenance costs in some circumstances namely;
ƒ When the quality of the gravel is poor.
ƒ Where no sources of gravel are available within a reasonable haul distance.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 41

ƒ On road gradients greater than about 6%.


ƒ In areas of high and intense rainfall.

In these circumstances spot improvements will almost certainly be justified, and, in some cases, it may
prove to be more economical to build a fully paved road at the outset.
The structural design procedure for major gravel roads: The design procedure consists of the following
steps:
ƒ Determine the traffic volume and traffic loading (Section B.3.4).
ƒ Determine the strength of the sub-grade at the appropriate moisture condition (Section B.5.1.1).
ƒ Establish the quality of the gravel that is to be used (Section B.5.2.3). If only very poor gravel is
available, blending with another gravel or soil to improve its properties may be an option (Section
B.5.2.5).
ƒ Determine the thickness of gravel base that is necessary to avoid excessive compressive stresses
in the sub-grade from Tables B.6.3 (a), (b) and (c).
ƒ Calculate the thickness of the wearing course based on the expected rate of gravel loss and a
realistic choice of the frequency of re-gravelling.

Table B.6.3 (a): Gravel base thickness for major gravel roads – strong gravel (G45)

Subgrade Traffic Classes (mesas)


Strength Class LV1 LV2 LV3 LV4 LV5
CBR (%) (<0.01) (0.01-0.1) (0.1-0.3) (0.3-0.5) (0.5-1.0)
S2 (3-4) 175 225 250 300 350
S3 (5-7) 150 200 225 250 300
S4 (8-14) 100 150 200 200 250
S5 (15-29) 100 125 150 175 200

Table B.6.3 (b): Gravel base thickness for major gravel roads – medium gravel (G30)

Subgrade Traffic Classes (mesas)


Strength Class LV1 LV2 LV3 LV4 LV5
CBR (%) (<0.01) (0.01-0.1) (0.1-0.3) (0.3-0.5) (0.5-1.0)
S2 (3-4) 175 250 290 325 370
S3 (5-7) 150 200 250 275 325
S4 (8-14) 125 175 200 220 275
S5 (15-29) 100 100 150 175 200

Table B.6.3 (c): Gravel base thickness for major gravel roads – weak gravel (G15)

Subgrade Traffic Classes (mesas)


Strength Class LV1 LV2 LV3 LV4 LV5
CBR (%) (<0.01) (0.01-0.1) (0.1-0.3) (0.3-0.5) (0.5-1.0)
S2 (3-4) 225 325 375 NA NA
S3 (5-7) 200 250 325 350 NA
S4 (8-14) 150 225 275 300 NA
S5 (15-29) 150(1) 150(1) 200(1) 200(1) NA
Note:
1. This is the additional depth of compacted sub-grade material

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 42

The following points should be noted:


ƒ The thicknesses required increases considerably if the gravel is weak hence stronger gravels should
generally be used if they are available at reasonable cost.
ƒ On relatively weak subgrades (S2 and S3), the use of strong gravels (G45) should be avoided
because of the poor “balance” of such pavements. Instead, the use of an improved subgrade layer
should be considered (Section B.5.1.4).
ƒ Where the available gravel is not homogeneous, it will be necessary to substitute a particular
class of gravel with one or more different classes of gravel of appropriate thickness. The following
conversion factors may be used for this purpose.
G45 = 1.5 x G15
G30 = 1.3 x G15
Thus, a 200mm layer of G45 material could be substituted with a 300 mm layer of G15 material.

For effective compaction of the gravel layer, it is necessary to restrict the loose thickness of gravel to a
maximum lift of about 200 mm. Thus, any of the gravel layers that require a compacted thickness of more
than 150 mm must be compacted in more than one 200 mm lift.

Determination of wearing course thickness: The wearing course thickness depends on the annual gravel
loss and the number of years between re-gravelling operations. The predicted annual gravel loss is given
in Table B.6.4.

Table B.6.4: Typical gravel loss

Typical gravel loss (mm/


Material Quality Zone(1) Description of Material Quality
yr/100vpd)
Zone A Satisfactory 20
Zone B Poor 45
Zone C Poor 45
Zone D Marginal 30
Zone E Good 10
Note:
1. See Figure B.5.2

The rates of gravel loss increase significantly on gradients greater than about 6% and in areas of high and
intense rainfall. On some gradients, the increase could be greater than 50% depending on the steepness
of the gradient and material quality. Spot improvements should be considered on these sections.

Re-gravelling should take place before the sub-base is exposed. The re-gravelling frequency, R, is typically
in the range 5 - 8 years. This decreases considerably if poor quality gravels have to be used. For example,
if the gravel quality is in zones B or C, the loss rate will be 45mm per year per 100vpd. Therefore a class
DC4 gravel road carrying 200vpd will lose 90mm per year and require re-gravelling every two years

The wearing course thickness = R x GL


R = regravelling frequency in years
GL = annual gravel loss.

6.3.2 Minor gravel roads

The approach to the design of minor gravel roads is as follows:


ƒ The design chart (Table B.6.5) is based on the AADT (not the cumulative esas) of the road and
assumes the traffic includes approximately 30% of vehicles of classes 3 and above (as defined in
Table B.3.1);
ƒ The subgrade materials need not necessarily comply with the requirements of a low volume
sealed road;
ƒ A nominal wearing course thickness of 150 mm of G15 is assumed for all road classes and sub-
grade conditions with the sub-base thickness being influenced by the sub-grade class;

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 43

ƒ Drainage, but not necessarily geometry, is upgraded to acceptable minimum levels during
construction. As for Class DC3 and DC4 roads, this can be achieved by building up the formation
to an appropriate height to achieve the hmin requirements given in Table B.6.2.
ƒ The recommended sub-base thicknesses and wearing course material strengths for different sub-
grade and traffic conditions are shown in Table B.6.5.

Table B.6.5: Design Chart for minor gravel roads

Traffic Classes (AADT)


Subgrade Strength
Class CBR (%) DC1/DC2(1)
(< 75)
150 WC
S2 (3-4)
200 G15 (2)
S3 & S4 (5-14) 150 WC
S5 (15-29) Earth Road
Notes:
1. If more than 10 heavy vehicles per day, design as a major gravel road
2. If a G30 material is available the thickness can be reduced to 150 mm

6.4 Surfacing options and design standards for paved roads


The types of surfacing options and a rational procedure for selecting appropriate surface options is
contained in Part D, Chapter 7, along with the advantages and potential concerns regarding each option.
Some surface options are not appropriate for the higher traffic categories and are marked accordingly.

6.4.1 Bituminous surfaced roads

The design standards for paved roads with a bituminous surface assume a flexible pavement with a
granular base/sub-base. Table B.5.5 shows the material types for the various structural layers used in the
catalogues. For sub-bases, G30 and G25 materials are both suitable but G30 is preferred.

The design charts for roads with bituminous road surfaces are shown in Tables B.6.6 and B.6.7. The use
of the charts is described as follows.

Climatic zones N < 4


(a) Where the total sealed surface is 8 m or less, use Pavement Design Chart 1 (Table B.6.6). No road
base materials adjustments are allowed.
(b) Where the total sealed surface is 8 m or more, use Pavement Design Chart 2 (Table B.6.7). The limit
on the plasticity modulus of the road base may be increased by 20 per cent.
(c) (Where the total sealed surface is less than 8m but the pavement is on an embankment in excess of
1.2 m in height, use Pavement Design Chart 2 (Table B.6.7). The limit on the plasticity modulus of
the road base may be increased by 20 per cent.
(d) If the engineer deems that other risk factors (eg poor maintenance and/or construction quality) are
too high, then Pavement Design Chart 1 should be used.

Climatic zones N > 4


Use Pavement Design Chart 2 (Table B.6.7).
(a) Where the total sealed surface is less than 8 metres, the limit on the plasticity modulus of the road
base may be increased by 40%.
(b) Where the total sealed surface is over 8 metres and when the pavement is on an embankment in
excess of 1.2 metres in height, the plasticity modulus of the road base may be increased by up to
40% and the plasticity index by 3 units.

Once the quality of the available materials and haul distances are known, the flow chart shown in Figure
D.6.22 of Part D and the design charts can be used to review the most economical cross-section and

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 44

pavement; this involves assessment of design traffic class, design period, cross-section and other
environmental and design considerations.

When the project is located close to the border between the two climatic zones, the lower N-value should
be used to reduce risks.

When the design is close to the borderline between two traffic design classes, and in the absence of more
reliable data, the next highest design class should be used.

It may be more economical to use a wider cross-section in the seasonal tropical and wet climate zone and
then use Pavement Design Chart 2 rather than to design a narrow cross-section and a pavement using
Pavement Design Chart 1.

The design charts do not cater for weak subgrades (CBR < 3%) and other problem soils. Design guidance
for these conditions is given in Part D, Section 6.19.2.

Table B.6.6: Bituminous Pavement Design Chart 1

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
150 G65 150 G65 150 G65 175 G80 200 G80
S2 (3-4%) 150 G15 125 G30 150 G30 175 G30 175G30
130 G15 175 G15 175 G15 200 G15
125 G65 150 G65 150 G65 175 G65 200 G80
S3 (5-7%) 150 G15 100 G30 150 G30 150 G30 150 G30
100 G15 150 G15 150 G15 150 G15
175 G45 150 G65 150 G65 175 G65 200 G80
S4 (8-14%)
120G30 200 G30 200 G30 200 G30
175 G45 125 G65 175 G65 175 G65 175 G80
S5 (15-29%)
125 G30 150 G30 150 G30 150 G30
S6 (>30%) 150 G45 150G65 175 G65 175 G65 200 G80

Table B.6.7: Bituminous Pavement Design Chart 2

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
150 G45 150 G65 150 G80 175 G80 200 G80
S2 (3-4%) 150 G15 120 G30 150 G30 150 G30 175 G30
120 G15 150 G15 150 G15 175 G15
125 G45 150 G55 175 G65 200 G65 200 G65
S3 (5-7%)
125 G15 150 G30 175 G30 200 G30 250 G30
150 G45 150 G45 150 G55 175 G55 175 G65
S4 (8-14%)
100 G30 150 G30 175 G30 200 G30
S5 (15-29%) 150 G45 175 G55 175 G55 175 G55 175 G65
S6 (>30%) 150 G45 150 G45 150 G55 150 G55 175 G65

6.4.2 Non bituminous surfaced roads

Table B.6.8 lists the non-bituminous pavement (NBP) options with their respective design charts.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 45

Table B.6.8: Non-bituminous pavement surfacing options

NBP Option Code Ref. Table


Water-bound and Dry-bound Macadams WBM and DBM B.6.10
Hand-Packed Stone HPS B.6.11
Stone Setts or Pavé SSP and MSSP B.6.12
Cobblestone/ Dressed Stone CS, DS & MCS, MDS B.6.12
Fired Clay Brick CB, MCB B.6.12
Non reinforced Concrete NRC B.6.13
Ultra-thin Reinforced Concrete UTRCP B.6.14

In Tables B.6.10 to B.6.14, unbound gravel material is used for capping, subbase and road base. In many
cases the specifications for the strength of these materials is flexible and, depending on the materials
available, substitutions can be made. It is indicated in the Tables where substitutions are allowed and
where they are restricted. Table B.6.9 defines the allowable substitutions. Table B.6.9 is used by simply
taking the ratio of thicknesses of the material to be used and the material designated in the thickness
designs in Tables B.6.10 to B.6.14 and scaling the thickness given in the Tables appropriately. For example,
if the thickness of a G45 material is given as 150mm in the Tables and a G80 material was more readily
available the thickness required becomes:
150 x 65/80 = 122mm

Table B.6.9: Substitution of pavement layer material

Material Material CBR Required thickness


Designation (%) (mm)
G15 15 100
G30 30 90
G45 45 80
G65 65 70
G80 80 65

Water-bound and Dry-bound Macadam (WBM and DBM)


A Macadam layer consists of a stone skeleton of single sized coarse aggregate in which the voids are filled
with finer material. The stone skeleton, because it is a single size large material, contains considerable
voids which are filled by fine aggregate which is washed or ‘slushed’ into the coarse skeleton with
water. Dry-bound macadam is a similar technique to the original WBM, however instead of water and
deadweight compaction being used in the consolidation of fine material, a small vibrating roller is used.
WBM or DBM are commonly used as layers within a sealed flexible pavement, but in the appropriate
circumstances may be used as an unsealed option with a suitably cohesive material being used as the
fines component. The WBM or DBM may be constructed as a low cost, initial surface to be later sealed
and upgraded in a ‘stage construction’ strategy.

WBM is suitable for labour based construction and should provide a relatively high quality surface layer
similar to a good quality natural gravel surface. However, like gravel, it is worn away by traffic and rainfall
and therefore requires similar maintenance.

The structural designs for WBM are similar to those required for a gravel road as shown in Table B.6.10
with the WBM itself acting as the wearing course. A capping layers and a sub-base are required as
indicated but thicknesses can be reduced if stronger material is available.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 46

Table B.6.10: Thickness designs for WBM pavements

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
150 WBM 150 WBM 150 WBM
S2 (3-4%) 150 G30 150 G30 175 G20 NA NA
150 G15 200 G15
150 WBM 150 WBM 150 WBM
S3 (5-7%) 125 G30 125 G30 150 G30 NA NA
100 G15 150 G15
150 WBM 150 WBM 150 WBM
S4 (8-14%) NA NA
100 G30 150 G30 200 G30
150 WBM 150 WBM 150 WBM
S5 (15-29%) NA NA
NOTE NOTE NOTE
150 WBM 150 WBM 150 WBM
S6 (>30%) NA NA
NOTE NOTE NOTE
Notes:
1. The capping layer of G15 material and the subbase layer of G30 material can be reduced in thickness if stronger
material is available (Table B.6.9)
2. On sub-grade > 15%, the material should be scarified and re-compacted to ensure the depth of material of in situ CBR
>15% is in agreement with the recommendations in Figure D.6.7 and Table D.6.7.

Hand-Packed Stone (HPS)


HPS paving consists of a layer of large broken stone pieces (typically 150 to 300mm thick) tightly packed
together and wedged in place with smaller stone chips rammed by hand into the joints using hammers
and steel rods. The remaining voids are filled with sand or gravel. A degree of interlock is achieved and
has been assumed in the designs shown in Table B.6.11. The structures also require a capping layer when
the subgrade is weak and a conventional sub-base of G30 material or stronger.

The HPS is normally bedded on a thin layer of sand (SBL). An edge restraint or kerb constructed, for
example, of large or mortared stones improves durability and lateral stability.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 47

Table B.6.11: Thicknesses designs for Hand Packed Stone (HPS) pavement (mm)

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
150HPS 200 HPS 200 HPS 250 HPS
50 SBL 50 SBL 50 SBL 50 SBL
S2 (3-4%) NA
175 G30 125 G30 150 G30 150 G30
150 G15 200 G15 200 G15
150 HPS 200 HPS 200 HPS 250 HPS
50 SBL 50 SBL 50 SBL 50 SBL
S3 (5-7%) NA
125 G30 200 G30 150 G30 150 G30
150 G15 150 G15
150 HPS 200 HPS 200 HPS 250 HPS
S4 (8-14%) 50 SBL 50 SBL 50 SBL 50 SBL NA
100 G30 150 G30 200 G30 200 G30
150 HPS 200 HPS 200 HPS 250 HPS
S5 (15-29%) 50 SBL 50 SBL 50 SBL 50 SBL NA
NOTE NOTE NOTE NOTE
150 HPS 200 HPS 200 HPS 250 HPS
S6 (>30%) 50 SBL 50 SBL 50 SBL 50 SBL NA
NOTE NOTE NOTE NOTE
Notes:
1. The capping layer of G15 material and the subbase layer of G30 material can be reduced in thickness if stronger
material is available (Table B.6.9)
2. On sub-grades > 15%, the material should be scarified and re-compacted to ensure the depth of material of in situ CBR
>15% is in agreement with the recommendations in Table B.5.3

Stone Sett or Pavé Pavements (SSP or MSSP).


Stone sett surfacing or Pavé consists of a layer of roughly cubic (100mm) stone setts laid on a bed of
sand or fine aggregate within mortared stone or concrete edge restraints. The individual stones should
have at least one face that is fairly smooth to be the upper or surface face when placed. Each stone sett
is adjusted with a small (mason’s) hammer and then tapped into position to the level of the surrounding
stones. Sand or fine aggregate is brushed into the spaces between the stones and the layer is then
compacted with a roller. Suitable structural designs are shown in Table B.6.12.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 48

Table B.6.12: Thicknesses designs for various discrete element surfacings (mm)

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
100 SSP 100 SSP 100 SSP 100 SSP 100 SSP
25 SBL 25 SBL 25 SBL 25 SBL 25 SBL
S2 (3-4%) 100 G65 150 G80 150 G80 150 G80 150 G80
100 G30 150 G30 150 G30 175 G30 200 G30
100 G15 175 G15 175 G15 200 G15 200 G15
100 SSP 100 SSP 100 SSP 100 SSP 100 SSP
25 SBL 25 SBL 25 SBL 25 SBL 25 SBL
S3 (5-7%) 125 G65 125 G80 125 G80 150 G80 150 G80
100 G30 125 G30 125 G30 150 G30 175 G30
150 G15 150 G15 150 G15 175 G15
100 SSP 100 SSP 100 SSP 100 SSP 100 SSP
25 SBL 25 SBL 25 SBL 25 SBL 25 SBL
S4 (8-14%)
150 G65 150 G80 150 G80 150 G80 175 G80
150 G30 200 G30 200 G30 225 G30
100 SSP 100 SSP 100 SSP 100 SSP 100 SSP
25 SBL 25 SBL 25 SBL 25 SBL 25 SBL
S5 (15-29%)
125 G65 125 G80 150 G80 150 G80 150 G80
125 G30 125 G30 125 G30 150 G30
100 SSP 100 SSP 100 SSP 100 SSP 100 SSP
25 SBL 25 SBL 25 SBL 25 SBL 25 SBL
S6 (>30%)
125 G65 150 G80 150 G80 150 GBO 175 G80
NOTE NOTE NOTE NOTE NOTE
Notes:
1. The capping layer of G15 material and the subbase layer of G30 material can be reduced in thickness if stronger
material is available (Table B.6.9)
2. The capping layer can be G10 provided it is laid 7% thicker
3. The road base layers (G65 and G80) must not be weaker
4. The subbase layers can be material stronger than G30 and laid to reduced thickness as shown in Table B.6.9
5. On sub-grades > 15%, the material should be scarified and re-compacted to ensure the depth of material of in situ CBR
>15% is in agreement with the recommendations in Table B.5.3.

Cobblestone or Dressed Stone Pavement (CS, DS, MCS or MDS)


Cobble or Dressed Stone surfacing consists of a layer of roughly rectangular dressed stone laid on a bed
of sand or fine aggregate within mortared stone or concrete edge restraints. The individual stones should
have at least one face that is fairly smooth, to be the upper or surface face when placed. Each stone is
adjusted with a small (mason’s) hammer and then tapped into position to the level of the surrounding
stones. Sand or fine aggregates is brushed into the spaces between the stones and the layer then
compacted with a roller. Cobble stones are generally 150 mm thick and dressed stones generally 150-
200mm thick. These options are suited to homogeneous rock types that have inherent orthogonal stress
patterns (such as granite) that allow for easy break of the fresh rock into the required shapes by labour
based means.

The thickness designs are given in Table B.6.12 except that the thickness of the cobblestone is generally
150mm instead of 100mm shown in the Table.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 49

Fired Clay Brick Pavement


Fired Clay Bricks are the product of firing molded blocks of silty clay. The surfacing consists of a layer of
edge-on engineering quality bricks within mortar bedded and jointed edge restraints, or kerbs, on each
side of the pavement. The thickness designs are as shown in Table B.6.12 for LV1 and LV2. Fired clay brick
surfacings are not suitable for traffic classes above LV2.

Mortared options
In some circumstances (eg on slopes in high rainfall areas and volume susceptible sub-grade) it may be
advantageous to use mortared options. This can be done with Hand-packed Stone, Stone Setts (or Pavé),
Cobblestone (or Dressed Stone), and Fired Clay Brick pavements. The construction procedure is largely
the same as for the un-mortared options except that cement mortar is used instead of sand for bedding
and joint filling. The behaviour of mortared pavements is different to that of sand-bedded pavements
and is more analogous to a rigid pavement than a flexible one. There is, however, little formal guidance
on mortared option, although empirical evidence indicates that inter-block cracking may occur. For this
reason the option is currently only recommended for the lightest traffic divisions up to LV2 (Tables B.6.12)
until further locally relevant evidence is available.

Non Reinforced Concrete (NRC)


The non-reinforced cement concrete option for LVRs involves casting slabs of 4.0 to 5.0 metres in length
between formwork with load transfer dowels between them. In some cases, where continuity of traffic
demands it, these slabs may be half carriageway width.

Table B.6.13: Thicknesses (mm) - Non-Reinforced Concrete Pavement (NRC)

LV1 LV2 LV3 LV4 LV5


SG CBR
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0
160 NRC 170 NRC 175 NRC 180 NRC 190 NRC
S2 (3-4%)
150 G30 150 G30 150 G30 150 G30 150 G30
150 NRC 160 NRC 165 NRC 170 NRC 180 NRC
S3 (5-7%)
125 G30 125 G30 125 G30 125 G30 125 G30
150 NRC 150 NRC 160 NRC 170 NRC 180 NRC
S4 (8-14%)
100 G30 100 G30 100 G30 100 G30 100 G30
150 NRC 150 NRC 160 NRC 170 NRC 180 NRC
S5 (15-29%)
100 G30 100 G30 100 G30 100 G30 100 G30
S6 (>30%) 150 NRC 150 NRC 160 NRC 170 NRC 180 NRC
Notes:
1. Cube strength = 30 MPa at 28 days
2. On sub-grades > 30%, the material should be scarified and re-compacted to ensure the depth of material of in situ CBR
>30% is in agreement with the recommendations in Table B.5.3

Ultra-thin Reinforced Concrete Pavement (UTRCP)


An Ultra-thin Reinforced Concrete Pavement (UTRCP) option has been developed in South Africa for a
low maintenance surfacing suitable for LVRs. A thin (50mm) layer of reinforced concrete is used as a rigid
structural surfacing over a good sub-base layer comprising well compacted good quality material, the
top 150mm of which should have an effective CBR of 80%. The pavement layers below a UTRCP slab
must contribute significantly to the strength of the pavement as a whole.

It should be emphasised that the formal design approach for this option is still under development and
that its use within an Ethiopian LVR road environment should be undertaken with caution.
Areas where the use of UTRCP can be considered include:
ƒ Surfacing of a new road or the rehabilitation/upgrading of an existing road;
ƒ All traffic and road classes from low-volume urban streets to inlays, to “provincial” roads where
typical traffic volumes are below 2 000 vehicles per day with less than 5% heavy vehicles (at this
stage);

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 6 - 50

ƒ Areas of steep grades and stop/start heavy traffic;


ƒ Areas where regular maintenance is unlikely.

The concrete is only 50mm thick and therefore tolerances are critical. The success of the UTRCP process is
therefore dependent on attention to detail. This applies not only to the concrete layer (concrete strength,
thickness, placing, curing) but also to the placing, supporting and joining of the steel mesh panels, as
well as the tolerances of the layer supporting the UTRCP. The need for meticulous monitoring and control
during construction cannot be over-emphasised. Competent site staff must be intensively involved in all
the processes associated with and control of all the construction activities.

Table B.6.14: Ultra-Thin Reinforced Concrete Pavement (UTRCP) Design

Traffic(2)
SG CBR%
Low Medium High
(1)
50 RC 50 RC 50 RC
S2 (3-4%) 150 G80 150 G80 150 G80
200 G30 250 G30 350 G30
50 RC 50 RC 50 RC
S3 (5-7%) 150 G80 150 G80 150 G80
125 G30 150 G30 200 G30
50 RC 50 RC
50 RC
S4 (8-14%) 150 G80 150 G80
150 G80
100G30 150 G30
50 RC 50 RC 50 RC
S5 (15-29%)
100 G80 125 G80 150 G80
50 RC 50 RC 50 RC
S6 (>30%)
75 G80 100 G80 100 G80
Notes:
1. Concrete must have a 28-day cube strength of 30MPa
2. The currently suggested traffic divisions are
L A 30kN wheel load division suggested for urban streets,
M A 40kN wheel load division for bus routes, and
H A 60kN wheel load division for provincial roads carrying up to 2000 vpd (10% heavy)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 51

7. DRAINAGE AND EROSION CONTROL

The maximum water flow in a watercourse can be estimated using the following methods:
ƒ Direct observation of the size of watercourse, erosion and debris on the banks, history and local
knowledge;
ƒ The Rational Method.
ƒ The SCS method (USA Soils Conservation Services, TR-55)

A combination of these methods should be used to provide the maximum level of reliability.

7.1 Size of watercourse


Watercourses enlarge to a size sufficient to accommodate the maximum water flow. The cross-sectional
area of the water course is measured and a cross-sectional area of apertures of the structure provided that
is equal to that of the water course. If the return period of the maximum flow is much longer than that for
which the structure is being designed, the typical high water level can be estimated from lateral erosion
on the banks or debris caught in the branches of trees. The cross-sectional area of the water course to
this level is calculated and a structure provided with cross-sectional area of apertures equal to this area.
Future high water levels can also be estimated from recorded history, including measurements taken in
the watercourse or from the recollections of local residents.

7.2 The Rational Method


The flow of water in a channel, q, is calculated from the following equation.

q = 0.278 x C x I x A (m3/s) Equation B.7.1

Where:
C = the catchment runoff coefficient
I = the intensity of the rainfall (mm/hour)
A = the area of the catchment (km2)

7.2.1 Catchment runoff coefficient, C

C is obtained from Table B.7.1 and Table B.7.2.

Table B.7.1: Runoff coefficient: Humid catchment

Soil Permeability
Average
Ground Slope Very low Low Medium High
(rock & hard clay) (clay loam) (sandy loam) (sand & gravel)
Flat 0-1% 0.55 0.40 0.20 0.05
Gentle 1-4% 0.75 0.55 0.35 0.20
Rolling 4-10% 0.85 0.65 0.45 0.30
Steep >10% 0.95 0.75 0.55 0.40

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 52

Table B.7.2: Runoff coefficient: Semi-arid catchment

Soil Permeability
Average
Ground Slope Very low Low Medium High
(rock & hard clay) (clay loam) (sandy loam) (sand & gravel)
Flat 0-1% 0.75 0.40 0.05 0.05
Gentle 1-4% 0.85 0.55 0.20 0.05
Rolling 4-10% 0.95 0.70 0.30 0.05
Steep >10% 1.00 0.80 0.50 0.10

7.2.2 Rainfall intensity, I (mm/hour)

The intensity of rainfall (I) is obtained from the Intensity-Duration-Frequency charts in Annex A. The storm
duration is equal to the Time of Concentration (Tc). Tc is the time taken for water to flow from the farthest
extremity of the catchment to the crossing site.

Tc = Distance from farthest extremity (m) / Velocity of flow (m/s) Equation B.7.2

The velocity of flow depends on the catchment characteristics and slope of the watercourse. It is
estimated from Figure B.7.1.

The storm design return period is taken from Table B.7.3. If the route is of strategic importance, or if
the alternative route in the event of a drainage failure is more than an additional 75km or if there is no
alternative route, Table B.7.4 should be used.

Figure B.7.1: Velocity of flow

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 53

Table B.7.3: Storm design return period (years)

Geometric design standard


Structure type
DC4 DC3 DC2 DC1
Gutters and inlets 2 2 2 1
Side ditches 10 5 5 2
Ford 10 5 5 2
Drift 10 5 5 2
Culvert diameter <2m 15 10 10 5
Large culvert diameter >2m 25 15 10 5
Gabion abutment bridge 25 20 15 -
Short span bridge (<10m) 25 25 15 -
Masonry arch bridge 50 25 25
Medium span bridge (15 – 50m) 50 50 25 -
Long span bridge >50m 100 100 50 -

Table B.7.4: Storm design return period (years) for severe risk situations

Geometric design standard


Structure type
DC4 DC3 DC2 DC1
Gutters and inlets 5 5 5 2
Side ditches 15 10 10 5
Ford 15 10 10 5
Drift 15 10 10 5
Culvert diameter <2m 25 20 20 10
Large culvert diameter >2m 50 25 20 10
Gabion abutment bridge 50 25 20 -
Short span bridge (<10m) 50 50 25 -
Masonry arch bridge 50 50 25
Medium span bridge (10 – 50 m) 100 100 50 -
Long span bridge >50m 100 100 100 -

7.2.3 Catchment area, A (km2)

The area of the drainage catchment should be estimated from a map or an aerial photograph.
In the Rational Method it is assumed that the intensity of the rainfall is the same over the entire catchment
area. The consequence of applying the method to large catchments is an over-estimate of the flow and
therefore a conservative design.

A simple modification can be made to take into account the spatial variation of rainfall intensity across a
larger catchment. The effective area of the catchment is reduced by multiplying by the areal reduction
factor (ARL) given by the following equation:

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 54

ARL = 1 - 0.04 x t-1/3 x A1/2 Equation B.7.4

Where,
t = storm duration in hours
A = catchment area in km2

7.3 The SCS method


The SCS method for calculating rates of runoff requires much of the same basic data as the Rational
Method namely catchment area, a runoff factor, time of concentration, and rainfall. However the SCS
method also considers the time distribution of the rainfall, the initial rainfall losses to interception and
storage, and an infiltration rate that decreases during the course of a storm. It is therefore potentially
more accurate than the Rational Method and is applicable when the catchment area is larger than 50
hectares.

7.3.1 Catchment area

The catchment area is determined from topographic maps and field surveys. For large catchment areas
it might be necessary to divide the area into sub-catchment areas to account for major land use changes.

7.3.2 Rainfall

The SCS method is based on a 24-hour storm event. The characteristics of storms are defined in terms
of the relationship between the percentage of the total storm rainfall that has fallen as a function of
time. Three basic types of storm are defined for three levels of maximum intensity, Type I being the least
intense and Type III being the most intense. In Ethiopia a Type II distribution is used (see ERA’s Drainage
Design Manual – 2002 or the revised version when available).

A relationship between accumulated rainfall and accumulated runoff was derived by SCS for numerous
hydrologic and vegetative cover conditions. The storm data included total amount of rainfall in a calendar
day but not its distribution with respect to time. The SCS runoff equation is therefore a method of
estimating direct runoff from 24-hour or 1-day storm rainfall.

The equation is:


Q = (P - Ia)2 / [(P - Ia) + S] Equation B.7.5
Where:
Q = accumulated direct runoff, mm.
P = accumulated rainfall (ie, the potential maximum runoff), mm
Ia = initial abstraction including surface storage, interception, and infiltration prior
to runoff, mm.
S = potential maximum retention, mm.

S is related to the soil and cover conditions of the catchment area through the Curve Numbers, CN,
described below.
S = 25.4(1000/CN – 10) Equation B.7.6
The relationship between Ia and S was found to be;
Ia = 0.2S = 50.8.(100/CN-1) Equation B.7.7
Substituting into Equation B.7.5,
Q = [P – 50.8(100/CN - 1)]2/[P + 203.2(100/CN - 1)] Equation B.7.8

Figure B.7.2 shows a graphical solution which enables Q, the direct runoff from a storm, to be obtained
if the total rainfall and catchment area curve number are known.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 55

Figure B.7.2: Relationship between Precipitation, Direct Runoff and CN

7.3.3 Runoff and Curve Numbers

The physical catchment area characteristics affecting the relationship between rainfall and runoff (ie the
CN values) are land use, land treatment, soil types, and land slope.

Land use is the catchment area cover and it includes agricultural characteristics, type of vegetation,
water surfaces, roads and roofs. Land treatment applies mainly to agricultural land use, and it includes
mechanical practices such as contouring or terracing and management practices such as rotation of
crops. The SCS method uses a combination of soil conditions and land-use to assign a runoff factor to
an area. These runoff factors or curve numbers (CN), indicate the runoff potential of an area. The higher
the CN, the higher is the runoff potential.

Soils are divided soils into four hydrologic groups (Groups A, B, C, and D) based on infiltration rates
(Table B.7.5). These groups are described in detail in the ERA Drainage Design Manual.

Table B.7.5: Hydrological characteristic soil groups

Soil Group General Description


A Well drained, sandy High infiltration, low runoff
B Sandy loam, low plasticity
C Clay loam, medium plasticity
D Highly plastic clay Low infiltration, high run off

Runoff curve numbers also vary with the antecedent soil moisture conditions, defined as the amount of
rainfall occurring in a selected period preceding a given storm. In general, the greater the antecedent
rainfall, the more direct runoff there is from a given storm. A five-day period is used as the minimum for
estimating antecedent moisture conditions.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 56

Table B.7.6 gives runoff curve numbers for various land uses. (NB: More comprehensive tables are given
in the ERA Drainage Design Manual). This Table is based on an average antecedent moisture condition
(ie soils that are neither very wet nor very dry when the design storm begins). Table B.7.7 gives conversion
factors to convert average curve numbers to wet and dry curve numbers. The recommended antecedent
moisture conditions (AMC) in Ethiopia are shown in Table B.7.8.

Table B.7.6: Runoff Curve Numbers (CN)

Land use A B C D
Without conservation treatment 72 81 88 91
Cultivated land
With conservation treatment 62 71 78 81
Pasture land Poor condition 68 79 86 89
Good condition 39 61 74 80
Meadow 30 58 71 78
Thin stand, poor cover, no mulch 45 66 77 83
Wood or forest
Good cover 25 55 70 77
Good condition, grass cover >75% of
Open spaces, lawns, 39 61 74 80
area
parks
Fair condition, grass on 50-75% 49 69 79 84
Commercial and business areas, 85%
89 92 94 95
Urban districts impervious
Industrial districts, 70% impervious 81 88 91 93
Average lot size Average % impervious
< 0.05 hectares 65 77 85 90
0.1 hectares 38 61 75 83
Residential
0.2 hectares 25 54 70 80
0.4 hectares 20 51 68 79
0.8 hectares 12 46 65 77
Paved roads with curbs and storm drains, paved parking areas, roofs. 98 98 98 98
Gravel roads 76 85 89 91
Earth roads 72 82 87 89
Open water 0 0 0 0

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 57

Table B.7.7: Conversion from average to wet and dry antecedent moisture conditions

CN values
Average conditions Dry Wet
95 87 98
90 78 96
85 70 94
80 63 91
75 57 88
70 51 85
65 45 82
60 40 78
55 35 74
50 31 70
45 26 65
40 22 60
35 18 55
30 15 50

Table B.7.8: Antecedent moisture conditions

Region(1) Antecedent moisture conditions


D Dry
B Wet
All other regions Average
Bahir Dar area Although in region A, use Wet
Notes:
1. See Appendix B.1 for regional map

7.3.4 Time of concentration

The next step in the SCS Method is to determine the Time of Concentration. This is the time it takes water
to flow from the edge of the catchment area to the point of interest. It is a combination of three values;
A sheet flow,
B shallow concentrated flow, and
C open channel flow.
The type that occurs is a function of the conveyance system and is determined by field inspection. It is
often a combination of these so that the total travel time is the sum of the time taken for the water to pass
through all of the segments of the catchment.
Travel time is the ratio of flow length to flow velocity:
T = L/(3600V) Equation B.7.9

Where:
T = travel time, hr
L = flow length, m
V = average velocity, m/s
3600 = conversion factor from seconds to hours.

Sheet flow
Sheet flow is flow over plane surfaces. It usually occurs in the headwater of streams. With sheet flow,
the friction value (Manning’s n) is an effective roughness coefficient that includes the effect of raindrop

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 58

impact; drag over the plane surface; obstacles such as litter, crop ridges, and rocks; and erosion and
transportation of sediment. These n values are for very shallow flow depths of about 0.03m or so. Table
B.7.9 gives Manning’s n values for sheet flow for various surface conditions.

Table B.7.9: Manning’s roughness coefficients for sheet flow

Surface n1
Smooth surfaces: concrete, asphalt, gravel or bare soil 0.011
Fallow (no residue) 0.05
Cultivated soils
Residue cover < 20% 0.06
Residue cover > 20% 0.17
Grasses
Short grass 0.15
Dense grass 0.24
Range 0.13
Woods(1)
Light underbrush 0.4
Dense underbrush 0.8
Note:
1. Consider cover to a height of 30mm. This is the only part of the cover that will affect sheet flow

For sheet flow of less than 100 metres Manning’s kinematic solution should be used to compute the travel
time T,
T = [0.091 (n.L)0.8.8/ (P2_)0.5S0.4] Equation B.7.10
Where:
T = travel time, hr
n = Manning’s roughness coefficient (Table B.7.9)
L = flow length, m
P2 = 2-year, 24-hour rainfall, mm
S = slope of hydraulic grade line (land slope), m/m

Shallow Concentrated Flow


After a maximum of 100 metres, sheet flow usually becomes shallow concentrated flow. The average
velocity for this flow can be determined from the following equations in which average velocity is a
function of watercourse slope and type of channel.
Unpaved
V = 4.918 (S)0.5
Paved
V = 6.196 (S)0.5
Where:
V = average velocity, m/s
S = slope of hydraulic grade line (watercourse slope), m/m

After determining average velocity, the travel time for the shallow concentrated flow segment is calculated
from Equation B.7.9.

Open Channel flow


Open channels are assumed to begin where surveyed cross section information has been obtained,
where channels are visible on aerial photographs, or where blue lines (indicating streams) appear on
Ethiopian Mapping Authority topographic maps (1:50,000). Average flow velocity is usually determined
for bank-full elevation. Manning’s equation or water surface profile information can be used to estimate

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 59

average flow velocity. When the channel section and roughness coefficient (Manning’s n) are available,
then the velocity can be computed using the Manning Equation.
V = (R2/3. S1/2)/n Equation B.7.11
Where:
V = average velocity, m/s
R = hydraulic radius, m (equal to a/Pw)
a = cross sectional flow area, m2
Pw = wetted perimeter, m
S = slope of the hydraulic grade line, m/m
n = Manning’s roughness coefficient (Table B.7.10)

After the average velocity is computed, the travel time for the segment can be calculated using Equation
B.7.9.

Reservoir or Lake
Sometimes it is necessary to compute a time of concentration for a catchment area having a relatively
large body of water in the flow path. The travel time is computed using the equation:
Vw = (g.Dm)0.5 Equation B.7.12
Where:
Vw = the wave velocity across the water, m/s
g = 9.81 m/s2
Dm = mean depth of lake or reservoir, m

This equation only deals with the travel time across the lake, not the time at the inflow or outflow channels.
The times for these are generally very much longer and must be added to the travel time across the lake
(see ERAs Drainage Design Manual). Equation B.7.12 can be used for swamps with much open water, but
where the vegetation or debris is relatively thick (less than about 25% open water), Manning’s equation
is more appropriate.

7.3.5 Steps in the SCS procedure

The steps in using the SCS method are as follows:


1. Ethiopia has been divided into five regions based on rainfall characteristics (Annex A). Determine
which region is appropriate.
2. Determine the catchment area, A, and its soil and land use characteristics.
3. Determine the ‘curve runoff number, CN, from Table B.7.6 and any adjustment based on the likely
antecedent soil moisture conditions (Tables B.7.7 and B.7.8).
4. Calculate the value of Ia from equation B.7.7.
5. Choose the appropriate design storm recurrence frequency. This is based on the class of road and
the drainage structure being designed.
6. For the recurrence frequency chosen, determine the 24-hour rainfall (P) for the appropriate rainfall
region from Figure B.7.3.
7. Determine the direct runoff (Q) for the rainfall (P) and curve number (CN) obtained in steps 3 and 5
from Figure B.7.2
8. The catchment must be divided into uniform areas for the purpose of determining the Time of
Concentration, Tc. The flow lengths for sheet flow, shallow concentrated flow, and channel flow
must be determined and the relevant equations in Section B.7.3.4 used to calculate the total time of
concentration.
9. At this stage the following data have been obtained,
Ia – the initial abstraction based on the curve number CN (step 4)
P – the design storm precipitation (step 6)
Q – the accumulated direct runoff (step 7)
Tc – the Time of Concentration (step 8)
The next step is to determine the unit peak discharge, Qu and this is done using Figure B.7.4, the
value of Ia/P and the Time of Concentration.
10. The final step is to compute the actual peak discharge from the unit value as follows,
Design peak discharge = Qu x Q x A

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 60

Where Q is in mm and A is in units of 100 hectares.

Frequency Interval (years)


Region 2 5 10 25 50 100
24-hour depth in mm
A1, A4 60 79 93 113 127 142
A2, A3 52 67 79 95 107 118
B, C 65 84 98 118 132 147
D 67 89 105 127 144 161
Lake Tana 74 106 131 163 187 211

Figure B.7.3: 24 hour depth-frequency curves

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 61

Figure B.7.4: Unit peak discharge (Type II rainfall)

7.4 Design of culverts


A fall of 3 - 5% should be allowed on culverts to ensure that water flows without depositing silt and
other debris. All pipes should have a minimum diameter of 600mm to ensure that they can be manually
cleaned.

7.4.1 Nomograph method for culvert sizing

The required size of a culvert opening is estimated using the nomographs in Figure B.7.5 to Figure B.7.7.
These figures apply to culverts with inlet control where there is no restriction to the downstream flow of
the water.

In flat terrain, where there is a high risk of silting, a factor of safety of 2 should be allowed in the design
of the culvert.

7.4.2 Correlation with successful practice

If a high proportion of structures along a road or in a region have been in operation for a number of years
without overtopping, it is reasonable to assume that the relationship between catchment area, catchment
characteristics, rainfall intensity and maximum water flow used in their design is valid. The design of new
culverts can be based on simply the catchment area using the same relationships.

7.4.3 Design of drifts and fords

Drifts and fords are designed for water to flow over the running surface. It is not expected that vehicles
can use them at all times. The following criteria should be considered when designing drifts:
ƒ The level of the drift should be as close as possible to the existing river bed level.
ƒ The normal depth of water should be a maximum of 150mm and the maximum 5 year flow should
be 6m3/second on the drift to allow traffic to pass.
ƒ Approach ramps should have a maximum gradient of 10% (7% for roads with large numbers of
heavy trucks).

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 62

7.5 Components of External Drainage

7.5.1 General principles

Those parts of the natural slope drainage system that experience increased run-off as a result of road
construction should be strengthened through:
ƒ Control of road surface drainage;
ƒ Design of culverts or drifts that convey water and debris load efficiently;
ƒ Optimised frequency of drainage crossings to prevent excess concentration of flow;
ƒ Protection of drainage structures and stream channels for as far downstream as is necessary to
ensure their safety and to prevent erosion of land adjacent to the water course;
ƒ Planting of vegetation on all new slopes and poorly-vegetated areas, around the edges of drainage
structures and appropriately along stream courses.

7.5.2 Side drains

Side drains serve two main functions: collection and removal of surface water from the road and the
immediate vicinity of the road, and prevention of sub-surface water from adversely affecting the road
pavement structure. Side drains can be constructed in three forms: V-shaped, rectangular or trapezoidal.
The trapezoidal cross-section facilitates maintenance and improved traffic safety. Trapezoidal drains can
be constructed and maintained by hand. In flat terrain and reasonable soils it may be best to use wide
unlined drains with high capacity yet low flow velocity. The minimum recommended width of the side
drain is 500mm.

Design volumes of run-off in side drains and other channels are estimated using the Rational Method.
The cross sectional area of the drain must be sufficient to accommodate the expected flow of water, Q,
where:
Q = AV

Flow velocity is calculated from the Manning equation.

V = velocity in m/s
R = hydraulic depth (the area for the stream
flow divided by the wetted perimeter)
S = hydraulic gradient ( the slope of the
river bed over a reasonable distance either
side of the crossing point)
n = roughness coefficient (see TableB.7.10)
Definition of hydraulic depth

Figure B.7.5: Manning’s Formula

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 63

Figure B.7.6: Headwater depth and capacity for corrugated metal pipe
culverts with inlet control (Adapted from FHWA, 1998)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 64

Figure B.7.7: Headwater depth and capacity for concrete pipe


culverts with inlet control (Adapted from FHWA, 1998)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 65

Figure B.7.8: Headwater depth and capacity for concrete box


culverts with inlet control (Adapted from FHWA, 1998)

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 66

Table B.7.10: Roughness coefficient (n) for drains

Material in the drain Roughness coefficient


Sand, loam, fine gravel, volcanic ash 0.022
Stiff clay 0.020
Course gravel 0.025
Conglomerate, hard shale, soft rock 0.040
Hard rock 0.040
Masonry 0.025
Concrete 0.017

Side drains (as well as the road itself) should have a minimum longitudinal gradient of 0.5%, except on
crest and sag curves. Reduction of the side drain gradient in the lower reaches of a long length of drain
should be avoided in order to prevent siltation.

7.5.3 Erosion control in the side drain

Limiting values for the velocity of flow to prevent scour in excavated drains are given in Table B.7.11.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 67

Table B.7.11: Permissible flow velocities (m/sec) in excavated ditch drains

Soil type Clear water Water carrying fine silt Water carrying sand and fine gravel
Fine sand 0.45 0.75 0.45
Sandy loam 0.55 0.75 0.6
Silty loam 0.6 0.9 0.6
‘Good’ loam 0.75 1.05 0.7
Lined with
established
1.7 1.7 1.7
grass on good
soil
Lined with
bunched
grasses 1.1 1.1 1.1
(exposed soil
between plants)
Volcanic ash 0.75 1.05 0.6
Fine gravel 0.75 1.5 1.15
Stiff clay 1.15 1.5 0.9
Graded loam to
1.15 1.5 1.5
cobbles
Graded silt to
1.2 1.7 1.5
cobbles
Alluvial silts
0.6 1.05 0.6
(non colloidal)
Alluvial silts
1.15 1.50 0.9
(colloidal)
Coarse gravel 1.2 1.85 2.0
Cobbles and
1.5 1.7 2.0
shingles
Shales 1.85 1.85 1.5
Rock Negligible scour at all velocities

Drain erosion is controlled by building scour checks or lining the drain.

Scour checks reduce the speed of water and help prevent it from eroding the road structure. The scour
check acts as a small dam. When the scour check is naturally silted up on the upstream side, it effectively
reduces the gradient of the drain on that side, and therefore the velocity of the water. There must be
sufficient cross-sectional area in the drain above the scour check (ie where the water has been slowed
down) to accommodate the maximum design flow.

The distance between scour checks depends on the road gradient and the erosion potential of the soils.
Table B.7.12 shows recommended values. These should be modified for erodible soils.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 68

Table B.7.12: Spacing between scour checks

Scour check interval


Road gradient (%)
(metres)
3 Not required
4 17
5 13
6 10
7 8
8 7
9 6
10 5
12 4

When constructing a channel lining it is important to reproduce or exceed the dimensions of the original
channel. A curved shaped cross-section to the lining is preferable to a rectangular cross-section. Measures
must be taken to control erosion downstream of the drain outlet.

Dry stone pitching for drain lining is usually only suitable where the discharge is lower than 1 m/sec per
metre width, and where sediment load is relatively fine-grained.

7.5.4 Mitre drains or turnouts

Water from the side drains should be discharged as frequently as possible. If the water can be discharged
on the same side of the road as the drain, a turnout or mitre drain is used to lead the water onto adjacent
land. Low volumes of flow and low velocities should be achieved at each discharge point to minimise
erosion. Table B.7.13 shows the maximum spacing of mitre drains related to gradient.

Table B.7.13: Maximum spacing of mitre drains

Road gradient (%) Maximum mitre drain interval (m)


12 40
10 80
8 120(1)
6 150(1)
4 200(1)
2 80(2)
<2 50(2)
Notes:
1. A maximum of 100m is preferred but not essential.
2. At low gradients silting becomes a problem.

A block-off is required to ensure that water flows out of the side drain into the mitre drain. The angle
between the mitre drain and the side drain should preferably be 30 degrees, but not greater than 45
degrees.

The desirable slope of the mitre drains is 2%. The gradient should not exceed 5% otherwise there may
be erosion in the drain or on the land where the water is discharged. The drain should lead gradually
across the land, getting shallower and shallower. Stones may need to be laid at the end of the drain to
help prevent erosion.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 69

In flat terrain, a small gradient of 1% or even 0.5% may be necessary to discharge water, or to avoid very
long drains. These low gradients should only be used when absolutely necessary. The slope should be
continuous with no high or low spots.

7.5.5 Wet lands

Special drainage or construction methods are needed if wet areas must be crossed. An embankment is
normally required. The embankment should include multiple drainage pipes or coarse permeable rock-
fill to keep the flow dispersed. Sub-grade reinforcement with coarse permeable rock, filter layers and
geotextiles may also be required. The objective is to maintain the natural groundwater level and flow
pattern dispersed across the wet land and, at the same time, provide a stable, dry roadway surface.

Subsurface drainage, through use of under-drains, interceptor trenches or aggregate filter blankets, is
used in localized wet or spring areas to remove the groundwater and keep the roadway sub-grade dry.
In localised areas subsurface drainage is often more cost-effective than adding a thick structural section
to the road. In extensive swamp or wet areas subsurface drainage is often less effective than raising the
roadway platform or providing a thick aggregate layer to support the road pavement.

7.5.6 Subsoil Drains

Longitudinal subsoil drains can be used to locally lower a water table. They normally consist of porous
concrete, open jointed or perforated pipe laid in a trench and backfilled with a free draining material such
as graded crushed stone or gravel. The pipe size should not be less than 15cm internal diameter. The
trench should be at least 60cm wide and 1.5m deeper than the formation level of the road.

7.5.7 Filters

A filter is as a transitional layer of small gravel or geotextile placed between a structure, such as riprap
or gabions, and the underlying soil. Its purpose is to prevent the movement of soil behind the structure
or into under-drains. Filters allow groundwater to drain from the soil without building up pressure. A
sand or gravel filter layer is typically about 150 to 300 mm thick. In some applications, two filter layers
may be needed between fine soil and very large rock. Geotextiles are commonly used to provide filter
zones between materials of different size and gradation. The geotextile can be a woven monofilament
or a needle punched non-woven geotextile, but it must be permeable. The geotextile should have an
apparent opening size of 0.25 to 0.5mm. A 200g/m2 needle-punched non-woven geotextile is commonly
used for soil filtration and separation applications.

7.5.8 Interceptor, cut-off or catch-water drains.

These drains are constructed to prevent water flowing into vulnerable locations by ‘intercepting’, ‘cutting
off’ or ‘catching’ the water flow and diverting it to a safe point of discharge, usually a natural watercourse.
Interceptor drains above cut faces should have a gradient of 2% on their full length and should be at least
3 to 5 m from the cut face. If steeper gradients in the drain are unavoidable then scour checks should be
installed or the drain should be lined. The drain should also be lined where seepage will weaken the cut
slope. Alternatively the drain should be replaced by a vegetated earth bund.

Interceptor drains should be 60 cm wide, 40 cm (minimum) deep with sides back-sloped at 3:1 (vertical:
horizontal).

7.5.9 Chutes

Chutes are structures intended to convey a concentration of water down a slope that, without such
protection, would be subject to scour. Since flow velocities are very high, stilling basins are required
to prevent downstream erosion. The entrance of the chute needs to be designed to ensure that water
is deflected from the side drain into the chute, particularly where the road is on a steep grade. On
embankments it may be necessary to lead water to the top of chutes using kerbing.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 7 - 70

7.5.10 Slope protection

Recommended cut and fill slopes for LVRs to avoid excessive erosion are given in Table B.7.14 and Table
B.7.15.

Table B.7.14: Common cut-slope ratios for LVRs.

Soil and rock condition (Horizontal : Vertical)


Most hard rocks (without adverse jointing) 1:4 – 1:2
Closely fractured rock 1:2 – 1:1
Well consolidated highly to completely
1:2 – 1:1
weathered rock (residual soil)
Dense coarse granular soils 1:2 – 1:1
Loose coarse granular soils 1.5:1
Plastic clay soils 2:1 – 3:1
Low cuts (less than 3 m high) 2:1

Table B.7.15: Common fill slope batters for LVRs.

Soil and rock condition (Horizontal : Vertical)


Soft clays or fills on wet areas and swamps (>3m) 2:1 - 3:1
Fills of most soils (free draining granular fill >3m) 1.5: 1 - 2:1
Fills of hard, angular rock (rock fill >3m) 1.5:1 – 1.25:1
Low fills (less than 3m high) 2:1 or flatter (for vegetation)

Techniques for slope protection against erosion are:


ƒ Intercepting ditches at the top and bottom of slopes with gutters and spillways used to control the
flow of water down the slope;
ƒ Stepped or terraced slopes to reduce the height of the slope;
ƒ Riprap or rock facing material embedded in a slope face, sometimes with planted vegetation;
ƒ Topsoiling and grassing: the slope is covered with 75mm of fine topsoil and planted with a suitable
indigenous grass;
ƒ Retaining structures such as gabions and other retaining walls;
ƒ Reinforced earth, where the embankment walls build up as the earth fill is placed, within anchors
compacted into the fill material;
ƒ Shotcreting and geotextiles: techniques that are usually expensive and should therefore be
carefully considered before being used for specific applications.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 8 - 71

8. WATER CROSSINGS AND ASSOCIATED STRUCTURES

The design of water crossings and associated structures for low volume are covered in Part E Chapter 8 of
the manual. This Part of the manual includes bridges up to a span of 10m. For detailed design of bridges
with spans greater than 10m, the designer should consult the ERA Bridge Design Manual – 2011.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 9 - 72

9. ROAD FURNITURE AND SIGNAGE

The ERA Geometric Design Manual provides the requirements for road furniture and signage. The main
elements are:
ƒ Traffic signs provide essential information to drivers for their safe and efficient manoeuvring on
the road;
ƒ Road markings to delineate the pavement centre line and edges to clarify the paths that vehicles
should follow (paved roads);
ƒ Marker posts to indicate the alignment of the road ahead.

Not all of these measures are needed on low volume roads.

9.1 Traffic Signs


Traffic signs are of three general types:
ƒ Regulatory Signs: indicate legal requirements of traffic movement and are essential for all roads.
ƒ Warning Signs: indicate conditions that may be hazardous to highway users
ƒ Information Signs: convey information of use to the driver

Warning signs should be provided where there are unexpected changes in the driving conditions, for
example where:
ƒ The geometric standards for a particular class of road have been changed along a short section
of road, for example a sharp bend, a sudden narrowing of the road, or an unexpectedly steep
gradient;
ƒ A bend occurs after a long section of straight road;
ƒ There is an unexpected school crossing;
ƒ A drift or other structure is not clearly visible from a safe distance;
ƒ The driver is approaching traffic calming measures such as speed humps.

Hazard warnings that are done by means of road markings on paved roads must be done by means of
traffic signs on unpaved roads.

Information signs are less important on lower classes of road frequented primarily by local people.

9.2 Road Markings


Pavement markings consist primarily of centre lines, lane lines, no overtaking lines and edge lines. Not all
of these are necessary on low volume roads, but on a paved two lane road a centre line is recommended.

9.3 Marker Posts


Marker posts have the function of controlling traffic to encourage safe operation of the road. There are
two types of marker posts: guideposts and kilometre posts.

Guideposts are intended to make drivers aware of potential hazards such as abrupt changes in shoulder
width, abrupt changes in the alignment, approaches to structures etc. For changes in shoulder width
and approaches to structures, guide posts should be placed at 50m intervals.

The required spacing of guideposts on curves is shown in Table B.9.1.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


B - Chapter 9 - 73

Table B.9.1: Spacing of Guide Posts at Curves

Curve Radius (m) Guide Post Spacing (m)


500 35
200 20
100 12
50 8
30 5

Kilometre posts are a requirement for all trunk and link roads and are therefore only likely to be needed
on some roads of class DC4. Details are given in the section on “Road Furniture and Markings” in the
Geometric Design Manual-2011.

9.4 Safety barriers


Safety barriers are expensive and seldom justified on low volume roads. The geometric design of such
roads should be done to eliminate the need for such barriers but sometimes they might be required in
highly dangerous situations, for example, on some bends on an escarpment road that cannot be made
safe by other means.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


Appendix B.1 - 74

APPENDIX B.1 - RAINFALL INTENSITY-


DURATION-FREQUENCY CHARTS

Rainfall Regions
Note:
Rainfall data used in the preparation of this figure was collected from Ministry of Water Resources
meteorology stations and analysed during the preparation of the 2002 ERA Drainage Design Manual.
The information is subject to review, and future data may indicate the need for a further refinement in
both values and regional boundaries.

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


Appendix B.1 - 75

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


Appendix B.1 - 76

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


Appendix B.1 - 77

PART B: DESIGN STANDARDS FOR LOW VOLUME ROADS


)('(5$/'(02&5$7,&5(38%/,&2)
('(5$/'(02&5$7,&5(38%/,&2)

(7+,23,$152$'6$87+25,7<

,7 <
( 7+

25
,2

,$
15 87
+
3

2$'6 $

'(6,*10$18$/)25/2:92/80(52$'6
3$57$3$57%$1'3$57&
),1$/'5$)7$35,/
Part C
COMPLEMENTARY INTERVENTIONS

Context and application of complementary


interventions

Planning, identification and implementation of


complementary interventions

Employment and human resource issues

Contract provisions to support complementary


interventions
Part C
Complementary
Interventions
Supporting small scale contractors

Supervision consultant’s contract

Supervision consultant’s contract

Management, monitoring and enforcement


C - iii

C TABLE OF CONTENTS

C. TABLE OF CONTENTS ......................................................................................................C.III


C. LIST OF TABLES ............................................................................................................... C.IV
C. LIST OF PLATES ................................................................................................................ C.V
1. CONTEXT AND APPLICATION OF COMPLEMENTARY INTERVENTIONS .........................C.1
2. PLANNING, IDENTIFICATION AND IMPLEMENTATION OF COMPLEMENTARY
INTERVENTIONS ...............................................................................................................C.3
2.1 Planning ...........................................................................................................................C.3
2.2 Identification and selection of complementary interventions ........................................C.5
3. EMPLOYMENT AND HUMAN RESOURCE ISSUES ............................................................C.7
3.1 Encouraging participation of marginalised groups .........................................................C.8
3.2 Meeting existing obligations ...........................................................................................C.9
4. CONTRACT PROVISIONS TO SUPPORT COMPLEMENTARY INTERVENTIONS ...............C.10
4.1 Design services contract................................................................................................C.10
4.2 Works contract...............................................................................................................C.11
4.2.1 Options for inclusion of complementary interventions in
works contracts – measurement and payment ...............................................C.12
4.3 The works contract evaluation .......................................................................................C.15
5. SUPPORTING SMALL SCALE CONTRACTORS.................................................................C.16
5.1 Utilising the lower level bidding documents: ................................................................C.16
5.2 Utilising sub-contracting clauses ...................................................................................C.16
6. SUPERVISION CONSULTANT’S CONTRACT ....................................................................C.17
7. MANAGEMENT, MONITORING AND ENFORCEMENT ...................................................C.18
8. REFERENCES ...................................................................................................................C.19
C. APPENDIX C.1 - COMPLEMENTARY INTERVENTIONS – PLANNING,
FEASIBILITY AND PRELIMINARY DESIGN STAGES..........................................................C.20
C. APPENDIX C.2 - INDICATIVE EXAMPLES OF CIS BY CATEGORY AND THEME ..............C.23

PART C: COMPLEMENTARY INTERVENTIONS


C - iv

C LIST OF TABLES

Table C.4.1: Works bidding/contract documents............................................................................ C.12


Table C.4.2: Example Bill of Quantities entries ............................................................................... C.13
Table C.4.3: Example of Provisional Sum entries ............................................................................ C.13
Table C.4.4: Lump Sum item for completed output related to RS .................................................. C.14
Table C.4.5: Example of schedule of rates where quantities are unknown ..................................... C.14
Table C.4.6: Example of Schedule of rates where quantities are known ........................................ C.15

PART C: COMPLEMENTARY INTERVENTIONS


C-v

C LIST OF PLATES

Plate C.1.1: Community participation meetings............................................................................... C.3


Plate C.3.1: Promotion of labour-intensive methods ........................................................................ C.7
Plate C.3.2: Promotion of Intermediate Technology option ............................................................. C.7
Plate C.3.3: Participation of women in road works ........................................................................... C.9

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 1 - 1

1. CONTEXT AND APPLICATION OF COMPLEMENTARY INTERVENTIONS

Part C describes some concepts and practical issues relating to the planning, design and implementation
of potential complementary interventions on low volume road projects.

Complementary interventions are:


ƒ Actions that can, if desired, be included in and implemented through the roads project or the road
works contract;
ƒ Targeted toward the communities that lie within the influence corridor of the road and are affected
by the road itself, by road users or by the road works;
ƒ Intended to optimise the benefits brought by the road and to extend positive, and mitigate
negative, impacts of road projects on local communities;
ƒ Not mandatory. They are included at the discretion of the client;
ƒ Not designed to remove or replace the responsibilities of the contractor, client or other authorities
or institutions.

The concept of complementary interventions is motivated by opportunity. In its simplest terms


complementary interventions take advantage of the presence of the road project to build in aspects that
will enhance the social, environmental and safety situation of communities affected by the road. These
are additional to the normal social, environmental and safety obligations of the contractor and do not
replace or share the contractors normal obligations.

The opportunity to take advantage of complementary interventions will only be available during the
period of the works and when the contractor is on site. As such, any initiative or series of actions planned
under a complementary interventions umbrella is short term.

As mentioned, complementary interventions are not mandatory. Any action will have a cost consequence,
which will be borne by the client, unless other arrangements are made. The extent to which the client
wishes to include complementary interventions within a project would be communicated to the design
engineer through his terms of reference or by site instruction. In some cases, forward thinking contractors
may wish to absorb all or part of such costs as a good will gesture. This can be reflected in the bill of
quantities.

Complementary interventions are opportunistic and reflect the presence of a road contractor in the
project area for a certain period of time. The contractor will have access to physical, human and financial
resources that may not normally be readily available in the project area.

The presence of the contractor and road contract provides a rare opportunity to widen the positive
impacts of the road project for little additional cost or effort. A contractor will mobilise skills, materials,
plant, labour and supplies and will have access to substantial transport mechanisms. Without major
disruption to the contractor’s programme of works, it may be possible to make some of these resources
available to the communities affected by the road project.

Existing Federal, Regional, Wereda and Kebele structures and administrations are central to the successful
application of the complementary intervention approach. If complementary interventions are to be
included in the works, the relevant administrations must have been communicated with at the feasibility
stage of the project and have agreed in principle the extent, type and approach for any inclusion of
actions and initiatives (Appendix C.1). Complementary interventions are demand driven, reflect the needs
expressed by local communities themselves, and are agreed through normal institutional structures and
plans. This may also require the beneficiary community adding some additional effort and resources.

With an outline framework, agreed at the feasibility stage, the task of the design engineer is to materialise
these desires and agreements into the works bidding documents and eventual contract.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 1 - 2

Part C aims to explain the concept of complementary interventions, provide examples of the types of
intervention that may be provided, and explain how the detailed design and bidding documents should
be used to clearly define the requirements for implementation, measurement, and monitoring.

Complementary interventions can be grouped into three categories:


Cat.I) Management Interventions – simple actions that enhance the road project itself and are well
within the normal skills of the road contractor. These aim to improve the wider impacts of the
project itself and build on or extend the normal socio-environmental and safety obligations of
the contractor.
Cat.II) Opportunity Interventions – actions that are beyond the scope of traditional road projects but
are within the technical and management skills of the road works contractor.
Cat.III) Enhancement Interventions – actions that utilise the provisions of the contract but extend
beyond the normal skills and experience of road works contractor. These actions would normally
be implemented by other parties with the relevant skills.

One of the main advantages for including complementary interventions through a contractor already
mobilised for a road project is that they can be completed far more quickly, efficiently and at a lower
cost than if implemented separately. By minimising the transaction costs, more can be achieved for less.
Not only does it bring more benefits to the local community, it improves the economic rate of return on
the road investment and, depending on the type of actions, can improve the prospect for local socio-
economic growth and empowerment.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 2 - 3

2. PLANNING, IDENTIFICATION AND IMPLEMENTATION


OF COMPLEMENTARY INTERVENTIONS
The identification and development of complementary interventions should take into account current
national, regional and sector policies; legal instruments; international conventions and treaties; guidelines and
procedures relating to public consultation/participation; local development planning and implementation.

2.1 Planning
Complementary interventions need to be considered early in the project development (Appendix C.1)
and be an integral part of project planning, from project identification to feasibility study. It is important
that the client and key stakeholders (for example those who identify the need for the project and local
authority representatives in the project area) work together to develop an outline plan for inclusion of
complementary interventions in the road project/programme to a sufficient level of detail for their further
development during the feasibility study and detailed design.

The outline plan and budget for complementary interventions, and an assessment of the potential
impacts, should be included in the economic analysis of road projects as they may raise the economic
rate of return of the road investment, despite any initial additional costs.

Key issues for consideration during the planning, feasibility and preliminary design stages of a project are
considered more fully in Appendix C.1.

When approaching the design of complementary interventions it is necessary to consider actions that:
ƒ Are demand driven;
ƒ Are agreed in principle by all of the relevant local and other authorities;
ƒ Will have a high level of participation and involvement from the local authorities and communities
during implementation (See Plate C.1.1);
ƒ Are matched with and complement actions within existing local plans, such as, Wereda development
plans.

Plate C.1.1: Community participation meetings

In developing a detailed design for complementary interventions it is essential that the design engineer
works through the client and with the right local institutions and structures. The client will use existing
consultation frameworks, structures and plans to be advised on the identification and development of
any proposed complementary interventions.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 2 - 4

For detailed design a high level of consultation with affected communities will also be needed. The
design engineer will work through the client to ensure that the correct local procedures are adopted
and that the right formalities are followed. Decisions, prioritisation methods and approvals for planned
initiatives would be introduced by the client and achieved through the clients interaction with the existing
and appropriate local level structures. The client may require the design consultant to assist with the
identification and prioritisation process. Where this occurs the design engineer will need to communicate
with local authorities and communities. If identification and prioritisation are within the design consultants
mandate these should be carried out under the clients guidance and using participatory processes.

Participation considerations
In promoting consultations and participation the following should be noted:
ƒ Beneficiaries should not be viewed as a passive element. Beneficiaries should be active
participants.
ƒ Situations should be avoided that override existing and legitimate decision-making processes and
structures
ƒ Due care and attention should be given to group decision-making processes that may reinforce
existing power structures at the expense or exclusion of vulnerable groups.

Effective participatory decision making processes have already been developed on many rural infrastructure
projects and integrated development projects in Ethiopia and within client road organisations (e.g. ERTTP
2003). The design consultant should refer to the guidelines for these projects and participatory approach
manuals from other line ministries (MoFED, 2006) in developing their own detailed methodology for
identification and selection of complementary activities for each low volume road project.

The design engineer should also be familiar with existing local and regional development plans, potential
sources of complementary financing or resources that may be allocated to the complementary activities,
willingness of local communities to make other contributions, work by local NGO and CBO. Typically,
consultations and participatory decision making may involve the following organisations:

Federal level Regional level Wereda level Kebele level


ERA: Planning & ICT Regional Roads Wereda Kebele
Directorate; Regional Authority: General Administrations: Administrations:
Directorates Rural Manager or Regional Wereda Development Kebele Administrator,
Roads Technical Road Engineer Committee, Kebele Development
Support Branch; and Regional Transport Administrator and Committee. Kebele
Environment Branch Authority: General Wereda Sector Development Units
Ministries: Ministry of Manager Department/office and cooperatives.
Transport; Ministry of chiefs NGO and CBO.
Agriculture Bureaus: Regional
Federal Transport Bureau Heads, Wereda Road Desk:
Authority Administrators and Wereda Road Desk
Sectoral chiefs Engineer and Public
Works Engineer

The consultant, with guidance from the client and key stakeholders, may be required to help establish
and provide support to a ‘Complementary Intervention Oversight Committee’ at the different levels.
These would become the main contact points for the design consultant and may assist in the community
consultation and participatory decision making processes. If the establishment of a separate committee
is not required, a main contact point within the existing local administration would need to be nominated
or identified through the client.

The participatory process will require a multi-disciplinary design team to ensure an appropriate
consideration of technical and financial aspects of the proposals identified, to develop appropriate

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 2 - 5

designs and implementation mechanisms that meet the expressed needs of the communities, and to
ensure adequate coverage in the construction contracts.

Environmental and social safeguard requirements


Complementary interventions are about creating opportunities through road project contracts to
further enhance the positive impacts of low volume roads. They are not about mitigating or minimising
negative impacts brought about by the road project. The requirements for assessing and addressing
Environmental and Social Safeguards during project planning, design and construction are clearly
addressed through other existing documents. These documents, and their subsequent revisions or
replacements, should be followed where appropriate.

A current list of the relevant existing documents is maintained by ERA and is available from ERA or the
ERA website. www.era.gov.et

2.2 Identification and selection of complementary interventions


In theory, complementary interventions may include almost anything that can be implemented through a
road works contract and which contributes to the socio-economic, environmental or safety of communities
affected by the road. They have been divided into three categories to help clarify the different types of
activities and how they relate to the traditional role of the road works contractor or works contract.

Category I - Management interventions


The works contractor will already have clear cut environmental, safety and employment obligations set
out under the provisions of the works contract. These would be captured within the relevant design and
implementation management documents (including EIA and EMP). Management interventions add to
and extend the normal obligations of the contractor. These could include interventions such as items
relating to improving road and resident access; reinstatement/improvement of areas used temporarily
during construction; or provision of facilities and services disrupted by construction activities. These
activities would be included as Bill of Quantity items (Measured Works) in the contract. These small scale
activities could be captured within other items or provided directly as a contribution by the contractor.

Category II – Opportunity interventions


Opportunity interventions go beyond the normal scope of a road works contract, but which are within
the technical and management skills of a road works contractor. Opportunity interventions could
include support for provision or repair of community infrastructure (such as a clearance of a market
area); or provision of material, labour or supplies for small community works such as rehabilitation or
repair of community facilities (community buildings, hand pumps, irrigation infrastructure and the like).
Provision of technical training to Wereda and Kebele administrations and staff by the contractor is also
possible (eg vehicle maintenance; financial management; road construction and maintenance; building
and/or sanitation management). Similar technical support could also be provided to local enterprises
and cooperatives. These activities would be included as Bill of Quantity Items (Measured Works) in the
contract or could be established through a separate or parallel agreement between the contractor and
the Wereda/Kebele Administrations or with the community.

Category III – Enhancement interventions


Enhancement interventions extend beyond the skills and experience of a normal road contractor and
would require specific arrangements through the contract with other skilled parties. Such parties may
include local government offices, NGOs, private sector organisations, community based organisations
and cooperative societies who are better placed and skilled to implement the proposed interventions.
The role of the contractor would be to manage the activity through the contract and provide physical
support, if necessary or appropriate, provide financing to the organisation implementing the activity to
include such activity under monthly site reporting, and to present the activity for payment through his
normal certificate for payment. Verification of activities would be undertaken by the supervising consultant
or client. Such activities could include awareness raising and education campaigns; establishment of

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 2 - 6

new or improved livelihoods options; building of facilities; life skills training; or provision of supplies and
training to local service providers. These activities would be included as provisional sums in the contract.

As well as dividing complementary interventions into three categories, they can also cover a range of
themes; for example road safety, road corridor environment, road transport services, and community
development. Appendix C.2 shows some indicative examples of complementary interventions by
category and theme.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 3 - 7

3. EMPLOYMENT AND HUMAN RESOURCE ISSUES

Road works provide an opportunity for temporary employment. Employment opportunities can be
maximised by promoting labour (Plate C.3.1) or the use of intermediate equipment (Plate C.3.2) rather
than employing heavy equipment technologies.

Plate C.3.1: Promotion of labour-intensive methods

Plate C.3.2: Promotion of intermediate technology option

The client will guide the design engineer on any special emphasis with regards the approach to be
adopted. It is reasonable to assume that for the foreseeable future a mix of labour and intermediate
equipment approaches will be favoured for provision of community, wereda and some regional roads.

The principles of complementary interventions can be used to enhance access to employment


opportunities for the wider community, where unemployment or under-employment are local issues.
Road works can provide paid employment opportunities and such opportunities should be used as far as
possible for the local people.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 3 - 8

There are many unskilled and semi-skilled works activities that could easily be carried out by the local
people and with little training. Employment of local people also provides a mechanism for training and
support under Category II and III type interventions that relate to skills development and future income
generating activities.

Typical activities include:


ƒ Crushing and screening rock for aggregate or hand screening gravels;
ƒ Grubbing and clearing;
ƒ Digging ditches and foundations;
ƒ Making gabion baskets;
ƒ Seedling and sapling planting and maintenance;
ƒ Cooking, cleaning, and managing work sites and camps.

The contract should include employment clauses and provisions that require or provide an incentive
to the contractor to use local labour as much as possible. Typical contractual mechanisms may include,
for example, minimum targets for percentage of labourers on specific activities employed from local
communities. Payment for the relevant activities would reflect the extent to which the employment
targets have been met. This should also be a topic for review at monthly site meetings.

More simply, an overall target can be set in terms of person-days worked either as a total or percentage
of the total by local people. Again payments would reflect the extent to which the target has been met.

When developing such clauses, or enhancing existing clauses, the following issues should be given due
consideration:
ƒ Boundaries for defining local labour – eg living within a designated number of kilometres from the
project road; or living within certain administrative areas;
ƒ Availability of skilled and semi-skilled labour within the project area, and within sections of the road
project;
ƒ General health and physical condition of the locally available labour (especially in areas where
there is significant out-migration for work, drought, food insecurity or endemic health issues);
ƒ Social, religious or traditional barriers that may prevent some social groups from accessing labour
opportunities and how these might be overcome;
ƒ Incentives or penalties to be used to encourage the contractor to meet local employment targets;
ƒ Measurement, monitoring and recording mechanisms to accurately report on local employment.

The design consultant will need to carry out a labour survey during the detailed design stage to develop
appropriate local employment contract clauses and targets that accurately reflect conditions along the
project corridor.

3.1 Encouraging participation of marginalised groups


Every society has groups that are, for whatever reason, disadvantaged or excluded from participation in
employment. Often these groups can benefit most from temporary employment in road works projects.

Typically excluded groups include:


ƒ Women in general and especially mothers with young children;
ƒ Physically or mentally disadvantaged;
ƒ Ethnic or religious minorities;
ƒ HIV/AIDS affected or infected.

It has to be recognised that the labour required on the road site usually requires physically hard work
and construction sites can be relatively dangerous places. It is not, therefore, appropriate to require
contractors to employ physically or mentally disadvantaged people, though there are of course some
jobs, for example at the works camp, that could be appropriate.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 3 - 9

It should be possible for the design engineer to develop an understanding of the barriers to participation
in employment by women and minority groups, and find ways to help them access employment - without
causing conflict or concern amongst the wider community (Plate C.3.3).

In many areas, for example, barriers to women’s participation in road works is caused by their own need
to collect water, fire wood, animal fodder and look after the home and children.

Plate C.3.3: Participation of women in road works

ƒ Measures that can be easily introduced to help women overcome such barriers include:
ƒ Allowing women to form work groups and share the work load between them. This provides for
flexible working hours for individual women and protection in numbers. This also allows them to
share child care responsibilities through a rotating crèche or similar.
ƒ Ensuring women work in areas where they feel safe and secure – due consideration should be
given to issues of isolation (eg distance from other work groups), appropriate areas or facilities for
nature calls and distance from home/crèche.
ƒ Removal of time constraints preventing women from accessing employment opportunities. This
can be done by, for example, providing firewood or alternative fuel or potable water through the
contract. This would require the contractor to bring fuel or potable water to site for distribution. In
some cases this could also be used as part of payment for work done (by men as well as women).
Importantly, it should only be part payment.

3.2 Meeting existing obligations


There are already existing legal requirements relating to employment of temporary labourers defined by
the Ministry of Labour and social affairs that should not be neglected in the contract (see Box).

In general, contract clauses should include provisions that ensure:


ƒ Minimum wage rates per unit of time or unit of work are adhered to;
ƒ Equal pay for equal work;
ƒ Equal access to employment opportunities (taking into account local traditions, with respect to
male / female jobs, for example);
ƒ Effective monitoring and recording of work done and wages paid (transparency);
ƒ Health and safety of employees and affected communities (due attention to sanitation,
waste disposal, disease control and prevention, first aid and emergency health care, accident
management and reporting);
ƒ Safe and appropriate temporary accommodation and recreational facilities;
ƒ Effective grievance procedures for temporary employees;
ƒ Limitations on use of child labour (according to local laws, regulations and customary practices);
ƒ Use of ILO guidelines and regulation if it is an internationally funded contract.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 10

4. CONTRACT PROVISIONS TO SUPPORT


COMPLEMENTARY INTERVENTIONS

The key to successful implementation of complementary interventions will be ensuring that clearly defined
requirements and adequate provisions are included in contract documents and at all stages of the project.
The following sections describe how the different contracts (for project design and supervision services
and works) may be used to address complementary interventions.

4.1 Design services contract


During project planning, the client will need to determine the approximate budget and scope of the
project, including the budget and scope for complementary interventions. This then needs to be
reflected in the Request for Proposals (RFP), in particular the terms of reference, for consulting services
for the detailed design. The RFP should specifically include appropriate inputs of key personnel with the
requisite skills to meet the requirements of the client with regards complementary interventions.

The client may require the consultant to undertake some participatory approach to assist with the
development of the complementary intervention package. With regards to Category III interventions, it
would be far more efficient and appropriate for the client to undertake the identification and preliminary
selection using the existing appropriate government structures and plans.

The development of Category I and II intervention packages are relatively straight forward and easily
within the skill area of a multi-disciplinary design team. However, the RFP still needs to be well developed
and thought out (see Box).

The feasibility study will have developed the preliminary options and budget/cost estimates for
the complementary interventions. The detailed design will require preparation of the finalised list of
complementary interventions, detailed designs and engineers cost estimates.

The RFP should include:


ƒ Clearly defined and appropriate inputs for key personnel to be involved in developing
complementary interventions;
ƒ Reference to this Part C or alternative guidance on identifying, selecting and designing
complementary interventions;
ƒ Requirement for organisation of a mobilisation workshop that draws all stakeholders together on
site;
ƒ Requirement to review participatory decision making practices and develop project specific
methodology that best reflects local decision making structures (formal and informal);
ƒ Clearly defined duties and responsibilities of the parties to the contract and any external
organisations with respect to development of the complementary interventions;
ƒ Requirement to review national, regional and local development plans in the development of
complementary interventions;
ƒ Requirement to include complementary interventions in any further economic analysis;
ƒ Requirement to consult with local government and community representatives to enable them
to participate effectively in the decision making process when developing complementary
interventions;
ƒ Requirement to clearly define and specify requirements for complementary interventions in the
preparation of bidding documents.
The RFP may require the consultant to provide guidance on:
ƒ How to decide which potential CIs will be the most effective or efficient use of resources and how
this can be measured;
ƒ How to ensure that the contractor does not become overburdened by complementary
interventions; that the proposed interventions are suitable to the scope of works; and provide
guidance on the anticipated experience and skills of the contractors to be procured.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 11

4.2 Works Contract


For the purposes of including Complementary Interventions within works bidding documents (See Table
C.4.1 for guidance), the key documents requiring attention are:

Instructions to Bidders (ITB) and the Bid Data Sheet (BDS): For a LVR project the client should include
an additional item that will draw the attention of the bidder to any requirements for complementary
interventions.

Standard Technical Specifications: The Standard Technical Specifications form a separate volume in the
ERA 2011 series. These capture some of the proposed interventions mentioned in Series 11000 Ancillary
Works. Other complementary interventions should be included in the Particular Specification.

Particular Specifications: This is where any detailed technical requirements and specifications, and
implementation mechanisms specific to the designed set of complementary inventions should be clearly
defined for the project.

Bills of Quantities or Schedules of Rates: This should be linked by item number to the Standard Technical
Specifications and to the Particular Specifications; and is where the schedule of activities and estimated
quantities for the complementary interventions are set out for the bidder to price.

Drawings: Some standard detailed drawings may be applied directly eg provision of trail bridges.
Supplementary drawings, linked with the Particular Specifications, may also be required where new or
special complementary approaches are included. Where trail bridges are included within the contract, as
a complementary intervention, a separate volume of drawings is provided. For Category III interventions
reference should also be made to any standard drawings used by line ministries for infrastructure under
their control.

Conditions of Contract: This includes standard provisions for execution of the contract and unless
amended in the Conditions of Particular Application, these will apply. In some cases modification of
some clauses may be required to reflect the desired approach and will need due consideration. For
example, where there is a desire to support small and medium enterprises and small scale contractors
then due consideration should be given to the clauses referring to Performance Security, Performance
Program, Insurances, Cash Flow, Plant, Equipment & Workmanship, Payments, Retention and Advances,
Price Adjustment and currency restrictions.

Conditions of Particular Application: This is where any Provisions in the General Conditions of
Contract may be amended, as required, to make them more appropriate. This may apply to some of the
complementary intervention envisaged. Where assets are involved, the document should be clear on the
responsibilities for asset transfer. Due consideration should also be given to strengthening clauses aimed
at promoting sub-contracting/assignment; local employment and conditions (particularly for women);
rights and insurances; and for strengthening complementary interventions.

Important documents which are not part of the Works Contract are the User Guides. These are guidelines
to the design Consultant or the client organisation on the preparation of the Works Bidding documents
and include reference to provision of complementary interventions.

The main issues for a contractor are to fully understand the scope of all of the works, including the
complementary interventions, and the fundamental issues of measurement and payment.

The main issue for the client is to achieve the objectives of the intervention, to get value for money and
for these interventions not to require a disproportionate amount of supervision or monitoring.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 12

Table C.4.1: Works bidding/contract documents

General Notes for Guidance


The ITB is generally a standard document which may vary depending on
Instructions to
the procedures of different clients. For a low volume road project the client
Bidders (ITB) and
should include an additional item that will draw the attention of the bidder
the Bid Data Sheet
to the low volume road approach. The BDS is linked to the ITB and provides
(BDS)
specific project information.
These will define the scope of the technical requirements of the contract,
including the type and quality of materials and equipment, the standards of
workmanship. The Standard Technical Specifications form a separate volume
Standard Technical
in the ERA 2011 series. The ERA Standard Technical Specifications includes
Specifications
the works to be undertaken, general obligations information on the format
of Bill Items for the Bill of Quantities, on item coverage and the method of
payment.
This is where any detailed technical requirements and specifications, and
implementation mechanisms specific for the project, should be clearly
defined. Particular technical specifications add further detail to complement
Particular
or replace those stated in the Standard Technical Specifications. The
Specifications
particular technical specification should also include any specifications and
limitations on the freedom of choice for the contracting company related to
the execution of works.
This should be linked by item number to the Standard Technical
Bill of Quantities Specifications and to the Particular Specifications; and is where the schedule
of activities and estimated quantities are set out for the bidder to price.
Some standard detailed drawings may be applied directly for low volume
roads works (eg cross-sections, standard culvert design and road signs).
Drawings
Supplementary drawings, linked with the Particular Specifications, may also
be required where new, innovative or special approaches are included.
This includes standard clauses or provisions for contracting requirements,
obligations and legal commitments which, unless amended in the
Conditions of Particular Application, will apply for the project. In some cases
Conditions of
modification of some clauses will need due consideration and modification
Contract
may be required to reflect the desired approach for the low volume road
works (see below).

Conditions This is where any Provisions in the General Conditions of Contract may be
of Particular amended as required, to make them more appropriate to non-standard
Application project requirements, including complementary interventions.
These documents will provide detailed guidelines to the design consultant
User Guides
on the preparation of the works bidding documents.

4.2.1 Options for inclusion of Complementary Interventions in Works Contracts –


Measurement and Payment

Many complementary interventions will involve provision of work items, activities or services beyond the
usual core activities associated with a traditional road works contracts.

Bill of Quantity Items (Measured Works)


Where the scope and detailed design of a complementary intervention is well defined and within the
scope of activities normally expected of a road contractor (ie Category I and II interventions), the preferred

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 13

option would be to include these activities as items within the Bill of Quantities for the contractor to price.
This approach would apply to any extension and an example of typical entries is provided in Table C.4.2.

Table C.4.2: Example Bill of Quantities entries

No Item Description Unit Quantity Rate Amount


Clear and prepare hard standing
for market, in accordance with
1* m2 500
Specification Ref: (……….) and
Drawing: (……….)
Rehabilitate school block in
accordance with Specification
2* Item 1
Ref: (……….) and Drawing:
(……….)
Note:
* Item No. relates to item in Standard Technical Specifications or Particular Specification. Shaded area for bidder to price

Provisional Sums
If the intervention is not fully developed and agreed (eg enhancement of temporary works areas such as
borrow pits to small dam or fish pond; or provision training if it is outside the scope of activity normally
expected of the contractor in some category II and III interventions), then it is probably better to describe
the activity briefly and to include a ‘Provisional Sum’ item in the Bill of Quantities.

What information is known could be included in the Particular Specifications or Drawings, with a requirement
for the Contractor and/or Supervision Consultant to further develop and define the intervention later.
Generally the Provisional Sum Item is only one or two lines included in the Bill of Quantities. A ‘cost
estimate’ is inserted by the Client to cover the cost of these items.

The provisional sum is an estimated cost for the intervention based on the information available at the
time of bidding document preparation. The actual cost of the intervention may change, with the final
design and price for the intervention being agreed between the parties to the contract once all the
necessary information is known.

Provisional Sums are flexible and are used ‘at the discretion’ of the Employer Engineer.

When using provisional sums, provision must be made for the administration and management of these
funds by the contractor, in the form of a % fee or adjustment to the provisional sum item.

An example of Provisional Sum entries in the Bill of Quantities is shown in Table C.4.3.

Table C.4.3: Example of Provisional Sum entries

No Item Description Unit Quantity Rate Amount


Provision of road maintenance training
1* PS 125,000
to Wereda staff
Allow percentage for administrative fee
% 1
on provisional sum, item (....)
Provision of construction supplies to
2* PS 75,000
Kebele Administration
Allow percentage for administrative fee
% 1
on provisional sum, item (....)
* Item No. relates to item in Standard Technical Specifications or Particular Specification. Shaded area for bidder to price

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 14

Output Related Items


Output related approaches are most applicable for procurement of supplies or services. Measurement
and payment will normally include stages and be linked with outputs or ‘deliverables’ (eg an activity
report following a training programme). If the Terms of Reference are complete for such an intervention
then bill items can be included for the contractor to price, as in the example in Table C.4.4. If such
information is not available to the contractor then the intervention should be included as a Provisional
Sum item and an estimated cost included in the Bill of Quantities by the client.

Table C.4.4: Lump Sum item for completed output related to RS

No Item Description Unit Quantity Rate Amount


Provide Training Materials and
deliver training Community and
1* Schools Road Safety Education Lump Sum 1
in accordance with Terms of
Reference (Annex TBA)
*Item No. relates to item in Standard Technical Specifications or Particular Specification. Shaded area for bidder to price

Any items for supplies or services that are outsourced to appropriate implementation partners will require
a detailed Terms of Reference to be developed and included in the sub-contract agreement. Such sub-
contractors may be nominated by the client or selected by the contractor as a ‘domestic’ sub-contractor
(the client would still need to approve the choice of sub-contractor). These ToR will need to clearly define
the scope of work, identify the beneficiaries to be reached and set out the detailed specifications if the
sub-contract is for supplies. For a services sub-contract the projected person month inputs and minimum
staff requirements will be included. ToR should also include a payment schedule either based on inputs
or, most likely, linked to deliverable.

Schedule of rates
This is a particular form of pricing mechanism that is used where the activities or procurement items are
known, but the quantities are not (See Table C.4.5). The schedule lists the items to be provided giving
the unit of measure but not the quantities, or if quantities are included it is made clear that such quantities
are nominal. The bidder will then submit a rate (rate only) against each item in the schedule.

Table C.4.5: Example of schedule of rates where quantities are unknown

No Item Description Unit Quantity Rate Amount


Provide animal drawn carts in
accordance with Specification
1* No 1 Rate Only
Ref: (……….) and Drawing:
(……….)
*Item No. relates to item in Standard Technical Specifications or Particular Specification. Shaded area for bidder to price

If standard unit rates are known then they can be inserted by the design engineer (See Table C.4.6) and
will then normally form the basis of a nominated sub- contract. This could be used, for example, for the
provision of school furniture from a nominated supplier, or a national procurement office, whose rates
have already been agreed with the Ministry of Education. If such fixed rates are to be used they will
generally include for contractor overheads and profit.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 4 - 15

Table C.4.6: Example of Schedule of rates where quantities are known

Quantity
No Item Description Unit Rate Amount
(Nominal)
Provide school desks in
accordance with Specification
1* No 50 100 5000.00
Ref: (……….) and Drawing:
(……….)
*Item No. relates to item in Standard Technical Specifications or Particular Specification

Dayworks
The dayworks schedule is intended for the pricing and payment for small scale ‘incidental works’ and should
not be relied on as a mechanism for the measurement and payment of complementary interventions.

4.3 The Works Contract evaluation


Although the complementary interventions component of a road project may be a relatively small part
of the construction budget, it must be included within the evaluation of the bid process. The aim should
be to ensure that the contractor understands the complementary intervention requirements and has
consulted with the relevant partners to provide considered and accurate price estimates.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 5 - 16

5. SUPPORTING SMALL SCALE CONTRACTORS

Promoting development of local small scale and emergent contractors and enterprises can be considered
as a complementary intervention. This can be achieved in two ways:

5.1 Utilising the lower level bidding documents


The following lower level bidding documents are available:
ƒ Standard Bidding Document for the Procurement of Minor Works – Maximum contract value USD3
Million
ƒ Standard Bidding Document for the Procurement of Micro Works – Maximum contract value
USD300,000

Smaller scale works contracts are more forgiving to the emergent contractor and protect the client from
high levels of financial risk.

When contracting work to small scale contractors, small enterprises or community based organisations
that are not so used to undertaking road related works, the following issues need to be considered and
a position agreed with the client:
ƒ It should be understood that emergent contractors may lack contract experience. The client and
supervisor will need to be forgiving in some instances where full compliance with normal contract
provisions could have a negative impact on bidding, the performance of the contract or the
existence of the contractor.
ƒ Ensuring the requirements of the contract, allocation of responsibilities, technical and performance
standards, payment terms and conditions are clearly understood by the tenderer/contractor.
The project could include elements of contract management and supervision support, technical
assistance and/or management training.
ƒ Cash flow: It is unlikely that small sub-contractors will have sufficient cash or resources to start the
activity without some form of advance payment (in cash or in kind) and it is likely that they would
need more frequent payments (eg bi-weekly rather than monthly).
ƒ Advances: The advance amount, repayment conditions, and regular payments need to be defined
in a payment schedule in the agreement. Advance guarantees used on larger contracts may not
be appropriate or even possible for small scale works.
ƒ Performance guarantees and bonds: It is mandatory to include performance guarantees for the
works. Where small scale enterprises are sub-contracted, the main contractor guarantees will cater
for the SME. Where SME are contracted directly, the provisions set out in the respective model
tender/contract documents should be responsive to the size of the contract/contractor and to local
constraints.
ƒ Performance Programme: The period of the works should also take account of the likely size of
the contractor and the approach adopted (labour, intermediate equipment, heavy equipment etc).

5.2 Utilising sub-contracting clauses


By utilising sub-contracting, risk is carried by the main contractor who has qualified and met the minimum
criteria set by the client. The client may set a minimum percentage or type of works that should be sub-
contracted to small scale enterprises. In doing so the client needs to recognise the risk that is carried by
the contractor and to ensure that the performance programme and payment clauses recognise that the
works contract has a capacity building component.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 6 - 17

6. SUPERVISION CONSULTANT’S CONTRACT

The Request for Proposals (RFP), in particular the Terms of Reference, for consultancy services for
construction supervision need to reflect the role of the Engineer in supervising and administering
payments for the complementary interventions.

As with the main engineering design, the supervision consultant will be required to review the agreed list
and detailed designs of the complementary interventions, and to consult with the necessary stakeholders
to ensure that priorities remain unchanged and that the interventions are still appropriate to the needs of
the beneficiary communities.

The RFP also needs to take into account the need to further develop and negotiate detailed agreements
for some of the interventions as the construction works progress, for example interventions relating
to reinstatement or temporary works areas. The Employer will probably also need to be involved in
monitoring grievances and dispute resolution, should the need arise. In such cases, the Engineer shall
provide the required assistance.

The supervision and monitoring of complementary interventions is likely, to require specialist skills beyond
the normal engineering supervision team. The RFP should reflect this, and should include:
ƒ Clearly defined and appropriate inputs for key personnel to be involved in monitoring and
supervising complementary interventions;
ƒ Reference to this part C of the low volume design manual or alternative guidance on designing
and monitoring complementary interventions;
ƒ Clearly defined duties and responsibilities of the parties to the contract and any external organisations
with respect to development and implementation of the complementary interventions;
ƒ Requirement to consult with and provide appropriate training to implementation partners to assist
them in fulfilling any obligations as defined in agreements / sub-contracts for complementary
interventions;
ƒ A requirement to include complementary interventions in their progress meeting agendas and
progress reports.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 7 - 18

7. MANAGEMENT, MONITORING AND ENFORCEMENT

In general, the complementary intervention aspects of the contract should be managed monitored and
enforced using the normal provisions of the contract documents.

Complementary interventions should be included in the contractors detailed work plan and the payment
schedule. Progress and performance should be reported through the monthly site meetings and progress
reports. It may be appropriate to prepare specific reports for local communities and their leaders on the
progress of complementary interventions in their area. The frequency of such reports would depend on
the nature and scale of complementary interventions being implemented in that area, which should be
determined during the detailed design stage and uses provisions made in the reporting sections of the
works and supervision contracts.

While it is the Contractors responsibility to manage and implement the complementary interventions
according to the contract, it is the Engineer’s responsibility to ensure complementary interventions are
monitored regularly and that technical and performance standards are met.

Monitoring and enforcement should be closely linked to the contractors payments. It is essential that
measurement and payment for complementary interventions and any incentives or penalties are clearly
defined in the works contract.

Payment for sections of the road should only be fully paid once all aspects of the contract have been
completed in that section. This includes completion of environmental mitigation measures and the
complementary interventions, to the required standards. This should be reflected in the payment schedule.
It may be worth using a form similar to the Environmental Clearance Certificate (see examples from ERA
quality management system) for complementary interventions when processing payment requests, which
makes provision for retaining a certain amount from the contractors payment to ensure that sufficient funds
are available to the client to complete the mitigation measures/complementary interventions should the
contractor fail to do so.

As the complementary interventions may involve small scale contractors or community groups, an element
of community based monitoring should also be included. A mechanism should be established by which
the affected communities are clearly informed about what is to be done and by whom, and to monitor
implementation by all parties to the complementary intervention agreements. This can be done through
formal monitoring groups, or by putting up information on a notice board that allows all members of the
community to see what should be happening.

A mechanism for reporting on performance from the wider community would also need to be established.
This could be through the appointment of a local inspector and/or contact person, or through supervision
visits by the supervision consultant’s team.

PART C: COMPLEMENTARY INTERVENTIONS


C - Chapter 8 - 19

8. REFERENCES

Ethiopian Rural Travel and Transport Programme, ERTTP (2003). Planning, Monitoring & Evaluation
Manual, Volume 2. Ethiopian Roads Authority.

Ministry of Finance and Economic Development, MOFED (2006). Guidelines for the Preparation of
Public Sector Projects. January, 2006.

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.1 - 20

APPENDIX C.1 - COMPLEMENTARY INTERVENTIONS –


PLANNING, FEASIBILITY AND PRELIMINARY DESIGN STAGES
A) Key issues to consider during early planning stages
Complementary interventions need to be treated as an integral part of the planning process, in much the
same way as environmental and social safeguards. Provision for them should be included in the long and
medium term budgets to prevent them being removed due to inadequate budgeting or fund allocation.

The client should consult with key stakeholders during project identification stages. Key stakeholders at
the early planning stage could include those who identified the need for the project and representatives
from local administrations. The outcome of the consultations should be an outline or indicative plan for
inclusion of complementary interventions in the road project/programme.

It is the responsibility of the client to decide on the extent of complementary interventions that are to be
considered during feasibility study and prepare the outline plan. The outline plan should describe the
category, type and scale of the complementary interventions to be developed further.

Road projects, by their nature, cover relatively long distances, cross many local administrative boundaries,
and affect a number of different communities. Identification of complementary interventions is not
therefore a simple task of consulting one community to identify their development needs and priorities,
but requires consultation and negotiation with many communities – each with their own internal structures
and cultures, needs and priorities; each group thinking another is being allocated a better ‘share’ than
they are, etc.

The key stakeholders will be expected to have a deeper understanding of the beneficiary communities
than the client, and should be able to provide guidance on the most locally acceptable means for
engaging with local communities and appropriate participatory decision making methods.

The developing an outline plan the client should consider and provide guidance on the following issues:
ƒ Is the client interested in including complementary interventions in the project?
ƒ What category of complementary interventions are appropriate, bearing in mind the cost/budget,
timeframe, scope and complexity of the project?
ƒ How to define the boundaries of where complementary interventions can be implemented – area
of influence of the road project?
ƒ What proportion of the road project budget may be set aside for complementary interventions?
ƒ How to determine the level of willingness of local authorities and communities to participate in
development and implementation of complementary interventions?
ƒ Who, at the local level, could best assist the client and his service providers to develop and
implement the complementary interventions? Is there a need to establish ‘Complementary
Intervention Oversight Committees’ or can existing committees or administrators take on this role?
ƒ What are the best methods for raising awareness of the opportunities for, and identification and
prioritisation of, complementary interventions? Eg large meetings or lots of small meetings with
different groups? Which language(s) to work in? Who are participants and key speakers? How best
to present information (in writing or pictures)? Which analysis and decision making tools are most
appropriate? How to consolidate different needs identification and prioritisation results?
ƒ What may be the best method for identifying and selecting the proposed complementary
interventions – for example, forming a committee with representatives from regional and wereda
levels to determine the shortlist? Or each wereda presents its shortlist to the consultant/client for
consideration?
ƒ Can key stakeholders, based on current local development plans and initial consultation with
local authorities within the project area, develop an indicative list of potential complementary
interventions?
ƒ What additional funding or resources may be available from other sources to allocate to
complementary interventions?

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.1 - 21

ƒ To what level does the participation process need to extend? This partly depends on the budget
and category of intervention agreed on for the specific project. It will also depend on the number
and size of communities that live along the road.
ƒ How is each community defined? What is the formal and informal structure for decision making
purposes? Who are the key stakeholders – those likely to influence the decisions made and those
likely to influence the success of implementation and sustainability of interventions?
ƒ How to ensure vulnerable and/or excluded groups within a community are included in the
participation process? (eg women, elderly, children, physically and mentally disadvantaged, ethnic
or religious minorities, etc.)

B) Key issues to consider during feasibility study and preliminary design


During Feasibility Study, the consultant, with guidance from the client and key stakeholders, takes on the
responsibility to further investigate the options for and develop preliminary designs and cost estimates for
complementary interventions, based on the outline plan previously prepared.

The investigations at this stage are aimed at developing the complementary interventions to a sufficient
level of detail to enable reasonably accurate cost estimates to be prepared and impacts to be assessed
for complementary interventions. Complementary interventions are to be included as an integral part of
the options analysis and economic analysis of the road project. Complementary interventions, despite
involving additional costs, may well raise the economic rate of return of the road investment.

The Feasibility Study consultant will need to continue and expand upon the consultations already
undertaken through a detailed participation strategy. Initial awareness should raise the outline road
project to the local communities. Information to be given including the approximate timeframe for the
project
ƒ Potential route options (the final route selection may mean some communities do not fall within the
road project corridor);
ƒ Nature of the works;
ƒ Potential for complementary interventions.

Following on from the awareness raising programme, the client or his representative, should work
closely with the local authorities and communities to identify and prioritise the potential complementary
interventions.

The detailed participation strategy will need to define how decision making may be devolved to the local
level, whilst maintaining an overview and consistency in approach along the road project. It will need
to clearly define decision making methodologies and allocate responsibilities for decision making at all
levels.

The outcome of the identification and prioritisation process should be a list of potential interventions.
Ideally, this should be presented as an impact analysis table, with each intervention having an estimated
cost, an estimate of the number of people and extent of impact on different groups of people positively
or negatively affected by the intervention, and an attempt at quantifying the benefits of the intervention.
The location of potential complementary interventions and the intended beneficiaries should be clearly
defined.

A preliminary or outline design will also be necessary at this stage to enable relatively accurate cost
estimates for each potential complementary intervention. The consultant should also identify any sources
of additional financing or resources that can be used to implement the complementary interventions. This
could include funds coming from the Client through the road project contract, financing promised from
other sectors or local authorities, contributions of labour or materials promised from local authorities and
local communities inputs such as labour.

The local authorities, based on the guidance given by the client, should review the list of proposed
interventions and select those which they propose for inclusion in the project. This will be a difficult task
and needs to balance allocated budgets against meeting the prioritised needs of the most people in an

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.1 - 22

equitable manner. The decision making process must be transparent to minimise the potential for conflict
and complaint.

Due consideration needs to be given by local authorities to the client on:


ƒ How to divide the CI resources between different communities along the road?
ƒ How to divide the CI resources among the different categories of CI? Is this defined in the policy
or is it project specific?
ƒ If interventions are to be demand driven – whose demands are more important? Are all groups
equal? What are the implications for decision making and implementation? Are their any
development policies that prioritise certain social groups or types of intervention?
ƒ How to coordinate ideas of local communities with development plans of local authorities and
higher level sector organisations (line ministries)?

It is likely that the final selection of complementary interventions will be an iterative process, to ensure
ownership at all necessary levels. Conflict management between different groups of one community,
between adjacent communities, between different sectors and authorities should be resolved by the
appropriate Regional or next level up authorities.

The feasibility report should capture the selection methodology, analysis of potential complementary
interventions and recommendations for those to be selected for inclusion in the project, along with an
outline budget.

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.2 - 23

APPENDIX C.2 - INDICATIVE EXAMPLES OF


CIS BY CATEGORY AND THEME

Category 1 Management Category 2 Opportunity Category 3 Enhancement


Theme
Interventions Interventions Interventions
ƒ Provide illumination/ ƒ Distribution of reflective ƒ Road safety awareness
additional marking in strips for pedestrian, NMT : schools/community
dangerous areas and IMT road users. road safety education
ƒ Extend provision or ƒ Provide boards warning campaigns – Community
Road and Site Safety

maintenance of access community of construction theatre, TV and radio etc


to specific services and and road hazards ƒ Provide Road Safety
facilities for pedestrians ƒ Provide refresher/first aid equipment, teaching
and IMTs training for local health aids or additional
ƒ Provide road safety officials equipment
education to employees
ƒ Provide access to first aid
training for community
representatives and to
facilities in emergency
ƒ Rigorously enforce speed
limits of equipment and
plant
ƒ Provide temporary and ƒ Plant productive (eg fruit, ƒ Establish nurseries
permanent accesses to nuts, fuel) trees and plants for supply of trees
Road Corridor Environment (including climate change adaptation measures)

homes, tracks and paths along roadside and in and shrubs for bio-
ƒ Provide water and hand reinstatement of borrow engineering, fruit
sprinkler systems to local areas orchards, wood lots etc.
communities to control ƒ Establish landfill/waste ƒ Provide protective
dust on road sections management sites, utilising
tubing for saplings,
near their properties as borrow and quarry areas
needed where appropriate, or covers for seedlings and
ƒ Reinstatement of areas designated by local water supply
diversion roads – authorities ƒ Extend productive
consider transferring ƒ Repair to areas suffering planting to other areas
ownership for use by previous erosion or siltation identified by local
IMTs (particularly in busy/ damage community
dangerous areas) ƒ Improve access to the road ƒ Support programmes to
ƒ Reinstatement of – access roads, trail bridges, eradicate invasive plant
temporary work areas – footpaths species
eg provide designs for ƒ Improve access from the ƒ Supply fingerlings for
utilisation of borrow pits road to local community
borrow areas upgraded
to dams or fish ponds facilities
ƒ Provide opportunity ƒ Provide road maintenance to fish ponds
for community to training eg to lengthmen (or ƒ Build community/village
claim spoiled materials other technical training) to assets – school rooms,
including wood from local administrations/SME health or veterinary
grubbing, topsoil or ƒ Rehabilitation and repair posts, storage facilities,
oversize to community/village training/meeting rooms
ƒ Provide additional assets: roads, market areas, etc
soil protection and meeting areas, sanitation/ ƒ Utilise road drainage
road/structure erosion water supply facilities, systems to provide
protection in vulnerable drainage systems, etc. water-harvesting facilities
areas

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.2 - 24

Category 1 Management Category 2 Opportunity Category 3 Enhancement


Theme
Interventions Interventions Interventions
ƒ Ensure adequate ƒ Supply IMT to cooperatives/ ƒ Provide awareness
Transport Services physical access for associations training on options for
Pedestrian, NMT, IMT ƒ Provide IMT maintenance rural transport services
and normal RTS is training to cooperatives ƒ Provide seed financing
maintained ƒ Provide technical skills for establishment of
ƒ Provide adequate bus- training to local transport rotating funds for supply
bays and shelter service operators and maintenance of IMT
ƒ Make available mechanical ƒ Provide animal
workshops for IMT/RTS husbandry training
repairs
ƒ Provide vehicles and ƒ Provide HIV/AIDS testing ƒ Provide classroom
temporary emergency/ and counselling services furniture (desks and
first aid services for local along the road corridor for chairs)
communities whose construction workers and ƒ Promote use of ICT
access to mainline local communities in schools through
services is hindered by ƒ Distribution of first aid improved electrical
the construction works supplies to health posts and communications
ƒ Supply local health ƒ Assist with the repair, installations, provision of
Support Service Sectors

centres with ARVs, and rehabilitation or computers, etc.


other drugs relating to maintenance of health ƒ Provide mosquito nets
communicable disease and education facilities and mattresses to
control centres (incl. hospices & orphanages, hospices
orphanages) and nurseries
ƒ Provide support to
initiatives supporting
community education
and awareness (health,
safety, livelihoods and
income generation)
ƒ Provide water supply/
construct sanitation
facilities for roadside
communities
ƒ Maximise employment ƒ Provide advisory services ƒ Provide office furniture,
opportunities for local to local administration accommodation and
communities, including with regards construction, sanitation facilities for
women – provide crèche rehabilitation or community facilities
and other support maintenance of community ƒ Skills enhancement -
facilities infrastructure Train casual and other
Community Development

ƒ Provide ground water local labourers (eg better


recharge schemes, water livestock management;
harvesting or small micro- agricultural methods etc)
irrigation schemes ƒ Provide life skills training
ƒ Provide materials, (eg literacy, numeracy,
equipment and training to basic accounting, kitchen
support establishment and gardening, sanitation
development of local SMEs and hygiene, etc.) to
ƒ Supply materials, local community groups
equipment, labour, etc for and SMEs
community projects (eg
pipes, cement, steel, timber,
wiring, tractors, excavators,
skilled labourers, etc)

PART C: COMPLEMENTARY INTERVENTIONS


Appendix C.2 - 25

Category 1 Management Category 2 Opportunity Category 3 Enhancement


Theme
Interventions Interventions Interventions
ƒ Inclusion of Road ƒ Investigate different ƒ Provide technical

Research, Demonstration
Authority personnel on approaches to CI design and and management
contractors team for implementation to improve advice and training to
and Training professional training/ future provision and contract local authorities and
experience conditions administrations on key
ƒ Inclusion of trial/ ƒ Provide technical training issues
demonstration sections to local mechanics,
for new technical options electricians, plumbers,
carpenters, masons, etc. (eg
through employment and
maintenance at camp /work
sites)

PART C: COMPLEMENTARY INTERVENTIONS


)('(5$/'(02&5$7,&5(38%/,&2)
('(5$/'(02&5$7,&5(38%/,&2)

(7+,23,$152$'6$87+25,7<

,7 <
( 7+

25
,2

,$
15 87
+
3

2$'6 $

'(6,*10$18$/)25/2:92/80(52$'6
3$57'
),1$/'5$)7$35,/
PART D
EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN

Explanatory notes for low volume road design


general introduction

Site investigation for route selection and design

Roadside slope stabilisation

Part D
Explanatory notes for Geometric design
low volume road design

Drainage

Materials and pavement design

Surfacing
D-i

D TABLE OF CONTENTS

D. TABLE OF CONTENTS ....................................................................................................... D.I


D. LIST OF TABLES ............................................................................................................... D.VI
D. LIST OF FIGURES ............................................................................................................. D.IX
D. LIST OF PLATES ................................................................................................................ D.X
1. GENERAL INTRODUCTION .............................................................................................. D.1
2. SITE INVESTIGATION FOR ROUTE SELECTION AND DESIGN ......................................... D.2
2.1 Introduction .................................................................................................................... D.2
2.2 Stages of Site Investigation ............................................................................................ D.3
2.2.1 Desk Studies .................................................................................................... D.3
2.2.2 Identification and General Planning ................................................................ D.6
2.2.3 Pre-feasibility Study ......................................................................................... D.6
2.2.4 Feasibility Study or Preliminary Engineering Design ....................................... D.6
2.2.5 Scope of Investigations ................................................................................... D.7
2.3 Principal considerations for route selection ................................................................... D.8
2.3.1 General considerations and best practice ....................................................... D.8
2.3.2 Special Considerations in mountainous areas (see also Chapter D.3): ............ D.9
2.4 Site investigation techniques........................................................................................ D.10
2.5 Site investigation for final engineering design ............................................................ D.11
2.5.1 Characterisation of alignment soils and in-situ materials (subgrade) ............ D.11
2.5.2 Problem soils ................................................................................................. D.13
2.5.3 Location and characteristics of construction materials,
volumes available and haulage ..................................................................... D.17
2.5.4 Earthworks – Cut and fill investigation .......................................................... D.23
2.5.5 Water crossings.............................................................................................. D.24
2.5.6 Water sources ................................................................................................ D.24
2.6 References .................................................................................................................... D.25
3. ROADSIDE SLOPE STABILISATION ................................................................................ D.26
3.1 Introduction .................................................................................................................. D.26
3.2 Roadside slope instability ............................................................................................ D.26
3.3 Slope instability above the road ................................................................................... D.26
3.3.1 Types of slope failure above the road ........................................................... D.26
3.3.2 Cut slopes ...................................................................................................... D.29
3.4 Slope Instability below the road ................................................................................... D.30
3.4.1 Types of slope failures below the road .......................................................... D.30
3.4.2 Fill slopes ....................................................................................................... D.33
3.5 Drainage ....................................................................................................................... D.34
3.6 Retaining walls .............................................................................................................. D.35
3.6.1 Design of Retaining walls .............................................................................. D.35
3.6.2 Gabion walls .................................................................................................. D.36
3.6.3 Dry stone walls............................................................................................... D.36
3.6.4 Mortared masonry walls................................................................................. D.36
3.7 Bio-engineering methods............................................................................................. D.37
3.7.1 Selection of appropriate plant species .......................................................... D.37
3.7.2 Site preparation ............................................................................................. D.38
3.7.3 Recommended techniques ............................................................................ D.38
3.8 Useful dos and don’ts ................................................................................................... D.41
3.9 References: ................................................................................................................... D.42
4. GEOMETRIC DESIGN ..................................................................................................... D.43
4.1 Introduction .................................................................................................................. D.43
4.1.1 Principal factors affecting geometric standards ............................................ D.43
4.1.2 How the Standards are used ......................................................................... D.44
4.2 The principal factors determining the choice of geometric standard ......................... D.46
4.2.1 Traffic volume ................................................................................................ D.46

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - ii

4.2.2 The design vehicle ......................................................................................... D.47


4.2.3 Traffic composition – proportion of heavy vehicles ....................................... D.48
4.2.4 Traffic composition - use of Passenger Car Units (PCUs) ............................... D.48
4.2.5 Terrain ............................................................................................................ D.49
4.2.6 Roadside population (open country or populated areas) .............................. D.50
4.2.7 Pavement type ............................................................................................... D.50
4.2.8 Soil type and climate ..................................................................................... D.51
4.2.9 Safety ............................................................................................................. D.51
4.2.10 Construction technology ............................................................................... D.52
4.2.11 Administrative function .................................................................................. D.54
4.2.12 Matrix of standards ........................................................................................ D.54
4.3 Cross sections ............................................................................................................... D.54
4.3.1 Slopes of shoulders, side slopes, embankments ........................................... D.54
4.3.2 Roadside ditches ........................................................................................... D.56
4.3.3 Clear zones .................................................................................................... D.56
4.3.4 Right-of-way................................................................................................... D.56
4.3.5 Shoulders ....................................................................................................... D.57
4.3.6 Single lane roads and passing places ............................................................ D.58
4.3.7 Width standards ............................................................................................ D.58
4.4 Design speed and geometry ........................................................................................ D.59
4.4.1 Stopping sight distance ................................................................................. D.60
4.4.2 Stopping sight distance for single lane roads (meeting sight distance) ........ D.60
4.4.3 Intersection sight distance ............................................................................. D.61
4.4.4 Passing sight distances .................................................................................. D.61
4.4.5 Camber and cross-fall ................................................................................... D.61
4.4.6 Adverse cross-fall ........................................................................................... D.61
4.4.7 Super-elevation .............................................................................................. D.62
4.5 Horizontal alignment .................................................................................................... D.63
4.5.1 Curve length .................................................................................................. D.64
4.5.2 Curve widening ............................................................................................ D.64
4.6 Vertical alignment ......................................................................................................... D.65
4.6.1 Crest curves ................................................................................................... D.65
4.6.2 Sag curves...................................................................................................... D.66
4.6.3 Gradient ......................................................................................................... D.66
4.6.4 Hairpin stacks................................................................................................. D.66
4.7 Harmonisation of horizontal and vertical alignment..................................................... D.67
4.7.1 Situations to avoid ......................................................................................... D.67
4.7.2 Balance .......................................................................................................... D.68
4.7.3 Phasing .......................................................................................................... D.68
4.7.4 Junctions and Intersections ........................................................................... D.68
4.8 Safety ............................................................................................................................ D.69
4.8.1 Traffic calming ................................................................................................ D.70
4.8.2 Road markings, signage and lighting ............................................................ D.71
4.9 Traffic Signs .................................................................................................................. D.71
4.9.1 Warning signs ................................................................................................ D.71
4.9.2 Information signs ........................................................................................... D.72
4.10 Road Markings.............................................................................................................. D.72
4.10.1 Pavement markings ....................................................................................... D.72
4.10.2 Object markers ............................................................................................. D.72
4.10.3 Road studs ..................................................................................................... D.72
4.10.4 Marker Posts .................................................................................................. D.72
4.11 Lighting ........................................................................................................................ D.73
4.12 Safety barriers............................................................................................................... D.73
4.12.1 Segregating vulnerable road users ................................................................ D.73
4.12.2 Crash barriers................................................................................................. D.73
4.12.3 Safety audits .................................................................................................. D.73
4.13 Using the standards ...................................................................................................... D.74

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - iii

4.13.1 Upgrading an existing road ........................................................................... D.74


4.13.2 Designing a road to replace an existing track ............................................... D.74
4.13.3 Designing a new road .................................................................................... D.74
5. DRAINAGE ..................................................................................................................... D.75
5.1 Introduction .................................................................................................................. D.75
5.2 Summary of standards and departures from standards................................................ D.76
5.2.1 Design standards and storm return period.................................................... D.76
5.2.2 Methods of design......................................................................................... D.78
5.2.3 Departures from standards ............................................................................ D.78
5.3 Hydrology: estimating maximum flow for drainage design ......................................... D.79
5.3.1 General principles ........................................................................................ D.79
5.3.2 Storm severity and maximum flow ............................................................... D.79
5.3.3 Simple estimation methods for maximum flow ............................................. D.80
5.3.4 Normal flow rates .......................................................................................... D.85
5.3.5 Direct flow methods ...................................................................................... D.85
5.4 Components of external drainage ............................................................................... D.85
5.4.1 General principles.......................................................................................... D.86
5.4.2 Sources of water ............................................................................................ D.86
5.4.3 Road surface drainage ................................................................................... D.86
5.4.4 Side drains .................................................................................................... D.87
5.4.5 Erosion control in the side drain .................................................................... D.90
5.4.6 Mitre drains or turnouts ................................................................................. D.93
5.4.7 Wet lands ....................................................................................................... D.95
5.4.8 Interceptor, cut-off or catch-water drains. .................................................... D.96
5.4.9 Chutes............................................................................................................ D.98
5.5 Erosion control ............................................................................................................. D.98
5.5.1 Identifying and assessing potential erosion problems .................................. D.99
5.5.2 Mitigation measures to control erosion and scour ........................................ D.99
5.6 Particular drainage problems in severe terrain ........................................................... D.103
5.6.1 Drainage of hairpin stacks ........................................................................... D.103
5.6.2 Road construction along valley floors .......................................................... D.105
5.6.3 Freeboard .................................................................................................... D.105
5.6.4 Flood plain scour and embankment protection. ........................................ D.106
5.6.5 Cross drainage and tributary fan crossings.................................................. D.106
5.7 Potential references and/or bibliography ................................................................... D.109
6. MATERIALS AND PAVEMENT DESIGN ........................................................................ D.110
6.1 Introduction ................................................................................................................ D.110
6.2 Underlying principles.................................................................................................. D.110
6.2.1 Approach to low volume road pavement design ........................................ D.110
6.2.2 Pavement structure and function ................................................................. D.111
6.2.3 Pavement and surfacing options ................................................................. D.112
6.2.4 Road environmental factors ......................................................................... D.112
6.3 Climate ....................................................................................................................... D.113
6.3.1 Temperature and solar radiation.................................................................. D.113
6.3.2 Rainfall ......................................................................................................... D.114
6.3.3 Moisture in the road pavement ................................................................... D.115
6.3.4 Winds ........................................................................................................... D.116
6.4 Surface/Sub-surface hydrology .................................................................................. D.116
6.5 Subgrade .................................................................................................................... D.117
6.5.1 Subgrade Classification .............................................................................. D.117
6.5.2 Specifying the design subgrade class ......................................................... D.118
6.5.3 Material depth ............................................................................................ D.118
6.5.4 Dealing with poor subgrade soils ................................................................ D.119
6.5.5 Improved subgrade layers ........................................................................... D.120
6.6 Traffic .......................................................................................................................... D.120
6.6.1 Estimating design traffic loading ................................................................. D.120
6.6.2 Equivalent standard axles per vehicle class ................................................. D.121

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - iv

6.6.3 Design traffic classes .................................................................................. D.124


6.7 Construction Materials ............................................................................................... D.125
6.7.1 Performance characteristics of the pavement materials ............................. D.125
6.7.2 Pavement material types ............................................................................. D.127
6.7.3 Materials requirements for road base .......................................................... D.129
6.7.4 Material requirements for sub-base ............................................................. D.132
6.7.5 Material requirements for gravel wearing course ........................................ D.133
6.7.6 Material Improvement ................................................................................. D.136
6.8 Terrain ......................................................................................................................... D.139
6.9 Construction regime ................................................................................................... D.139
6.10 Maintenance regime................................................................................................... D.140
6.11 Road safety regime..................................................................................................... D.140
6.12 The ‘green’ environment ............................................................................................ D.140
6.13 Environmentally optimised design (EOD) .................................................................. D.140
6.14 Structural design ....................................................................................................... D.141
6.14.1 Reliability and terminal condition ................................................................ D.141
6.14.2 Design and analysis period.......................................................................... D.142
6.14.3 Upgrading Strategy ..................................................................................... D.142
6.15 Design of earth roads ................................................................................................. D.143
6.15.1 Rationale for ENS......................................................................................... D.144
6.15.2 Design criteria.............................................................................................. D.144
6.15.3 Estimating traffic capacity ............................................................................ D.145
6.15.4 Construction ................................................................................................ D.146
6.15.5 Maintenance ............................................................................................... D.148
6.16 Design of gravel roads ............................................................................................... D.149
6.16.1 Design method ........................................................................................... D.150
6.16.2 Major and minor gravel roads. ................................................................... D.151
6.16.3 The structural design procedure.................................................................. D.151
6.16.4 Moisture regime and determination of material strength .......................... D.152
6.16.5 Material classification .................................................................................. D.152
6.16.6 Selection of appropriate pavement structure ............................................. D.152
6.16.7 Determination of wearing course thickness................................................. D.153
6.16.8 Minor gravel roads ....................................................................................... D.154
6.17 Structural design of paved roads ............................................................................. D.155
6.17.1 Design methods .......................................................................................... D.155
6.17.2 Design method for Bituminous surfaced roads ........................................... D.155
6.17.3 Use of the design charts .............................................................................. D.157
6.17.4 Design method for non-bituminous surfaced road ..................................... D.158
6.18 Drainage and Shoulders ............................................................................................ D.159
6.18.1 Sources of Moisture Entry into a Pavement ................................................. D.159
6.18.2 Permeability ................................................................................................. D.160
6.18.3 Achieving effective internal drainage .......................................................... D.160
6.19 Problem Soils .............................................................................................................. D.165
6.19.1 Performance risk .......................................................................................... D.165
6.19.2 Expansive soils ............................................................................................. D.165
6.19.3 Collapsible soils ........................................................................................... D.169
6.19.4 Dispersive/erodible soils.............................................................................. D.170
6.19.5 Saline soils ................................................................................................. D.171
6.19.6 Micaceous soils ............................................................................................ D.172
6.19.7 Low-strength soils ........................................................................................ D.173
6.20 Construction Issues .................................................................................................... D.173
6.20.1 Subgrade compaction ................................................................................. D.173
6.20.2 Quality Attainment ...................................................................................... D.176
6.21 References: ................................................................................................................. D.177
7. SURFACING .................................................................................................................. D.179
7.1 Introduction ................................................................................................................ D.179
7.2 Types of Surfacings ..................................................................................................... D.179

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D-v

7.2.1 Basic Surfacings ........................................................................................... D.180


7.2.2 Stone Paving ................................................................................................ D.180
7.2.3 Fired Clay Brick ........................................................................................... D.184
7.2.4 Bituminous surfacings .................................................................................. D.184
7.2.5 Concrete Surfacings..................................................................................... D.195
7.3 Choice of pavement and surfacing ............................................................................ D.197
7.3.1 Evaluation framework .................................................................................. D.198
7.3.2 SDMS Procedure.......................................................................................... D.199
D. APPENDIX D.1 ............................................................................................................. D.208
D. APPENDIX D.2 ............................................................................................................. D.257
1.1 Basic .......................................................................................................................... D.261
1.2 Stone Paving .............................................................................................................. D.261
D. APPENDIX D.3 ............................................................................................................. D.261
1.3 Bituminous Surfacings ................................................................................................ D.263
1.4 Concrete .................................................................................................................... D.264

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - vi

D LIST OF TABLES

Table D.2.1: Existing data sources relevant to road construction .....................................................D.4


Table D.2.2: Summary of survey requirements for route selection and design .................................D.7
Table D.2.3: Classification of expansive soils according to US Bureau of Reclamation ..................D.15
Table D.2.4: Qualitative assessment of collapse potential ..............................................................D.16
Table D.2.5: Typical laboratory tests used to assess the suitability of alignment soils
and pavement materials ..............................................................................................D.20
Table D.2.6: Weathering grades for road design and construction ................................................D.22
Table D.3.1: Stabilisation and protection measures for slope failures and erosion above the road
(based on Hunt et al 2008) ..........................................................................................D.28
Table D.3.2: Common cut-slope ratios for LVRs. .............................................................................D.29
Table D.3.3: Stabilisation and protection measures for slope failures below the road (Hunt et al.,
2008) ...........................................................................................................................D.32
Table D.3.4: Common fill slope batters for LVRs .............................................................................D.33
Table D.3.5: Recommended bio-engineering techniques for slopes above the road. ...................D.39
Table D.3.6: Bio-engineering methods useful to improve slope stability below the road ..............D.40
Table D.3.7: Plants useful for erosion control in the surroundings of roads ....................................D.41
Table D.4.1: Design vehicle characteristics .....................................................................................D.48
Table D.4.2: Design vehicle for each LVR class ...............................................................................D.48
Table D.4.3: PCU values ..................................................................................................................D.49
Table D.4.4: Safety of slopes (ratios are vertical:horizontal) ............................................................D.56
Table D.4.5: Right of way widths .....................................................................................................D.57
Table D.4.6: Design speeds ............................................................................................................D.60
Table D.4.7: Stopping sight distances (m) .......................................................................................D.60
Table D.4.8: Passing sight distances (m)..........................................................................................D.61
Table D.4.9: Adverse cross-fall to be removed if radii are less than shown ....................................D.62
Table D.4.10: Super-elevation development lengths ........................................................................D.63
Table D.4.11: Recommended minimum horizontal radii of curvature: paved roads (m) ...................D.64
Table D.4.12: Recommended minimum horizontal radii of curvature: unpaved roads (m) ...............D.64
Table D.4.13: Widening recommendations (m) .................................................................................D.65
Table D.4.14: Minimum values of L/G for crest curves ......................................................................D.65
Table D.4.15: Minimum values of L/G for sag curves ........................................................................D.66
Table D.5.1: Design storm return period (years) .............................................................................D.77
Table D.5.2: Design storm return period (years) for severe risk situations ......................................D.78
Table D.5.3: Advantages and disadvantages of the Rational Method ............................................D.82
Table D.5.4: Catchment area and total cross-sectional area of an example structure
for a set of standard conditions based on the Rational method .................................D.83
Table D.5.5: Catchment area, total cross-sectional area and number of standard arch
culverts for a set of standard conditions based on the Rational Method ...................D.84
Table D.5.6: Spacing between scour checks ...................................................................................D.91
Table D.5.7: Permissible flow velocities (m/sec) in excavated ditch drains .....................................D.92

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - vii

Table D.5.8: Maximum spacing of mitre drains ...............................................................................D.94


Table D.5.9: Comparison of various slope erosion mitigation measures ......................................D.100
Table D.6.1: Ethiopia’s climatic zones............................................................................................D.113
Table D.6.2: Subgrade classes ......................................................................................................D.117
Table D.6.3: Design CBR values related to Traffic Classification ...................................................D.118
Table D.6.4: Typical material depth by road category ..................................................................D.119
Table D.6.5: Load equivalency factors for different axle load groups (esas) .................................D.123
Table D.6.6: Average equivalency factors for different vehicle types ..........................................D.124
Table D.6.7: Factors for design traffic loading ..............................................................................D.124
Table D.6.8: Traffic classes for pavement design ..........................................................................D.125
Table D.6.9: Pavement material categories and relative characteristics .......................................D.126
Table D.6.10: Variation of CBR with moisture content ....................................................................D.127
Table D.6.11: Pavement material types and abbreviated nominal specifications used in
the paved and unpaved catalogue of designs..........................................................D.128
Table D.6.12: Particle size distribution for natural gravel base .......................................................D.129
Table D.6.13: Plasticity requirements for natural gravel road base materials..................................D.130
Table D.6.14: Guidelines for the selection of lateritic gravel road base materials ..........................D.131
Table D.6.15: Typical particle size distribution for sub-bases ..........................................................D.133
Table D.6.16: Plasticity characteristics for granular sub-bases ........................................................D.133
Table D.6.17: Typical standardised gravel loss ................................................................................D.135
Table D.6.18: Recommended material specifications(1,3) for unsealed rural roads ..........................D.135
Table D.6.19: Recommended material specifications for unsealed ‘urban’ roads .........................D.136
Table D.6.20: Required minimum height (hmin) between road crown and invert level of drain in
relation to climate .....................................................................................................D.150
Table D.6.21a: Catalogue for major gravel roads – strong gravel (G45)............................................D.152
Table D.6.21b: Catalogue for major gravel roads – medium gravel (G30) ........................................D.152
Table D.6.21c: Catalogue for major gravel roads – weak gravel (G15)..............................................D.153
Table D.6.22: Typical standardised gravel loss ................................................................................D.153
Table D.6.23: Typical standardised gravel loss ................................................................................D.155
Table D.6.24: Design traffic classes .................................................................................................D.157
Table D.6.25: Typical causes of water ingress to, and egress from a road pavement .....................D.159
Table D.6.26: Typical material permeabilities (Lay, 1998) ................................................................D.160
Table D.6.27: Classification of road drainage..................................................................................D.161
Table D.6.28: Countermeasures for dealing with expansive soils ...................................................D.167
Table D.6.29: Minimum compaction requirements .........................................................................D.176
Table D.7.1: Typical bituminous surfacing service lives1 ...............................................................D.188
Table D.7.2: Factors affecting choice of bituminous surface treatments.......................................D.189
Table D.7.3: Typical prime application rates in relation to pavement surface type.......................D.190
Table D.7.4: Grading of sand for use in sand seal .........................................................................D.191
Table D.7.5: Binder and aggregate application rates for sand seals.............................................D.191
Table D.7.6: Gives a nominal slurry seal mix. ................................................................................D.191
Table D7.7: Aggregate requirements for Chip Seals....................................................................D.192

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - viii

Table D.7.8: Binder and application rates for Chip Seals ..............................................................D.193
Table D.7.9: Nominal Application rates for single Chip Seals .......................................................D.194
Table D.7.10: Choice of binder in relation to traffic and grading ...................................................D.194
Table D.7.11: Nominal binder application rates for Otta Seal ........................................................D.195
Table D.7.12: Nominal aggregate application rates .......................................................................D.195
Table D.7.13: Definition of Indicative Traffic Regime ......................................................................D.203
Table D.7.14: Definition of Erosion Potential ..................................................................................D.203
Table D.7.15: Preliminary engineering filter - surfacing ..................................................................D.204
Table D.7.16: Primary engineering filter - pavement layers / shoulders..........................................D.205
Table D.7.17: Primary engineering filters (continued) - surfacing....................................................D.206
Table D.7.18: Secondary engineering filters - pavement layers / shoulders ..................................D.207

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - ix

D LIST OF FIGURES

Figure D.2.1: Guide to collapsibility and expansion based on in-situ dry density and liquid limit
(after Mitchell and Gardner, 1975 and Gibbs, 1969) ...................................................D.14
Figure D.2.2: A simple way of testing the dispersive nature of soils ................................................D.17
Figure D.3.1: Slope failures that occur above the road (modified from Hunt et al., 2008). ..............D.27
Figure D.3.2: Slope instabilities expected below the road (modified from Hunt et al., 2008). ........D.31
Figure D.4.1: Selection procedure for appropriate geometric standards .........................................D.45
Figure D.4.2: Examples of effects of engineering design on road safety ........................................D.53
Figure D.4.3: Details of the road edge .............................................................................................D.55
Figure D.4.4: Removal of adverse camber........................................................................................D.62
Figure D.4.5: Preferred intersection design ......................................................................................D.69
Figure D.5.1: Example of a standard arch culvert .............................................................................D.83
Figure D.5.2: Inadequate side drains ................................................................................................D.87
Figure D.5.3: Inadequate side drains and subsurface drainage .......................................................D.88
Figure D.5.4: Proper interception of surface runoff and subsurface seepage ..................................D.88
Figure D.5.5: Side drains ..................................................................................................................D.88
Figure D.5.6: Typical design of scour checks ....................................................................................D.91
Figure D.5.7: Angle of mitre drain ....................................................................................................D.94
Figure D.5.8: Mitre drain angle greater than 45 degrees .................................................................D.95
Figure D.5.9: Typical sub-surface drain .............................................................................................D.96
Figure D.5.10: An interceptor, cut-off or catch-water drain ................................................................D.97
Figure D.5.11: Size of stone that will resist displacement for various velocities of water flow and
side slopes ................................................................................................................D.101
Figure D.5.12: Use of riprap 1...........................................................................................................D.102
Figure D.5.13: Use of riprap 2...........................................................................................................D.102
Figure D.5.14: Suggested apron details for side drain turnout ........................................................D.104
Figure D.6.1: Pavement design system ..........................................................................................D.111
Figure D.6.2: Wheel load transfer through pavement structure .....................................................D.112
Figure D.6.3: Road environment factors .........................................................................................D.113
Figure D.6.4: Rainfall pattern in Ethiopia ........................................................................................D.114
Figure D.6.5: Climatic N-value map for Ethiopia ............................................................................D.116
Figure D.6.6: Illustration of CBR strength cumulative distribution..................................................D.118
Figure D.6.7: Material depth...........................................................................................................D.119
Figure D.6.8: Procedure to determine design traffic loading .........................................................D.121
Figure D.6.9: Material quality zones ...............................................................................................D.134
Figure D.6.10: Ternary diagram for blending unsealed road materials ............................................D.138
Figure D.6.11: Application of the principle of environmentally optimised design ...........................D.141
Figure D.6.12: Design reliability in relation to road category and terminal surface condition ........D.142
Figure D.6.13: Components of a typical EOD designed LVR ..........................................................D.143
Figure D.6.14: Carrying capacity of soils .........................................................................................D.145

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D-x

Figure D.6.15: Road built above ground level – incorrect ................................................................D.147


Figure D.6.16: The “high level method” - correct ............................................................................D.147
Figure D.6.17: Road built at ground level - incorrect .......................................................................D.147
Figure D.6.18: (a) – Erosion in wheel tracks ......................................................................................D.148
Figure D.6.18: (b) – Erosion in side drains ........................................................................................D.148
Figure D.6.19: (a) - Incorrectly maintained ENS road (Odier et al, 1971) .........................................D.149
Figure D.6.19: (b) - Correctly maintained ENS road (Odier et al, 1971) ...........................................D.149
Figure D.6.20: Gravel road - Typical pavement layers .....................................................................D.149
Figure D.6.21: Typical gravel road cross section in flat terrain. ........................................................D.150
Figure D.6.22: Paved road - Typical pavement layers .......................................................................D.155
Figure D.6.23: Flow chart for bituminous surfaced sealed road pavement design process .............D.156
Figure D.6.24: Traffic loading versus dominant mechanism of pavement distress (Schematic only) D.157
Figure D.6.25: Recommended drainage arrangements ..................................................................D.162
Figure D.6.26: Typical drainage deficiencies associated with pavement shoulder construction (adapted
from Birgisson and Ruth, 2003 ................................................................................. D.163)
Figure D.6.27: Moisture zones in a typical LVR .................................................................................D.164
Figure D.6.28: Moisture movements in expansive soils under a paved road ...................................D.166
Figure D.6.29: Location of side drains in expansive soils .................................................................D.168
Figure D.6.30: Construction on expansive soils (use of pioneer layer) .............................................D.168
Figure D.6.31: Construction on expansive soils (embankment height <2m) ...................................D.169
Figure D.6.32: Construction on expansive soils (embankment height >2m) ....................................D.169
Figure D.6.33: Manner of additional settlement due to collapse of soil fabric ................................D.170
Figure D.6.34: Illustration of concept “compaction to refusal” ........................................................D.174
Figure D.6.35: Deflection-life relationship and benefits of “compaction to refusal”........................D.174
Figure D.6.36: Extended compaction curve for low moisture contents (Parsons, 1992) ..................D.175
Figure D.6.37: Air voids in dry and well compacted soil (Parsons, 1992) .........................................D.175
Figure D.7.1: Examples of typical surface treatments.....................................................................D.185
Figure D.7.2: Differing mechanisms of performance of surface treatments .................................D.187
Figure D.7.3: General road surface selection factors ......................................................................D.197
Figure D.7.4: Overview of the SDMS procedure ............................................................................D.198
Figure D.7.5: Overview of SDMS procedure...................................................................................D.199
Figure D.7.6: Decision Flow Chart for the Preliminary Consideration of
LVR Surface Options for a road section – STEP 1......................................................D.200
Figure D.7.7: Decision Flow Chart for the Preliminary Consideration of
LVR Surface Options for a road section – Step 1 continued .....................................D.201
Figure D.7.8: Decision Flow Chart for the Preliminary Consideration of
LVR Surface – Step 1 continued ................................................................................D.202

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - xi

D LIST OF PLATES

Plate D.6.1: Severe erosion of road side slopes in high rainfall area ............................................D.115
Plate D.6.2: Grid roller ..................................................................................................................D.136
Plate D.6.3: Rock Buster ...............................................................................................................D.137
Plate D.6.4: Typical unformed earth road .....................................................................................D.144
Plate D.6.5: Typical formed earth road .........................................................................................D.144
Plate D.6.6: Manual excavation of side ditch material to form ENS camber
(prior to spreading and compaction) ........................................................................D.146
Plate D.6.7: Example of a well-drained pavement where the drainage is
classified as “good”, ie the drainage factor DF (d x h) >7.5 ....................................D.161
Plate D.6.8: Infiltration of water through a permeable surfacing and subsequent outflow to an
impermeable shoulder ..............................................................................................D.164
Plate D.6.9: Expansive “black cotton” soil exhibiting wide-spaced shrinkage cracks ..................D.166
Plate D.6.10: Typical longitudinal cracking and pavement deformation caused by an
expansive soil subgrade ............................................................................................D.166
Plate D.6.11: Collapse settlement in excess of 150mm following impact compaction ..................D.170
Plates D.6.12 and D.6.13: Examples of severe erosion in erodible/dispersive soils in Ethiopia........D.171
Plate D.6.14: An example of severe distress to a road surfacing due to salt attack
resulting in damage within two years of its construction (Botswana) ........................D.172
Plate D.6.15: Salt damage may appear in the form of “blistering”, “heaving” and
“fluffing” of the prime surfacing. ..............................................................................D.172
Plate D.6.16: Three-sided impact compactor .................................................................................D.175
Plate D.7.1: An example of an Engineered Natural Surface (ENS) ..............................................D.180
Plate D.7.2: An example of a Natural Gravel Surfacing (Cinder gravel) .......................................D.180
Plate D.7.3: An example of a Waterbound/Drybound Macadam .................................................D.181
Plate D.7.4: Hand-packed Stone ..................................................................................................D.181
Plate D.7.5: Stone Setts (Linear pattern) .......................................................................................D.182
Plate D.7.6: Stone Setts (Radial pattern).......................................................................................D.182
Plate D.7.7: Mortared Stone .........................................................................................................D.183
Plate D.7.8: Dressed Stone/Cobble Stone ...................................................................................D.183
Plate D.7.9: Fired clay brick ..........................................................................................................D.184
Plate D.7.10: Concrete slab surfacing.............................................................................................D.196
Plate D.7.11: Ultra Thin Reinforced Concrete Paving .....................................................................D.197

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 1 - 1

1. GENERAL INTRODUCTION

This Part of the manual provides the following chapters of explanatory notes and supporting information
that should be considered during the design process and provides background to the standards shown
in Part B of the manual:
ƒ Chapter 2: Site investigation for route selection and design
ƒ Chapter 3: Roadside slope stabilisation
ƒ Chapter 4: Geometric design
ƒ Chapter 5: Drainage
ƒ Chapter 6: Materials and pavement design
ƒ Chapter 7: Surfacing

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 2 - 2

2. SITE INVESTIGATION FOR ROUTE SELECTION AND DESIGN

2.1 Introduction
Site investigation is an integral part of the location, design and construction of a road and provides
essential information on the alignment soil characteristics, construction materials availability, topography,
land use, environmental issues (including climate) and socio-political considerations for the client related
to the following:
ƒ Selection of the route/alignment of the road;
ƒ Location of water crossings and drainage structures;
ƒ Design information for the road pavements, bridges and other structures;
ƒ Areas for specialist geotechnical investigation;
ƒ Areas of potentially problematic soils requiring additional investigation and treatment;
ƒ Location and assessment of suitable, locally available borrow and construction material.

This list indicates that the main component of site investigations is focussed on what is generally described
as ‘engineering’ or, more precisely, ‘geotechnical engineering’. However, various other types of survey
are required. Hydrological surveys are required to determine the water flows that determine the drainage
design of the road; traffic surveys are required to estimate the numbers of vehicles, both motorised and
non-motorised, that will use the road; surveys are required to evaluate environmental impacts and how to
control them; surveys are required in which the local communities are consulted about the road project;
and so on. This chapter deals primarily with the engineering surveys. Surveys required for these specialist
purposes are described in the chapters dealing with those topics.

Information captured during the site investigation is used by the design engineer to prepare and refine
the detailed engineering design. This information is usually captured within a series of documents that
are prepared by the design engineer, initially for consideration by the client and ultimately to develop
the tender and draft contract documents. These documents would normally include separate volumes
dealing with the following design aspects:
ƒ Alignment Survey and Geometric Design;
ƒ Traffic and Traffic Loading;
ƒ Materials and Subgrade Design;
ƒ Pavement Design;
ƒ Hydrology, Drainage and Water Crossings;
ƒ Ground Stability and Geotechnical Design;
ƒ Environmental (EIA and outline EMP);
ƒ Social and Complementary Activities;
ƒ Engineer’s Cost Estimate.

Road projects fall into one of the following categories:


ƒ A new road following the general alignment of an existing track or trail;
ƒ Upgrading a lower class of road to a higher class;
ƒ A completely new road where nothing currently exists.

Some realignment, and therefore site investigation, will almost certainly be necessary when upgrading an
existing road and considerably more will be required when converting a track into an all-weather route.
Major site investigations are usually only needed when designing and building a completely new road.
In all cases the extent and quality of any investigation has a strong influence on the selection of the most
cost-effective route and road design.

Low volume roads, of all standards, require sufficient investigation to provide enough data and information
that enables the engineer to optimise the design. In this respect, it is the job of the design engineer to
ensure that a well-designed and organised site investigation is undertaken. The design engineer must
therefore specify a site investigation programme for the site investigation teams (survey, materials,

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 2 - 3

geotechnical, socio-environmental) that will provide adequate information and data to examine the
feasibility of all the routes and designs under consideration.

The focus of this chapter is to provide guidance on the appropriate type and level of site investigation
that is required for route selection and subsequent design of low volume roads. The chapter also provides
practitioners with the necessary tools to develop suitable site investigation programmes and in-situ
testing schedules and with assistance in interpreting the data obtained.

Site investigation techniques encompass a large range of methods and the amount and type of exploration
that is needed for a specific road will depend on the nature of the proposed project and the environment
in which it is to be built. It is not the purpose of this chapter to explain individual site investigation
techniques. Where information on the type, use and interpretation of site investigation techniques is
needed, the reader is referred to the Site Investigation Manual (ERA 2011) that covers site investigation
procedures for all roads. This complementary manual provides information on the general distribution
of local materials in Ethiopia and an explanation of the physiography, geology, terrain, climate, and soil
distribution in the country.

This Chapter only deals with the investigation of the site in terms of gathering relevant and appropriate
engineering information for the selection of the most suitable route and the subsequent design of the
road along that route. Site investigations for low level water crossings are dealt with in Part E of the
manual. Where complimentary interventions are required to increase the positive impact of the road
project under consideration, the design consultant should be familiar with the requirements in Part C of
the manual.

2.2 Stages of Site Investigation


Some form of site investigation is required at all stages in the development of a road project. In general
there are four stages leading up to and including Final Engineering Design. These are:
ƒ Identification and general planning
ƒ Pre-feasibility study
ƒ Feasibility Study or Preliminary Engineering Design
ƒ Final Engineering Design

These stages are described briefly in the following sections together with details of the site investigations
associated with each stage. The final or detailed engineering design is dealt with in section 2.5. Not all
stages will be required for all projects, particularly for projects for upgrading a road from a lower class to
a higher class.

2.2.1 Desk Studies

Before any ground survey is carried out and, indeed, before such a survey can be planned and designed,
it is vital to study all the relevant information that is available about the project area. This is done through a
systematic desk study which entails the collection of detailed information for review and analysis. It allows
checking the suitability of all environmental and engineering conditions along different route options.
Studying existing documents, including site investigations from earlier project phases, and examining
maps and aerial photographs often eliminates an unfavourable route from further consideration, thus
saving a considerable amount of time and money. Topographic maps give essential information about
the relief of an area, and whether or not there are any existing routes. Aerial photographs provide a quick
means for preparing valuable sketches and overlays for reconnaissance/field surveys.

There are a number of very helpful sources of information in Ethiopia that can and should be used for this
purpose. Table D.2.1 provides names of federal and other local agencies in Ethiopia where data relevant
for site investigation can be obtained.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 4

Table D.2.1: Existing data sources relevant to road construction

Information
Functional Use Location Examples
Resource
ƒ Additional data and ƒ Ethiopian Roads ƒ Site Investigation
information on aspects Authority (ERA). procedures detailed
of geometry, drainage, in ERA 2011 Site
pavement and materials Investigation Manual;
ERA Design
and structural design ƒ Data, Information
Manuals
of roads and bridges in and Maps on
Ethiopia climate, terrain, soils,
construction materials
etc.
ƒ Previous road (and ƒ Ethiopian Roads ƒ Road Design and
other engineering) Authority (ERA); Rehabilitation
investigations in the ƒ Regional / Rural Reports;
locality will provide Roads Authorities; ƒ Maintenance
a range of data and ƒ Wereda and other Planning and
information that can local Administrations; Activity Reports may
Road and supplement project ƒ Office of the Road provide geological,
Other design such as: soil Fund; hydrogeological,
Engineering and rock type, strength ƒ Transport and geotechnical
Reports parameters, hydro- Construction Design information for the
geological issues, Share Company general area that may
construction materials (TCD) and other reduce the scope
etc; local agencies/ or better target the
ƒ Information on local engineering firms. nature of the site
road performance and investigation.
issues.
ƒ Identifies manmade ƒ Ethiopian Mapping ƒ Evaluating a series of
structures, potential Agency (EMA); aerial photographs
borrow source areas; ƒ Information Network may save time during
ƒ Provides geologic Security Agency construction material
Aerial and hydrological (INSA); survey.
Photographs information which can ƒ Other International
be used as a basis for Agencies such as
site reconnaissance; Quick Bird; IKONOS
ƒ Track site changes over and Google Earth.
time.
ƒ Provides good index ƒ Ethiopian Mapping ƒ Engineer identifies
map; Allows estimation Agency (EMA); access areas and
of site topography; ƒ Google Earth. restrictions, identifies
Topographic ƒ Identifies physical areas of potential
Maps features; Can be slope instability; and
used to assess access can estimate cut and
restrictions fill before visiting the
site.
ƒ Provides information ƒ Geological Survey of ƒ A report on regional
on nearby soil and rock Ethiopia (GSE); geology identifies
Geologic
type and characteristics, ƒ Ethiopian Mapping rock types, fracture
Reports and
ƒ Hydro-geological Agency (EMA). and orientation and
Maps
issues, groundwater flow
ƒ Environmental concerns patterns.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 5

Information
Functional Use Location Examples
Resource
ƒ Local soil types; ƒ Ministry of ƒ The local soil survey
ƒ Permeability of local Agriculture; provides information
soils; ƒ Geological Survey of on near-surface
ƒ Climatic and geologic Ethiopia (GSE); soils to facilitate
Soil maps information. ƒ Ethiopian Mapping preliminary borrow
Agency (EMA); source evaluation.
ƒ Local soil
conservation and
research institutes.
ƒ Mean Annual/Monthly; ƒ National ƒ Climate controls the
Rainfall and distribution Metrological Agency degree and type
ƒ Maximum and minimum of Ethiopia. of weathering and
Meteorological
temperatures; influences the type
and Climatic
ƒ Evaporation rates; of soils and materials
data
ƒ Weinert-N value and in the locality.
Thornthwaite Moisture
Index.
ƒ Distributional and type ƒ Ministry of ƒ Land use or land
of : Agriculture and local cover maps assist to
‚ Soils; Administrations; identify the physical
‚ Road; ƒ Universities and and biological
Land use /
‚ Drainage and water research institutes. cover over the land,
land cover
courses; including water,
‚ Agriculture and vegetation, bare
Forest. soil, and artificial
structures.
ƒ Traffic classification, ƒ Regional, Wereda ƒ Identification of
Seasonal traffic and Kebele specific problems
variation, road user Administrations. and hazards along
demand, hazards and proposed alignment;
ground instability, local ƒ Local sources
Local
road performance and of materials
Knowledge
maintenance history, and previous
accident black spots, performance.
water sources, local
weather conditions and
drainage characteristics.
ƒ Population data and ƒ Regional, Wereda ƒ Statistical data,
demographics; and Kebele information and maps
ƒ Village distribution and Administrations; to optimise route
pattern; ƒ Central Statistical alignment;
Statistics and
ƒ Socio-Economic and Agency. ƒ Recognition of
Future Plans
household survey planned future
information; activities within
ƒ Development Plans. vicinity of planned
road corridor.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 6

2.2.2 Identification and General Planning

This is the stage at which the need for the project is identified and projects that do not meet selection
criteria defined by the appropriate authorities are rejected. For DC1 and DC2 low volume roads this will
usually be done at a relatively local level and will be the output of a planning process. It is likely that only
a desk study and possibly a simple reconnaissance survey will be carried out.

2.2.3 Pre-feasibility Study

This is the stage where a broad economic and engineering assessment is made. It is at this stage that
the main engineering problems and any other issues affecting the route are identified (for example,
environmental and cultural issues) and likely corridors for the proposed road selected.

As part of the pre-feasibility study stage it is important to identify and investigate the major technical,
environmental, economic and social constraints through a reconnaissance survey in order to obtain a
broad appreciation of the viability of the competing alignment options. For low volume roads, one of the
most important aspects of the pre-feasibility study is communication with the people who will be affected
by the road. Their views are vital for the completion of a successful project and interacting with them is
essential right from the outset.

A reconnaissance survey provides data that enables specialists to study the advantages and disadvantages
of a variety of routes and then to determine which routes should be considered for further investigation.
It is an opportunity for checking the actual conditions on the ground and for noting any discrepancies
in the maps or aerial photographs. During this survey, it is necessary to make notes of soil conditions,
especially potentially soil problems; availability of construction materials; unusual grade or alignment
problems, water crossings and potential drainage problems; and requirements for clearing and grubbing.
It is also very useful to take photographs or make sketches of reference points, structure sites, landslides,
washouts, or any other unusual circumstances.

A desk study comprises the first step of the site investigations followed by a reconnaissance survey plus
some additional testing to confirm the scale of any significant engineering problems within the potential
corridors.

For the lower classes of roads (DC 1 and 2), predominantly unpaved earth or gravel roads, the information
from the pre-feasibility study may be the only available data to assist in the design of the road due to
financial constraints, hence, it is important to bear this in mind when designing the survey that is to be
undertaken. The outputs of the pre-feasibility study for DC1 and DC2 gravel roads should be a single
selected alignment for possible further investigation at the feasibility stage if required. For paved roads
and most DC3 and DC 4 gravel roads, more than one viable alignment option should be available.

Although not covered in this part of the manual (see Part C), the importance of community participation
at this stage, and throughout the project, cannot be over-emphasised as an input to the route selection,
design and development of complimentary interventions related to a project.

2.2.4 Feasibility Study or Preliminary Engineering Design

At this stage sufficient data are required to identify the final choice of route and the structural design of
the road. The feasibility study survey consists, essentially, of mapping the terrain along the centre-line
of the viable route or routes identified at the pre-feasibility stage. Data are required that are sufficient
to obtain likely costs to an accuracy of better than about 25%. General costs for similar roads that have
been built recently may be used for much of the assessment but the costs for major structures such as
bridges and major earthworks need to be estimated sufficiently accurately hence the extent of the site
investigation programme is dictated by these requirements.

After the Feasibility Study there should be sufficient information for the final route alignment to be
selected. Minor adjustments to the route alignment(s) may still be necessary during design, but the
number of iterations needed to establish the best alignment and confirm the choice of the route should
decrease significantly.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 7

In some cases the choice of final route alignment might depend on factors other than just the engineering
factors. Considerations such as environmental issues, numbers of people within a minimum distance from
the road, proximity to historic, religious or other cultural sites and so on (also see Part C) might override
the basic economic analysis. Decisions based on some form of multi criteria analysis are available and
could be used by those responsible if required.

2.2.5 Scope of Investigations

Table D.2.2 summarises the level of detail generally required for the site surveys and other investigations
for low volume roads at each stage of the road design process.

Table D.2.2: Summary of survey requirements for route selection and design

Stage of
Study DC1 and DC2 DC3 and DC4
design
Very broad brush approach based Probably a road or track in
on historic records. Usually a desk existence already. Very broad
Engineering study only. brush approach based on historic
records. Major engineering
problems identified
Identification

The need for the road will have been based on the current planning
process at regional or local level. Social assessment based on desk
Social
study information and concentrated on major issues such as land take
and resettlement.
Assessment based on desk study information but concentrated on
Environment
major issues such as land take and re-instatement.
Cost Historic data only. Based principally on terrain and number of
estimation structures. Accurate to only ±100%
In most cases there will be only Each option is broadly specified in
one route option identified terms of alignment, geometric and
through dialogue with the local pavement design and structures.
community and the design team. Limited geotechnical surveys may
Any major problem areas must be need to be undertaken together
Engineering identified and avoided if possible. with historic surveys to identify
basic designs and availability
of materials. Limited evaluation
of drainage conditions is also
Pre -feasibility

required to identify likely numbers


and sizes of drainage structures.
Essential to engage local communities in dialogue concerning the
Social impact of the road. Details now required of land-take and resettlement
for each option
Many common environmental issues associated with major roads are
Environment unlikely to be significant for Low volume roads but attention must be
paid to borrow and spoil areas and likely changes in drainage patterns.
Largely based on historic records
An accuracy of ±50% or better
Cost but now supplemented with
should be possible with the data
estimation more detail about the scale (and
available and historic information
therefore likely cost) of structures.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 8

Stage of
Study DC1 and DC2 DC3 and DC4
design
In most cases there will be only Detailed geotechnical surveys may
one option identified through need to be undertaken in unstable
dialogue with the local community mountainous terrain and for major
hence no additional survey data structures. A hydrological study
are required to select options. may be needed if substantial
rivers are to be crossed. The
Engineering
availability of materials should be
confirmed. The data should be
sufficient for the preferred option
to be selected and specified in
terms of alignment, geometric and
pavement design and structures.
Feasibility

Social The main social and environmental


issues should have been identified
and a preliminary assessment
Environment made. Any additional data should
be obtained if required.
Data are required that are
sufficient to obtain likely costs
to an accuracy of better than
Cost about 25%. Usually based on
estimation historic costs of similar roads
supplemented with any additional
costs if any expensive structures
are required
engineering

The pre-feasibility and feasibility studies will have identified all major issues and should
design
Final

also have provided information on any additional data that might be required for
completion of the final design and the required supporting documents outlined in
Section 2.1

2.3 Principal considerations for route selection


This section highlights most of the issues that require consideration when establishing and finalising
the route alignment. For upgrading existing roads many of the points raised will not be relevant. For
entirely new roads most of the issues should at least be considered and can act as a check list for the
road designer.

2.3.1 General considerations and best practice

Socio-Economic:
ƒ The road should be as direct as possible (within the bounds of the geometric standards for the
particular class of road) between the cities, towns or villages to be linked, thereby minimising road
user transport costs and probably minimising construction and maintenance costs as well.
ƒ The preferred alignment should be one that permits a balancing of cut and fill to minimise borrow,
spoil and haul.
ƒ The road should be close to sources of borrow materials and should minimise haulage of materials
over long distances.
ƒ The road should not be so close to public facilities that it causes unnecessary disturbance. Cultural
sites such as cemeteries, places of worship, archaeological and historical monuments should be

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 9

specifically protected. Although a road is designed to facilitate access to hospitals, schools and so
on, it should be located at a reasonable distance away for safety and to reduce noise.
ƒ Where the proposed location interferes with utility lines (eg over-head transmission cables and
water supply lines), the decision between changing the road alignment and shifting the utility line
should be based on a study of the feasibility and the relative economics.
ƒ The road should, as far as possible, be located along edges of properties rather than through
them to minimise interference to agriculture and other activities and to avoid the need for frequent
crossing of the road by the local people.
ƒ When the road follows a railway line or river, frequent crossings of the railway or river should be
avoided.
ƒ The location should be such as to avoid unnecessary and expensive destruction of trees and
forests. Where intrusion into such areas is unavoidable, the road should be aligned on a curve so
as to preserve an unbroken background.
ƒ The road should be ‘integrated’ with the surrounding landscape as far as possible. Normally, it is
necessary to study the environmental impact of the road and ensure that its adverse effects are
kept to the minimum. (see Part C)

Engineering:
ƒ The preferred alignment is one that is founded on strong sub-grades, thereby minimising pavement
layer thicknesses. Therefore marshy and low-lying areas and places having poor drainage and
weak materials should be avoided.
ƒ Problematic and erosion susceptible soils should also be avoided.
ƒ An important control point in route selection is the location of river crossings. The direction of the
crossings of major rivers should be normal to the river flow.
ƒ When an alignment passes near to a river, flood records for the past 50 years must be reviewed, if
these are available. Areas liable to flooding and areas likely to be unstable due to toe-erosion by
rivers should be avoided (see Chapter D.5)

Other:
ƒ Where possible, the road should be located such that the road reserve can be wide enough to
allow future upgrading to a wider carriageway.
ƒ Areas of valuable natural resources and wildlife sanctuaries should remain protected.

2.3.2 Special Considerations in mountainous areas (see also Chapter D.3):

General principles:
ƒ The location should, as far as possible, facilitate easy grades and curvatures.
ƒ High fills should be avoided and special attention should be paid to the compaction of all fills.
ƒ The alignment should minimise the number of hairpin bends. Where unavoidable, the bends and
switchbacks should be located on stable ground. A series of hairpin bends on the same face of the
hill should be avoided (Chapter D.5).
ƒ In relatively stable slopes, half cut and half fill cross-sections should be adopted to minimise the
disturbance to the natural ground.
ƒ Natural terrain features such as stable benches, ridge-tops, and low gradient slopes should be
utilized. If a ridge top is considered, roads should be located far enough above convergent gully
headwalls or confluences to provide a buffer, otherwise a structure is needed to intercept moving
sediment below the road.
ƒ In crossing mountain ridges, the location should be such that the road preferably crosses the ridge
at the lowest elevation.
ƒ Needless rise and fall should be avoided, especially where the general purpose of the route is to
gain elevation from a lower to a higher point.
ƒ Locations along river valleys have the inherent advantage of comparatively gentle gradients,
proximity to inhabited villages, and easy supply of water for construction purposes. However,
there are also disadvantages such as the need for large number of cross-drainage structures and
protective works against erosion.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 10

ƒ To minimise the adverse effect of moisture on the road environment, an alignment that is
predominantly in sunlight should receive priority compared with one that is entirely or partially in
the shade throughout the day.

Unstable terrain:
ƒ If possible unstable slopes, areas having frequent landslide problems and benched agricultural
fields should be avoided.
ƒ Mid-slope locations on long, steep, or unstable slopes should be avoided. If an unstable area
such as a headwall must be crossed, end-hauling excavated material rather than using side-cast
methods should be considered.

Erosion potential:
ƒ Erosion is a serious problem in much of Ethiopia. If possible, it is best to avoid areas of high erosion
potential. If not, considerable attention is required to dissipate flow in road drainage ditches and
culverts and reduce surface erosion (Chapter D.5). It is also advisable to consult local agricultural
experts during the process of route selection to ensure that the selected alignment has a minimum
potential for soil erosion and that the project design provides sufficient erosion control measures.
ƒ In selecting the best location for the road, the engineering measures designed to minimise erosion
will add to the construction costs but ongoing maintenance to deal with debris, blockage and
siltation will be required and no erosion protection system is guaranteed.

2.4 Site investigation techniques


The choice of methods for site investigation is determined by the type of road project and the practical
problems arising from site conditions, the terrain and climate. A wide variety of methods are used for site
investigation. The ERA Site Investigation Manual (ERA 2011) describes the most frequently employed
techniques for all aspects of the road design. Only those specialist techniques that can be used for site
investigation on low volume roads are described in this chapter.

If an investigation is to be effective, it must be carried out in a systematic way, using techniques that
are understood by the industry, relevant to the project in hand, reliable and cost-effective. For low
volume roads, investigations should employ relatively standard and simple engineering methods. More
sophisticated and expensive procedures should only be employed when a severe geotechnical problem
is encountered. Under such circumstances it is advisable to seek specialist assistance.

It is the decision of the design engineer to determine frequency and type of testing necessary for the
specific road project and to assess when bulk samples should be taken for laboratory testing in accordance
with the appropriate standard.

Site investigation techniques for unpaved DC1 and DC2 roads can utilise relatively simple sampling
and testing techniques. These include visual inspection and description of test pits along the proposed
alignment, use of dynamic cone penetrometer testing to assign uniform sections and use of simple
material testing kits to assess the grading and plasticity of in-situ soils and borrow materials.

The benefit of utilising the materials test kits is that a large number of simple tests can be conducted in
the field relatively quickly and cheaply and the frequency of testing will not be compromised. Verification
tests in the laboratory will also be required. Strength, compaction and other types of test can only be
conducted by appropriate sampling and laboratory testing.

A detailed explanation on the application and use of the test kits is provided in the ASIST Technical
Brief Number 9: Material Selection and Quality Assurance for labour-based unsealed road projects,
reproduced in Annex D.2.

Site investigation techniques for DC3 and DC4 paved and unpaved roads should, in general, follow the
traditional techniques used for site investigation on road projects as set out in the ERA Site Investigation
Manual (ERA 2011).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 11

2.5 Site investigation for final engineering design


The final engineering design requires sufficient design data for preparation of the tender and draft
contract documents. This stage requires the most rigorous site investigation and considerably more data
will be required than hitherto. An estimate of the requirement for detailed site investigation should be
made as part of the feasibility study. The entire process of project design should now be completed with
sufficient accuracy to minimise the risk of changes being required after the contracts has been awarded.
Detailed investigation will be required to provide technical data on the following:
ƒ Topography
ƒ Traffic count and loading
ƒ Alignment soils and construction materials (fill, gravels, rock, potential aggregate, sand and water)
– including potential haul and quantities
ƒ Hydrology and drainage
ƒ Ground stability, geotechnics and the characteristics of water crossings
ƒ Socio-environmental considerations

The scope and extent of the site investigation for final engineering design will depend on the characteristics
of the alignment and the type of road under consideration. For DC1 and for many DC2 class low volume
roads the design of the feasibility study should be such that most of the information obtained should be
sufficient for final design. The data obtained at the feasibility stage will not be so comprehensive and will
not be so robust from a statistical point of view as that obtained from site investigations for DC3 and DC4
roads. However, it should be adequate and reliable and sufficient to provide a competent design for DC1
and DC2 low volume roads. It is likely that some additional detailed survey will be required, particularly
for water crossings, within areas of problem soils and unstable terrain.

The quality and level of the site investigation for final design should not be compromised to provide
cost savings nor should the level of investigation be necessarily reduced to reflect an anticipated
low design class.

The sub-surface investigation for the final design stage is typically performed prior to defining the
proposed structural elements or the specific locations of culverts, embankments or other structures.
Accordingly, the investigation process includes techniques sufficient to define soil and rock characteristics
and the centreline sub-grade conditions.

An important assumption is that the topographic survey, based on the preliminary route alignment has
been completed prior to the detailed site investigation. It is only against the topographic model that
locations of structures and other features of the design can be fixed and estimates made of quantities,
haulage and ultimately construction costs.

In general, the site investigation for final design will focus on sampling and testing of materials to provide
information on the following reports:
ƒ Characteristics of alignment soils and in-situ materials
ƒ Location and characteristics of construction materials, volumes available and haulage
ƒ Earthworks investigations – cut and fill
ƒ Water crossings
ƒ Water sources

2.5.1 Characterisation of alignment soils and in-situ materials (subgrade)

The subgrade can be defined, in terms of location, as the upper 600mm of the road foundation. The
subgrade is required to resist repeated stressing by traffic and to be stable to the stresses imposed by
varying climatic and moisture influence.

The character of the subgrade is determined by the geological and weathering characteristics of the
rocks that produce the soil and the interaction with the local climate, moisture and drainage regime
prevailing in the area. As a general “rule of thumb” better subgrades are found in well drained areas.
Clayey soils often predominate in flat areas and along valley floors.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 12

The design of a paved or unpaved road is very strongly dependant on the characteristics of the subgrade
and, therefore, so is its potential performance. The desirable properties of a good subgrade include high
strength, high stiffness, good drainage characteristics, ease of compaction and low compressibility. A
good subgrade is strong enough to resist shear failure and has adequate stiffness to minimize vertical
deflection. Stronger and stiffer materials provide a more effective foundation for the pavement layers
and are more resistant to stresses from repeated loadings and environmental (moisture) conditions. Most
importantly, the stronger the subgrade, the thinner the pavement layers above need to be. Unfortunately
the designer usually has very little choice about the subgrade for most of the route.

Because the road design is so dependent on the subgrade, it is vital that the characteristics of the
subgrade along the alignment are measured in some detail and understood. In cases where the subgrade
materials are unsuitable, either cost effective methods of improving the existing conditions must be
identified (eg. improving drainage or stabilisation) or the road alignment must be altered to avoid such
areas completely.

DCP surveys
The most cost effective method for obtaining sub-surface information to a depth of approximately
800mm is by using the DCP test. The use of the DCP helps to delineate homogenous subgrade sections
along the road and to identify soft spots of the subgrade for further investigation using pits and trenches.
The advantage of the DCP is that information can be gathered without disturbing the in-situ material.
Using this test, strength characteristics of the subsurface soils at field moisture and density conditions
can be obtained directly. The equipment is light and portable and is also useful for investigating the
characteristics of all the pavement layers of existing roads for rehabilitation projects. In addition, for road
widening and upgrading projects, DCP tests along the main pavement can be compared to those from
shoulders. DCP tests can also be used for quality control during construction.

DCP tests are quick and simple. Tests can be carried out every few hundred metres along the chosen
alignment to delineate uniform sections. Where obvious changes of surface conditions occur more
frequently, the frequency of the tests should be modified to include the changes. Similarly, where surface
conditions are uniform, the frequency of testing may be reduced. As a minimum standard, four DCP tests
per kilometre should be used for DC1 and DC2 roads, whilst ten DCP tests per kilometre should be used
for DC2 paved roads and all DC3 and DC4 roads.

A number of correlations exist to link the DCP penetration rate (mm/blow) to the subgrade strength
parameters required for a pavement design. These correlations are based on either soaked or unsoaked
CBR values versus DCP penetration rates measured in different soil types. It is important to make sure
that the correlation being used is the correct one for the purposes of the study. In general, the correlation
should be between the DCP penetration rate and the actual CBR of the material being tested (i.e. the
CBR at the density and moisture content of the material at that time). In this way the in-situ strengths can
be determined.

Test pits and trenches


Test pits and trenches are used to provide samples for testing and information on the in-situ subgrade
soil and potential fill material.

The location, frequency and depth of pits and trenches for characterizing the subgrade depend on the
type of the road and the general characteristics of the project area (the soil type and variability). In
addition, the DCP testing carried out to assist delineation of uniform sections can be used to target areas
for pitting and trenching. Spacing will decrease when the subsurface soils demonstrate more variability.
In these areas, pits can also to be staggered left and right of the centerline to cover the full width of the
road formation.

The depth of pits and trenches is determined by the nature of the subsurface. In pavement design, the
depth of influence is related to the magnitude and distribution of traffic loads. Current AASHTO and
many other standards limit this depth to 1.5m below the proposed subgrade level. For the purpose of
sampling and description, pits should be dug to at least 0.5 m below the expected natural subgrade
level. In cut sections, the depth can be reduced to 0.3 m. For upgrading and rehabilitation projects there

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 13

is usually vehicular access hence pits can be excavated using a backhoe through all the existing pavement
layers. In these circumstances the depth could be increased to 1.5 m below the subgrade if required, but
this will rarely be necessary for such projects.

For a new alignment, the depth of any pit should not be less than 2m unless a rock stratum is encountered.
Some problem subgrade conditions may require deeper exploration. Greater depths may also be needed
for high embankment design. This is also true for boulder identification as buried basaltic boulders are
common in the highlands of Ethiopia. A limited number of deep pits may also be needed to ascertain
groundwater influence and irregular bedrock.

The location of each test pit should be precisely determined on the preliminary route alignment and all
layers, including topsoil, should be accurately described and their thicknesses measured. All horizons,
below the topsoil should be sampled. This will promote a proper assessment of the materials excavated in
cuts to be used in embankments. The samples should be taken over the full depth of the layer by taking
vertical slices of materials.

It is sometimes impossible to dig trial pits to the depth of all layers of soil or weathered rocks affected by
foundation loads. In this case it is recommended that hand or power augers are used for identification
(AASHTO T203). Borings could also be necessary to investigate the materials that lie below pavement
layers. This is especially true in areas where thick problem soils and soft deposits exist, and when the road
alignment passes through landslide zones, solution cavities, and unconsolidated soils.

Standard laboratory testing


Samples collected from the test pits are used to provide the following basic information on the properties
of the in-situ materials and subgrade along the alignment:
ƒ Soil Profile: Overburden thickness, layer/horizon thickness, visual description; in situ moisture
content
ƒ Index Tests: Particle size distribution (AASHTO T88); Plasticity/Atterberg limits (AASHTO T89 and
T90); Linear shrinkage
ƒ Compaction: Density and Moisture relationship (Standard AASHTO T99)
ƒ Strength: CBR and swell (AASHTO T99 or T180)

Most of the subgrade test samples should be taken from as close to the top of the subgrade as possible,
extending down to a depth of 0.5 m below the planned subgrade elevation. Potential fill materials should
be sampled to a greater depth.

For the design of DC1 and DC2 low volume roads, a presumptive design CBR could be assigned on
the basis of previous test data and the performance of soils in similar environments. Some regional road
authorities have considerable experience and performance data on specific soil types in the local climate
and topographic conditions. Use of this information can supplement and reduce (but not replace) the
overall requirement for subgrade evaluation. The approach involves the assessment of subgrades on the
basis of local geology, topography and drainage, together with regular routine soil classification tests.

2.5.2 Problem soils

Soils which can cause foundation problems and decrease the performance of roads are common in many
parts of Ethiopia. These soils are collectively called problem soils and comprise among others; expansive;
collapsible and compressible; and dispersive soils. The identification of such soils is crucial during site
investigation so that appropriate designs can be established at the outset. Failure to recognise problem
soils at the design stage could result in claims and overruns if identified later during construction or
detrimental impact on the long term performance of the road.

Expansive soils
Problems are associated with the presence of expansive soils on the road alignment include:
ƒ A very low bearing capacity when wet.
ƒ Variable and seasonal moisture distribution leading to differential volumetric movement, settlement
and cracking.
ƒ Difficulty in locating any associated natural gravels for sub-bases and bases.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 14

Expansive soils are typically clayey soils that undergo large volume changes in direct response to moisture
changes in the soil. Although the expansion potential of a soil can be related to many factors (eg soil
structure and fabric; and environmental conditions), it is primarily controlled by the clay mineralogy (eg
smectites and montmorillonite).

Known as vertisols in agricultural soil classifications, expansive soils are found in the central, north-western
and eastern highlands of Ethiopia; in the western lowlands around Gambella; and in some parts of the
rift valley. Local deposits of these soils are also present throughout the country near rivers; water logged
areas; and in drainage restricted localities. Damage caused by expansive clays is particularly prevalent
around Addis Ababa.

Volume changes in expansive soils are confined to the upper few metres of a soil deposit where seasonal
moisture content varies due to drying and wetting cycles. The presence of surface desiccation cracks or
fissures in a clay deposit are indications of expansion. The zone within which volume changes are most
likely to occur is defined as the active zone. The active zone can be evaluated by plotting the in situ
moisture content with depth for samples taken during the wet and dry seasons. The depth at which the
moisture content becomes nearly constant is the limit of the active zone. This is also referred to as the
depth of seasonal moisture change.

Several empirical relationships have been developed to identify expansive soils, although a standard
classification procedure does not exist. Generally, soils with a plasticity index (PI) of less than 15% and
liquid limit below 55% do not exhibit expansive behaviour. For soils with a plasticity index greater than
15%, the clay content of the soil should be evaluated in addition to the Atterberg limits.

Figure D.2.1 relates expansion potential and collapsibility to liquid limit and in-situ dry density. Additional
tests for the qualitative assessment of expansion potential include percentage swell calculated from the
CBR test (ASTM D4429), the free swell test, and the expansion index test (ASTM D4829). Such correlations
are semi-empirical but can be used for an initial assessment of the expansion potential of a soil.

Figure D.2.1: Guide to collapsibility and expansion based on in-situ dry density
and liquid limit (after Mitchell and Gardner, 1975 and Gibbs, 1969)

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 15

For classification purposes, the US Bureau of Reclamation developed a correlation between observed
volume changes and colloidal content, plastic index, and shrinkage limit as shown in Table D.2.3. The
measured volume changes were taken from oedometer swell tests using a surcharge pressure of 7 KPa
from air-dry to saturation conditions.

Table D.2.3: Classification of expansive soils according to US Bureau of Reclamation

Colloid content PI SL Potential Degree of


% < 1µm (%) (%) expansion (%) expansion
<15 <18 >15 <10 Low
13-23 15-28 10-16 10-20 Medium
20-31 25-41 7-12 20-30 High
>28 >35 <11 >30 Very high

Appropriate designs that can be used to reduce the effect of expansive soils on pavements are provided
Chapter D.6. Treatment options and recommendations are also provided in the ERA Geotechnical Design
Manual (ERA 2011).

Collapsible soils
Collapsible soils are generally described as soils that undergo a relatively significant, sudden and
irreversible decrease in volume upon wetting. These types of soils predominantly consist of silt and sand
with some clayey material. Deposits of collapsible soils are usually associated with regions of moisture
deficiency, such as those in arid and semi-arid regions.

Collapsible soils are present in the southern part of the Omo River and in the central and southern
part of the rift valley. Often, their existence around Zeway, Shashemene, and Awassa is manifested by
the occurrence of ground cracks and a series of potholes after heavy rains in spring. In the Afar region,
collapsible soils are present in the form of sand dunes.

Collapsible soils usually exist in the ground at very low values of dry unit weight (density) and moisture
content. In their natural conditions, collapsible soils can support moderate loads and undergo relatively
small settlements. They are also moderately strong and exhibit a slight but characteristic apparent
cohesion. Usually, this cohesion is the result of calcareous clay binder that holds the silt particles together.
The clay coating and the silt create a very loose soil structure with little true particle-to-particle contact.
Upon wetting, however, the cohesion is lost and large settlements can occur even if the load remains
constant.

For rapid identification, the liquid limit can be used. If under normal circumstances the void ratio of
a given soil is higher than that at its liquid limit, on absorbing water the soil will lose strength. Before
saturation is achieved, the soil will undergo considerable structural collapse accompanied by a reduction
in volume. If this is the case, then laboratory testing of undisturbed samples should be performed to
quantify the magnitude of volume reduction. Silts containing collapsible soils are often also characterized
by being extremely erodible.

Typical pit excavation and disturbed sampling procedures can be used to obtain soil samples for sieve
analysis, hydrometer, soil classification and Atterberg limits. For samples to be collected at shallow
depths, it may be prudent to obtain block samples from trenches or test pits. Unlike expansive soils,
where volume change occurs in the top few metres, the depth of collapse can be much higher and may
involve loose soils in the region. In the rift valley, there are indications that the thickness of potential
collapsible soils is greater than 8 m. In this case auger sampling or shallow boring can be considered.

For situations in which it is necessary to construct a road on collapsible soils, it is of primary importance
to estimate the magnitude of potential collapse that may occur if the soil becomes wet. The amount of

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 16

collapse normally depends on the initial void ratio, stress history of the soil, thickness of the collapsible
soil layer and magnitude of the applied stress.

The collapse potential (CP) is calculated as the percentage collapse of a soil specimen using the change
in void ratio before and after saturation. It is an index value used to compare the susceptibility of collapse.
Table D.2.4 provides a relative indication of the degree of severity for various values of CP. Additional
information is provided in the Site Investigation Manual - 2011.

Table D.2.4: Qualitative assessment of collapse potential

Collapse Potential
Severity of Problem
(CP)
0 - 1% None
2 - 5% Slight
6 - 10% Moderate
11 - 20% Severe
> 20% Very severe

Appropriate designs that can be used to reduce the effect of collapsible soils on pavements are provided
Chapter D.6. Treatment options and recommendations are also provided in the ERA Geotechnical Design
Manual (ERA 2011).

Dispersive soils
Soils in which the clay particles detach from each other and from the soil structure without a flow of water
and go into suspension, are termed dispersive soils. These soils deflocculate in the presence of relatively
pure water to form colloidal suspensions and are, therefore, highly susceptible to erosion and piping.
Normally, they contain a higher content of sodium in their pore water than other soils. However, there are
no significant differences in the clay contents of dispersive and non-dispersive soils although soils with
high exchangeable sodium such as Na-montmorillonite clays tend to be more dispersive than others.

Dispersive soils tend to develop in low-lying areas with gently rolling topography and relatively flat slopes.
Their environment of formation is also characterized by an annual rainfall of less than 850 mm. Dispersive
soils have low natural fertility. Often, they are calcareous with a PH value of about 8. Suspicion of their
presence is indicated by the occurrence of erosion gullies and piping.

In Ethiopia, dispersive soils exist in the rift valley, the southern and eastern lowlands, and Afar, Somali and
Tigray regions. Isolated occurrences of these soils can also be found in other parts of the country.

It is difficult to identify dispersive soils using conventional engineering index tests such as Atterberg limits,
gradation or compaction characteristics. Chemical properties can determine the dispersion potential of
soils by measuring the dissolved sodium in the pore water.

The pinhole (BS1377 Part 5 – clause 6.2) and Crumb Test (BS1377 Part 5 – clause 6.3) provides a relatively
simple way of identify dispersive soils in the field. The test starts by collecting soil aggregates (1-2 cm
diameter) from each layer in the soil profile. The aggregates are dried in the sun for a few hours and
placed in a small bowl of rain water. The aggregates are left in the water without shaking or disturbing for
2 hours. The samples are then observed for a milky ring around the aggregates and classified as shown
in Figure D.2.2. The soil is highly dispersive if discoloration and cloudiness extends throughout the jar or
bowl.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 17

Figure D.2.2: A simple way of testing the dispersive nature of soils

While laboratory tests such as the crump, dispersion, pinhole tests are a useful way of identifying dispersive
soils, much can be determined by observing the behaviour of the soils in the field. For instance, the
occurrence of deep erosion gullies, ‘worm channels’, and piping failure in existing road embankments
indicates the presence of dispersive soils. Erosion of road cuttings along ditches, gully lines or weathered
rock joints; cloudy water after rains; and high turbidity in ponds are also linked to the effect of dispersive
soils.

The geology of the area can also be a guide to the presence of dispersive soils. Many dispersive soils are
of alluvial origin. Soils derived from shale and clay-stone in sedimentary areas and pyroclastic sediments
in volcanic regions are also dispersive in nature.

2.5.3 Location and characteristics of construction materials, volumes available and haulage

Sources of road-building materials have to be identified within an economic haulage distance and they
must be available in sufficient quantity and of sufficient quality for the purposes intended. Previous
experience in the area may assist with this but additional survey is usually essential.

Two of the most common reasons for construction costs to escalate, once construction has started and
material sources fully explored, are that the materials are found to be deficient in quality or quantity. This
leads to expensive delays whilst new sources are investigated or the road is redesigned to take account
of the actual materials available.

The construction materials investigation often requires an extensive programme of site and laboratory
testing, especially if the materials are of marginal quality or occur only in small quantities.

The site investigation must identify and prove that there are adequate and economically viable reserves
of natural construction materials. The materials required are:
ƒ Common embankment fill;
ƒ Capping layer / imported subgrade;
ƒ Sub-base and road-base aggregate;
ƒ Road surfacing aggregate;
ƒ Paving stone (eg for cobblestone pavements);
ƒ Aggregates for structural concrete;
ƒ Filter/drainage material;
ƒ Special requirements (eg rock-fill for gabion baskets).

If the project is in an area where good quality construction materials are scarce or unavailable, alternate
solutions that make use of the local materials should be considered to avoid long and expensive haulage.
For example consideration should be given to:
ƒ Modifying the design requirements
ƒ Modifying the material (eg mechanical or chemical stabilization)
ƒ Material processing (eg crushing, screening, blending)
ƒ Innovative use of non-standard materials (particularly important for low traffic roads)

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 18

The materials investigations should also take into account any future needs of the road. This is particularly
important in the case of gravel roads where re-gravelling is normally needed every few years to replace
material lost from the surface.

Sources of good material could be depleted with the result that haul distances and costs will increase.
Furthermore, good quality material may be required at a later stage in the road’s life
when the standard needs to be improved to meet increased traffic demands.

The design engineer will need to ascertain the availability of sufficient suitable materials in the vicinity
of the road alignment. A comprehensive list of the location and potential borrow pits and quarries is
needed, along with an assessment of their proposed use and the volumes of material available. Apart
from quality and quantity of material, the borrow pits and quarries must be:
ƒ Accessible and suitable for efficient and economic excavation;
ƒ Close to the site to minimize haulage costs;
ƒ Of suitable quality to enable cost-effective construction with little or no treatment;
ƒ Located such that their exploitation will not lead to any complicated or lengthy legal problems and
will not unduly affect the local inhabitants or adversely affect the environment.

Exploration of an area to establish availability of materials has the following objectives:


ƒ Determination of the nature of the deposit, including its geology, history of previous excavation
and possible mineral rights;
ƒ Determination of the depth, thickness, extent and composition of the strata of soil and rock that
are to be excavated;
ƒ Analysis of the condition of groundwater, including the position of the water table, its variations,
and possible flow of surface water into the excavation ground;
ƒ Assessment of the property of soils and rocks for the purposes intended.

Records of roads already built can be a valuable source of data, not only on the location of construction
materials but also on their excavation, processing, placement and subsequent performance. Potential
problems with materials can also be identified. Construction records are kept by different departments of
ERA, regional road authorities, or by road design consultants and construction supervising organisations
and contractors.

Fill: In general, location and selection of fill material for low volume roads poses few problems. Exceptions
include organic soils and clays with high liquid limit and plasticity. Problems may also exist in lacustrine
and flood plain deposits where very fine materials are abundant.

Where possible, fill should be taken from within the road alignment (balanced cut-fill operations) or by
excavation of the side drains (exception in areas of expansive soils). Borrow pits producing fills should
be avoided, as far as possible, and special consideration should be given to the impacts of winning fill in
agriculturally productive areas where land expropriation costs can be high.

Improved subgrade: The subgrade can be made of the same material as any fill. Where in-situ and
alignment soils are weak or problematic, import of improved subgrade may be necessary. As far as
possible the requirement to import material, from borrow areas, should be avoided, due to the additional
haulage costs. However, import of strong (CBR>9) subgrade materials can provide economies with
regards the pavement thickness design (see Part D -6). Where improvement is necessary or unavoidable,
mechanical and chemical stabilisation methods can be considered.

Road base and subbase: Where possible, naturally occurring unprocessed materials should be selected
for sub-base and road base in paved low volume roads. However, under certain circumstances, mechanical
treatments may be required to improve the quality to the required standard. This often requires the use
of special equipment and processing plants that are relatively immobile or static. For this reason, the
borrow pits for road base and sub-base materials are usually spaced widely. In current practices, distances
between these pits of about 50km are not unusual. Main sources of sub-base and base materials are
rocky hillsides and cliffs, high steep hills, and river banks. In Ethiopia, sub-base materials have also been

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 19

extracted from cinder cones and lateritic deposits. Sub-base materials are expected to meet requirements
related to maximum particle size, grading, plasticity, and CBR.

The minimum thickness of a deposit normally considered workable for excavation for materials for
subgrade, sub-base and road base is of the order of one metre. However, thinner horizons could also be
exploited if there are no alternatives. The absolute minimum depends on material availability and the
thickness of the overburden. If there is no overburden, as may be the case in arid areas, horizons as thin
as 300mm may be excavated.

Hard Stone and aggregate: The ERA Site Investigation Manual (ERA 2011) provides some details on
the location and variety of rocks in Ethiopia that can be used as material sources for concrete aggregate,
bituminous road surfacing aggregate, masonry and cobble stone. In any area, a relatively fresh rock must
be encountered at some depth as there is a gradual transition from one weathering state to the other.
The recovery of a suitable material is, therefore, a matter of understanding the geological history and
weathering profile at the quarry site.

Prospecting and testing: The earlier feasibility and pre-feasibility studies will have likely have used
desk studies (topography, geology, soils, hydrology, vegetation, land-use and climate in the area) field
survey and possibly also laboratory testing programmes to make preliminary identification and location
of potential construction materials. This information will guide the verification process undertaken by the
design engineer in preparation of the detailed design.

Laboratory Testing
The quality of the testing programme depends upon the procedures in place to ensure that
tests are conducted properly using suitable equipment that is mechanically sound and calibrated
correctly. The condition of test equipment and the competence of the laboratory staff are therefore
crucial. There needs to be a robust Quality Assurance (QA) procedure (overseen by a competent
geotechnical engineer) in place that will reject data that does not meet acceptable standards of
reliability. There should be no compromise on the QA procedure or quality of testing data just
because the project is perceived as a low volume road.

Site investigation activities will include detailed prospecting for materials through surface mapping,
test pitting, boreholes, material sampling, and in situ testing. A variety of sub-surface sampling and
investigation procedures appropriate for different materials are used to recover the samples needed for
laboratory testing. The ERA Site Investigation Manual (ERA 2011) deals with the techniques required for
materials prospecting.

The laboratory testing programme should be part of a rational programme designed by the engineer
to give all of the information needed to adequately define the nature, use and volumes available of
construction materials.

Maximum use should be made of data and information compiled during earlier parts of the project design.
The construction materials used for low volume roads and the design philosophies that are adopted in
this manual, mean that it is important that the relationships between expected / in situ conditions and
laboratory conditions are considered when designing and developing the test regime.

Early phases of the laboratory test programme will generally concentrate on gaining clues to unusual soil
behaviour, eg swelling or collapse potential. Bearing in mind the difficulties of sample recovery, statistical
sample sizes and the cost of laboratory testing, most testing programmes will be based around relatively
simple classification tests that can be done quite quickly (see Table D.2.5). More sophisticated tests will
only be used if absolutely necessary.

However, even at the stage of final design, there is always the problem that natural materials show high
variability in their properties and therefore obtaining design parameters at the ideal level of statistical
reliability is very difficult. As a result, considerable engineering judgement and skill is required

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 20

For projects involving the use of aggregate processing there may be an additional requirement to
undertake quality assurance / laboratory tests on trials of the product produced using the expected
processing procedures (eg crushed aggregate for surfacing).

Projects involving significant fill and aggregate requirements will require mass-haul diagrams to be drawn
that augment cost-benefit decisions with respect to utilizing any alternative materials or treatments, for
example modifying the design requirements by modifying the material (stabilization) or by additional
material processing (eg crushing and screening).

The frequency of testing of borrow pit material needs to strike a balance between cost, time and
statistical validity. Where possible, the location and testing of borrow pit material should be done by
traditional methods using full laboratory facilities. In the absence of these facilities, testing using the test
kit described in Appendix D.1 should be substituted for DC1 and DC2 earth and gravel roads.

The frequency of testing will depend on the variability of the material in that the more homogeneous
the material, the less testing will be required. However, it is important to carry out sufficient tests to
quantify the variability of the material within the pit during the site investigation stage and prior to
construction. For low volume roads projects, irrespective of the testing techniques and methods used,
it is recommended that test samples are taken from at least five randomly selected locations per borrow
pit (covering the full depth of the layer to be used) to quantify the variability. The variability provides an
indication of the variation in material quality that can be expected during construction for process control.

Table D.2.5: Typical laboratory tests used to assess the suitability


of alignment soils and pavement materials

Subgrade Road Surfacing Wearing


Sub-base
& Fill Base Aggregate Course Gravel
Index Tests
ƒ Atterberg limits 9 9 9 9 (see note 1)
ƒ Linear Shrinkage 9 (see note 1)
ƒ Particle Size Distribution 9 9 9 9 (see note 1)
Compaction and Strength
Tests
ƒ Dry Density and Optimum
9 9 9 9
Moisture Content
ƒ CBR and CBR Swell 9 9 9 9
Particle Strength Tests
(aggregate dependent)
ƒ Durability Mill 9 9
ƒ 10% FACT, AIV, ACV, LAA 9 9
ƒ Glycol Soak (9) (9)
ƒ Water absorption 9
ƒ Specific Gravity 9 9
ƒ Flakiness and ALD 9
ƒ Affinity with bitumen -
9
Immersion tray test
Note 1:
1. Refer to appropriate test techniques and methods in Appendix D.1

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 21

Weathering
The quality and durability of borrow materials and crushed stones can be greatly affected by weathering
or alteration. The type and rate of weathering vary from one region to another. In the tropics, high
temperatures associated with high humidity often produce physical and chemical changes to a considerable
depth in surface rocks. In dry areas, weathering is predominantly physical, and rock masses disintegrate
by alternate heating and cooling, but still keep their general appearance. In more humid areas, chemical
weathering proceeds quite rapidly and rock masses may be partially or completely weathered.

Weathering effects generally decrease with depth, although zones of differential weathering can occur in
many outcrops. Examples of differential weathering are:
ƒ Resulting from compositional or textural differences;
ƒ At contact zones associated with thermal effects within volcanic rocks;
ƒ Along permeable joints, faults, or contacts where weathering agents can penetrate more deeply
into the rock mass;
ƒ Within a single rock unit due to relatively high permeability; and
ƒ Resulting from topographic effects.

Table D.2.6 presents a system for describing and classifying the states of weathering in borrow or quarry
materials. In this table, the degree of weathering is divided into categories that reflect definable physical
changes that could result in modified engineering properties. The general descriptions cover ranges in
bedrock conditions and are intended for a rapid assessment of the use of borrow and quarry materials for
different purposes in pavement construction.

Weathering tables may generally be applicable to all rock types. However, they are easier to use in
igneous and metamorphic rocks that contain ferromagnesian minerals. Weathering in many sedimentary
rocks will not always conform to the criteria in Table D.2.6. In addition, weathering descriptions and
categories may have to be modified to reflect site-specific conditions, such as fracture openness and
filling, and the presence of groundwater.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Table D.2.6: Weathering grades for road design and construction
D - Chapter 2 - 22

Chemical weathering
Descriptive General characteristics
Grade Fracture Texture
term Body of rock (strength, excavation, etc)
surfaces
Hammer rings when crystalline rocks are struck.
No
Almost always excavation involves rock except for
W1 Fresh No discoloration, not oxidized discoloration No change
naturally weak or weakly cemented rocks such as
or oxidation
siltstones or shales.
Minor to
Discoloration or oxidation is Hammer rings when crystalline rocks are struck.
complete
Slightly limited to surface of, or short Body of rock not weakened. With few exceptions,
W2 discoloration Preserved

PART D: EXPLANATORY
weathered distance from, fractures; some such as siltstones or shales, classified as rock
or oxidation of
feldspar crystals are dull. excavation.
most surfaces
Hammer does not ring when rock is struck.
Discoloration or oxidation extends All fracture
Body of rock is slightly weakened. Depending

PART D: EXPLANATORY
Moderately from fractures, usually throughout; surfaces are Generally
W3 on fracturing, usually considered as rock during
weathered Fe-Mg minerals are “rusty,” discoloured or preserved
excavation except in naturally weak rocks such as
feldspar crystals are “cloudy.” oxidized
siltstone or shales.

NOTES FORNOTES
Discoloration or oxidation Dull sound when struck with hammer; can be
All fracture
throughout; all feldspars and Fe- broken with moderate to heavy pressure or by
surfaces are
Highly Mg minerals are altered to clay to Texture altered light hammer blow without reference to planes of

LOW VOLUME
W4 discoloured
weathered some extent; or chemical alteration by hydration weakness such as incipient or hairline fractures,
or oxidized,

FOR ROADS
produces in situ disaggregation, rock is significantly weakened. Usually easy for
surfaces friable
see grain boundary conditions. excavation.
Resembles a
soil, partial

ROAD DESIGN
or complete
Discoloured or oxidized
remnant rock
throughout, but resistant minerals Can be granulated by hand. Always easy for
structure may
W5 Decomposed such as quartz may be unaltered; ---- excavation. Resistant minerals such as quartz may
be preserved;
all feldspars and Fe-Mg minerals be present as grains
leaching
are completely altered to clay
of soluble
minerals usually
complete
D - Chapter 2 - 23

2.5.4 Earthworks – Cut and fill investigation

Natural slopes, road cuts and existing embankment fill in the vicinity of the planned project provide
evidence of expected ground stability and likely requirement for detailed surface and subsurface
investigations.

These investigations should consider; the types of materials in the cut; slope stability and the different
types of movements that may occur. Scars, anomalous bulges, odd outcrops, broken contours, ridge top
trenches, fissures, terraced slopes, abrupt changes in slope or in stream direction, springs or seepage
zones all indicate the possibility of past ground movements.

The first indication of possible instability problems can be obtained from a study of the topography.
Topographic maps and aerial photographs provide useful data on whether instability is likely to occur
or has occurred in the past. Moreover, an understanding of the local geology is essential. Slope failure
along road cuts is often associated with pre-existing planes. Survey of the orientation and characteristics
of joints and weak zones is therefore essential. In addition, the degree of weathering along these joints
should be inspected.

When a visual survey is not enough, it is often useful to excavate a trench. In deep cuts, where interference
with existing stability and groundwater conditions is expected, a trench across the face of the slope
provides a better understanding of the geology of the area. Trenches are preferable to pits to inspect cuts
because of their dimension. Depending on the geology and degree of weathering, up to five trenches
are normally enough to investigate a 100m long slope cut. The trenches should be located at places
where material changes are expected and range between 1m and 3m in depth. Additional information
on performance of slopes can also be obtained by inspecting soil and rock exposures along existing road
cuts in the region.

A particular difficulty in steep terrain is the disposal of excess material (spoil), therefore every effort should
be made to balance the cut and fill. Where this is not possible, suitable stable areas for the disposal of
spoil must be identified. Spoil can erode, or may become very wet and slide in a mass. Material is carried
downslope and may cause scour of watercourses or bury stable vegetated or agricultural land. Material
may choke stream beds causing the stream to meander from side to side, undercutting the banks and
creating instability.

High level embankment foundation investigation should, as a minimum, consider; the range of materials
and settlement potential; side-slope stability; groundwater; moisture regime and drainage requirements;
erosion resistance; haul distance; and environmental impact.

Settlement problems are unlikely if rock is encountered at a shallow depth. However, if the underlying
foundation is covered by transported soils, problems are likely as the material may vary from soft alluvial
clays to collapsing silts (sands) or expansive clays. It is therefore important to understand the particular
transportation history and mechanism and the result that this has on the nature of the soil and its behaviour.

The type of field investigation will depend on the types of soils encountered. If soils are predominately
cohesive, the primary design issues will be bearing capacity, side slope stability, and long-term settlement.
These design issues will usually require the collection of undisturbed soil samples for laboratory strength
and consolidation testing. The vane shear test can provide valuable in-situ strength data, particularly in
soft clays (for more information on vane shear tests, see the Site Investigation Manual - 2011).

Where embankments cross alluvial deposits, there will probably be a stream requiring a structure.
Therefore investigations should assess the interaction between these structures, the embankment and
the in-situ material. Most embankment problems at streams are a direct result of poor drainage and
consequent high pore pressures. During the site investigation it is important that all sources of water
along the alignment are identified and their impact on the design assessed.

If groundwater is not identified and adequately addressed early, it can significantly impair constructability,
road performance and slope stability. Claims related to unforeseen groundwater conditions often form

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 24

a significant proportion of contractual disputes. Many of these claims originate from a failure to record
groundwater during site investigation.

Groundwater is frequently encountered along road cuts in many parts of Ethiopia. In areas where springs
and seepages are present, there are several good indicators that may be used to determine the height
that groundwater may rise in a slope and roughly how long during the year that the slope remains
saturated. For example, in the highland areas where weathered basaltic lava flows are common, iron
containing soils within the slope usually oxidize when in contact with groundwater and turn rusty red or
bright orange and give the soil a mottled appearance. The depth below the ground surface where these
mottled appearances first occur indicates the average maximum height that the fluctuating water table
rises in the slope.

At locations where the water table remains relatively stable, iron compounds reduce chemically and
give the soil a grey or bluish-grey appearance. The occurrence of these greyed soils indicates a slope
that is saturated for much of the year. Occasionally, mottles may appear above greyed subsoil, which
indicates a seasonally fluctuating water table above a layer that is subjected to a prolonged saturation.
The practitioner should be aware of the significance of mottled and greyed soils exposed during road
construction. These soil layers give clues to the need for drainage or extra attention concerning the
stability of the road cut.

The ERA Site Investigation Manual (ERA 2011) and ERA Geotechnical Design Manual (ERA 2011) can be
referred for more information on investigation of major road cuts and embankment. The basic principle
for low volume road engineering design is to minimise cost. As far as possible this requires minimising
the earthworks cut and fill operations. In some geotechnically fragile areas, increased earthworks can lead
to an increased risk of landsliding.

2.5.5 Water crossings

Site investigation techniques needed for appropriate low level structures for water crossings are explained
in Part E of this manual. Additional information is also provided in the ERA Site Investigation Manual (ERA
2011).

There is no compromise on the design of major structures, such as bridges, where these are placed on
low volume roads. Design procedures follow the ERA Bridge Design Manual (ERA 2011) and thus the
site investigation techniques and procedures described in the ERA Site Investigation Manual (ERA 2011)
must be followed.

2.5.6 Water sources

Water is a vital construction resource. Many projects have been delayed because of an underestimate of
the quantity of water that is conveniently available for construction. Sources of water must therefore be
identified at the design stage and due attention should be given to the phasing of construction if best
use is to be made of the natural moisture in the materials.

In certain areas, water may be scarce for construction purposes and, in particular, for providing proper
moisture content during compaction of the soils and pavement layers. Since this problem is serious in
some regions of Ethiopia, it is important to search for water sources, their yields and the distances from
the construction site. In regions where water is scarce, such as the Afar and Somali regions, a separate and
dedicated hydro-geological study may be needed. Alternatively dry compaction could be considered for
some types of materials (see Chapter D.6). Data from the field reconnaissance can indicate if surface
water is a critical problem.

In the rift valley, water sources for construction need to be chemically analysed to assess the concentration
of chloride and sulphate, which could be deleterious to performance of concrete.

The Ethiopian Geological Survey has a collection of hydro-geological maps for various parts of Ethiopia.
There are also some private and state agencies which are responsible for water resource studies in the
country. Additional information can be obtained from these sources.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 2 - 25

When water is a concern for the road construction, it is advisable to contact the Geological Survey of
Ethiopia or the Regional Water Resources Bureau. It is also recommended that all water resource surveys
are carried out by specialist practitioners.

2.6 References
Gibbs, H.J. (1969). “Discussion, Proceedings of the Specialty Session No. 3 on Expansive Soils and
Moisture Movement in Partly Saturated Soils.” Seventh International Conference on Soil Mechanics and
Foundation Engineering, Mexico City, Mexico.

International Labour Organisation (ILO), Advisory Support, Information Services, and Training (ASIST)
Technical Brief No. 9 (1998). Material selection and quality assurance for labor based unsealed road
projects. Author: Paige-Green P., Editor: Mason David, Nairobi, 50 pp.

Mitchell, J.K. and Gardner, W.S. (1975).“In Situ Measurement of Volume Change Characteristics.” State-
of-the-Art Report, Proceedings, ASCE Specialty Conference on In Situ Measurement of Soil Properties,
Raleigh, North Carolina, Vol. II, 333 pp.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 26

3. ROADSIDE SLOPE STABILISATION

3.1 Introduction
Unstable natural slopes, cuts and embankments in the highlands of Ethiopia often disrupt traffic flow
and create a considerable problem to road users during rainy seasons. These slope failures or landslides
typically occur where a natural slope is too steep, a cut slope in soil and/or weathered rock contains weak
materials or adverse joints, or fill material is not properly compacted. The first two of these factors are
interdependent and cannot be considered as separate entities. In all three cases, a rise in groundwater
will lead to an overall reduction in stability and possibly failure. Once slope failures are initiated they can
expand rapidly, causing even greater traffic holdups and maintenance costs.

Erosion can also take place on unprotected cut and fill slopes and in river channels, especially downstream
of culverts, bridges and roadside turnouts. In addition, uncontrolled runoff can erode the roadside drain,
road pavement and the edge of the road. Sediment derived from the erosion not only impacts on the
road, but also on the wider environment. Large landslides can also contribute significant volumes of
debris to watercourses, with significant downstream effects.

Slope stabilisation and erosion control usually employs a number of methods to reduce the causes of
failure, together with measures to improve the stability of slopes. During design, it is important to select
affordable methods that are relevant to the class of the road, the type of landslide, the materials involved
and the extent of the slope instability problem. Techniques commonly used to prevent the occurrence of
landslides and to stabilise the existing slope failures include earthworks (cuts and fills), retaining structures
and revetments, surface and sub-soil drainage and bioengineering. Stabilisation methods that involve
more substantial engineering works involve anchoring, piling and deep subsurface drainage, but these
are rarely used on low volume roads. For LVRs, a combination of bioengineering, low cost retaining
walls such as gabions, dry-stone and mortared masonry walls, and surface drainage structures is a cost-
effective method to stabilise slopes.

It should be noted that whilst this chapter is based on international best practice, the recommendations
may not be the only options needed to stabilise slopes and mitigate the effects of landslides in all regions
of Ethiopia. The engineer should combine these recommendations with existing local practice.

For most minor slope failures, any remedial work will often rely on visual observation and the use of
simple measures such as debris clearance, trimming, and the removal of overhangs. While this may be
acceptable for many situations along low volume roads, there are occasions when additional measures are
required. These include a combination of bioengineering, low-cost retaining structures, and drainage.
For more complex slope stability and geotechnical issues, ERA’s Geotechnical Design Manual should be
consulted.

3.2 Roadside slope instability


Slope instability associated with roads is classified as that which occurs either above or below the road
(Hunt et al, 2008). The slope failures above the road involve the road cut and the natural hill slope above
it. Those below the road can affect both fill slopes and the valley slope and the natural ground below.

3.3 Slope instability above the road

3.3.1 Types of slope failure above the road

The types of slope failures and erosion processes commonly observed above a road alignment include
(Figure D.3.1):
ƒ Erosion of cut slope surface;
ƒ Failures in cut slope;
ƒ Failures in cut slope and hill-slope;

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 3 - 27

ƒ Failures in hill-slope;
ƒ Deep failure in original ground.

Once a cut has been formed the slope is exposed to erosion. The effect of erosion is normally to remove
the weathered mantle. This may lead to the formation of gullies, and ultimately to slope failures. Erosion
is especially common on cuts in weak rocks such as shale, marl, ash and tuff.

Failures in cut slopes usually occur due to a change in the external forces acting on the soil, a change in
the geometry of the cut, or the introduction of water. Many of the failures in cut slopes occur in the form
of minor slips. Sometimes, however, these small slips may develop into major failures and affect the entire
road corridor. In cut slopes, these failures often develop at the toe of the slope and propagate upward.
The majority of failures affecting the natural hillside above roads are caused by the removal of support in
the road cutting, combined with the effects of groundwater rise during heavy or prolonged rainfall.

Figure D.3.1: Slope failures that occur above the road (modified from Hunt et al, 2008).

Failures in hill slopes usually develop over long periods of time through geological and geomorphological
processes. These slopes are only stable if the soil has sufficient strength to resist gravitational forces.
Changes in pore water pressure conditions, slope geometry and engineering works may cause these
natural slopes to fail.

Deep-seated failures are defined, geologically, as those that involve bedrock. Generally in the tropics
and sub-tropics they can be regarded as failures deeper than 5 metres. These landslides are usually
caused by adverse joints in the underlying rock mass, or by the development of water pressures at depth
in soils or within jointed rock. These types of failures often occur naturally on hillsides. If located on the
slopes below a road they can quickly expand upslope until the road is affected. If a road is constructed
across these landslides, it is likely that ground movements affecting the road will take place during rainfall.
Sometimes, slope excavation for road construction can initiate deep-seated landslides that can reactivate
later and result in frequent road blockage.

The techniques commonly used to protect and stabilise slopes and prevent the occurrence of landslides
above a road alignment, especially along low volume roads, are summarized in Table D.3.1. The use of
these techniques depends on site-specific conditions such as the size of the slide, soil type, road use, and
the cause of failure.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Table D.3.1: Stabilisation and protection measures for slope failures and erosion above the road (based on Hunt et al 2008)
D - Chapter 3 - 28

Instability Stabilisation options Drainage options Protection options


1. Erosion of None ƒ Usually none; ƒ In most cases, bio-engineering is
the cut slope ƒ Occasionally a cut-off drain above the adequate, usually grass slip planting;
surface cut slope can reduce water runoff; ƒ Where gullies are long or slopes are
however, these are difficult to maintain very steep, small check dams may be
and can contribute to instability if required;
blocked or otherwise disturbed. ƒ Sometimes a revetment wall at the
toe helps to protect the side drain.
2. Failures in cut ƒ Reduce the slope grade and if ƒ A subsoil drain may be required behind ƒ Bio-engineering is usually important
slope this is feasible, then add erosion a wall if there is evidence of water to prevent surface erosion and
protection; seepage; increase the resistance of the surface

PART D: EXPLANATORY
ƒ A retaining wall to retain the sliding ƒ Herringbone surface drains may soil.
mass; be required if the slope drainage is
ƒ For small sites where the failure impeded.
is not expected to continue, a

PART D: EXPLANATORY
revetment might be adequate.
3. Failures in cut ƒ Reduce the slope grade, and if this is ƒ A subsoil drain may be required behind ƒ Bio-engineering is usually important

NOTES FORNOTES
slope and hill feasible, then add protection; a wall if there is evidence of water to prevent surface erosion and
slope ƒ A retaining wall to retain the sliding seepage; increase the resistance of the surface
mass. This may need to be quite ƒ Herringbone surface drains may soil
large, depending on the depth of be required if the slope drainage is

LOW VOLUME
the slip plane. impeded.

FOR ROADS
4. Failures in hill ƒ Reduce the slope grade, and if this is ƒ A subsoil drain may be required behind ƒ Bio-engineering is usually important
slope but not feasible, then add protection; a wall if there is evidence of water to prevent surface erosion and
cut slope ƒ A retaining wall to support the seepage; increase the resistance of the surface
sliding mass, as long as foundations ƒ Herringbone surface drains may soil

ROAD DESIGN
can be found that do not surcharge be required if the slope drainage is
or threaten the cut slope. impeded.
5. Deep failure ƒ Consider re-alignment of road away ƒ Ensure road-side drainage is controlled ƒ Bio-engineering is usually important
in the original from instability to prevent surface erosion and
ground ƒ If slow moving, short term option increase the resistance of the surface
underneath may be to repave or gravel the road. soil.
the road
D - Chapter 3 - 29

3.3.2 Cut slopes

A major cause of cut slope failures is related to the release of stress within the soil upon excavation.
This includes undermining the toe of the slope and over-steepening the slope angle, or cutting into
heavily over-consolidated clays. Cut slope failures also occur where adverse joints within the underlying
rock mass become exposed in the excavation. Careful consideration should be given to prevent these
situations by designing cut slope angles to be stable (ie avoiding over-steep angles), or using retaining
walls (typically gabion or masonry) to provide the necessary support.

Sometimes cut slope failures occur where erosion in side drains removes sufficient support. This can be
prevented by constructing lined drains, keeping the base of the slope as supported as possible, or by
locating drainage ditches a suitable distance away from the toe of the cut. Consideration should also be
given to establishing vegetation on the slope to prevent long-term erosion.

Minor cut slopes are generally designed in a prescriptive way based on past experience with similar soil
and rock materials. Cut slopes greater than 3m in height usually require a more detailed engineering
geological assessment depending on the complexity of the ground conditions. This would include an
assessment of the strength of the soil or the orientation of joints in the rock. This assessment can be done
in a descriptive way with plots of some representative sections.

The slope angles indicated in Table D.3.2 have been provided as a guide for LVRs. Note that these angles
cannot be applied without due consideration of the ground conditions. In the design phase, the slope
angles can be used to establish the arrangement between sections of road on the same slope (eg hairpin
sections). During construction and maintenance, the slope angles are useful indicators to maintain the
stability of slopes and reduce the possibility for the occurrence of landslides.

Cuttings in strong rocks can often be very steep where adverse joints are not present, but in weathered
rocks and soils it is necessary to use shallower slopes. In heterogeneous slopes, where both weak and
hard rock occur, the appropriate cut-slope angle can be determined on the basis of the type of geological
materials exposed, the stratigraphy, and the variations in permeability between the different horizons.
One of the simplest ways to decide upon a suitable cut slope is to survey existing cuttings in similar
materials along other roads or natural exposures in the surrounding areas. Generally, new cuttings can be
formed at the same slope as stable existing cuttings if they are in the same material with the same overall
structure. Excavation of rock slopes should be undertaken in such a way that disturbance, for example
due to blasting, is minimised. It should also be undertaken in a manner to produce material of such size
that allows it to be placed in embankments in accordance with the requirements.

Table D.3.2: Common cut-slope ratios for LVRs.

Soil and rock condition (Horizontal : Vertical)


Most hard rocks (without
1:4 – 1:2
adverse jointing)
Closely fractured rock 1:2 – 1:1
Well consolidated highly to
completely weathered rock 1:2 – 1:1
(residual soil)
Dense coarse granular soils 1:2 – 1:1
Loose coarse granular soils 3:2
Plastic clay soils 2:1 – 3:1
Low cuts (less than 3 m high) 2:1

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 30

In general, single-sloped cuts along LVRs should be limited to a height of 5m. For deeper cuts, the
construction of benches should be considered to intercept falling debris and control the flow of water.
There is no hard rule regarding the dimension of benches, but a preliminary approach is to provide
bench widths that are one third of the height of the cut immediately above. Outward sloping benches are
generally not recommended because this may concentrate and erode channels through the bench if the
bench is in weathered rock or soil. If the bench is in strong, unweathered rock then this erosion will not
occur and outward sloping benches are permitted. In weaker materials the water should be encouraged
to drain along the bench to a discharge point rather than over it. Maintenance of these drains is important
to prevent water accumulating on the bench.

In upgrading or rehabilitation projects, the remediation of cut slope failures usually requires the removal
of failed material from the road and side drain and any overhangs and potentially unstable masses. A
decision has to be made as to whether the slope can be stabilised through earthworks, or whether a
retaining structure isrequired. In some cases the removal of slip debris can serve to undercut the slope
above causing further failure. In such cases a gabion or masonry retaining wall can be constructed to
support the slope. Most cut slope failures are shallow and the most common forms of treatment comprise
removal, trimming and drainage. Bio-engineering should be considered as an integral part of the solution.
Any cut slope where failure would result in large rehabilitation costs or would threaten public safety
should be designed using more rigorous techniques. Situations that warrant more in-depth analysis
include large cuts, cuts with complex geological structure (especially if weak zones are present), cuts
where high groundwater or seepage pressures are likely, cuts involving soils with low strength, cuts in
landslide debris, and cuts in formations susceptible to landslides.

3.4 Slope Instability below the road

3.4.1 Types of slope failures below the road

The type of slope failures and erosion processes commonly observed below a road, as shown in Figure
D.3.2, can be categorized into five classes as follows:
ƒ Erosion of fill slope surface;
ƒ Failures in fill slope;
ƒ Failures in fill slope and original valley slope;
ƒ Failures in original valley slope;
ƒ River under-cutting.

Most earth embankments or fill slopes face light to moderate erosion problems arising from rainfall
splash, surface runoff from the road and agricultural activities. The problem along roads is especially the
occurrence of surface runoff which results in sheet erosion and the formation of rills and gullies on poorly
compacted and unprotected embankment shoulders and slopes. The degree of erosion is normally a
function of material type and its compaction, rainfall or runoff intensity, slope angle, length of slope, and
vegetation cover.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 31

Figure D.3.2: Slope instabilities expected below the road (modified from Hunt et al, 2008).

Slope failures in fill slopes are often in the form of small-scale shallow translational slides, where the
failure is contained entirely within the embankment side slopes and maximum depth of rupture does not
exceed 2m. Generally, embankment stability is dependent on fill type and compaction, presence of water
or drainage provision, shrink and swell cycles, vegetation, slope angle and height, construction method
and type of foundation. Failure in embankments during and after construction can occur at the interface
between the natural ground and the fill, if the natural ground is incorrectly prepared.

In some cases, fills and embankments can serve to overload the natural slope upon which they are
constructed, thus creating instability that involves both the fill and valley slopes. This failure can extend
for some distance down-slope, depending on topography and the underlying geology. High groundwater
levels and the presence of perched or trapped water can often exacerbate the situation.

Sometimes when valley slopes are exposed to high stresses exerted by fill materials or rock wastes,
they may fail. This is especially true with unbalanced cut and fills, where excess material from the cut is
dumped onto the slopes below. The accumulation of this material together with any ponding in the rainy
season creates conditions that can cause valley slopes to fail.

When rivers erode their side-slopes, an undercut or over-steepened condition occurs, and the process
may give rise to slope failure, including the failure of any retaining walls or fill embankments constructed
on the slope.

Table D.3.3 gives slope stabilisation and protection options appropriate for fill and valley slopes below a
road. As with measures recommended for failures above the road, designs to stabilise and protect slopes
below the road are also typically site specific and may require local information from engineers.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Table D.3.3: Stabilisation and protection measures for slope failures below the road (Hunt et al., 2008)
D - Chapter 3 - 32

Instability Stabilisation options Drainage options Protection options


1. Erosion of the fill slope surface ƒ None ƒ Ensure road-side ƒ Bio-engineering is usually important
drainage is controlled to prevent surface erosion and
increase the resistance of the surface
soil.
2. Failures in fill slope ƒ Re-grade or remove, replace and ƒ Ensure road-side ƒ Bio-engineering is usually important
compact fill; drainage is controlled to prevent surface erosion and
ƒ Before replacing fill, cut steps increase the resistance of the surface
in original ground to act as key soil.
between fill and original ground;

PART D: EXPLANATORY
ƒ A new road retaining wall may be
the only option
3. Failure in fill slope and original ƒ Re-grade or remove, replace and ƒ Ensure road-side ƒ Bio-engineering is usually important
valley slope compact fill; drainage is controlled to prevent surface erosion and

PART D: EXPLANATORY
ƒ Before replacing fill, cut steps increase the resistance of the surface
in original ground to act as key soil.
between fill and original ground;

NOTES FORNOTES
ƒ A new road retaining wall may be
the only option.

LOW VOLUME
4. Failure in original valley slope ƒ Re-grade if sufficient space between ƒ Ensure road-side ƒ Bio-engineering is usually important
road and valley side; drainage is controlled to prevent surface erosion and

FOR ROADS
ƒ A new road retaining wall may be increase the resistance of the surface
the only option. soil.

5. Removal of support from ƒ May need extensive river training ƒ None ƒ Slope protection (walls and rip-rap

ROAD DESIGN
below by river erosion works to prevent further erosion. etc) may be necessary
D - Chapter 3 - 33

3.4.2 Fill slopes

Slope failures in fill slopes occur when there are changes in slope profile that add driving weight at the
top of the fill, or reductions in resisting load at the base. Stability is also reduced as a result of an increase
of pore water pressure resulting in a decrease in frictional resistance in cohesion-less soils, or swell in
cohesive soils. Water can also cause a progressive decrease in shear strength due to weathering, erosion,
leaching, opening of cracks and fissures, and softening and overstressing of the foundation soil. Usually,
short-term stability of embankments on soft cohesive soil is more critical than long-term stability, because
the foundation soil gains shear strength as the pore pressures dissipate. It may be necessary to check the
stability for various pore pressure conditions.

A wide range of slope stabilisation measures is available to address slope stability problems in fill slopes.
However, in most cases, the construction of more gentle slopes requires adequate compaction and
improved drainage to eliminate routine instability problems.

In general, for LVRs, fill slopes 3m or less in height with 2H:1V or flatter side slopes, may be designed
based on the strength characteristics of a compacted free draining granular fill and engineering judgment.
This is provided there are no known problem soils such as expansive or collapsible soils, organic deposits,
of soft and loose sediments.

Fill slopes over 3m in height or any embankment on soft soils, in unstable areas, or those comprised of
light weight fill require site-specific engineering geological assessment depending on specific ground
conditions. Embankments with side slope inclinations steeper than 2H:1V also require special attention.
Moreover, any fill placed near or against a bridge abutment or foundation, or that can impact on a nearby
structure, will likewise require stability analysis by a specialist engineer.

When deciding on the fill slope batters for design, it is advisable to consider the type of material that is
going to be used. For example, over-steep fill slopes formed in loose side-cast materials may continue
to ravel with time. Similarly, flat fill slopes formed by soft materials may be exposed to flooding and
settlement. Design should also consider the stability of existing fill slopes in the surroundings of the
project site. Observation of existing slopes should include vegetation, in particular the types of plants that
may indicate wet soil. Subsurface drainage characteristics may be indicated by the vegetative pattern. It
is also advisable to assess whether tree roots are providing anchoring of the different types of soils.

Generally, fill slope batters commonly used in different materials are summarized in Table D.3.4. Ideally,
fill slopes should be constructed with a 2H:1V or flatter slopes to promote vegetation growth. A suitable
combination of plants can help to stabilise fill slopes that are prone to very shallow translational slides.
Failures deeper than 1m would be out of the zone of influence of most plant rooting systems. Besides,
terraces, benches or bunds are desirable on large fill slopes to break up the flow of surface water.

Table D.3.4: Common fill slope batters for LVRs

Soil and rock condition (Horizontal : Vertical)


Soft clays or fills on wet areas and swamps (>3m) 2:1 - 3:1
Fills of most soils (free draining granular fill >3m) 3: 2 - 2:1
Fills of hard, angular rock (rock fill >3m) 3:2 – 5:4
Low fills (less than 3m high) 2:1 or flatter (for vegetation)

In addition to the selection of appropriate fill slope batters and the use of vegetation, the construction of
toe berms can also improve the stability of fill slopes by increasing the resistance along potential failure
planes. Toe berms are typically constructed of granular materials with relatively high shear strength that
can be placed quickly with minimum compaction. As the name indicates, toe berms are constructed near
the toe of slopes where stability is a concern. The sides of toe berms are often flatter than the slopes.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 34

In general, toe berms increase the shearing resistance by:


ƒ Adding weight and thus increasing the shear resistance of granular soils below the toe area of the
fill slope;
ƒ Adding high strength material for additional resistance along potential failure surfaces that pass
through the toe berm;
ƒ Creating a longer failure surface, thus more shear resistance, as the failure surface must pass below
the toe berm.

It is generally preferable to avoid the construction of fill slopes in swampy areas. When this cannot
be avoided, fill slopes through swampy areas must be constructed on a stable foundation. This helps
reduce problems which may occur within a short time after construction. The manner in which this is
accomplished, and the problems to overcome, depends largely on the type and depth of materials that
exist in the swamp.

In upgrading and rehabilitation projects, it may be more difficult to determine accurately the depth of
failure surface in fill slopes. Careful site inspection is therefore required. Failed fill material should be
excavated and stockpiled while the ground beneath is prepared. Organic, weathered and weak materials
should be removed, benched profiles created for a shear key, etc.) The fill should then be replaced and
compacted in layers until the final slope profile is achieved. Planting and drainage may be necessary.

During operation and maintenance, should embankments fail, they can either be:
ƒ Excavated and re-compacted if traffic flow and safety allows;
ƒ Excavated and replaced with granular materials if traffic flow and safety allows;
ƒ Supported by retaining walls where foundation stability and allowable bearing pressures permit;
ƒ Isolated from the road by construction of a road edge retaining wall founded beneath the zone of
movement.

3.5 Drainage
Slope instability is controlled to a large extent by water. Rainfall infiltrating through the ground surface
can concentrate within the slope. Excess water may decrease pore suction in the underlying soil and
create pore water pressure, reducing the effective stress and hence the stability of the slope. Hence, the
construction of surface and sub-surface drainage structures is vital to ensure that this excess water can be
intercepted and conveyed to a safe location where it will not create further instability problems. If water
flow in the slope cannot be controlled properly, then retaining structures may be ineffective.

Several surface drainage systems can be constructed on cuts, fills and natural slopes depending on
slope geometry and ground materials. The engineering design of surface drainage systems on the slope
is based on the amount of run-off. The runoff is a function of the catchment area, concentration time,
rainfall intensity, slope geometry, and surface conditions.

The types of surface drainage systems that can be used in low volume roads include open ditches
(rectangular, U-shaped, trapezoidal and semi-circular) and gravel filled trench drains. Trench drains are
in almost all cases rectangular. Open ditches should collect runoff from the catchment area and convey
water as efficiently as possible. Abrupt changes in flow direction may cause splashing, turbulence, and
erosion of ditch walls. Attempts should be made to avoid this. Wherever the flow velocity is high, energy
dissipators should be placed inside the channel.

On most slopes, it is common to construct down-slope channels with steps (stepped channels or
cascades). Like slope ditches, these stepped channels can be constructed in different shapes. Although
these channels are divided into different steps, they should be constructed as a single unit. Otherwise,
when sections are built independently, the flow kinetic energy can cause openings between sections,
which may result in water infiltration and differential settlements. It is worthwhile to note that the steps
may not provide an efficient energy dissipator when the volume of discharge is high. Hence, stepped
channels should not be built on slopes steeper than 50o because shooting (high-speed) flow will occur
whereby individual steps are bypassed. Where these drainage structures are constructed in slopes
steeper than 50%, energy dissipators such as up-stands should be introduced.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 35

Cut-off drains intercept surface and shallow subsurface flow and are normally placed at the crest of
the slope. They are internally filled with granular materials, perforated pipes, wrapped in geosynthetic
materials for drainage and filtering purposes. The draining material must collect and conduct the water
and have adequate permeability to allow flow to take place along the line of the drain rather than across
it. Free-draining inert material is used as backfill. Filter fabric is designed to allow water to penetrate, but
prevent fines from washing through. It must also be designed to prevent clogging.

A cut-off drain can be excavated with a minimum width of 0.5m at the bottom and 0.6m at the top. Its
depth normally varies between 1.0 and 1.5m. For deeper excavation, care should be taken with the
stability of walls. The choice of filling material and pipe perforations must satisfy filter design criteria.
When small quantities of seepage are to be removed, a single layer of well-graded moderately permeable
material meeting grain-size requirements may serve the dual purpose of filter and drain. But when large
quantities of seepage are to be removed, a filter layer is usually needed for the prevention of piping, and
a coarse layer for the removal of water. In areas where erosion is high, a graded filter material at the top
and an impermeable liner at the base could be needed. The rest of the drain is filled with free-draining
granular material. A slotted pipe at the base facilitates drainage, and it may need to be filter-wrapped.

In general, the use of surface drainage structures together with bio-engineering techniques and retaining
walls may suffice to reduce the occurrence of erosion and landslides on LVRs. Where there is an indication
that slope failure is partially or entirely due to groundwater, then the use of subsurface (sub-soil) drainage
systems such as deep cut-off and horizontal drains may be considered. Refer to the ERA Geotechnical
Design Manual for explanations on sub-surface drainage systems and their use in slope stabilisation.

3.6 Retaining walls

3.6.1 Design of Retaining walls

Retaining walls are used to resist the lateral pressure of soil where there is a desired change in ground
elevation that exceeds the angle of repose of the soil. This normally occurs when it is necessary to gain
roadway space without making large cuts into the hillside or constructing wide fills below the road.

Retaining walls must be designed to withstand the pressure exerted by the retained material attempting
to move forward down the slope due to gravity. The lateral earth pressure behind the wall depends on
the angle of internal friction and the cohesive strength of the retained material. Lateral earth pressures
are smallest at the top of the wall and increase towards the bottom. The total pressure may be assumed
to be acting through the centroid of a triangular load distribution pattern, one-third above the base of
the wall. The wall must also withstand pressure due to material placed on top of the fill behind the wall
(“surcharge”). Groundwater behind the wall that is not dissipated also exerts a horizontal hydrostatic
pressure on the wall and must be taken into account in the design. Dissipation of ground water is
normally achieved by constructing horizontal drains behind the wall with weep holes.

Gravity walls depend on their weight to resist pressures from behind the wall that tend to overturn the
wall or cause it to slide. A factor of safety of 1.5 should be applied to the calculations of overturning and
sliding. Gravity walls are normally designed with a slight “batter” to improve stability by leaning the wall
back into the retained soil. The foundations should be wide enough to ensure that excessive pressure is
not applied to the ground.

This manual covers the design and construction of gravity retaining walls, including gabion walls, dry
stone walls and mortared stone walls. These options are described below. Construction details are given
in Part E of the manual.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 36

3.6.2 Gabion walls

Gabion walls are built from gabion baskets tied together. A gabion basket is made up of steel wire mesh
in a shape of rectangular box. It is strengthened at the corners by thicker wire and by mesh diaphragm
walls that divide it into compartments. The wire should be galvanized, and sometimes PVC coated for
greater durability. The baskets usually have a double twisted, appropriate size, hexagonal mesh, which
allows the gabion wall to deform without the box breaking or losing its strength.

There are two types of gabions with different uses:


ƒ Gabion boxes are the heavier, more rigid form, with larger stones used in bank protection, aprons,
and retaining walls. Usually 1.0m high boxes are used for these purposes but sometimes 500mm
boxes may be used for rigid aprons.
ƒ Gabion mattresses are thinner using smaller stones and mesh and, therefore, more flexible so that
they will fold down to protect scour holes. The maximum depth of the box for this purpose is often
300mm.

Gabion walls are cost effective because they employ mainly locally available rock and local labour. Gabion
structures are commonly used for walls of up to 6m high. Gabion walls are usually preferred where the
foundation conditions are variable, the retained soils are moist, and continued slope movements are
anticipated.

Because of their inherent flexibility, they are not preferred as retaining walls immediately below and
adjacent to sealed roads due to the likelihood of movement of the backfill behind the wall and subsequent
pavement cracking. Where gabion walls are used to support a sealed road, care should be taken to locate
the base of the wall on a good foundation, in order to reduce the potential for movement.

Gabion walls have the following advantages:


ƒ Gabions can be easily stacked in different ways, with internal or external indentation to improve
the stability of the wall;
ƒ They can accommodate some movement without rupture;
ƒ They allow free drainage through the wall;
ƒ The cross section can be varied to suit site conditions;
ƒ They can take limited tensile stress to resist differential horizontal movement.

Their disadvantages include:


ƒ Gabion walls need large spaces to fit the wall base (this base width normally occupies about 40%
to 60% of the height of the wall);
ƒ The high degree of permeability can result in a loss of fines through the wall. For road support
retaining walls this can result in potentially problematic settlement behind the wall, although this
can be prevented by the use of a geo-textile (filter fabric) between the wall and the backfill.

3.6.3 Dry stone walls

Dry-stone walls are constructed from stones without any mortar to bind them together. The stability of
the wall is provided by the interlocking of the stones. The great virtue of dry stone walls is that they are
free-draining. The durability of dry-stone walls depends on the quality and amount of the stone available
and the quality of the work. In a slope management situation, they are useful as revetments for erosion
protection and as a means of supporting soil against very shallow movement. Dry stone walls should not
exceed 5m in height.

3.6.4 Mortared masonry walls

As with gabion walls and dry stone walls, a mortared masonry wall design uses its own weight and
base friction to balance the effect of earth pressures. Masonry walls are brittle and cannot tolerate large
settlements. They are especially suited to uneven founding levels but perform equally well on a flat
foundation. Mortared masonry walls tend to be more expensive than other gravity wall options. If the
wall foundation is stepped along its length, movement joints should be provided at each change in wall
height so that any differential settlement does not cause uncontrolled cracking in the wall.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 37

Mortared masonry walls require the construction of weep-holes to prevent build up of water pressure
behind the wall. Weep holes should be of 75mm diameter and placed at 1.5m centres with a slope of 2%
towards the front of the wall. A filter of lean concrete or geo-textile should be placed at the back of the
weep holes to permit free drainage of water.

3.7 Bio-engineering methods


In the design of LVRs, both cut and fill slopes should be constructed so that they can be vegetated.
Hill and valley slopes should be protected using plants. Certain types of plants, arranged in particular
configurations, can be used to control erosion and reduce the likelihood of shallow landslides. Vegetation
is unlikely to have a significant impact on slope stability where slip planes are deep, due to the shallow
rooting nature of many species. However, the condition of the upper layer, stabilised by plant roots, may
be sufficient to maintain the stability of the slope.

Different types of plants and planting materials give rise to a variety of rooting patterns, with the result
that the surface layer of soil will be bound together and have its resistance to deformation, failure and
landslides increased.

Grasses are very quick growing plants that offer a dense protective ground cover. Since their stem is
at the ground level, moderate harm to the area by humans or domestic animals does not cause lasting
damage to them. In many cases, grass re-grows quickly if damaged. Grasses with their dense network of
shallow roots are useful in protecting sites from surface erosion. However, some species such as vetiver
have very deep root systems and can offer greater potential for shallow slope stabilisation.

Herbs have little or no woody tissue. They tend to grow closer to the ground providing a dense ground
cover with shallow root systems. Together with other plants such as grasses and shrubs, they can protect
the slope from erosion.

Woody plants and shrubs are other types of vegetation that can increase stability in slopes. A woody
plant has a perennial woody stem and supports vegetative growth. Shrubs are defined as low-growing
woody plants with multiple stems. Shrubs can vary in height from 0.2m to 6m. In areas where visibility is
essential, such as road curves, shrubs are preferred to trees as they do not grow as large and are easier
to control and maintain. Although root systems may not spread as deep and as far as tree root systems,
the tensile strength developed in soils as a result of the presence of shrubs is normally sufficient to hold
soil materials together.

Trees are perennial woody plants having a main stem and root. Tree roots can extend to considerable
depth and over a large area if sub-soil stratigraphy permits. Therefore, trees are often considered
suitable for reinforcing slopes with deep soil profiles. Trees have been classified as having three main
root system types: plate, heart and tap. Plate root systems have large lateral roots and vertical sinkers,
heart systems possess many horizontal, oblique and vertical roots, and tap systems have one large central
root and smaller lateral roots. Trees having heart and tap root systems have been classified as being the
most resistant to uprooting. Individual roots within a system may be further classified into sub-groups
depending on their morphology and function. Extensive roots are those which grow to large depths and
spread over large areas, while intensive roots are short, fine roots and tend to be much more localised.

3.7.1 Selection of appropriate plant species

The main factors to be addressed when selecting the particular species for use in bio-engineering works
are as summarized as (Hunt et al, 2008):
ƒ The plant must be of the right type to undertake the bio-engineering technique that is required.
The possible categories include:
• A grass that forms large clumps,
• A shrub or small tree that can be grown from woody cuttings,
• A shrub or small tree that can grow from seed in rocky sites,
• A tree that can be grown from a potted seedling,
• A large bamboo that forms clumps.
ƒ The plant must be capable of growing in the location of the site.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 38

Other points to be considered are:


ƒ There is no single species or technique that can resolve all slope protection problems. It is always
advisable to use local species which don’t invade and harm the environment, and were able to
protect the slope from sliding in the past.
ƒ Large trees are suitable on slopes of less than 3H:2V or in the bottom 2m of slopes steeper than
3H:2V. Maintaining a line of large trees at the base of a slope can help to buttress the slope and
reduce undercutting by streams.
ƒ Grasses that form dense clumps generally provide robust slope protection in areas where rainfall
is intense. They are usually best for erosion control, although most grasses cannot grow under the
shade of a tree canopy.
ƒ Shrubs (ie woody plants with multiple stems) can often grow from cuttings taken from their
branches. Plants propagated by this method tend to produce a mass of fine, strong roots. These
are often better for soil reinforcement than the natural rooting systems developed from a seedling
of the same plant.
ƒ In most cases the establishment of full vegetation cover on unconsolidated fill slopes may take one
to two rainy seasons. Likewise, the establishment of full vegetation on undisturbed cut slopes in
residual soils and colluvial deposits may need 3 to 5 rainy periods. Less stony and more permeable
soils have faster plant growth rates, and drier locations have slower rates.
ƒ Plant roots cannot be expected to contribute to soil reinforcement below a depth of 500mm.
Plants cannot be expected to reduce soil moisture significantly at critical periods of intense and
prolonged rainfall.
ƒ Grazing by domestic animals can destroy plants if it occurs before they are properly grown. Once
established, plants are flexible and robust. They can recover from significant levels of damage (eg
flooding and debris deposition).

3.7.2 Site preparation

Before bio-engineering treatments are applied, the site must be properly prepared. The surface should
be clean and firm, with no loose debris. It must be trimmed to a smooth profile, with no vertical or
overhanging areas. The object of trimming is to create a semi-stable slope with an even surface, as a
suitable foundation for subsequent works.
ƒ Trim soil and debris slopes to the final desired profile, with a slope angle of between 30° and 60°
(in certain cases the angle will be steeper, but review this carefully in each case). Trim off excessively
steep sections of slope, whether at the top or bottom. In particular, avoid slopes with an over-steep
lower section, since a small failure at the toe can destabilise the whole slope above.
ƒ Remove all small protrusions and unstable large rocks. Eradicate indentations that make the
surrounding material unstable by trimming back the whole slope around them. If removing
indentations would cause an unacceptably large amount of work, excavate them carefully and
build a buttress wall. Remove all debris from the slope surface and toe to an approved tipping site.
If there is no toe wall, the entire finished slope must consist of undisturbed material.
ƒ When materials form the lower parts of slopes to be trimmed, the debris can be used for backfilling.
In this case, compact the material in layers, by ramming it thoroughly with tamping irons. This must
be done while the material is moist.

3.7.3 Recommended techniques

Table D.3.5 provides the different types of bio-engineering techniques recommended for various kinds
of slopes and soil materials above the road structure. Similarly, the techniques useful for slopes below the
road are summarized in Table D.3.6. The types of plants commonly used for erosion control at any site in
the surroundings of the right-of-way of the road are given in Table D.3.7. Many of these techniques, as
they are compiled by Hunt et al (2008), have been used successfully in many countries notably in Asia.

In Ethiopia, different regions use different types of plants for erosion control and slope stabilisation.
These plants, among many others, include Chrysopogon zizanioides (Vetiver grass), Agave sisalana,
(Sisal, Kacha), Aloe vera (Eret), Euclea schimperi (Dedeho), Calpurnia subdecandra (Laburnum, Digita),
Juniperus procera (Cedar, Tid), Hyparrhenia rufa (Senbelet), and Oxytenanthera abyssinica (Lowland
Bamboo).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 39

Table D.3.5: Recommended bio-engineering techniques for slopes above the road.

Site characteristics Recommended techniques


Cut slope in soil, very highly to completely
weathered rock or residual soil, at any grade
up to 1H:2V
Cut slope in colluvial debris, at any grade
up to 1H:1V (steeper than this would need a
retaining structure) Grass planting in lines, using slip cuttings.

Trimmed landslide head scarps in soil, at any


grade up to 1H:2V
Roadside lower edge or shoulder in soil or
mixed debris
Cut slope in mixed soil and rock or highly
weathered rock, at any grade up to about
1H:4V
Direct seeding of shrubs and trees in crevices.
Trimmed landslide head scarps in mixed soil
and rock or highly weathered rock, at any
grade up to about 1H:4V

Chrysopogon zizanioides (Vetiver grass) is a perennial grass with short rhizomes and massive, finely
structured root system that grows very quickly in some places. Its root depth reaches 3-4 m in the first
year. The deep root makes vetiver grass extremely drought tolerant and very difficult to dislodge when
exposed to a strong water flow and subsequent erosion. It is also highly resistant to pest, disease and fire.

According to Mekonnen (2000), vetiver grass was first introduced to Ethiopia from India at the start of
the 1970s by the State Coffee sector and the Ethiopian Institute of Agricultural Research. Since then, the
grass is increasingly used for soil and water conservation and stabilisation of steep slopes on different
slope classes. The upper slope limit lies between 40% and 45%. The vertical intervals between units vary
from one slope class to the other. The vertical classes recommended in the country for slope classes
between 3-15%, 16-25%, and greater than 25% are 2m, 1.5m and 1m, respectively. For very steep slopes,
it is advisable to plant grass units closer right after the rainy season. The effective elevation limit for its
use is about 2,800 m.

Agave sisalana (Sisal, locally called Kacha) is a plant that yields a stiff fibre used in some countries to
make twine, rope and dartboards. Depending on context, the term may refer either to the plant or the
fibre. Sisal plants consist of a rosette of sword-shaped leaves about 1.5m to 2m tall. They are common in
the northern, western and eastern part of Ethiopia. In the east, sisal resists the semi-arid and arid climate
where the average rainfall is around 500mm per year and grows on rocky slopes. In Tigray, sisal is used
on road cuts and natural slopes to control erosion and landslides. Sisal is valued for making geotextiles
because of its strength, durability, and ability to stretch. These textile fabrics are used in geotechnical
engineering for soil erosion control. Biodegradable products create a stable temporary slope environment
in which plants can develop and then break down to add nutrients to the soil.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 40

Table D.3.6: Bio-engineering methods useful to improve slope stability below the road

Site characteristics Recommended techniques


Fill slopes and backfill above walls without a
Brush layers (live cuttings of plants laid into
water seepage or drainage problem; these
shallow trenches with the tops protruding) using
should first be re-graded to be no steeper than
woody cuttings from shrubs or trees
3H:2V
Debris slopes underlain by rock structure, so Palisades (the placing of woody cuttings in a line
that the slope grade remains between 1H:1V across a slope to form a barrier) from shrubs or
and 4H:7V trees
Other debris-covered slopes where cleaning
Brush layers using woody cuttings from shrubs or
is not practical, at grades between 3H:2V and
trees
1H:1V
Fill slopes and backfill above walls showing Fascines (bundles of branches laid along shallow
evidence of regular water seepage or poor trenches and buried completely) using woody
drainage; these should first be re-graded to be cuttings from shrubs or trees, configured to
no steeper than about 3H:2V contribute to slope drainage
Large and less stable fill slopes more than 10m
from the road edge (grade not necessarily Truncheon cuttings (big woody cuttings from
important, but likely eventually to settle trees)
naturally at about 3H:2V)
Large bamboo planting; or tree planting using
The base of fill and debris slopes
seedlings from a nursery

Like sisal, Aloe vera (Eret) is commonly found in warm, fertile regions of Ethiopia where it is capable of
withstanding very long periods of drought. There are also many species of Aloe vera in the dry parts of
Ethiopia that are commonly used for erosion control. The plant takes four to five years to reach maturity,
at which time its leaves, which grow from a short stem, are about 60cm in length and 8-10cm wide at the
base. Being perennial in behaviour, Aloe vera has a lifespan of about 12 years, and if planted early, it may
protect cut and fill slopes throughout most of the design life of the road.

Euclea schimperi (Dedeho) is a widespread Afromontane shrub or small tree common in overgrazed
rolling hills or fallow land. It is seldom consumed by domestic animals and can be used to stabilise cuts
and fills as its roots can hold soil particles together. Similarly, Calpurnia subdecandra (Laburnum, Digita),
and Juniperus procera (Cedar, Tid) are also traditionally used for protection of slopes from erosion.
Besides, hill slopes and the toe of road cuts are covered and stabilised by Hyparrhenia rufa (Senbelet).

Oxytenanthera abyssinica (Lowland Bamboo) is a clump forming, solid stemmed plant that is widely
distributed in the dry regions of the western part of Ethiopia. Traditionally, it has been used as a raw
material for building and making numerous household utensils, basketry, and handicrafts. Bamboo is very
drought resistant, sustains itself with minimal rainfall, and has a very economical water uptake. It thrives
in very poor, shallow soils unsuitable for most trees. This characteristics and its capability to grow in steep
areas makes it ideal to stabilise slopes. Besides, its ability to resist high tensile stresses makes the stem
useful for building retaining walls.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 41

Table D.3.7: Plants useful for erosion control in the surroundings of roads

Site characteristics Recommended techniques


Stream banks where minor erosion is Local plants including grasses, shrubs and bushes,
possible bamboos, etc
Gullies or seasonal stream channels with Live check dams using woody cuttings of shrubs and
occasional minor discharge trees
Gullies or seasonal stream channels with Stone pitching, probably vegetated. Gabion check
regular or heavy discharge dams may also be required
Other bare areas, such as on the land
above landslide head scars, on large Tree planting using potted seedlings from a nursery
debris heaps and stable fill slopes

When reinforcing soils against shallow slope instability, the most important criteria to consider are the
number, diameter, and tensile resistance of roots crossing the slip surface. Therefore, root systems
composed of deep taproots and sinker roots are ideal for this purpose. In contrast, as root tensile strength
is greater in thin woody roots, a large number of small diameter roots would provide a root-soil matrix
that resisted shear better. Vetiver grass is often used for vegetating shallow slope failures, due to its
relatively deep and fibrous root system. It is advisable to consult local agricultural experts to select plants
appropriate for different bio-engineering purposes.

At the top or toe of the slope, it is also be necessary to have roots that provide lateral reinforcement.
Thus, the ideal root morphology in shrubs and trees would be a heart root system, with deep sinkers and
wide-spreading lateral roots. In planar slides, tap-rooted species could be planted as the slip surface
would most likely be parallel to the slope. In many embankments and cuts, the slip surface is circular and
the root network should have sufficient depth to interact with the slip surface.

In active rock-fall corridors, mechanical properties of stem are more useful than root system morphology
for determining tree resistance. Nevertheless, if trees are well-anchored with a deep taproot, they will less
likely to uproot when hit by a failing rock.

3.8 Useful dos and don’ts


The following advice is given when considering slope stability problems and solutions for low volume
roads:
ƒ Always determine the cause and extent of the slope instability or erosion problem;
ƒ Prioritise each alignment and alignment section in order to identify critical areas;
ƒ Regularly inspect side slopes, culvert outlets and side ditches or drains to identify potential
problems;
ƒ Identify areas where land use activities are adversely affecting engineering performance and take
steps to rectify the situation;
ƒ Identify areas where engineering works are adversely affecting land use;
ƒ When scheduling retaining structures in unstable areas try to locate rock or firm ground for
foundation purposes;
ƒ Ensure adequate bearing capacity for all structures using in situ tests (eg DCP);
ƒ When building gabion (or masonry) structures in stream channels or on colluvial deposits, ensure
adequate foundation beneath potential scour depths or loose stratum and key structure adequately
into channels banks or side slopes;
ƒ Drainage control is a major factor in the maintenance of natural and man-made slopes. Drainage
must be diverted away from vulnerable areas and every attempt made to slow down but not
impede or pond drainage;
ƒ Select appropriate areas for spoil disposal.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 3 - 42

3.9 References:
Hunt T., Hearn G., Chonephetsarath X., and Howell J. (2008). Slope maintenance manual. Ministry of
Public Works and Transport, Roads Administration Division, Laos.

Mekonen A., 2000. Erosion control in agricultural areas: an Ethiopian perspective, IFSPEthiopia.
Proceedings 2nd International Vetiver Conference (ICV), Thailand.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 43

4. GEOMETRIC DESIGN

4.1 Introduction
Geometric design is the process whereby the layout of the road through the terrain is designed to
meet the needs of all the road users. The geometric standards are intended to meet two important
objectives namely to provide minimum levels of safety and comfort for drivers by provision of adequate
sight distances, coefficients of friction and road space for manoeuvres; and to minimise earthworks to
reduce construction costs.

Geometric design covers road width; cross-fall; horizontal and vertical alignments and sight lines; and the
transverse profile or cross-section. The cross-sectional profile includes the design of the side drainage
ditches, embankment heights and side slopes, and is a vital part of geometric design for low volume
roads. The cross-section essentially adapts the pavement to the road environment and is part of the
drainage design. For example, wide, sealed shoulders and high camber or cross-fall can significantly
improve the operating environment for the pavement layers by minimising the ingress of surface water.
Sub-surface water is a problem in low-lying flood-prone areas and whenever the road is in cut. Again, the
height of embankment and the depth and type of drainage ditch have very significant effects. Some of
these aspects are dealt with in the Drainage Chapter (Chapter 5).

4.1.1 Principal factors affecting geometric standards

The principal factors that affect the appropriate geometric design of a road are:
ƒ Cost and level of service;
ƒ Terrain;
ƒ Safety;
ƒ Pavement type;
ƒ Traffic volume and composition;
ƒ Roadside population (open country or populated areas);
ƒ Soil type;
ƒ Climate;
ƒ Construction technology; and
ƒ Administrative or functional classification.

The cost of a road is usually the most critical factor. It is also the most difficult to include in the setting of
the design standards. The standard of a road is essentially an index of its ‘service level’ but ‘service level’
is a rather imprecise term that means different things to different people. However, most would agree
that its main components include; speed of travel, safety, comfort, ease of driving, stopping and parking,
and reliable trafficability or passability. The chosen service level is directly associated with traffic volume
and, hence, is not treated as a separate variable. The standards for service level simply increase from the
lowest road class to the highest, remaining relatively constant within each class. (see Table A.3.1 in Part A)

Since these factors differ for every road, the geometric design of every road could, in principle, be
different. This is impractical and it is therefore normal practice to identify the main factors and to design
a fixed number of geometric standards to cope with the range of values of these key factors.

For LVRs in Ethiopia four basic standards are defined based on traffic levels (see Part B - Chapter 3).
These are then modified, sometimes quite considerably, to cater for the other key factors. The most
important of these are:
ƒ Terrain;
ƒ Traffic composition (including pedestrians and non-motorised vehicles);
ƒ Roadside land use – activities and population density;
ƒ Safety;
ƒ Pavement type (paved or unpaved).

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 4 - 44

Varying standards of geometric design do not exist to cater specifically for climate and soil type. However,
these factors are taken into account in the design of the drainage features of the road (Chapter D.5) and
affect the road cross-section thereby contributing to the geometric design.

The designer, therefore, has a very wide range of standards from which to choose, ensuring that a suitable
standard is available for almost all situations. However, there will be cases where it is impossible to meet
any of the standards, often because of extremely severe terrain conditions. Under such circumstances the
standards must be relaxed and road users must be warned of the reduction with suitable and permanent
signage.

4.1.2 How the Standards are used

A national ‘standard’ is not a specification, although it could, and often is, incorporated into specifications
and contract documents. Rather, a standard is a specific level of quality that should be achieved at
all times and nationwide. Amongst other things this ensures consistency across the country. For the
geometric standards, this means that road users know exactly what to expect. Drivers, for example, are
not ‘surprised’ by unexpected changes in quality. Thus they will not unexpectedly find that a road is too
narrow, or that they have to alter their speed drastically to avoid losing control of their vehicle. Thus
standards are a guarantee of a particular quality level and, for roads, this is vital for reasons of safety.

It is important to note that there is no reason why a higher standard than the standard appropriate to the
traffic and conditions should not be used in specific circumstances. For example, for reasons of national
and international prestige or for strategic or military reasons, a road may be built to a higher standard
than would normally be justified eg a road to an international sports facility (where the traffic is low for
most of the time but can be quite high for short periods), the road to an airport, and roads to military
establishments. Thus higher standards can be used if required but lower standards should not be used
except in exceptional circumstances, for example, in particularly difficult terrain.

Figure D.4.1 shows how the appropriate geometric standards are selected.

Step 1: The first step is to determine the basic traffic level because this defines the road class (see Section
4.2.1 and 4.2.3). At this point, the proportion of heavy vehicles in the traffic stream is also determined.
This step is not specific to the geometric design and will usually have been done by the time it is necessary
to determine the geometric characteristics of the road. However, more details of the traffic are required
for the geometric design in terms of the other road users such as pedestrians, bicycles, motor cycles,
motor cycle taxis and animal drawn vehicles. These are taken into account in Step 5.

Step 2: The numbers and characteristics of all the other road users are considered (see Section 4.2.4).
It is here that the road layout may be altered and additional widths provided for safety and to improve
serviceability for all road users (eg reduce congestion caused by slow moving vehicles).

Step 3: The terrain class; flat, rolling, mountainous and escarpment is determined (see Section 4.2.5).

Step 4: The ‘size’ of the villages through which the road passes is evaluated to determine whether they
are large enough to require parking areas and areas for traders (see Section 4.2.6).

Step 5: For most road classes there are options for road type and therefore the next step is to decide
which type will be built (Section 4.2.7) In many cases the adoption of an EOD policy will mean that
different parts of the road may be designed with a different surfacing. The choice of road type is described
in Chapter D.6.

Step 6: From the available data the widths of carriageway and shoulders should be determined (see
Section 4.3.7). At this stage additional factors that affect the geometric standards are also considered
such as additional road safety features and the construction technology to be employed.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 45

Step 1 Step 1
Determine AADT of motorised Determine AADT of heavy
traffic trucks (3-axles or more)
(Section 4.2.1) (Section 4.2.3)

Select Road
Class

Step 2 Step 4
Step 3
Determine daily PCUs of non Determine nature of roadside
Determine terrain class
-motorised traffic population
(Section 4.2.5)
(Section 4.2.4) (Section 4.2.6)

Step 5
Select Road Type
or Types
Section 4.2.7

Step 6
Select widths of
carriageway & shoulders
Section 4.3.7

Determine a trial alignment


using the parameters
selected.

Figure D.4.1: Selection procedure for appropriate geometric standards

Once the basic parameters have been determined, the appropriate Table from Tables B.3.11 to Table
B.3.17 is selected. This provides details of the other geometric factors that are needed to carry out the
geometric design.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 46

The completion of the process is the design of a trial alignment as a check to ensure that all the standards
have been met. If not, alternative alignments should be tried. In extreme conditions it may not be possible
to adhere to all the standards at all points along the road. In such cases engineering judgement or additional
technical advice may be needed.

The pre-feasibility study should have shown that the costs of the road are likely to be acceptable. However,
at this stage it may be found that the engineering problems are more costly to solve than anticipated. This
needs to be checked and a final alignment selected. If the costs are too high then the project will need to
be reviewed.

4.2 The principal factors determining the choice of geometric standard

4.2.1 Traffic volume

Roads are designed to provide good service for many years and therefore the traffic level to be used in
the design process must take into account traffic growth. Designing for the current traffic will invariably
lead to inadequate standards in the future unless the traffic growth rate is extremely low. To deal with
these uncertainties it is generally expected that there is a strong correlation between traffic level, traffic
growth rates and the functional classification of a road and therefore such a classification is often seen as
a suitable alternative to represent traffic.

However, although traffic levels often increase in line with the functional classification, this is not always
true and, furthermore, the traffic levels and growth rates are likely to differ considerably between different
areas and different regions of the country. For example, the traffic on a ‘collector’ road in one area of the
country might be considerably more than on a ‘main access’ road in another area. The design of the road,
and therefore the standards adopted, should reflect the traffic level. In addition, traffic growth rates are
often expected to be considerably higher on roads connecting district (wereda) centres than on roads
connecting villages but this is not always the case.

In general it is expected that growth rates on roads that do not have ‘through’ traffic (essentially feeder
roads) will have lower traffic growth rates than the higher classes of road but each situation should be
treated on its own merits taking into account any expected future developments.

For geometric design it is the daily traffic that is important. The approach recommended for estimating
the traffic for geometric design purposes is based on the estimated traffic level at the middle of the
design life period and this therefore requires an estimate of the traffic growth rate This method eliminates
the risk of under-design that may occur if the initial traffic is used and the risk of over-design if traffic at
the end of the design life is used. A design life of 15 years is recommended for paved roads and 10 years
for unpaved roads.

Normally a general growth rate is assumed or is provided by government based on the growth in
registered vehicles during previous years. However, local development plans may indicate higher growth
rates in some places.

Where there is no existing road, estimating the initial traffic is difficult and estimating future traffic even
more so. However, in many cases where a new road is proposed there is likely to be pedestrian traffic
and therefore some information on the likely vehicular traffic after the road is constructed. In some cases
an economic evaluation may have been carried out to justify the road in the first place. This will have
provided an estimate of the amount of goods transported by pedestrians and the likely amount that will
be carried by vehicles. In the unlikely event that there is no information available, the lowest class of
engineered road (DC1) should be designed. Historical growth rates of similar roads in any specific area
should be considered if available.

It should be noted that the issue of road classification to determine the standards to be applied is not
difficult. A maximum of four different standards are defined for LVRs (DC1-DC4 – Tables B.3.1 and B.3.2)
and each will be applicable over a specific traffic range. These ranges are therefore quite wide and little

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 47

difficulty should normally be experienced in assigning a suitable standard to a new road project. Where
the expected traffic is near to a traffic boundary, it would be prudent to use a higher classification.

4.2.2 The design vehicle

For geometric design it is the physical dimensions of a vehicle that are also important. A truck requires
more space than a motorcycle, for example, and this does not depend on whether the truck is empty or
fully loaded.

The way that vehicle size influences the geometric design of low and high volume roads is fundamentally
different. When the volume of traffic is high, the road space occupied by different types of vehicle is an
essential element in designing for capacity (ie the number of vehicles that the road can carry in a unit of
time -vehicles per hour or per day). For example, at the highest traffic levels, when congestion becomes
important, traffic volume dictates how many traffic lanes need to be provided.

For LVRs the volume of traffic is sufficiently low that congestion issues do not arise from traffic volume but
from the disparity in speed between the variety of vehicles and other road users which the road serves.
In other words the traffic composition is the key factor; traffic capacity is not the problem. Nevertheless it
is the size of the largest vehicles that use the road that dictates many aspects of geometric design. Such
vehicles must be able to pass each other safely and to negotiate all aspects of the horizontal and vertical
alignment. Trucks of different sizes are usually used for different standards – the driver of a large 5 or
6-axle truck would not expect to be able to drive through roads of the lowest standards.

In some countries the truck population in rural areas is predominantly one or two types and sizes of
vehicle. This makes it relatively easy to select a typical vehicle for setting geometric standards. Conversely,
some countries have a wide variety of truck sizes and selecting a suitable truck size for geometric design
is more difficult.

Good information on the vehicle fleet in Ethiopia is lacking but, in view of the low density of roads and,
hence, lack of alternative routes, together with the limited choice of vehicle for many transporters, it is
prudent to be conservative in choosing the design vehicle for each class of road so that the maximum
number of vehicle types can use them. In Ethiopia four different design vehicles have been used as
shown in Table D.4.1 and Table D.4.2. However there is very little difference between design vehicles
DV2 and DV3. Roads designed for the single unit truck will be suitable for the bus provided the front
and rear overhangs of the bus are taken into account when designing curves; and this can be done with
suitable curve widening where required as described later. The standard for only the lowest class of road
is insufficient for DV2 and DV3.

Diagrams showing the full minimum swept out path of the design vehicle are shown in ERA’s Geometric
Design Manual (2011).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 48

Table D.4.1: Design vehicle characteristics

Minimum
Front Rear
Design Height Width Length Wheelbase turning
Designation overhang overhang
vehicle (m) (m) (m) (m) radius
(m) (m)
(m)
4x4
DV1 1.3 2.1 5.8 0.9 1.5 3.4 7.3
Utility

Single
unit DV2 4.1 2.6 11.0 1.5 3.0 6.5 12.8
truck
Single
unit DV3 4.1 2.6 12.1 2.1 2.4 7.6 12.8
bus
Truck
and 4.8+8.4
DV4 4.1 2.6 15.2 1.2 1.8 13.7
semi- =13.2
trailer

Table D.4.2: Design vehicle for each LVR class

Design standard Design vehicle


DC4 DV4
DC3 DV3
DC2 DV3
DC1 DV1

4.2.3 Traffic composition – proportion of heavy vehicles

The density of roads in Ethiopia is quite low and one of the consequences of this is that the proportion
of heavy vehicles in the traffic stream on LVRs is often quite high. Design standards DC2, DC3 and DC4
include a modification to cater for this.

For DC4, if the number of ‘large’ vehicles, defined as 3-axled (or more) trucks with GVWs (Gross Vehicle
Weights) potentially greater than 12 tonnes, is greater than 40 the width of the paved surface is increased
to 7.0m and the shoulders reduced to 1.0m. If there are more than 80 large vehicles then the standard for
DC5 (as defined in the Geometric Design Manual – 2011 series) should be used instead of DC4.

For DC3, if the number of large vehicles is greater than 25, design standard DC4 should be used and, for
DC2, if the number of large vehicles exceeds 10 then DC3 should be used.

4.2.4 Traffic composition - use of Passenger Car Units (PCUs)

In order to quantify traffic for normal capacity design the concept of equivalent PCUs is often used. Thus
a typical 3-axle truck requires about 2.5 times as much road space as a typical car hence it is equivalent to
2.5 PCUs. A motor cycle requires less than half the space of a car and is therefore equivalent to 0.4 PCU’s.

The PCU concept is very useful where traffic congestion is likely to be a problem and it was not originally
intended for use in the geometric design of LVRs. However, vehicles that move slowly cause congestion
problems because of their speed rather than because of their size. In effect, they can be considered

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 49

to occupy more road space than would be expected from their size alone. The actual PCU rating of a
vehicle is affected by the function of a road (i.e. the nature of the other traffic) and varies as the traffic
mix varies and as the traffic volume and traffic speeds vary. Nevertheless, in situations where the number
of slow moving vehicles, both motorised and non-motorised, is significant, in order to retain the level of
service appropriate to the traffic level of motorised vehicles, the road standard should be improved by
reducing congestion and this is best done by widening the shoulders. Thus when the PCU level of the
slow moving and intermediate forms of transport reaches a certain level, shoulder widening is justified.

The PCU concept is also useful for identifying the need for additional safety features where the numbers
of pedestrians and slow moving vehicles are high.

The PCU values for Ethiopia are shown Table D.4.3. Motorcycle taxis (eg bajaj) are becoming popular
in urban situations and it is only a matter of time before these spread to more rural areas and become
adapted for freight as well as for passenger transport.

Table D.4.3: PCU values

Vehicle PCU value


Pedestrian 0.15
Bicycle 0.2
Motor cycle 0.25
Bicycle with trailer 0.35
Motor cycle taxi (bajaj) 0.4
Motor cycle with trailer 0.45
Animal drawn cart 0.7
Bullock cart 2.0
All based on a passenger car = 1.0

4.2.5 Terrain

Terrain has the greatest effect on road costs therefore it is not economical to apply the same standards in
all terrains. Fortunately drivers of vehicles are familiar with this and lower standards are expected in hilly
and mountainous terrain.

Four categories have been defined which apply to all roads as follows:
Flat 0-10 five-metre contours per km. The natural ground slopes perpendicular to the
ground contours are generally below 3%.
Rolling 11-25 five-metre contours per km. The natural ground slopes perpendicular to the
ground contours are generally between 3 and 25%.
Mountainous 26-50 five-metre contours per km. The natural ground slopes perpendicular to the
ground contours are generally above 25%.
Escarpment Escarpments are geological features that require special geometric standards
because of the engineering problems involved. Typical gradients are greater than
those encountered in mountainous terrain.

The following information is particularly relevant to LVR design:


Hilly terrain
An important aspect of geometric design concerns the ability of vehicles to ascend steep hills. Roads
that need to be designed for very heavy vehicles or for animal drawn carts require specific standards to
address this, for example, special climbing lanes. Fortunately the technology of trucks has improved

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 50

greatly over the years and, provided they are not grossly overloaded (which is a separate problem) or
poorly maintained, they do not usually require special treatment. On the other hand, animal drawn
vehicles are unable to ascend relatively low gradients and catering for them in hilly and mountainous
terrain is rarely possible. Climbing lanes cannot be justified on LVRs and nor can the provision of very
low maximum gradients. The maximum gradients allowable for different road classes are shown in Tables
B.3.11 to Table B.3.17.

Mountainous and escarpment terrain


In mountain areas the geometric standard for LVRs takes account of the constraints imposed by the
difficulty and stability of the terrain. This design standard may need to be reduced locally in order to
cope with exceptionally difficult terrain conditions. Every effort should be made to design the road so
that the maximum gradient does not exceed the standards shown in Tables B.3.11 to Table B.3.17 but
where higher gradients cannot be avoided, they should be restricted in length. Gradients greater than
12% should not be longer than 250m and relief gradients are also required as indicated in the Tables.
Horizontal curve radii of as little as 13m may be unavoidable, even though a minimum of 15m is specified.

4.2.6 Roadside population (open country or populated areas)

The more populated areas in village centres are not normally defined as ‘urban’ but in any area having
a reasonable sized population or where markets and other business activities take place, the geometric
design of the road needs to be modified to ensure good access and to enhance safety. This is done by
using:
ƒ A wider cross section;
ƒ Specifically designed lay-byes for passenger vehicles to pick up or deposit passengers;
ƒ Roadside parking areas.

The standards are specified in the Geometric Design manual - 2011 and are not specific to LVRs. However,
they are repeated here for completeness.

The additional width depends on the status of the populated area that the road is passing through. If the
road is passing through a Wereda seat or a larger populated area, an extra paved carriageway of 3.5m
width and to the same structural design as the main carriageway, is provided in each direction for parking
and for passenger pick-up and a 2.5m pedestrian footpath is also specified. The latter is essentially the
shoulder. In addition, the main running surface is paved and is 7.0m wide. Thus the road in such areas is
similar to Class DC4 but with an additional wide parking/activities carriageway and a footpath on each
side. The pavement structure of the wide parking should be identical to the pavement of the running
surface.

When passing through a Kebele seat a 2.5m paved shoulder is specified but no additional footpath;
although one could easily be provided if required. The carriageway is also increased to 7.0m and therefore
the standard is very similar to DC4 but with wider shoulders.

These standards are not justified for the lower traffic levels of DC2, which is a single carriageway, unless
the road is passing through a particularly well populated area that is not classified as a Kebele or Wereda
seat but where additional traffic may be expected. In such circumstances the shoulders should be widened
to 2.5m for the extent of the populated area.

4.2.7 Pavement type

For a similar ‘quality’ of travel there is a difference between the geometric design standards required for
an unsealed road (gravel or earth) and for a sealed road. This is because of the very different traction and
friction properties of the two types of surfaces and the highly variable nature of natural materials. Higher
geometric standards are generally required for unsealed roads. A road that is to be sealed at a later date
should be designed to the higher, unsealed, geometric road standards.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 51

4.2.8 Soil type and climate

Soil type affects the ideal geometric design, principally in terms of cross-section rather than in terms
of the width of the running surface or road curvature. With some problem soils the cross-section can
be adjusted to minimise the severity of the problem by, for example, minimising the speed of water
flow; minimising the likelihood of excessive water inundation or penetration into the carriageway; and/
or moving problems areas further away from the carriageway itself. These aspects are dealt with in the
drainage section (Chapter 5) of the manual and in the pavement design section (Chapter 6).

Ideally maximum gradients for unpaved roads should also depend on soil types but this is usually
impracticable because, in most climatic regions, almost any gradient causes problems for unpaved roads.
Recent research has demonstrated that gravel-surfaced roads are unsustainable in many more situations
than has been thought previously and this applies equally to earth roads. Consequently every effort
is being made to introduce or to develop more sustainable surfacings for use where unpaved roads
deteriorate too quickly. Such surfacings cannot usually be justified for long stretches of road where they
are not essential hence the concepts of spot improvements and environmentally optimised design (EOD)
are being developed and refined.

4.2.9 Safety

Experience has shown that simply adopting ‘international’ design standards from developed countries
will not necessarily result in acceptable levels of safety on rural roads. The main reasons include the
completely different mix of traffic, including relatively old, slow-moving and usually overloaded vehicles;
a large number of pedestrians, animal drawn carts and, possibly, motorcycle-based forms of transport;
poor driver behaviour; and poor enforcement of regulations. In such an environment, methods to improve
safety through engineering design assume paramount importance.

Although little research has been published on rural road safety in Ethiopia, the following factors related
to road geometry are known to be important:
ƒ Vehicle speed;
ƒ Horizontal curvature;
ƒ Vertical curvature;
ƒ Width of shoulders.

These factors are all inter-related and part of geometric design. In addition, safety is also affected by:
ƒ Traffic level and composition;
ƒ Inappropriate public transport pick-up/set-down areas;
ƒ Poor road surface condition (eg potholes);
ƒ Dust (poor visibility);
ƒ Slippery unsealed road surfaces.

The last three factors are related to structural design covered in Chapter 5.

Conflicts between motorised vehicles and pedestrians are always a major safety problem on many rural
roads where separation is generally not economically possible. The World Bank Basic Access document
(World Bank, 2001) considers that there are sound arguments based on safety for keeping traffic speeds
low in mixed traffic environments rather than aiming for higher design speeds, as is the case for major
roads. The use of wider shoulders is also suggested. These considerations have been incorporated into
this manual.

Traffic level and composition are both considered. A considerable number of conflict situations can arise
when the number of PCUs of non-motorised traffic is large even though the number of two (or more)-
axled motorised traffic is quite low. Furthermore, the proportion of heavy vehicles on the LVRs of Ethiopia
can also be high, leading to more serious conflict situations. The overall traffic class standards are based
on the number of two (or more)-axled motorised vehicles but additional safety features are based on:
ƒ the number of PCUs of non 2-axled motorised vehicles and pedestrians; and
ƒ the proportion of heavy vehicles in the motorised stream.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 52

Pedestrians (and draft animals) find it very uncomfortable to walk on poorly graded gravel shoulders
containing much oversized material, especially in bare feet. They usually choose to walk on a paved
running surface, if available, despite the greatly increased safety risk. Thus, provision of a wider unsurfaced
shoulder does not ensure greater safety. On the approaches to market villages, where the pedestrian
traffic increases greatly on market days, provision of a separate footpath is the best solution provided
that the soil is suitable.

A checklist of engineering design features that affect road safety is given in Figure D.4.2. Not all are
suitable for rural roads but the general philosophy of design for safety is emphasised.

The following factors should be considered when designing for safety:


ƒ Wherever possible, non-motorised traffic should be segregated by physical barriers, such as raised
kerbs (through villages and peri-urban areas).
ƒ Designs should include features to reduce speeds in areas of significant pedestrian activity,
particularly at crossing points. Traffic calming may need to be employed (see Section 4.8.1).
ƒ To minimize the effect of a driver who has lost control and left the road, the following steps should
be taken.
• Steep open side-drains should be avoided since these increase the likelihood that vehicles
will overturn. (See Section 4.3.1).
• Trees should not be planted immediately adjacent to the road.
ƒ Guard rails should only be introduced at sites of known accident risk because of their high costs of
installation and maintenance.
ƒ Junctions and accesses should be located where full safe stopping sight distances are available
(see Section 4.4).

4.2.10 Construction technology

In a labour-abundant economy it is usually beneficial to maximise the use of labour rather than rely
predominantly on equipment-based methods of road construction. In such a situation the choice of
technology might affect the standards that can be achieved, especially in hilly and mountainous areas.
This is because:
ƒ Maximum cuts and fills will need to be small;
ƒ Economic haul distances will be limited to those achievable using wheel-barrows;
ƒ Mass balancing will need to be achieved by transverse rather than longitudinal earth movements;
ƒ Maximum gradients will need to follow the natural terrain gradients;
ƒ Horizontal alignments will need to be less direct.

The standards in hilly and mountainous terrain are always lower than in flat terrain but this reduction in
standards need not necessarily be greater where labour-based methods are used. Following the contour
lines more closely will make the road longer but the gradients can be less severe. Every effort should be
made to preserve the same standards in the particular terrain encountered irrespective of construction
method.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 53

Undesirable Desirable Principle applied

Route
Major routes should by-pass
location
towns and villages

Road Gently-curving roads have


geometry lowest accident rates

Prohibit direct frontal access


to major routes and use
service roads

Roadside Use lay-bys or widened


access shoulders to allow villagers
to sell local produce

Use lay-bys for buses and


taxis to avoid restriction and
improve visibility

Seal shoulder and provide


rumble divider when
pedestrian or animal traffic is
significant

Segregate
Construct projected footway
motorised and
for pedestrians and animals
non-motorised
on bridges
traffic

Fence through villages and


provide pedestrian crossings

Figure D.4.2: Examples of effects of engineering design on road safety

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 54

4.2.11 Administrative function

In many countries it is necessary to take account of the administrative or functional classification of roads
because a certain standard may be expected for each functional class of road irrespective of the current
levels of traffic. Generally the hierarchy of administrative classification broadly reflects the traffic levels
observed but anomalies are common where, for example, traffic can be lower on a road higher in the
hierarchy. It is recommended that the standards selected should be appropriate to the task or traffic level
of the road in question but a minimum standard for each administrative class can also be defined if it is
policy to do so.

4.2.12 Matrix of standards

For each of the basic standards based on traffic level there are four standards to cope with terrain (flat,
rolling, mountainous and escarpment), a further two standards (for DC4, DC3 and DC2) to cater for
roadside population/activities and three standards to cater for traffic composition, essentially the number
of PCUs of non-motorised traffic (including pedestrians) and the percentage of heavy vehicles in the
traffic stream. These additional standards for traffic composition and roadside activities are essentially
standards to enhance safety.

Once these factors have been taken into account, safety alone no longer affects the number of road
standards because an acceptable level of safety must be applied to each road class. This will differ
between classes (greater safety features for higher traffic) but not within classes. The administrative
classification does not add to the number of standards either. If the traffic level indicates that a lower
standard than would normally be acceptable based on administrative classification is sufficient, the road
can be built to the minimum standard appropriate to its administrative classification.

Aspects of geometric design outlined in the following sections require particular consideration because
they have a major influence on the life-cycle costs of rural roads. The basis for developing the standards
is also discussed in these sections.

In contrast to the judgements required for quantifying traffic, the standards themselves are largely dictated
by the selected design speed and form a continuous range as design speed increases.

4.3 Cross sections


The cross section of a road is essentially a geometric design feature but is also intimately related to
drainage issues as well as slope stability and erosion problems in hilly and mountainous areas. The cross
section includes the shape and size of the running surface; shoulders; the side slopes of embankments;
side slopes to drainage ditches; drainage ditches themselves; and slopes to the batter.

The basic cross sections for LVRs are shown in Part B, Section 4.5. Some aspects of cross sectional design
are concerned with drainage and further details concerning this aspect are discussed in Chapter D.5.

The cross-section of a road may need to vary over a route but it is essential that any changes take place
gradually over a transition length. Abrupt and isolated changes lead to increased hazards and reduced
traffic capacity.

A common situation arises at bridge and water crossing points where the existing structure is narrower
than desired. In such situations warning signs must be erected to alert drivers. Fortunately many such
crossings are visible well in advance but if not, extra signage may be required.

4.3.1 Slopes of shoulders, side slopes, embankments

Side slopes should be designed to ensure the stability of the roadway and to provide a reasonable
opportunity for recovery if a vehicle goes out of control across the shoulders. In addition, the position of
the side drain invert should be a reasonable distance away from the road to minimise the risk of infiltration
of water into the road if the drain should be full for any length of time.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 55

Figure D.4.3 illustrates the general cross section and defines the various elements.

Figure D.4.3: Details of the road edge

The side slope is defined as ‘recoverable’ when drivers can generally recover control of their vehicles
should they encroach over the edge of the shoulder. Side slopes of 1:4 or flatter are recoverable. Research
has also shown that rounding at the hinge point and at the toe of the slope is also beneficial.

A non-recoverable slope is defined as one that is traversable but from which most drivers will be unable
to stop safely or return to the roadway easily. Vehicles on such slopes can be expected to reach the
bottom. Slopes of between 1:3 and 1:4 fall into this category.

A critical slope is one on which the vehicle is likely to overturn and these will have slopes of greater than
1:3.

The selection of side slope and back slope is often constrained by topography, embankment height,
height of cuts, drainage considerations, right of way limits and economic considerations. For rehabilitation
and upgrading projects, additional constraints may be present such that it may be very expensive to
comply fully with the recommendations provided in this manual.

Table D.4.4 indicates the side slope ratios recommended for use in the design based on the type of
material and the height of fills and cuts. This Table should be used as a guide only. Where slope stability
problems are likely, slope configuration and treatment should be based on expert advice.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 56

Table D.4.4: Safety of slopes (ratios are vertical:horizontal)

Height of Side slope Safety


Material Back slope
slope (m) Cut Fill classification

0.0-1.0 1:4 1:4 1:3 Recoverable


Not
Earth(1) 1.0-2.0 1:3 1:3 1:2
recoverable
>2.0 1:2 1:2 1:1.5 Critical
Rock Any height Dependant on costs Critical

Expansive 0-2.0 n/a 1:6


Recoverable
clays(2) >2.0 n/a 1:4
Notes:
1. See Section
2. Certain soils may be unstable at slopes of 1:2. Geotechnical advice require
3. The drainage ditch should be moved away from the embankment

4.3.2 Roadside ditches

Detailed information concerning roadside ditches is provided in Chapter D.5, Section 5.4.4.

Side drains should be avoided when the road traverses areas of expansive clays. Water should be
discharged uniformly along the road. Where side drains cannot be avoided they should be a minimum
distance of 4m from the toe of the embankment and should be trapezoidal in shape.

4.3.3 Clear zones

The discussion in Section 4.3.1 has highlighted the safety aspects of embankment side slopes. However,
many accidents are made more severe because of obstacles that an out-of-control vehicle may collide
with. The concept of clear zones identifies these obstacles and attempts to eliminate such hazards.

The most common hazards are headwalls of culverts and road signage. The clear zone defined for high
volume roads is substantial (15m is typical) but for LVRs this is impractical. Ideally it should extend at least
to the toe of the embankment and should always be greater than 1.5 m from the edge of the carriageway.
At existing pipe culverts, box culverts and bridges the clear zone cannot be less than the carriageway
width. If this criterion cannot be met, the structure must be widened. New pipe and box culverts must be
designed with a 1.5m clearance from the edge of the shoulder. Horizontal clearance to road signs and
marker posts must also be an absolute minimum of 1.5m from the edge of the carriageway.

4.3.4 Right-of-way

Right-of-way (or the road reserve) is provided to accommodate road width and the drainage requirements;
to enhance safety; to improve the appearance of the road; to provide space for non-road travellers; and
to provide space for upgrading and widening in the future. The width of the right-of-way depends on
the cross-sectional elements of the highway, topography and other physical controls; plus economic
considerations. Although extended rights-of-way are convenient, right-of-way widths should be limited
to a practical minimum because of their effect on local economies.

Rights-of-way are measured equally each side of the centre line. Road reserve widths applicable for the
different road classes are shown in Table D.4.5. In mountainous terrain where large cuts are required, the
total width can exceed the right-of-way width.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 57

Table D.4.5: Right of way widths

Road Class Total Right of Way (m)


DC4 50
DC3 30
DC2 30
DC1 20

4.3.5 Shoulders

The shoulders of a road must fulfil the following functions:


ƒ A structural function;
ƒ Allow wide vehicles to pass one another without causing damage to the shoulder;
ƒ Provide safe room for temporarily stopped or broken down vehicles;
ƒ Allow pedestrians, cyclists and other vulnerable road users to travel in safety;
ƒ Allow water to drain from within the pavement layers;
ƒ Reduce the extent to which water flowing off the surface can penetrate into the pavement (often
done by extending a seal over the shoulder).

Shoulders have an important structural function which is often overlooked in the provision of LVRs. They
act as edge supports to contain the running carriageway; without adequate shoulders the road will move
laterally and deform. Therefore, there is a minimum width of shoulder that is required to perform this
function. Depending on the properties of the material and the traffic, this can range from 0.5 to 1.5m.

Shoulders also have to perform an important traffic carrying function for non-motorised vehicles and
pedestrians. Wider shoulders are required when this traffic is high enough. In addition, wider shoulders
are provided for some classes of road when the proportion of heavy vehicles in the traffic stream exceeds
certain values.

When the road passes through denser areas of population, additional width is provided for parking and
for other roadside activities. This widening may be considered to be shoulder widening although the
need to provide access to shops and market areas means that the construction is usually of an extra
carriageway.

Where the carriageway is paved, the shoulder may be gravel or may be sealed with a bituminous surface
treatment. The structural advantages of a sealed shoulder are discussed in Chapter 6, Section 6.18.
However, sealing the shoulders whenever the numbers of non-motorised traffic exceeds a critical value is
recommended in order to encourage the travellers to use the shoulders rather than the carriageway. On
the approaches to villages and towns the local traffic builds up quite quickly and therefore consideration
should be given to extending the sealed shoulders for considerable distances each side of the town/
village. No standard guidance can be given; each situation should be treated on its merits.

Shoulders constructed with the same material as the carriageway (earth or gravel) should have the same
cross-fall as the carriageway. If the shoulders are gravel and the carriageway is paved the cross-fall of the
shoulder should be 1.5-2.0 % steeper than that of the carriageway.

Shoulder widths in mountainous terrain and escarpments are reduced to minimise the high cost of
earthworks. Usually the design of the overall cross-section in such terrain will include significant drainage
and erosion control features and the shoulder will form an important component of this (Chapter 6,
Section 6.18.3.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 58

4.3.6 Single lane roads and passing places

There is good agreement internationally about the recommended carriageway width for single-lane
roads, namely 3.0m. Passing places will be required, depending on the traffic level and provision for
other traffic and pedestrians will need to be introduced (wider shoulders) if the numbers of other road
users exceed specified levels. The increased width should allow two vehicles to pass at slow speed and
hence depends on the design vehicle.

Passing places should normally be provided every 300m to 500m depending on the terrain and geometric
conditions. Care is required to ensure good sight distances and the ease of reversing to the nearest
passing place, if required. Passing places should be built at the most economic places rather than at
precise intervals provided that the distance between them does not exceed the recommended maximum.
Ideally, the next passing place should be visible from its neighbour.

The length of passing places is dictated by the maximum length of vehicles expected to use the road,
indicating the need to define a design vehicle. The design vehicle DV3 is 12.8m long therefore passing
places of twice this length should be provided. In most cases, a length of 25m will be sufficient for rural
roads.

A suitable width depends upon the width of the road itself. The criterion is to provide enough overall
width for two design vehicles to pass each other safely at low speed. Therefore, a total traffickable
minimum width of 6.3m is required (providing a minimum of 1.1m between passing vehicles). Allowing
for vehicle overhang when entering the passing bay, a total road width of 7.0m is suitable.

4.3.7 Width standards

Road width (running surface and shoulders) is one of the most important geometric properties since its
value is very strongly related to cost and to safety.

A review of international standards showed that some countries have adopted road widths that are
intermediate between single lane and two lane requirements. Such roads are considered to be dangerous
because vehicles try to pass each other at speed. Since there is not enough room to do so they are
forced onto the shoulder area and dangerously close to the road edge. If the road is paved, the edge of
the paved area becomes damaged very quickly.

The standards for Ethiopia do not include such intermediate widths. However, for all standards except
DC1, shoulders are widened if the number of road users other than 2-axled (and more) motorised vehicles
exceeds levels that cause too much interaction with the motorised traffic or if the proportion of heavy
vehicles in the traffic stream is high. For DC1 the traffic levels are so low that dangerous interactions will
be rare and drivers will expect other road users to have priority.

The basic vehicle classification in terms of traffic is shown in Table B.3.1.

Tables B.4.2 to B.4.8 show the standard road widths for each road class with the widths of running surface
and shoulder given for the paved road classes. The road width of unpaved roads shown in the tables
includes the widths of the shoulder as the surfacing of gravel or earth for these roads spreads across
the whole surface to the edge of the road making it difficult to define the shoulder for these roads. The
shoulder widths for paved roads are also varied for different terrains; for roadside population/activities;
and for traffic composition.

The lowest class (DC1) is a single lane road and the shoulder is effectively 0.75m wide. For DC2 the
minimum shoulder width is not specifically defined for the unpaved option but, for this class, the traffic
is less than 5 vehicles per hour in each direction which will invariably travel down the centre of the road
unless another vehicle is seen approaching. Therefore, the effective shoulder width for most of the time
is 1.5m. For this class, two vehicles can normally pass each other safely using the shoulder but if one of
the largest vehicles is involved the vehicles may need to slow down. If there are sufficient of these larger
vehicles, then class DC3 should be used.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 59

A similar argument applies to DC3. The unpaved option for DC3 is wider than DC2 and allows easier
passing but the level of service is not commensurate if there are a large number of the larger vehicles in
the traffic stream. In this case, the next higher class (DC4) is recommended. For DC4 the shoulders are
at least 1.0m wide.

Additional shoulder widths are provided if there is a high number of PCUs of non-motorised vehicles
(defined as more than 300 PCUs per day on average).

Other variations are also sometimes needed, for example, where certain problem soils are encountered
or in areas that are particularly wet and where the road is likely to be inundated and needs to be raised
on a higher embankment. These variations are discussed in Chapter D.2 Section 2.5.2 and Chapter D.6
Section 6.19.

Where spot improvements are made which involve a short length of paved surfacing (eg on a steep
incline) then the width used should be that shown in the respective Tables for paved surfaces.

The width standards for each classification are summarised in Part B, Section 4.3. For some of the cells in
the Tables the values quoted will never be a limitation. For example, in flat terrain there will be no need
to be concerned about the criteria for maximum gradient and paved road sections will be rare at the
lowest traffic levels. Nevertheless, for completeness all the cells have been filled.

4.4 Design speed and geometry


Design speed is defined as effectively the maximum (actually the 85th percentile) safe speed that can be
maintained over a specified section of road when conditions are so favourable that the design features
of the road govern the speed. Design speed is used as an index that essentially defines the geometric
standard of a road, linking many of the factors that determine the road’s service level, namely traffic level;
terrain; pavement type; safety/population density; and road function, to ensure that a driver is presented
with a consistent speed environment.

The concept of design speed is most useful because it allows the key elements of geometric design to be
selected for each standard of road in a consistent and logical way. For example, design speed is relatively
low in mountainous terrain to reflect the necessary reductions in standards required to keep road costs to
manageable proportions. The speed is higher in rolling terrain and highest of all in flat terrain.

In practice the speed of motorised vehicles on many roads in flat and rolling terrain will only be constrained
by the road geometry over relatively short sections but it is important that the level of constraint is
consistent for each road class and set of conditions.

In view of the mixed traffic that occupies the rural roads of Ethiopia and the cost benefit of selecting lower
design speeds, it is prudent to select values of design speed towards the lower end of the internationally
acceptable ranges. The recommended values are shown in Table D.4.6.

Changes in design speed, if required because of a change in terrain, should be made over distances that
enable drivers to change speed gradually. Thus changes should never be more than one design step
at a time and the length of the sections with intermediate standards (if there is more than one change)
should be long enough for drivers to realise there has been a change before another change in the same
direction is encountered (ie considerably more than one single bend). Where this is not possible, warning
signs should be provided to alert drivers to the changes.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 60

Table D.4.6: Design speeds

Design Design speed (km/h)


standard Flat Rolling Mountain Escarpment Urban
DC 4 70 60 50 40 50
DC 3 70 60 50 30 50
DC 2 60 50 40 30 50
DC 1 50 40 30 20 40

4.4.1 Stopping sight distance

In order to ensure that the design speed is safe, the geometric properties of the road must meet certain
minimum or maximum values to ensure that drivers can see far enough ahead to carry out normal
manoeuvres such as overtaking another vehicle or stopping if there is an object in the road.

The distance a vehicle requires to stop safely is called the stopping sight distance. It mainly affects the
shape of the road on the crest of a hill (vertical alignment) but if there are objects near the edge of the
road that restrict a driver’s vision on approaching a bend, then it also affects the horizontal curvature.

The driver must be able to see any obstacle in the road hence the stopping sight distance depends on
the size of the object and the height of the driver’s eye above the road surface. The driver needs time
to react and then the brakes of the vehicle need time to slow the vehicle down, hence stopping sight
distance is extremely dependant on the speed of the vehicle. The surface characteristics of the road also
affect the braking time so the values for unpaved roads differ from those of paved roads, although the
differences are small for design speeds below 60km/h.

The stopping distance also depends on the gradient of the road; it is harder to stop on a downhill
gradient than on a flat road because a component of the weight of the vehicle acts down the gradient in
the opposite direction to the frictional forces that are attempting to stop the vehicle.

Full adherence to the required sight distances is essential for safety reasons. On the inside of horizontal
curves it may be necessary to remove trees, buildings or other obstacles to obtain the necessary sight
distances. If this cannot be done, the alignment must be changed. In rare cases where it is not possible
and a change in design speed is necessary, adequate and permanent signage must be provided.

Recommended stopping sight distances for paved and unpaved roads at different design speeds are
shown in Table D.4.7.

Table D.4.7: Stopping sight distances (m)

Design speed (km/h) 20 30 40 50 60 70 80


Unpaved roads(1) 20 30 50 70 95 125 160
Paved roads(1) 18 30 45 65 85 110 135
Note:
1. In rolling and mountainous terrain these volumes should be increased by 10%.

4.4.2 Stopping sight distance for single lane roads (meeting sight distance)

For single lane roads, adequate sight distances must be provided to allow vehicles travelling in the
opposite direction to see each other and to stop safely if necessary. This distance is normally set at twice
the stopping sight distance (Table D.4.6) for a vehicle that is stopping to avoid a stationary object in the
road. An extra safety margin of 20-30 metres is also sometimes added.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 61

Although a vehicle is a much larger object than is usually considered when calculating stopping distances,
these added safety margins are used partly due to the very severe consequences of a head-on collision;
and partly because it is difficult to judge the speed of an approaching vehicle, which could be considerably
greater than the design speed. However, single lane roads will have a relatively low design speed, hence
meeting sight distances should not be too difficult to achieve.

4.4.3 Intersection sight distance

Intersection sight distance is similar to stopping sight distance except that the object being viewed is
another vehicle that may be entering the road from a side road or crossing the road at an intersection.
The required safe sight distance for trucks in metres is about 3 times the vehicle speed in km/hr. On
straight sections of road many vehicles will exceed the road’s design speed but, being straight, sight
distances should be adequate.

4.4.4 Passing sight distances

Factors affecting the safe sight distances required for overtaking are more complicated because they
involve the capability of a vehicle to accelerate and the length and speed of the vehicle being overtaken.
Assumptions are usually made about the speed differential between the vehicle being overtaken and the
overtaking vehicle but many road authorities have simply based their standards on empirical evidence.

For single lane roads, overtaking manoeuvres are not possible and passing manoeuvres take place only at
the designated passing places. On the lower classes of 2-lane roads, passing sight distances are based
on providing enough distance for a vehicle to safely abort a passing manoeuvre if another vehicle is
approaching. The recommended values are shown in Table D.4.8.

Table D.4.8: Passing sight distances (m)

Design speed (km/h) 30 40 50 60 70 80


Recommended values 75 110 160 205 260 320

4.4.5 Camber and cross-fall

Camber and cross-fall are essential to promote surface drainage. Ponding of water on a road surface
quickly leads to deterioration. There is general agreement that camber or cross-fall should be 3% on
sealed LVRs (2.5% is sometimes advocated but this is insufficient).

Drainage is less efficient on rough surfaces and therefore the camber or cross-fall needs to be higher on
earth and gravel roads. However, if the soil or gravel is susceptible to erosion, high values of camber or
cross-fall can cause erosion problems. Values that are too high can also cause driving problems but, on
the lower standards of rural roads where traffic is low and the road is a single carriageway, vehicles will
generally travel in the middle of the road. Therefore, high levels of camber are not as much of a problem
for drivers as high levels of cross-fall. The design of LVRs makes use of this fact so that higher camber
is used where appropriate. As a result, the optimum value of cross-fall/camber varies considerably but it
normally lies between 4% and 7% with 6% being the usual recommendation in the absence of additional
information concerning the erosion potential of the soil/gravel.

Shoulders having the same surface as the running surface should have the same slope. Unpaved shoulders
on a sealed road should have shoulders that are about 2% steeper, in other words 5% if the running
surface is 3%.

4.4.6 Adverse cross-fall

Adverse cross-fall arises on curves when the cross-fall or camber causes vehicles to lean outwards when
negotiating the curve. This affects the cornering stability of vehicles and is uncomfortable for drivers,
thereby affecting safety. The severity of its effect depends on vehicle speed, the horizontal radius of
curvature of the road and the side friction between tyres and road surface. For reasons of safety it is

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 62

recommended that adverse cross-fall is removed where necessary (see Table D.4.9) on all roads regardless
of traffic.

Table D.4.9: Adverse cross-fall to be removed if radii are less than shown

Design speed Minimum radii (m)


(km/h) Paved Unpaved
<50 500 700
60 700 1000
70 1000 1300
85 1400
100 2000

Some cross-fall is necessary for drainage and hence flat sections are not allowed. Instead, a single value
of cross-fall is designed in the proper direction (ie all camber is removed as shown in Figure D.4.4) such
that the cross sectional shape of the road is straight with the cross slope being the same as that of the
inner side of the cambered two-lane road (usually 3 or 4% for sealed roads). For unpaved roads the
recommended cross-fall should also be the same as the normal camber or cross-fall value of 6%.

To remove adverse cross-fall the basic cambered shape of the road is gradually changed as the road
enters the curve until it becomes simply cross-fall in one direction at the centre of the curve.

For sealed roads the removal of adverse camber may not be sufficient to ensure good vehicle control
when the radius of the horizontal curve becomes too small. In such a situation additional cross-fall may
be required. This is properly referred to as super-elevation but it has become common practice to refer
to all additional elevation as super-elevation and this convention will be used here.

Figure D.4.4: Removal of adverse camber

4.4.7 Super-elevation

Super-elevation on unsealed LVRs is not necessary. This is because it is recommended that adverse cross-
fall or camber is always removed on horizontal curves below 1000m radius. Since the recommended
cross-fall or camber is 6%, the effective ‘super-elevation’ when adverse cross-fall is removed will also be
6% and this therefore determines the minimum radius of horizontal curvature for each design speed in
the same way as for genuine super-elevation. In practice it may not be possible to maintain such a value
of cross-fall during the life of an unsealed road and therefore it is recommended that minimum radii are
based on the lower level of 4% cross-fall.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 63

For sealed roads the removal of adverse cross-fall will result in an effective super-elevation of 3% and
this should be used to determine minimum radii of curvature for such roads. However, if these radii are
difficult to achieve, genuine super-elevation of up to 7% (or, in exceptional circumstances, up to 10%) can
be used with a resulting decrease in horizontal radius of curvature.

The change from normal cross-section on straight sections of road to a super-elevated section should
be made gradually. The length over which super-elevation is developed is known as the super-elevation
development length. Two-thirds of the development length should be provided before the curve begins.
The development depends on design speed as shown in Table D.4.10. Between 50% and 75% of the
super-elevation should be achieved by the tangent point. 66% is usually used.

Table D.4.10: Super-elevation development lengths

Design speed (km/h) Development length (m)

30 25

40 30

50 40

60 55

70 65

80 80

4.5 Horizontal alignment


The horizontal alignment consists of a series of straight sections (tangents) connected to circular curves.
The horizontal curves are designed to ensure that vehicles can negotiate them safely. The alignment
design should be aimed at avoiding sharp changes in curvature, thereby achieving a safe uniform driving
speed. Transition curves between straight sections of road and circular curves whose radius changes
continuously from infinity (tangent) to the radius of the circular curve (R) are used to reduce the abrupt
introduction of centripetal acceleration that occurs on entering the circular curve. They are not required
when the radius of the horizontal curve is large and are normally not used on the lower classes of road. In
Ethiopia their use is confined to roads where the design speed is 80km/hr or greater and therefore they
are not required for LVRs.

In order for a vehicle to move in a circular path an inward radial force is required to provide the necessary
centripetal acceleration or, in other words, to counteract the centrifugal force. This radial force is provided
by the sideways friction between the tyres and the road surface assisted by the cross-fall or super-elevation.

The sideways friction coefficient is considerably less than the longitudinal friction coefficient. Its value
decreases as speed increases but there is considerable disagreement about representative values,
especially at the lower speeds. For paved roads it ranges from between 0.18 and 0.3 at 20km/h down
to between 0.14 and 0.18 at 80km/h. For unpaved roads it can be considerably less. The design speed
is therefore one of the main design parameters. Values for each class of road under each of its operating
conditions have been set as shown in Table D.4.6.

For both sealed and unsealed roads there are also constraints on the maximum cross-fall, as described
in Section 4.4.5 and 4.4.6. These constraints translate directly into minimum values of horizontal radii of
curvature.

The recommended values of horizontal curvature are shown in Table D.4.11 and Table D.4.12. As
indicated in the Tables, the use of a higher value of super-elevation makes it possible to introduce a

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 64

smaller horizontal curve based on the same design speed. This can be used for paved roads but not for
unpaved roads.

Table D.4.11: Recommended minimum horizontal radii of curvature: paved roads (m)

Design speed (km/h) 20 30 40 50 60 70 80


Minimum horizontal radius for SE = 4% 15 25 50 85 135 195 270
Minimum horizontal radius for SE = 7% 15 25 45 75 120 170 235
Minimum horizontal radius for SE = 10% 15 20 40 70 105 150 205

Table D.4.12: Recommended minimum horizontal radii of curvature: unpaved roads (m)

Design speed (km/h) 20 30 40 50 60 70 80


Minimum horizontal radius for SE = 4% 17 35 70 110 175 245 340
Minimum horizontal radius for SE = 6% 15 30 60 100 155 215 300

4.5.1 Curve length

For reasons of safety and ease of driving, curves near the minimum for the design speed should not be
used at the following locations:
ƒ On high fills, because the lack of surrounding features reduces a driver’s perception of the alignment.
ƒ At or near vertical curves (tops and bottoms of hills) because the unexpected bend can be extremely
dangerous, especially at night.
ƒ At the end of long tangents or a series of gentle curves, because actual speeds will exceed design
speeds.
ƒ At or near intersections and approaches to bridges or other water crossing structures.

There are conflicting views about curve lengths. One school of thought maintains that the horizontal
alignment should maximise the length of road where adequate sight distances are provided for safe
overtaking. Overtaking is difficult on curves of any radius and hence the length of curved road should
be minimised. This requires curve radii to be relatively close (but not too close) to the minimum for the
design speed to maximise the length of straight sections. This view is the currently accepted best practice
for roads except in very flat terrain but care should be exercised to ensure the curves are not too tight.

The alternative view is that very long straight sections should be avoided because they are monotonous
and cause headlight dazzle at night. A safer alternative is obtained by a winding alignment with tangents
deflecting 5 to 10 degrees alternately from right to left. Straight sections should have lengths (in metres)
less than 20 x design speed in km/h. Such ‘flowing’ curves restrict the view of drivers on the inside
carriageway and reduce safe overtaking opportunities, therefore such a winding alignment should only
be adopted where the straight sections are very long. In practice this only occurs in very flat terrain. The
main aspect is to ensure that there are sufficient opportunities for safe overtaking and therefore, provided
the straight sections are long enough, a semi-flowing alignment can be adopted at the same time. If
overtaking opportunities are infrequent, maximising the length of the straight sections is the best option.

For small changes of direction it is often desirable to use a large radius of curvature. This improves the
appearance and reduces the tendency for drivers to cut corners. In addition, it reduces the length of the
road segment and therefore the cost of the road provided that no extra cut or fill is required.

4.5.2 Curve widening

Widening of the carriageway where the horizontal curve is tight is usually necessary to ensure that the rear
wheels of the largest vehicles remain on the road when negotiating the curve; and, on two lane roads,
to ensure that the front overhang of the vehicle does not encroach on the opposite lane. Widening is

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 65

therefore also important for safety reasons. Any curve widening that is considered should only be applied
on the inside of the curve.

Vehicles need to remain centred in their lane to reduce the likelihood of colliding with an oncoming
vehicle or driving on the shoulder. Sight distances should be maintained as discussed above. The levels
of widening shown in Table D.4.13 are recommended except for roads carrying the lowest levels of traffic
(DC1). Widening should be applied on the inside of the curve and introduced gradually.

Widening on high embankments is often recommended for the higher classes of road. The steep drops
from high embankments unnerve some drivers and the widening is primarily for psychological comfort
although it also has a positive effect on safety. Such widening is not recommended for LVRs.

Table D.4.13: Widening recommendations (m)

Single lane roads Two lane roads


151- 301-
Curve radius 20 30 40 60 <50 51-150
300 400
Increase in width 1.5(1) 1.0 0.75 0.5 1.5 1.0 0.75 0.5
Notes:
1. See Section 3.6.4 dealing with hairpin stacks

4.6 Vertical alignment


The two major elements of vertical alignment are the gradient, which is related to vehicle performance
and level of service; and the vertical curvature, which is governed by safe sight distances and comfort
criteria.

The vertical alignment of a road seems more complicated than the horizontal alignment but this is simply
because of difficulties in presentation due to the inclusion of the algebraic difference in gradient (G %)
between the uphill and downhill sides. In addition, the equation of the vertical curve is a parabola rather
than a circle.

The required sight distance for safety is the basic stopping sight distance.

4.6.1 Crest curves

The minimum length of the curve (L metres) over the crest of the hill between the points of maximum
gradient on either side is related to G and to the stopping-sight distance; and therefore to the design
speed. Note that although drivers would like to overtake on hills, the required sight distance for safe
passing on crests is much too large to be economical on LVRs.

The minimum value of the L/G ratio can be tabulated against the stopping sight distance (Table D.4.6),
and therefore the design speed, to provide the designer with a value of L for any specific value of G. The
international comparisons give the values shown in Table D.4.14.

Table D.4.14: Minimum values of L/G for crest curves

Design speed (km/h) 30 40 50 60 70 80


Sealed roads 2 4 7 12 21 37
Unsealed roads 3 6 11 19 34 58

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 66

4.6.2 Sag curves

Sag curves are the opposite of crest curves in that vehicles first travel downhill and then uphill. In daylight
the sight distance is normally adequate for safety and the design criterion is based on minimising the
discomforting forces that act upon the driver and passengers when the direction of travel changes from
downhill to uphill. On rural roads such considerations are somewhat less important than road safety
issues. However, at night time the problem on sag curves is the illumination provided by headlights to
see far enough ahead. This depends on the height of the headlights above the road and the angle of
divergence of the headlight beams.

To provide road curvature that allows the driver to see sufficiently far ahead using headlights while driving
at the design speed at night is usually too expensive for LVRs. In any case, the driving speed should
be much lower at night on such roads. As a result of these considerations it is recommended that the
minimum length of curve is determined by the driver discomfort criterion. The results are shown in Table
D.4.15.

Table D.4.15: Minimum values of L/G for sag curves

Design speed (km/h) 30 40 50 60 70 80


Minimum L/G 0.7 1.3 2.2 3.5 4.8 7.5

In practice a minimum length of curve of 75m will cope with almost all situations on LVRs. For example,
on a steep down-hill of 10% followed by an up-hill of the same slope, the required minimum curve length
at a speed of 50km/h is 2.2 x (10 + 10) = 44m and 3.5 x (10+10) = 70m at 60km/h.

4.6.3 Gradient

For four-wheel drive vehicles, it is reported that the maximum traversable gradient is about 18%. Two-
wheel drive trucks can cope with gradients of 15%, except when heavily laden. Bearing in mind the
likelihood of heavily laden small trucks, international rural road standards have a general recommended
limit of 12%, but with an increase to 15% for short sections (< 250m) in areas of difficult terrain. Slightly
higher standards are recommended for DC4 with a preferred maximum of 10% and an absolute maximum
of 12% on escarpments where relief gradients of less than 6% are required for a distance of 250m following
a gradient of 12%.

For driving consistency, and hence safety, in terrains other than mountainous terrains and escarpments,
limiting values of gradient are also often specified. In flat terrain a maximum gradient of 7% is appropriate
for LVRs. In rolling terrain a maximum of 10% is appropriate.

Regional experience indicates that unsealed road sections in excess of 6% gradient are often unsustainable
in the medium to long-term. It is expected that the use of alternative surfacings will become more
common in Ethiopia to provide a more sustainable solution in critical areas. Therefore criteria need to be
developed to identify the critical areas where alternative surfacing are to be recommended. This is dealt
with in (Chapter D.7).

4.6.4 Hairpin stacks

Climbing sections on mountain and escarpment roads are often best designed using hairpin stacks. The
advantages are that the most favourable site for ascending the escarpment can be selected and a more
direct and therefore shorter route will often be possible. However there are several problems.

The limited space to construct cut and fill slopes necessitates either a reduction in geometric standards
or more expensive retaining structures. For LVRs the former solution should be adopted.

Lack of suitable sites for disposal of spoil and access difficulties for plant can pose difficulties during
construction.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 67

If there are problems of instability they may extend from one loop to another and so the advantage of
attempting to choose the most stable section of the escarpment is lost. This is a problem and is dealt
with in Chapter D.3.

Storm run-off will, of necessity, become very concentrated so, although the number of drainage structures
and erosion controls may be reduced, their capacity will need to be increased. The risk associated with
failure of the drainage is therefore correspondingly high and minimising this risk adds to the costs. If the
topography allows, some of the problems of stacked hairpins can be reduced by creating several stacks
that are offset from each other and staggered across the slope (ie not immediately above or below each
other). This will reduce drainage problems and limit the danger of instability to fewer hairpin loops.

The key aspect of their geometric design is that the curves should be as flat as possible and the tangents
should be used to achieve the ascent. This is because vehicle traction is much more efficient when the
vehicle is travelling in a straight line. The maximum gradient through the hairpin curve itself should be
4% for DC4 and DC3 and up to 6% for DC2 and DC1.

Considerable curve widening will be required where the curve radius is small to ensure that large vehicles
can negotiate the bends. Widening is also required for safety reason and, if space allows, to provide a
refuge area if a vehicle breaks down.

For LVRs it is recommended that the curves should be designed to allow the passage of the DV4 design
vehicle (Table D.4.1). This means that the curve radius at the centre line of the road should be an absolute
minimum of 13m and the road should be at least 8m wide.

4.7 Harmonisation of horizontal and vertical alignment

4.7.1 Situations to avoid

When designing the horizontal alignment of a road, the designer must ensure that the other elements
of the design are complementary to each other. It is therefore important to note that there are a number
of design situations that could produce unsatisfactory combinations of elements despite the fact that
the design standards have been followed for the particular class of road in question. These are designs
that could provide surprises for drivers by presenting them with unfamiliar conditions. They are therefore
comparatively unsafe.

Avoiding such designs is more important for the higher classes of road because design speeds are higher,
traffic is much greater and, consequently, any accidents resulting from poor design are likely to be more
severe and more frequent. However, in many cases, avoidance of such designs does not necessarily
impose a significant cost penalty and therefore the principles outlined below should be applied to roads
of all classes.

Multiple curves
In the more hilly and mountainous terrains, horizontal curves are required more frequently and have
small radii because the design speeds are low. The tangent sections become shorter and a stage can be
reached where successive curves can no longer be dealt with in isolation. There are three situations that
should be avoided if possible.

Reverse curves
A curve is followed immediately by a curve in the opposite direction. In this situation it is difficult for
the driver to keep the vehicle in its proper lane. It is also difficult for the designer to accommodate the
required super-elevation within the space available.

Broken back curves


This is the term used to describe two curves in the same direction connected by a short tangent. Drivers
do not usually anticipate that they will encounter two successive curves close to each other in the same
direction. There can also be problems fitting in the correct super-elevation in the space available.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 68

Compound curves
Compound curves occur when one curve connects to another of different radius. These can be useful in
fitting the road to the terrain but in some circumstances they can be dangerous. Drivers do not usually
expect to be confronted by a change in radius, and therefore in design speed, hence if, the change is too
great, some drivers are likely to be travelling too fast when entering the tighter part of the compound
curve from the larger one. Compound curves should be avoided where curves are sharp and where the
difference in radii is large. Thus, in any compound curve the smaller radius should not be less than 67%
of the larger one.

Isolated and long curves


An isolated curve close to the minimum radius connected by long straight sections is inherently unsafe.
Irrespective of the design speed, actual speeds on long straight sections will be relatively high and
therefore a curve of minimum radius will require a significant reduction in speed for most vehicles. It
is good practice to avoid the use of minimum standards in such situations. An added bonus is that,
provided no extra cutting of filling is required, the use of a larger radius of curvature results in a shorter
and less expensive road. Curve widening can help to alleviate this problem if a higher radius curve cannot
be used.

The same argument is true, but to a much lesser extent, for any small radius curve that is very long (i.e.
the road is turning through a large angle). Drivers can negotiate a short curve relatively safely at speeds
in excess of the design speed but they cannot do so if the curve is long hence a large radius should be
used in such situations.

4.7.2 Balance

It can be seen that there are several competing factors in providing the optimum horizontal alignment.
Small radii curves maximise the length of straight sections and optimise overtaking opportunities. This
should be the controlling factor where the terrain is such that overtaking opportunities are infrequent and
actual speeds are close to the design speeds. However, in more gentle terrain where overtaking is less of
a problem and vehicles generally travel at speeds higher than the design speed, the use of larger radius
curves is preferred for the reasons outlined previously.

In summary, engineering choice plays a part in the final design which is essentially a balance between
competing requirements.

4.7.3 Phasing

The horizontal and vertical alignment should not be designed independently. Hazards can be concealed
by inappropriate combinations of horizontal and vertical curves and therefore such combinations can be
very dangerous. Some examples of poor phasing are as follows:
ƒ A sharp horizontal curve following a pronounced crest curve. The solutions are to;
• Separate the curves;
• Use a more gentle horizontal curve;
• Begin the horizontal curve well before the summit of the crest curve.
ƒ Both ends of the vertical curve lie on the horizontal curve. If both ends of a crest curve lie on a
sharp horizontal curve the radius of the horizontal curve may appear to the driver to decrease
abruptly over the length of the crest curve. If the vertical curve is a sag curve the radius of the
horizontal curve will appear to decrease. The solution is to make both ends of each curve coincide
or to separate them completely.
ƒ A vertical curve overlaps both ends of a sharp horizontal curve. This creates a hazard because a
vehicle has to turn sharply while sight distance is reduced on the vertical curve. The solution is to
make both ends of each curve coincide or to separate them completely.

4.7.4 Junctions and Intersections

The result of an accident is likely to be that one or more vehicles will leave the road. Hence, where
possible, a safe ‘run-off’ environment should be created and good sight distances provided. Intersections

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 69

should therefore not be located on high embankments; near to bridges or other high level water crossings;
on small radius curves; or on super-elevated curves. To ensure good visibility, vegetation should be
permanently cleared from the area surrounding the junction.

It is also advisable to avoid building intersections on gradients of more than 3% or at the bottom of sag
curves. This is because:
ƒ Stopping sight distances are greater on downhill descents and drivers of heavy vehicles have more
difficulty in judging them; and
ƒ It is advantageous if heavy vehicles are able to accelerate as quickly as possible away from the
junction.

The ideal angle that intersecting roads should meet is 90o because this provides maximum visibility in
both directions but visibility is not seriously compromised as long as the angle exceeds 70o.

Where two roads have to cross each other, a simple X-cross junction is adequate for LVRs. However,
where possible, it is preferable to provide two staggered T-junctions as illustrated in Figure D.4.5 rather
than one X-cross junction since there is unlikely to be a cost penalty in doing so. The most heavily
trafficked road is retained as a direct through route. The minor road is then split so that traffic has to enter
the major road by making a left turn across the traffic stream onto the major road and then a right turn to
re-enter the minor road. This method halves the number of possible manoeuvres where the traffic from
the minor road has to cross the traffic stream on the major road. The entry points of the two arms of the
minor road should be spaced about 100m apart.

Figure D.4.5: Preferred intersection design

4.8 Safety
The road accident statistics in Ethiopia, in common with many other countries in Africa, show that death
rates from road accidents are 30 to 50 times higher than in the countries of Western Europe. The numbers
of serious injuries resulting from road accidents are equally alarming. Economic analysis has shown
conclusively that this high level of road accidents has economic consequences for the country that is
equivalent to a reduction of 2-3% of GDP. This is a very significant drain on the economy. Furthermore,
the consequences of the road accidents impose a great deal of grief and anguish on a considerable
proportion of the population. Every effort should therefore be made to reduce the number of serious
accidents.

The geometric design of the roads has an important part to play in this endeavour and road safety
aspects have been highlighted throughout this manual. Road and shoulder widths have been increased

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 70

to accommodate pedestrians, NMTs, and intermediate forms of transport (IMTs); moderate design
speeds have been used for elements of road alignment; parking places and lay-byes for buses have been
included in populated areas; account has been taken of reduced friction on unpaved roads; adequate
sight distances have been provided; and much more (see Figure D.4.2 for example).

However there are a number of other steps that could be taken to improve safety. These include:
ƒ Traffic calming measures to reduce speeds in populated area;
ƒ Road markings, signage and lighting;
ƒ Segregating pedestrians and motorised vehicles in populated areas;
ƒ Providing crash barriers at dangerous locations;
ƒ Providing a professional safety audit at the design stage.

4.8.1 Traffic calming

The seriousness of road accidents increases dramatically with speed and hence very significant
improvements to road safety are possible if traffic can be slowed down. This process is called traffic
calming. All such methods have their advantages and disadvantages and the effectiveness of the methods
also depends on aspects of driver behaviour that can vary considerably from country to country. Therefore
research needs to be carried out in Ethiopia to identify the most cost effective approaches.

The effect of any traffic calming measure on all the road users should be carefully considered before
they are installed. Some are unsuitable if large buses are part of the traffic stream; some are very harsh
on bicycles, motorcycles and motor cycle taxis; and some are totally unsuitable when there is any animal
drawn transport.

The three most common methods are:


ƒ Chicanes;
ƒ Rumble strips; and
ƒ Speed reduction humps.

Chicanes
These are designed to produce artificial congestion by reducing the width of the road to one lane for a
very short distance (3-5m) at intervals (typically 300m) along it. They are usually built on alternate sides of
the road. They cause drivers to slow down provided that the traffic level is high enough to make it very
probable that they will meet an oncoming vehicle. The method is obviously unacceptable if traffic flow is
high because the congestion that is causes will be severe. For safety, they must be illuminated at night.

Rumble strips
These are essentially a form of artificial road texture that causes considerable tyre noise and vehicle
vibrations if the vehicle is travelling too fast. They are used in two ways. The first is to delineate areas
where vehicles should not be. They are effectively a line running parallel to the normal traffic flow so that
if a vehicle inadvertently strays onto or across the line the driver will receive adequate warning. Secondly
they are used across the road where they are placed in relatively narrow widths of 2 to 4m but at intervals
along the road of typically 50 to 200 metres. They are uncomfortable to drive across at speed hence they
are usually effective in slowing down the traffic. They do not need to be illuminated at night.

Speed reduction humps and cushions


These are probably the most familiar measures used to slow traffic. They are essentially bumps in the road
extending uniformly from one side to the other. Unlike rumble strips, speed reduction humps are quite
high and, if they are designed badly, they can cause considerable vehicle damage. They are often used in
villages where they are placed at intervals of between 50m and 200m. They are very effective but usually
unpopular with drivers.

The shape of the hump is important to reduce the severity of the shock when a vehicle drives over it.
Ideally they should cause driver discomfort but not vehicle damage. The height of the bump is usually 50
or 75mm but the width should be at least 1.5m (2.0m is better) and the change in slope from the roadway

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 71

onto the hump should be gradual. The top of the hump can be rounded or flat. Pipes that are almost
buried are completely unsuitable.

Based on a similar principle to the speed hump, speed reducing cushions are more versatile. They are
essentially very similar to the speed hump but the hump is not continuous across the road. The width of
a two lane road is usually covered by two or three cushions with considerable gaps between them. The
idea is that large vehicles will not be able to pass without at least one wheel running over one of the
humps but bicycles and motorcycles can pass between them without interference. If suitably designed,
the wheels of animal drawn carts could also avoid the humps.

4.8.2 Road markings, signage and lighting

Theft of signs is a problem in some areas and therefore painting signs on permanent features such as
buildings, rocks and trees should be considered where necessary.

The extent to which road markings, signs and other road furniture is required depends on the traffic
volume, the type of road, and the degree of traffic control required for safe and efficient operation. For
low volume roads the primary purpose is to improve road safety hence not all of the features of road
furniture and signage described in the ERA’s standards (Road Furniture and Markings, Geometric Design
Manual-2011) will be used on such roads.

The main elements are:


ƒ Traffic signs provide essential information to drivers for their safe and efficient manoeuvring on the
road;
ƒ Road markings to delineate the pavement centre line and edges to clarify the paths that vehicles
should follow;
ƒ Marker posts to indicate the alignment of the road ahead and, when equipped with reflectors,
provide optical guidance at night;
ƒ Lighting to improve the safety of a road at night time.

4.9 Traffic Signs


Traffic signs are of three general types:
ƒ Regulatory Signs: indicate legal requirements of traffic movement and are essential for all roads;
ƒ Warning Signs: indicate conditions that may be hazardous to highway users;
ƒ Information Signs: convey information of use to the driver.

4.9.1 Warning signs

The physical layout of the road must sometimes be supplemented by effective traffic signing to inform
and to warn drivers of any unexpected changes in the driving conditions. Some of the common situations
are mentioned below but each situation is unique and the severity of any particular situation can vary
considerably. It is therefore recommended that the judgement of an experienced road safety expert is
obtained at the road design stage.

For an existing road that is to be upgraded, the hazardous locations should be identified at an early stage
and, ideally, should be corrected in the new design. If this is not possible, then suitable road signs should
be installed.

The most common situation occurs when the geometric standards for a particular class of road have
been changed along a short section of road. This is usually caused by a constraint of some kind that
has prevented the standard from being applied continuously and therefore causes an unexpected and
potentially dangerous situation. Examples are a sharp bend, a sudden narrowing of the road, or an
unexpectedly steep gradient.

A similar situation arises in easy terrain where, despite the fact that the geometric standard of the road
has been applied, a hazard such as a bend occurs after a long section of road where drivers are easily able
to exceed the design speed of the road by a considerable margin.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 72

As well as changes in the geometric standard of the road, many other relatively unexpected hazards can
occur and also need to be signed. For example an unexpected school crossing, a ford or other structure
that is not clearly visible from a safe distance – there are many examples too numerous to list. Once again,
engineering judgement is required.

A common situation occurs in populated areas where traffic calming measures have been introduced.
Speed humps are a particular problem because they are often not sufficiently visible from a reasonable
distance, and sometimes they have been badly designed and provide more of a jolt to the vehicle than
intended. It is therefore good practice to provide warning signs for these, especially on roads that are
likely to be used by traffic unfamiliar to the area. This will include classes DC3 and DC4 and many DC2
roads.

An important consideration on unpaved roads is that the road markings that are used on paved roads
to improve safety cannot be used on unpaved roads. This means that if drivers need to be warned of a
hazard that is traditionally done by means of road markings, on unpaved roads this will have to be done
by means of traffic signs.

4.9.2 Information signs

Information signs are less vital on the lower classes of road frequented primarily by local people. However,
for road classes DC3 and DC4 on which a considerable proportion of drivers will not be local, information
signs are desirable. They obviate the need for drivers to stop in populated areas to ask questions of
pedestrians and hence improve safety, but in most cases this effect is very marginal, especially if the road
standards that should be provided in populated areas have been applied. Hence the convenience of
some information signs is part of the provision of a particular level of service to the traveller.

4.10 Road Markings


Road markings either supplement traffic signs and marker posts or serve independently to indicate
certain regulations or hazardous conditions. There are three general types of road markings in use namely
pavement markings, object markings and road studs.

4.10.1 Pavement markings

Pavement markings consist primarily of centre lines, lane lines, no overtaking lines and edge lines. Not
all of these are possible, or justified, on low volume roads. However, on a paved, two lane a centre line
is desirable. Such a road is not likely to have been built unless the traffic justifies it and hence, for safety
reasons, a centre line is recommended.

Other pavement markings such as ‘stop’, pedestrian crossings and various word and symbol markings
may supplement pavement line markings. However, it is obvious that such markings can only be applied
to paved roads and then not to all surfacings. In cases where a warning is deemed necessary for safety
reasons but road markings cannot be used, road signs must be used instead if applicable.

4.10.2 Object markers

Physical obstructions in or near the carriageway should be removed in order to provide the appropriate
clear zone. Where removal is impractical, such objects should be adequately marked by painting or by
use of other high-visibility material.

4.10.3 Road studs

Road studs are used on more heavily trafficked roads and in urban areas. They are unlikely to be used
on low volume roads but if necessary advice can be found in Road Furniture and Markings, Geometric
Design Manual-2011.

4.10.4 Marker Posts

There are two types of marker posts in use namely guideposts and kilometre posts.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 4 - 73

Guideposts are intended to make drivers aware of potential hazards such as abrupt changes in shoulder
width and alignment, or approaches to structures for example. They are unlikely to be used on a low
volume road.

Kilometre posts are a requirement for all trunk and link roads and are therefore only likely to be needed
on some roads of class DC4. Details are given in Road Furniture and Markings, Geometric Design Manual
- 2011.

4.11 Lighting
Lighting of low volume rural highways is seldom justified except at intersections, railway level crossings,
narrow or long bridges, tunnels, sharp curves, and areas where there is activity adjacent to the road (eg
markets). Details are provided in Road Furniture and Markings, Geometric Design Manual-2011.

4.12 Safety barriers


Safety barriers are expensive and seldom justified on low volume roads. The geometric design of such
roads should be done to eliminate the need for such barriers but sometimes they might be required in
highly dangerous situations, for example, on some bends on an escarpment road that cannot be made
safe by other means. Expert advice should be sought.

4.12.1 Segregating vulnerable road users

Where possible, non-motorised vehicles and pedestrians should be physically segregated from the
motorised vehicles. While this is not specifically part of the geometric design of the road itself, if the
terrain and local conditions are suitable for the construction of parallel pathways wide enough for
NMTs, then some of the geometric features of the roadway designed to accommodate this traffic will
not be necessary and hence considerable savings may be possible. However if traffic does travel on
such pathways, sufficient connections need to be made to the roadway itself to enable access in either
direction.

4.12.2 Crash barriers

Crash barriers are designed to physically prevent vehicles from crossing them. They are an essential
feature of high speed roads to prevent vehicles travelling in opposite directions from colliding with each
other head-on, but their primary use on LVRs is simply to prevent vehicles from leaving the road at
dangerous places such as when the road comes to its end or where a vehicle could plunge over an edge
such as an escarpment. Good geometric design should prevent drivers from experiencing unexpected
situations where they might be in danger of losing control but sometimes crash barriers are required,
particularly at dangerous points on escarpments. However, they are expensive to install and they must
be installed properly otherwise they are not likely to be fit for purpose. They are rarely used on LVRs but
could justifiably be used on DC4 standard roads in some circumstances.

4.12.3 Safety audits

The subject of road safety is remarkably complex in that, although many unsafe practices are glaringly
obvious, there are many situations where it is difficult to identify what is likely to be unsafe, especially if
the project is a new road and one is working from drawings. The history of road safety is full of ideas that
were thought to improve road safety but often had no discernable effect or even made things worse.
The problem has always been lack of reliable data; there is no substitute for a systematic method of
recording the characteristics of road accidents and analysing the data when there is sufficient for reliable
conclusions to be drawn.

Professional road safety auditing is the next best thing and is regularly undertaken on every road project
in some countries in an attempt to improve the safety design from the very beginning. It is anticipated
that this practice will become increasingly common in Ethiopia, especially for road projects located in
populated areas.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 4 - 74

4.13 Using the standards


There are three design situations namely:
ƒ Upgrading from a lower class of road to a higher class;
ƒ Designing a road to replace an existing track; and
ƒ Designing a completely new road where nothing existed before.

4.13.1 Upgrading an existing road

The basic alignments will already exist but the standards of the existing road should be those applicable
to a road of lower class. The new road will require higher standards which may involve a wider cross
section, higher design speeds and therefore larger horizontal and vertical radii of curvature. In flat and
rolling terrain, larger horizontal radii of curvature are usually achieved by means of minor realignments at
the curves themselves. Larger vertical radii of curvature are usually more difficult but, depending on the
terrain, can often be achieved by additional fill rather than deeper cutting. In more severe mountainous
terrain it may be necessary to make substantial realignments to avoid deep cuts, for example, following
a contour more closely to avoid a steep hill with inadequate sight distances over a crest.

The most difficult aspect is likely to occur in mountainous terrain when substantial widening is required.
Under these circumstances it may not always be possible to meet the standards of the new road class
and therefore adequate warning signs will need to be employed to alert drivers to the lower standards.

In general, however, the main improvements, apart from overall widening, are essentially spot
improvements and do not require sophisticated design methods.

4.13.2 Designing a road to replace an existing track

In this case the existing geometric standards will be very much lower than those required hence some
substantial re-alignments may be necessary, especially in hilly and mountainous terrain. However,
the basic route selection has been carried out by virtue of the fact that there is an existing track and
the main control points along the alignment will already be defined. Although re-alignments may be
substantial, an experienced engineer could adopt a design-by-eye approach in many cases. However,
it is anticipated that, in general, the designs will be done with the help of computer programs based on
accurate topographical and other survey data (see Chapter 2 on route selection).

4.13.3 Designing a new road

Designing a geometric alignment for an entirely new road where nothing existed before is a considerably
more complex process because of the many different route alignments that are possible and the relative
lack of information available at the beginning of the process. In many cases there will need to be a pre-
feasibility study to identify possible corridors for the road and to decide whether the project is likely to
be viable. This will then need to be followed by a feasibility study to determine the best routes within the
best corridors and, finally, a detailed design study based on the route selected. The level of detail in this
process depends critically on the class of road being designed and the terrain through which it will pass.
Errors at this stage can be costly and, once the road is built, can also impose serious burdens in the future
if the road requires excessive maintenance.

The principles of route selection are described in detail in Chapter 2. They are based on surveys of
various kinds that provide information about all the likely technical engineering issues related to the new
road but also surveys concerned with environmental and social issues as well.
The final design will inevitably be a compromise between many competing factors and there is no formal
way of resolving all of them to everyone’s satisfaction. Engineering judgement and consensus will be
required to arrive at a satisfactory alignment.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 5 - 75

5. DRAINAGE

5.1 Introduction
Road drainage design is the general term that is applied to two separate topics namely:
ƒ Internal road drainage.
The process of minimising the quantity of water that remains within a road pavement by maximising
the ability of the road to lose water to an external drainage system. Sometimes this definition also
includes minimising the quantity of water that gets into a road pavement in the first place.
ƒ External drainage.
This consists of three components:
a. The process of determining the quantity of water that falls upon the road itself and its
associated works that needs to be channelled away from the road by the drainage system.
This is water that falls upon the road as rain.
b. The process of determining the quantity of water that flows in the streams, rivers and natural
drains that the road has to cross. This is water that falls as rainfall at locations away from the
road.
c. Design of the individual engineering features of the drainage system to accommodate the
flow of water.

This Chapter is concerned with the external drainage system and the drainage standards for roads
carrying less than 300 two-axled (and larger) motorised vehicles per day. The Chapter is essentially
a guide containing appropriate technical explanations of all the steps in designing the surface water
drainage system for LVRs.

Internal drainage is considered in Chapter D.6, Section 6.18 as part of the chapter on pavement design.
This Chapter does not deal with route surveying, site investigations, route selection or the actual structural
design of bridges and major water crossings; these topics are dealt with in other sections of this manual
or in the ERA 2011 manuals. The planning and structural design of river crossings of less than 10m span
and drainage structures for roads being considered in this Manual is given in Part E.

Neither rainfall nor rivers distinguish between roads carrying low and high volumes of traffic. Therefore,
the basic costs of protecting a road from the effects of water are essentially the same and largely
independent of traffic. Hence, for LVRs the cost of the drainage system can comprise a larger proportion
of the costs of the road.

There are, of course, different levels of protection associated with the risk of serious damage to the
road. For principal trunk roads little risk can be tolerated and so expensive drainage measures must be
employed. For LVRs the consequences of failure in the drainage system are correspondingly lower but,
within the range covered by LVRs, there are some significant differences depending on the length of the
road and the availability of an alternative route.

The challenge for the engineer is to choose a level of protection that is commensurate with the class
of road and the consequences of drainage failure. Thus a certain amount of engineering judgement is
required and a design manual such as this requires a consensus amongst road professionals.

Unfortunately, although it is possible to define the probability of specific storm events from extensive
rainfall records, if such records are available, it is practically impossible to define the overall level of risk
inherent in a drainage system design itself. This is because there are so many other factors that influence
its performance. First of all, simply calculating the water volumes flowing in the drainage system following
a specific storm involves several important assumptions.

Secondly, the drainage system is not a fixed, unchanging system despite every effort by the designer to
protect it and to make it so. Changes are always occurring as a result of aspects such as sedimentation,
erosion, the transport of debris, growth of vegetation and landslides. For example, sedimentation will

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 5 - 76

always occur in some places within the drainage system. This affects water flow and drainage capacities in
complex ways. Partial blockage by debris or landslides, a particularly important problem in mountainous
areas, can quickly lead to full blockage and catastrophic failures unless cleared by maintenance activities.

Erosion is also a formidable enemy of the drainage designer. Very erodible soils can be found extensively
in many parts of Ethiopia and catastrophic levels of erosion can arise from small perturbations in the
smooth flow of water leading to failure of the drainage system.

Naturally the designer attempts to minimise these affects but the effectiveness in doing so is directly
related to the cost and to the effectiveness of maintenance (also a function of cost). Hence different levels
of risk, and therefore cost, are applied to roads of different standard. This is the designers challenge
because such levels of risk are numerically very difficult to define.

5.2 Summary of standards and departures from standards

5.2.1 Design standards and storm return period

Once the drainage design has been completed, and provided maintenance is carried out to remove
potential blockages and repair minor damage, a road drainage system should operate successfully for
many years. However, drainage systems cannot be designed for the very worst conditions that might
occur on extremely rare occasions because it is too expensive to do so. The various standards for the
design of drainage are based on different levels of risk that are attached to the likely occurrence of the
different storm intensities for which they are designed, assuming that appropriate routine maintenance
is carried out.

Storm events are defined by the intensity and duration of rainfall and are extremely variable in nature over
periods of many years. Thus a statistical distribution of storm severities shows that very severe storms are
quite rare and that less severe storms are more common. The risk of a severe storm occurring is defined
by the statistical concept of its likely return period which is directly related to the probability of such a
storm occurring in any one year. Thus a very severe storm may be expected, say, once every 50 years but
a less severe storm may be expected every 10 years.

This does not mean that such storms will occur on such a regular basis. A severe storm expected once
every 50 years has, on average, a probability of occurring in any year of 1 in 50 (or 0.02 or 2%). Similarly
a storm of lower intensity that is expected to occur, on average, once every 10 years has a probability of
occurring in any one year of 1 in 10 (or 0.1 or 10%). The operative words here are “on average” and it
is salutary to realise that there is always a finite probability that the worst storm for 200 years may occur
tomorrow.

Most drainage structures are likely to be severely damaged if their capacity is exceeded for any length
of time hence their capacity is the most important aspect of their design. In general, the more severe the
storm for which the structure is designed, the more expensive it is to build; and the cost of designing for
the highest possible storm severity (ie zero risk) is prohibitive. Drainage standards are therefore defined
by the level of risk. This is done using the concept of return period of the maximum storm for which they
are designed.

There are three factors that determine the level of risk that is appropriate for each structure namely;
ƒ The standard of the road (ie the traffic level);
ƒ The cost of the drainage structure itself;
ƒ The severity of the consequences should the road become impassable because of a failure of the
drainage system.

If a drainage structure on a road carrying high levels of traffic is damaged or fails completely, the disruption
and associated costs to the traffic can be very high and therefore the structures on such a road are
designed for low risk (ie for storms of long return periods). They are therefore relatively expensive. On the
other hand, if a drainage structure should fail on a road carrying low levels of traffic, the likely disruption
to traffic and the associated costs are correspondingly less and hence the higher cost of designing the

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 77

drainage for low risk cannot be justified. The drainage is therefore designed for shorter storm return
periods.

Similarly, the cost of replacement or repair of large and expensive drainage structures is high and therefore
they are designed to minimise this risk by designing for very severe storms (ie storms with long return
periods). This increases their cost but reduces the risk of damage. Higher risks can be tolerated for smaller
and less expensive structures that are usually easier to repair; hence these are designed for less severe
storms (ie shorter return periods).

An overriding principle for the designer is to consider the consequences of a drainage failure. In situations
where the road is relatively short and an alternative route, albeit a longer one, is available, the social and
economic consequences of a drainage failure that makes the road impassable for any length of time are
not high. In contrast, there are also many situations in Ethiopia where there is no alternative route at all or,
if there is one, it is very long. Under these circumstances additional expenditure to reduce the risk of such
an occurrence is justified. This is done by designing for a larger storm (ie a longer storm return period).

It is difficult to calculate the exact trade-off between the cost of designing for low risk and the costs and
consequences of failure of a drainage structure. Furthermore, the precision with which design storms can
be calculated depends on the availability of detailed rainfall data that are required to have been collected
over a period of many years. Even with good rainfall data, there are other uncertain assumptions that
need to be made in carrying out the calculations. Thus, in most situations the accuracy of the calculations
of the required water flow capacity is not very high despite the apparent sophistication that is apparent in
some methods of drainage design and it is therefore prudent to include a factor of safety.

Because of these issues the drainage standards can only be based on a review of practices throughout
the world combined with local engineering judgement and consensus. Table D.5.1 indicates the design
standards for LVRs in Ethiopia. For strategic routes, routes of very high economic or social importance or
if the alternative route in the event of a drainage failure is more than an additional 75km or if there is no
alternative route suitable for vehicles, Table D.5.2 should be used instead.

Table D.5.1: Design storm return period (years)

Geometric design standard


Structure type
DC4 DC3 DC2 DC1
Gutters and inlets 2 2 2 1
Side ditches 10 5 5 2
Ford 10 5 5 2
Drift 10 5 5 2
Culvert diameter <2m 15 10 10 5
Large culvert diameter >2m 25 15 10 5
Gabion abutment bridge 25 20 15 -
Short span bridge (<10m) 25 25 15 -
Masonry arch bridge 50 25 25
Medium span bridge (15 – 50m) 50 50 25 -
Long span bridge >50m 100 100 50 -

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 78

Table D.5.2: Design storm return period (years) for severe risk situations

Geometric design standard


Structure type
DC4 DC3 DC2 DC1
Gutters and inlets 5 5 5 2
Side ditches 15 10 10 5
Ford 15 10 10 5
Drift 15 10 10 5
Culvert diameter <2m 25 20 20 10
Large culvert diameter >2m 50 25 20 10
Gabion abutment bridge 50 25 20 -
Short span bridge (<10m) 50 50 25 -
Masonry arch bridge 50 50 25
Medium span bridge (10 – 50 m) 100 100 50 -
Long span bridge >50m 100 100 100 -

5.2.2 Methods of design

An international review of LVRs indicated a wide range of practices. The simplest methods are usually
described in manuals specifically written for LVRs. These all use relatively short storm return periods to
keep costs low. The simplest method of all is essentially a stage construction process whereby simple
rules of thumb are used for the initial design with little or no calculation. The road is built and then, in the
following year or two, problems that arise where there is inadequate capacity in the drainage system are
rectified as quickly as possible. Such an approach is normally used only for very low volume roads such as
DC1 and may be applicable if the engineering resources are readily available during the required period
following initial construction.

Very few national standards specifically address the problem of designing drainage for LVRs. The
implication is that the methods used for all roads should be applied to LVRs but this is impracticable. The
methods described in this manual range from the simplest approach appropriate to the lowest standards
up to more comprehensive methods that could be used whenever sufficient data are available. The
manual does not include the full range of methods suitable for the higher road classes.

5.2.3 Departures from standards

It is fundamental to the concept of setting standards that they should be applied at all times. However,
the basic standards for drainage structures and drainage design cannot be precisely defined because
sufficient data may not be available to carry out the designs in the ideal way. As a result, the designer
must use simpler and apparently less accurate methods. Furthermore, even if data are available to allow
more sophisticated methods to be used, there are worrying large differences in the results that the
various methods give. Thus whether sophisticated numerical methods or simple methods are used,
different answers will arise from the various methods. The question therefore arises as to what is the
actual standard; what really is the true design storm for the selected return period? There is no answer
to this question. All that can be done is for the designer to use the methods available and to exercise a
degree of engineering judgement in selecting the result for the design.

The same arguments do not apply to the detailed engineering design of the components of the drainage
system once the maximum water flow has been estimated. For these, standard drawings and specifications
are provided. If the designer wishes to depart from these, then written approval will be required from

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 79

ERA. The designer must submit all proposals for departures from standards to the appropriate client
officer for evaluation as described in the Preface.

5.3 Hydrology: estimating maximum flow for drainage design


Before a drainage structure can be designed, it is necessary to determine the maximum likely flow of
water to be accommodated by the structure. Information may be needed on:
ƒ Water catchment area;
ƒ Rainfall characteristics;
ƒ Topography;
ƒ Vegetation and soils;
ƒ Catchment shape;
ƒ Stream and river flows if available;
ƒ Available storage in lakes and swamps;
ƒ Rural and urban development plans;
ƒ Water management plans (eg river basin master plans).

5.3.1 General principles

If the capacity of a water crossing structure is less than the volume of water flowing in the water course,
the upstream water level will rise. If this happens the water will flow through the structure at high speed.
This can erode the bed and banks of the water course. Furthermore, a high lateral pressure will be
exerted on the structure which can cause it to move and collapse. Eventually upstream water levels
may rise to the extent that water overtops the structure, damaging the road and blocking access. Water
crossing structures must therefore be designed to have a capacity equal to or greater than the maximum
water flow that is expected in the water course. This maximum flow depends on the rainfall pattern in the
area concerned and the characteristics of the particular catchment area on which the water, which will
eventually pass under the structure, falls.

Maximum flow is also important for the design of submersible structures such as drifts and causeways.
Although water flows over the surface, damage can still occur if the water rises higher than the level of
the approach slabs.

5.3.2 Storm severity and maximum flow

The cost of a drainage structure increases as its capacity increases; hence the first step is to determine the
maximum flow for which it is to be designed. To do this it is first necessary to decide on the return period
of the design storm for the structure (see Section 5.2.1). Table D.5.1 and D.5.2 show the return periods
for the design of typical structures on LVRs in Ethiopia.

Deciding on the return period defines the basic standard of the structure. The next step is to determine
how much water the structure must cope with when a storm of the design return period occurs. This
depends on:
ƒ The characteristics of the storm itself, namely the intensity, duration and the spatial extent of the
rainfall; and
ƒ The characteristic of the ground, or catchment, on which the rainfall falls.

Rainfall and storm characteristics depend on many factors and consequently vary widely between areas of
the world and within countries. The details of storms cannot be calculated theoretically and therefore all
rainfall information is based on experimental evidence. Thus, in order to determine the characteristics of
storms of a particular severity for drainage design purposes, good rainfall data need to be available. Such
data must include intensities and durations of all rainfall events over a period of time that is considerably
longer than the return period being used. If not, the number of occurrences of the return period storm
will be too low for statistically reliable information about its characteristics to be determined. In many
countries rainfall records are either not detailed enough or the records do not go back far enough for
accurate calculations. This is currently the situation in most of Ethiopia although the situation is improving
all the time. However, for LVRs, the situation is usually better because the return periods for design are
shorter than for the higher standard roads and there is a greater chance that rainfall data will be adequate.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 80

The characteristics of the storm are only the first part of the problem as these determine how much
rainfall falls onto the ground. The characteristics of the catchment area then determine how much of this
rainfall actually gets to the drainage structure and how long it takes to arrive. The situation is relatively
simple when dealing with the rainfall that runs off the road itself because the surface is uniform; its run off
characteristics are well understood; and the catchment area is known accurately. Where the catchment
is for a stream or river, the surface can be very variable with a variety of different run off characteristics,
thereby making run off estimation less accurate.

Provided rainfall records are available, there are several methods of calculating the maximum flow that
the drainage structures must cope with. Only one of them, the ‘Rational Method’ is really suitable for
LVRs and this is described below. However, before presenting details of this, there are several simple
observational or empirical methods of estimating the maximum flow in an existing water channel that
are suitable for LVRs although they need to be used with considerable care. Preferably as many of these
simple methods as possible should be used at the same time to provide the maximum level of reliability.

5.3.3 Simple estimation methods for maximum flow

The maximum water flow can be estimated in a number of simple ways requiring common sense and
observational techniques. The methods are based on:
ƒ Direct observation of the size of the watercourse;
ƒ Direct observation of erosion and debris;
ƒ History and local knowledge;
ƒ The Rational Method;
ƒ Correlation Tables;
ƒ The SCS method (US Soils Conservation Service, TR-55)
ƒ Successful practice.

Direct observation of the size of the watercourse


Most watercourses enlarge naturally to a size sufficient for the maximum water flows and no further. Thus
a simple way to design the structure is to measure the cross-sectional area of the water course itself. The
total cross-sectional area of the apertures of the structure should be equal to that of the water course so
that the flow is not constricted. Then upstream levels will not rise, and erosion, overtopping and collapse
will not occur.

This method assumes that the watercourse is free to drain and flows at a speed determined by the
gradient and nature of the channel. If, however, the water is flowing very slowly, for example, if the
water course is backing up from downstream, a reduced cross-sectional area will not cause upstream
levels to rise. In this case the structure does not need to have a cross-sectional area equal to that of the
water course. Further investigation is required to determine the maximum water level when water is
draining freely, but it is possible that a single aperture may be sufficient to ensure that the water is able
to gradually drain away as the downstream level falls.

The method also assumes that flood water levels do not rise beyond the watercourse and spread across
the surrounding land. If this is the case it is recommended that the total cross-sectional area of the
apertures of the structure should remain equal to that of the watercourse and that a series of smaller
structures, such as drifts, are constructed along the road on either side of the main structure.

The third assumption of this method is that the channel has not grown to its current size after years of
erosion by a much smaller flow of water. If this is the case, the channel size is not a guide to the maximum
water flow and is therefore not equal to the required total cross-sectional area of the apertures of the
structure. Additional investigation is required to check this.

This method must be supplemented with interviews with local residents.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 81

Direct observation of erosion and debris


Although most water courses enlarge to a size sufficient for maximum water flows, these maximum flows
may be very rare with a return period much longer than that for which some of the smaller structures may
be designed for LVRs In this case it is more sensible to find indications of typical high water levels. These
include lateral erosion on the banks of the water course and debris caught in the branches of trees.

It is unlikely that erosion or debris will remain visible for many years after a rare storm and therefore their
presence is a useful indicator of high water levels over a relatively short period of several years, similar
to the return period for which some of the smaller structures may be designed. As described above, the
total cross-sectional area of the apertures of the structure should be equal to that of the water course up
to the level of the erosion or debris.

This method must also be supplemented by interviews with local residents.

History and local knowledge


As suggested by the above two methods, past high water levels can be a reliable way of predicting
future levels. Past levels can be indicated either by evidence from the water course itself (as above) or
by recorded history, whether in measurements actually taken or from the recollections of local residents.

Measurements may be available but probably only if the site has been of interest to the authorities for
some time. This is unlikely to be the case for small water courses in most rural areas. Alternatively, rural
people tend to live in the same area for many years and have good memories of significant local events,
floods normally being very significant to those reliant upon the land. However, caution is required against
both the exaggeration of past levels and the belief that a single high flood many years ago was normal
rather than exceptional. Evidence should be sought from as many local residents as possible. It is usually
a good idea to make individual enquiries from people living on both banks of a river or stream and from
several locations along it. Alternatively a group may be asked to collectively agree on a maximum height
of the flood water and to agree on the frequency that such floods occur.

As in methods 1 and 2, it must be confirmed that the high water levels were for water which was flowing
rather than backed-up and stagnant. It must also be confirmed that water usage hasn’t changed since the
measurements were made. New upstream dams and land clearance can reduce or increase wet season
flows significantly.

Assuming that one can be confident that they were carefully made and that water usage has not changed,
recorded measurements are the most reliable way of predicting future levels. As described above, the
total cross-sectional area of the apertures of the structure should be equal to the cross-sectional area of
the water course up to the recorded or remembered level.

The Rational method


A number of numerical methods are available for estimating the maximum water flow which can be
expected in a water course. All methods require rainfall records and details of the catchment from which
rainfall flows into the water course, but some are too complex for use for LVRs. The Rational Method is
the most straightforward of these methods. It has a number of advantages and disadvantages shown in
Table D.5.3.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 82

Table D.5.3: Advantages and disadvantages of the Rational Method

Advantages Disadvantages
Not as accurate as more detailed methods
Simple equation; rapid estimation
(but within 30%)
Often over-estimates the flow
Very little data are required
(depends on catchment size)
The simplest form should not be used for catchment
greater than 3km2 but adjustments can be made for
larger catchments (see below)
Requires national records of storm duration and
intensity

The Rational Method estimates the maximum cross-sectional area of a water course by dividing the
maximum expected volume of water, q (m3/s), by the likely speed of the water, v (m/s).

q/v = a (m2) Equation 1


Where:
a = the cross-sectional area (m2) of the water when the water course is in flood.
The total cross-sectional area of the apertures of the structure should be equal to or
greater than a
q = the maximum expected volume of water (m3/s)

q is calculated from the following equation.

q = 0.278 x c x i x A (m3/s) Equation 2

Where:
c = the catchment coefficient
i = the intensity of the rainfall (mm/hour)
A = the area of the catchment (km2)

Details of the method are provided in Part B, Section 7.2.

Flow velocity and size of the drainage channel


Where appropriate, for LVRs, drainage structures that are not sensitive to exact predictions of flow such
as fords and drifts should be used, rather than culvert pipes and similar structures that have a fixed size.
However, where this is not appropriate, the required size of the drainage channel needs to be calculated.
This depends on the maximum flow at the location of the structure namely the value of q (m3/s) calculated
in equation 2. The maximum flow must be equal to the area of the channel multiplied by the likely
velocity of flow and hydraulic principles may have to be used to ensure that likely velocity is sufficient and
the design of the structure is adequate.

For example, for culverts in rolling and mountainous terrain it is often relatively simple to make sure that
their size is adequate by ensuring that their slop is 3-5% flow is not restricted on either the upstream
of downstream side. However, to ensure that the culvert is not significantly over designed it is usually
necessary to use hydraulic principles. The hydraulic design of drainage structures is dealt with in Part E
of this manual.

Correlation tables
It is possible to simplify the Rational Method into a correlation Table between catchment area and
the required total cross-sectional area of the structure. This is done for the conditions that prevail in a
particular region. The Table can then be used to design all structures in the same region in the future.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 83

In order to produce such a correlation table, the Rational Method must be used for a number of typical
situations and the following assumptions must be reasonably correct:
ƒ The storm intensity-duration pattern applies in the area in which a structure is to be constructed;
ƒ All catchments in the region are similar, for example, all are undulating with average soil and
average vegetation cover;
ƒ All similar structures are to be designed for the same return period.

Additional sophistication can be provided by developing separate correlation Tables for structures that
require different return periods. If necessary these can also be repeated for catchments with different
surface cover. Table D.5.4 is an example of a typical Table correlating the catchment area with required
total cross-sectional area of the structure for a set of average conditions.

Finally, the Table can be extended to show the correlation between catchment area and the required
number of standard culvert apertures. For example, assuming that a standard aperture is 600 mm wide,
with 600 mm walls onto which a semi-circular arch is constructed, its cross-sectional area is 0.50 m2 (Figure
D.5.1)

Table D.5.4: Catchment area and total cross-sectional area of an example


structure for a set of standard conditions based on the Rational method

Required cross-sectional area (m2)

Catchment area (ha) C=0.2 C=0.7


(rolling terrain, highly permeable (rolling terrain, arid area, low
cultivated soil) permeability soil)
10 0.35 1.0
25 0.75 2.0
50 1.25 3.5
75 1.75 4.7
100 2.2
125 2.6
150 3.0
175 3.5
Note:
The rainfall intensity is assumed to be 100mm/hr. This table is not applicable for higher intensities

Figure D.5.1: Example of a standard arch culvert

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 84

The full correlation Table between catchment area, required cross-sectional area and the required number
of standard apertures can now be completed (Table D.5.5). When more than three standard apertures are
required, it is recommended that the structure is designed in more detail. Although a series of standard
apertures may be the most suitable solution in a wide, shallow channel, if the channel is narrow and deep,
a box culvert, larger diameter pipes or even a small bridge may be more appropriate and cost effective.
The decision between these alternatives should be made carefully.

Table D.5.5: Catchment area, total cross-sectional area and number of standard
arch culverts for a set of standard conditions based on the Rational Method

Required cross- Required cross-


No. of No. of
Catchment area sectional area sectional area
standard standard
(ha) (m2) (m2)
apertures apertures
C=0.2 C=0.7
10 0.35 1 1.0 2
25 0.75 2 2.0 4
50 1.25 3 3.5 Carry out more
detailed design
75 1.75 4 4.7
100 2.2 Carry out more
detailed design
125 2.6
150 3.0
175 3.5

The SCS method


The SCS method for calculating rate of run off requires much of the same basic data as the Rational
method namely catchment area, a runoff factor, time of concentration and rainfall. However, the SCS
method also considers the time distribution of the rainfall, the initial rainfall losses to interruption and
storage and a filtration rate that decreases during the course of a storm. It is therefore practically more
accurate than the Rational method and is applicable when the catchment area is larger than 50 hectares.

The details of the SCS method are provided in Part B, Section 7.3

Successful practice
This method is similar in many respects to the correlation method above, except that instead of using
a theoretical basis for the correlation Table, the success of previous designs is used. Thus if a high
proportion of structures along a road or in a region have been in operation for a number of years without
overtopping or being damaged during wet season floods, it is reasonable to assume that the relationship
between catchment area, catchment characteristics, rainfall intensity and maximum water flow that was
used to design the structures, even if extremely simple, is reasonably valid. The only proviso is that care
must be exercised to ensure that all the structures have not been over-designed.

Within a region, catchment characteristics and rainfall intensity are normally consistent therefore it can be
assumed that a simple relationship between catchment area and maximum water flow also exists. If the
area of a number of catchments and the total cross-sectional area of the apertures of the structures are
measured, the relationship between them can be established in the form of a Table as shown above. For
new designs, it is a simple matter of measuring the area of the catchment and using the Table to establish
the required cross-sectional area, although it is necessary to be cautious if the catchment in question is
different in topography, soil or vegetation to the catchments used to derive the Table.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 85

5.3.4 Normal flow rates

Whilst information is required on the severity of different storm events in order to design the drainage
structures, for some structures it is also necessary to know something about the “normal” or less severe
flows that are likely to occur during each year.

For example, drifts and fords are designed for water to flow over the running surface and it is not expected
that vehicles can use them for 24 hours a day and for 365 days per year. For such structures, annual rainfall
records still need to be examined to determine the rare storms that could engulf the whole structure
and put it at risk, but under normal operating conditions water is expected to cover the “carriageway”
or running surface at much more frequent intervals. there will be days in the year where the water will
be too deep for several hours and it is necessary to design the structure so that it is not impassable for
more than, on average, a specified number of hours each year. If the volumes of water are too great for
a simple ford, then a vented ford of some other structure will need to be considered.

Similarly, the prevention of erosion damage caused by the “regular” flow of water through a structure
requires information about the range of flow velocities that occur every year. In other words, slow regular
attrition can be as damaging in the long term to some structures as the rare severe storm.

Whilst the hydrological principles described in this chapter apply to these conditions, the hydraulic
designs for the various structures are dealt with in Part E.

5.3.5 Direct flow methods

Methods of estimating flow discharges are divided into three categories;


ƒ Direct flow methods;
ƒ Run-off modelling;
ƒ Regionalised flood modelling.

Methods based on simple empirical assessments of stream flows and local recollections, are examples of
direct flow methods. They are clearly prone to considerable uncertainty and not suitable for estimating
long return period flows.

However, the same statistical techniques that are used to determine rain storms of different return periods
can also be applied to direct flow records of streams and rivers. Direct flow data or depth gauge data is
often available for regional and local streams and rivers and, together with measurements of the channel
geometry, can be used in conjunction with Manning’s Equation to determine flow velocity and thus flow
volume (discharge, or capacity) (see Part E). Statistical analysis can then be used to determine flows of
different return periods. As with rainfall data, the period during which such records have been kept needs
to be reasonably long for statistical reliability. If such data are available, direct flow methods are usually
the most accurate method of estimating flows for different return periods, especially for large catchments
where run-off modelling is not appropriate.

The Rational Method is an example of run-off modelling and, together with rainfall data and the statistical
analysis of storms provides a method of estimating longer return period flows for small catchments as
described above.

Flood modelling is the least accurate method and is not discussed further.

5.4 Components of external drainage


An effective external drainage system must fulfil several functions:
ƒ Prevent or minimise the entry of surface water into the pavement;
ƒ Prevent or minimise the adverse effects of sub-surface water;
ƒ Remove water from the vicinity of the pavement as quickly as possible;
ƒ Allow water to flow from one side of the road to the other.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 86

This must be achieved without endangering the road or adjacent areas through increased erosion or risk
of instability.

Thus an external drainage system consists of several complementary components;


ƒ Surface drainage to remove water from the road surface quickly.
ƒ Side drainage to:
• Take water from the road;
• Prevent water from reaching the road.
ƒ Turnouts to take the water in the side drains away from the road.
ƒ Cross drainage to allow the water in the side drains, and from any other sources, to cross the road
line by channelling it under or across the road.
ƒ Interceptor drains to collect surface water before it reaches the road.
ƒ Sub-surface drains to cut off sub-surface water and to lower the water table when required.
ƒ Erosion control (often simple scour checks) to slow down the water in the side drains and prevent
erosion in the drains themselves and downstream of drainage outlets or crossings.

All these types of drains have to work together in order to protect the road from being damaged by
water. Cross-drainage includes structures to allow permanent or seasonal water courses to cross the road
line and therefore includes bridges. The appropriate structures for low volume roads are dealt with in Part
E of the manual.

5.4.1 General principles

Conservation of the natural drainage system around the road alignment is one of the most important
concerns during design and construction. By effectively creating a barrier to natural surface drainage
that is only punctuated at intervals by constructed drainage crossings, road construction can lead to
significant local increases in catchment areas and increased water flows. Furthermore, in the case of
paved roads especially, road drainage reduces the time taken to reach maximum flow by shedding water
from impermeable surfaces relatively quickly. Therefore, in addition to constructing a drainage system to
convey the design run-off without surcharge, blockage by sediments, or scour, attention must be paid
to strengthening those parts of the natural slope drainage system that experience increased run-off, and
hence erosion potential, as a result of road construction. The main ways of doing this are to:
ƒ Control road surface drainage;
ƒ Design culverts or drifts that convey water and debris load efficiently;
ƒ Optimise the frequency of drainage crossings to prevent excessive concentration of flow;
ƒ Protect drainage structures and stream channels for as far downstream as is necessary to ensure
their safety and prevent erosion of land adjacent to the water course;
ƒ Plant vegetation on all new slopes and poorly-vegetated areas, around the edges of drainage
structures and appropriately along stream courses, without impairing their hydraulic efficiency or
capacity.

5.4.2 Sources of water

The main sources of water ingress to, and egress from, a typical pavement cross-section are listed in
Table D.6.20. This chapter is concerned with dealing with the water that flows outside the road prism.
Minimising the amount of water that gets into the structure of the road itself and minimising the damage
that it can cause are dealt with in Section D.6.18

5.4.3 Road surface drainage

Camber and cross-fall are part of the geometric design of the road. Their values are discussed in the
geometric design chapter of this Manual (Chapter D.4) but are repeated here for completeness.

For earth and gravel roads, the design cross-fall or camber should be a minimum of 4% and normally
about 6%, except in arid areas. This helps to prevent ponding on slack road gradients and longitudinal
scour on long, steep sections of alignment. In service, the cross fall should not be allowed to decrease
to less than 3% before maintenance is carried out to restore it. For paved roads cross-fall should be 3%
except on super-elevated sections. On structures the cross fall can be relaxed to 2-5% on condition that

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 87

the surface is durable and positive drainage (eg through scuppers) is provided and regular maintenance
is carried out.

In hilly, mountainous and escarpment terrain, an inward-sloping road carriageway (see cross sections in
Chapter D.4) is the normal means of shedding water from the road surface. The inward slope incorporates
an inherent factor of safety in retaining water that has accidentally escaped from the drainage system.
Occasionally, an outward-sloping road surface has been advocated on the grounds that, by allowing
water to disperse gently onto the hill slope along the whole length of the road, the potential for erosion
is reduced. In practice, the opposite is usually true. The method undoubtedly offers very large financial
savings in the reduction of drainage structures, but it is a highly hazardous form of design that cannot be
recommended except in areas of very low erosion potential. The design has the following weaknesses:
ƒ In practice it is impossible to design a road geometry for a distributed flow of water (topography
is the controlling factor).
ƒ Road settlement and the action of traffic, will, in time, change the design cross-fall
ƒ Road repairs will locally alter the cross-fall.
ƒ Partial blockage of the road by debris results in a change of the flow pattern of drainage water, and
instant local surcharge.
ƒ The outer edge of the road is particularly susceptible to erosion which can reach a disastrous level
before maintenance crews can be mobilised.
ƒ Slope protection from uncontrolled runoff requires a lengthy period of post construction monitoring
and remedial works that is usually not practicable for LVRs.
ƒ Vehicles can slide sideways uncontrollably across a wet road surface and over the edge unless
expensive partial safety barriers are provided.

An outward sloping cross section is therefore only suitable for pedestrian, NMT and IMT use and where
the longitudinal gradient is low and the surfacing material has low erosion potential.

5.4.4 Side drains

Side drains serve two main functions namely to collect and remove surface water from the immediate
vicinity of the road and, where needed, to prevent any sub-surface water from adversely affecting the
road pavement structure.

Seepage may occur where the road is in cut and may result in groundwater entering the sub-base or
subgrade layers as illustrated in Figures D.5.2 and D.5.3. Inadequate surface or subsurface drainage can
therefore adversely affect the pavement by weakening the soil support, and initiating creep or failure of
the downhill fill or slope. Localised seepage can be corrected in various ways but seepage along more
impervious layers, such as shale or clay, combined with changes in road elevation grades, may require
subsurface drains as well as ditches as shown in Figure D.5.4.

Figure D.5.2: Inadequate side drains

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 88

Figure D.5.3: Inadequate side drains and subsurface drainage

Figure D.5.4: Proper interception of surface runoff and subsurface seepage

If the road has effective side drains and adequate crown height, then the in situ subgrade strength will
stay above the design value. If the drainage is poor, the in situ strengths will fall to below the design
value. Crown height is discussed in Section D.6.18.

Side drains can be constructed in three forms (Figure D.5.5): V-shaped, rectangular or trapezoidal.

Figure D.5.5: Side drains

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 89

The choice of side drain cross-section depends on the required hydraulic capacity, arrangements for
maintenance, space restrictions, traffic safety and any requirements relating to the height between the
crown of the pavement and the drain invert (as discussed in Section D.618.

Design volumes of run-off are usually estimated using the Rational Method (Section 5.3.4). Flow velocities
are calculated from the Manning equation using roughness values shown in Table E.6.5. It should be
noted that most published roughness and velocity data are based on clean water flow. Along mountain
roads sediment-laden water is more common; hence flow velocities may be lower, but this errs on the
side of safety.

Under normal circumstances, the adoption of a trapezoidal cross-section will facilitate maintenance and
will be acceptable from the point of view of traffic safety. It is much easier and appropriate to dig and
clean a trapezoidal drain with hand tools and the risk of erosion is lower. The minimum recommended
width of the side drain is 500mm. This shape carries a high flow capacity and, by carefully selecting the
gradients of its side slopes, it will resist erosion.

The V-shape is the standard shape for a drainage ditch constructed by a motor-grader or towed grader.
It can be easily maintained by heavy or intermediate equipment but it has relatively low capacity
necessitating more frequent structures for emptying it. Furthermore the shape concentrates flow at the
invert and encourages erosion.

The rectangular shaped drain requires little space but needs to be lined with rock, brick or stone masonry,
or concrete to maintain its shape.

In very flat terrain and reasonable soils it is often best to use wide unlined “meadow drains”. These are
formed shallow and continuous depressions in the surface that avoid abrupt changes in surface profile.
When properly designed, their capacity is high and the flow velocity is low so that erosion should be
controlled.

When the subgrade is an expansive soil, changes in moisture content near to the road itself must be
minimised. The design of the side drains and side slopes for such conditions are described in Section
D.6.19.

As far as traffic safety is concerned, a wide and shallow drain for a given flow capacity is preferable to
a deeper one but, particularly on steep sidelong ground, the extra width required to achieve this may
be impracticable or too expensive. Side drain covers can be used to provide extra road width in places
where space is severely limited but their widespread use is not recommended. They are expensive and
the drains that they cover are then difficult to maintain.

Side drains (as well as the road itself) should have a minimum longitudinal gradient of 0.5%, except on
crest and sag curves. Slackening of the side drain gradient in the lower reaches of significant lengths of
drain should be avoided in order to prevent siltation.

For the construction of LVRs the spoil material from the construction of the side drain is usually used to
provide the formation of the road and its camber. When roads are built using labour-based methods
this is usually the only source of material (unless the road is to be built on an embankment) hence it is
important that the size of the drain is wide and deep enough to provide sufficient material. Failure to do
so is often the reason for the resulting low camber and early deterioration of gravel and earth roads. In
most circumstances a wide trapezoidal drain is the ideal solution.

Access across side drains for pedestrians, animals and vehicles needs to be considered. Community
representatives should be consulted with regard to locations, especially for established routes. The
methods that could be used are:
ƒ Widening the drain and taking its alignment slightly away from the road;
ƒ Hardening the invert and sides of the drain;
ƒ Beam/slab covers or small culverts.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 90

The arrangement must be maintainable and not risk blockage of the side drain. Failure to accommodate
these needs will usually result in later ad hoc arrangements that compromise the function of the side
drain.

Groundwater in the subgrade can be released either by using a drainage layer at sub-base level or by
incorporating gravel cross drains (grips) in the shoulder that exit via a weephole in the side drain backed
with a piece of filter fabric. The weepholes must be set at the correct level to take the water from the
appropriate pavement layer and also the drain must be sufficiently deep so that there is little possibility
of the water in the drain being of sufficient depth for it to flow back into the road.

Deeper drains, comprising a filter-wrapped perforated pipe within a graded gravel backfill, can be
constructed under very wet slope conditions to a depth of 1-1.5m below the level of the side drain invert,
and led to the nearest culvert inlet.

5.4.5 Erosion control in the side drain

When the water flows too fast, it will erode the bottom of the drain. The faster water flows, the more soil
it can erode and carry away. There are various methods of reducing erosion, the two most common being
to build simple scour checks or to line the drains.

Scour checks (sometimes called check dams) reduce the speed of water and help prevent it from eroding
the road structure. Typical designs are shown in Figure D.5.6. The scour check acts as a small dam and,
when naturally silted up on the upstream side, effectively reduces the gradient of the drain on that side,
and therefore the velocity of the water. The energy of the water flowing over the dam is dissipated
by allowing it to fall onto an apron of stones. Scour checks are usually constructed with natural stone,
masonry, concrete or with wooden or bamboo stakes. By using natural building materials available along
the road side, they can be constructed at low cost and be easily maintained after the road has been
completed.

There must be sufficient cross-sectional area above the scour check (ie where the water has been slowed
down) to accommodate the maximum design flow. Wide drains are also preferred to reduce the velocity
of the water and minimise erosion but space is at a premium in the type of terrain where scour checks are
required so wide drains may not always be practicable.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 91

Figure D.5.6: Typical design of scour checks

The distance between scour checks depends on the road gradient and the erosion potential of the soils.
Table D.5.6 shows recommended values but these may need to be modified for more erodible soils.

Table D.5.6: Spacing between scour checks

Scour check interval


Road gradient (%)
(metres)
3 Not required
4 17
5 13
6 10
7 8
8 7
9 6
10 5
12 4

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 92

After the basic scour check has been constructed, an apron should be built immediately downstream
using stones. The apron will help resist the forces of the waterfall created by the scour check. Sods of
grass should be placed against the upstream face of the scour check wall to prevent water seeping
through it and to encourage silting to commence on the upstream side. The long term goal is to establish
a complete grass covering over the silted scour checks to stabilise them.

Sections of side drain with scour checks cannot be maintained by motor grader or towed grader and will
need to be maintained by hand.

Depending on the strength of the material in which the drains are excavated and the velocity of runoff
they are expected to carry, side drains may also need to be lined. The controlling factor is the ease of
erosion of the soil. Table D.5.7 indicates the critical velocities for different materials. With velocities
greater than those shown, erosion protection measures will be required.

The drains may be lined with heavy duty polythene, or some other impermeable material, before masonry
pitching is applied. This will prevent water penetration if the masonry becomes cracked by movement.
The lining can also be extended up the banks to prevent lateral erosion. When the cross-sectional area is
less than about 0.1m2 and the gradient is gentle, drains can be lined with unbound masonry. Larger and
steeper drains are lined with mortared masonry, although they are considerably more expensive. Any
gap between the drain and the hill side must be filled with compacted impermeable material (eg clay)
sloping towards the drain to minimise infiltration behind it.

Table D.5.7: Permissible flow velocities (m/sec) in excavated ditch drains

Water carrying
Water carrying
Soil type Clear water sand and fine
fine silt
gravel
Fine sand 0.45 0.75 0.45
Sandy loam 0.55 0.75 0.6
Silty loam 0.6 0.9 0.6
‘Good’ loam 0.75 1.05 0.7
Lined with established grass on good soil 1.7 1.7 1.7
Lined with bunched grasses (exposed soil
1.1 1.1 1.1
between plants)
Volcanic ash 0.75 1.05 0.6
Fine gravel 0.75 1.5 1.15
Stiff clay 1.15 1.5 0.9
Graded loam to cobbles 1.15 1.5 1.5
Graded silt to cobbles 1.2 1.7 1.5
Alluvial silts (non colloidal) 0.6 1.05 0.6
Alluvial silts (colloidal) 1.15 1.50 0.9
Coarse gravel 1.2 1.85 2.0
Cobbles and shingles 1.5 1.7 2.0
Shales 1.85 1.85 1.5
Rock Negligible scour at all velocities

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 93

Channels can also be lined with gabion, dry stone pitching, rip-rap or vegetation.

When constructing a channel lining it is important to reproduce, as a minimum, the dimensions of the
original channel. A curved rather than rectangular shaped cross-section to the bed lining is preferable.
The main disadvantage with channel linings is that a lower channel roughness leads to an increase in flow
velocity and hence an increase in scour potential further downstream. In the case of masonry aprons,
or gabion mattresses with masonry screeds, some reduction in velocity can be achieved by cementing
protruding stones into the surface.

Masonry linings can be constructed to fit the stream bed much more closely than gabion. They are also
less easily abraded, but they cannot tolerate significant settlements, loss of support by seepage erosion
or high groundwater pressure.

Dry stone pitching is usually only suitable where discharges are lower than 1 m/sec per metre width, and
where sediment load is relatively fine-grained.

Grass can provide some resistance to channel erosion and may be used where flow velocities are not
expected to be too high. The introduction of grass will also tend to reduce flow velocities, although
channel vegetation should not be so widespread as to inhibit or divert flow, which could lead to bank
scour. Where immediate effective protection is required, a structural solution is preferable to a vegetative
one.

The winning of boulders and cobbles from gully beds for road construction materials can reduce the
armouring effect provided by coarse material. If the bed material appears to be weathered and static
for much of the time, then its removal could expose more erodible sediments beneath. In such cases,
extraction from the channel bed should be discouraged or prohibited. Conversely, where the entire bed
deposit is fresh and evidently mobile, the removal of material may not have a significant effect on channel
stability, especially if the quantities concerned are small compared to the volume of bed load.

Cascades or steps in the drain long-section can also be a useful means of reducing flow velocity, although
both scour checks and cascades can impede the transport of debris, increasing the risk of blockage.

5.4.6 Mitre drains or turnouts

It is normally best practice to discharge the water from the side drains as frequently as possible. If it
can be discharged on the same side of the road as the drain, a turnout or mitre drain is used to lead the
water away. Mitre drains simply lead the water onto adjacent land therefore care is required to design
them to ensure that problems associated with the road are not passed on to the farmer or landowner. It
is advisable to consult adjacent land users regarding the discharge of water onto their land to gain their
support and agreement and to avoid possible problems in the future.

The principle is to aim for low volumes and low velocities at each discharge point to minimise local
erosion and potential downstream problems. The maximum spacing of turnouts depends on the volume
of water flowing in the drain and therefore hydrological principles may need to be used to estimate this.
However, in many cases it is only water shed from the road itself that flows in the side drain and this is
relatively easy to estimate using the Rational Method (Section 5.3.4).

Where soils are very erodible, it may be preferable to increase side drain capacity to convey runoff to
the next available safe discharge point rather than to construct side drain turnouts or relief culverts on
erodible slopes. With the extra volumes of water that this entails, the design of these less frequent safe
discharge points will usually be more expensive.

In mountainous terrain the discharge of water is considerably more difficult and consequently more
expensive. This is discussed in more detail in Section D.5.6.

Table D.5.8 gives the maximum spacing. However, spacings of mitre drains should normally be more
frequent than this and values as low as one every 20 m may be required to satisfy landowners.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 94

Table D.5.8: Maximum spacing of mitre drains

Maximum mitre drain


Road gradient (%)
interval (metres)
12 40
10 80
8 120(1)
6 150(1)
4 200(1)
2 80(2)
<2 50(2
Notes:
1. A maximum of 100m is preferred but not essential
2. At low gradients silting becomes a problem

In order to ensure that water flows out of the side drain into the mitre drain, a block-off is required as
shown in Figure D.5.8. It is essential that the mitre drain is able to discharge all the water from the side
drain. If the slope of the mitre drain is insufficient, the mitre drain needs to be made wide enough to
ensure this.

The desirable slope of the mitre drains is 2%. The gradient should not exceed 5% otherwise there may
be erosion in the drain or on the land where the water is discharged. The drain should lead gradually
across the land, getting shallower and shallower. Stones may need to be laid at the end of the drain to
help prevent erosion.

In mountainous terrain, it may be necessary to accept steeper gradients. In such cases, appropriate soil
erosion measures should be considered.

In flat terrain, a small gradient of 1% or even 0.5% may be necessary to discharge water, or to avoid very
long drains. These low gradients should only be used when absolutely necessary. The slope should
be continuous with no high or low spots. For flat sections of road, mitre drains are required at frequent
intervals of 50m to minimise silting.

Angle of mitre drains


The angle between the mitre drain and the side drain should not be greater than 45 degrees. An angle
of 30 degrees is ideal (Figure D.5.7).

Figure D.5.7: Angle of mitre drain

If it is necessary to take water off at an angle greater than 45 degrees, it should be done in two or more
bends so that each bend is not greater than 45 degrees (Figure D.5.8).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 95

Figure D.5.8: Mitre drain angle greater than 45 degrees

5.4.7 Wet lands

Road crossings in wet areas, including damp meadows, swamps, high groundwater areas, and spring
sources, are problematic and undesirable. Wet areas are ecologically valuable and difficult for road
building. Soils in these areas are often weak and require considerable subgrade reinforcement. Drainage
measures are expensive and may have limited effectiveness. Therefore, if at all possible, such areas
should be avoided.

If wet areas must be crossed, special drainage or construction methods should be used to reduce impacts
from the crossing which will usually require an embankment. They include multiple drainage pipes or
coarse permeable rock fill to keep the flow dispersed, subgrade reinforcement with coarse permeable
rock, grade control, and the use of filter layers and geotextiles. The objective is to maintain the natural
groundwater level and flow patterns dispersed across the meadow and, at the same time, provide for a
stable, dry roadway surface.

Local wet areas can be temporarily crossed, or ‘bridged’ over, using logs, landing mats, tyres, aggregate,
and so on. Ideally, the temporary structure will be separated from the wet area with a layer of geotextile.
This helps to facilitate removal of the temporary material and minimizes damage to the site. Also, a layer
of geotextile can provide some reinforcement strength as well as provide separation to keep aggregate
or other materials from punching into the weak subgrade.

Subsurface drainage, through use of under-drains or aggregate filter blankets, is commonly used along
a road in localized wet or spring areas, such as a wet cut bank with seepage, to specifically remove the
groundwater and keep the roadway subgrade dry. A typical under-drain design uses an interceptor trench
1-2 meters deep and backfilled with drain rock, as shown in D.5.9.

Subsurface drainage is typically needed in local wet areas and is much more cost-effective than adding
a thick structural section to the road or making frequent road repairs. In extensive swamp or wet areas,
subsurface drainage will often not be effective. Here, either the roadway platform needs to be raised
well above the water table, or the surfacing thickness design may be based upon wet, weak subgrade
conditions that will require a relatively thick structural section. A thick aggregate layer is commonly used,
with the thickness based upon the strength of the soil and anticipated traffic loads.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 96

Figure D.5.9: Typical sub-surface drain

5.4.8 Interceptor, cut-off or catch-water drains.

As its names imply, such drains are constructed to prevent water flowing into vulnerable locations by
‘intercepting’, ‘cutting off’ or ‘catching’ the water flow and diverting it to a safe place.

For example, where the road is situated in sidelong ground on a hillside, a significant amount of rainwater
may flow down the hill towards the road. This may cause damage to the face of cuttings and even cause
landslips. Where this danger exists an interceptor drain should be installed to intercept this surface water
and carry it to a safe point of discharge; usually a natural watercourse (Figure D.5.10).

The interceptor drain should be located so that:


ƒ It drains at a satisfactory gradient throughout its length (2%)
ƒ It is not too close to the cut face. It should be at least 3-5m away so that it does not increase the
danger of a landslip.

If steep gradients in the drain are unavoidable then scour checks (Section 5.4.5) should be installed.

The material excavated to form the drain is usually placed on the downhill side to form a bund. Vegetation
cover should be established as soon as possible in the invert and sloping sides of the interceptor drain
and bund to resist erosion. However, where no seepage is tolerable, consideration should be given to
lining the drain so that it is truly impermeable thereby minimising the risk that water will weaken the cut
slope.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 97

Figure D.5.10: An interceptor, cut-off or catch-water drain

The interceptor drain should normally be 600mm wide, 400mm (minimum) deep with sides back-sloped
at 3:1 (vertical: horizontal).

Similar interceptor drains can be used whenever water is flowing towards the road; they are not restricted
to protecting cut slopes, but such drains are only useful when surface runoff rates are significant. Surface
runoff can be expected only during high intensity rainfall on moderate to steeply-inclined slopes, on
slopes of low permeability where vegetation is patchy, or where runoff from agricultural land becomes
concentrated onto un-vegetated soil slopes. If surface runoff is substantial, and there is a clear threat of
erosion or slope failure further downslope, the use of surface drains is justifiable. However, they are not
without problems. They are easily damaged or blocked by debris and are often not seen and therefore
not cleaned on a regular basis. In addition, differential settlement or ground movement will dislocate
masonry drains, leading to concentrated seepage, if they are constructed without polythene lining. If
there is any doubt about their effectiveness, or whether they can be maintained in the long term, it is
better not to build them than have them become forgotten and allowed to fall into disrepair, making
drainage and instability problems worse.

Factors to be considered in the design of surface drains are:


ƒ Water collected by the drain must be discharged safely in a manner that will not initiate erosion
elsewhere.
ƒ Construction of masonry-lined drains should be limited to undisturbed slope materials. Differential
settlement, which frequently occurs in made ground and particularly at the interface between
natural ground and fill, will lead to rupture.
ƒ Drain gradients should not exceed 15 %.
ƒ For ease of maintenance and to minimize erosion they should be wide and have sloped sides.
ƒ Where people have to cross the drain, easy side slopes should be provided so that the people will
not fill the drain to cross it.
ƒ Stepped drain outlets should be provided with a cascade down to the collection point.
ƒ Drains should discharge into a stream channel wherever possible, and preferably into channels that
already convey a sizeable flow in comparison to the drain discharge.
ƒ Low points in the drain system should be designed against overtopping by widening or raising the
side walls.
ƒ Lengths of drain should be kept short by the construction of frequent outlets in order to reduce
erosion potential should drain failure occur.

Where it is not practicable to discharge cut-off drainage into an adjacent stream channel, cascades can
be constructed down the cut slope to convey water into the side drain. However, these structures are

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 98

often vulnerable to the effects of side splash, undermining by seepage erosion and concentrated runoff
along their margins. They must be designed to contain the water, and their margins must be protected
with vegetation or stone pitching.

In very sensitive locations a simple earth bund can be constructed instead of a cut-off drain. The
disadvantage is that material may have to be excavated a little way from the bund and cast or transported.
However, the distinct advantage is that the soil surface is not disturbed at the bund and existing vegetation
can be encouraged to grow onto the bund to stabilise it. A range of bio-engineering measures can also
be used in sensitive areas and specialist advice should be sought on this.

5.4.9 Chutes

Chutes are structures intended to convey a concentration of water down a slope that, without such
protection, would be subject to scour. Since flow velocities are very high, stilling basins are required to
prevent downstream erosion. The entrance of the chute needs to be designed to ensure that water is
deflected from the side drain into the chute, particularly where the road is on the steep grade.

5.5 Erosion control


Erosion forces are one of the most destructive forces an engineer has to contend with in designing and
constructing roads but the problem of erosion can be minimised by providing suitable precautions at all
stages in the design.

The construction of a road often requires land clearing and levelling in the preparation stage. It involves
removal of shrubs and trees that are normally acting as wind-breaks, rainfall ‘sponges’ and soil stabilisers
and therefore, on removal, the soil erosion process is accelerated, especially in sloping areas.

After construction, erosion often appears in road embankments as gullies in the shoulders and embankment
slopes; as gouges in the side drains, which endangers the traffic; and in the actual road foundation. It
undermines fills and backslopes, initiating landslides, undermines bridge foundations and other road
structures, clogging drainage ditches, culverts and other waterways in the watershed.

It is often impossible to make reliable predictions concerning the full extent of erosion protection likely
to be required until the road drainage system is fully functioning and the slopes and drainage channels
have responded to the new drainage regime. Constructed roads interrupt the natural drainage of an
area, and concentrated water discharge through culverts and drains leads to soil erosion if the drainage
is not properly designed.

The general design philosophy of stream course protection is to dissipate as much water energy as
possible in the vicinity of the road itself, where erosion is likely to be worst, and protect outfall channels
down to a point where they are large enough or sufficiently resistant to withstand the increased flow.

Outfall channel protection usually consists of check dams, cascades and channel linings. It is not
uncommon to build protection works for 20-60m downstream of culverts, and there are instances where
they have been constructed for distances of 500m or more. If investment to this level of protection is
considered necessary, it is clearly important to be sure that the measures will be effective.

Protection of erodible channels upstream of culverts is usually accomplished by check dams and cascades
constructed over much shorter lengths, and usually within 20m of the inlet.

Good erosion control should preferably start at the top of the rainfall catchment with the objective being
the reduction of water run-off towards the road. The road should be designed with sufficient numbers
of culverts and mitre drains to avoid large concentrations of water discharging through the structures.
Below the road, water should be channelled safely to a disposal point (eg a stream) or dispersed without
causing damage to the land.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 99

Often the problem of erosion extends beyond the road environment itself and affects dams, slopes, rivers
and streams well away from the road. The steepness of the cut slopes and constructed embankments
together with a deficiency in drainage means that landslides may result.

The storage of spoil during road construction may kill local indigenous vegetation which can cause
erosion and slope stability problems. In mountainous regions large quantities of spoil can be generated
and the balance between cut and fill is difficult to maintain. Storage of spoil or disposal through haulage
may be difficult; therefore the process will involve more effective environmental management to avoid
erosion problems.

The channelling of run-off through new routes will result in changes within the natural equilibrium.
Excessive water flows may be generated when drainage ditches and other water control structures have
become blocked or damaged. The excessive flows will find new routes which will result in an enlargement
of the erosion problems.

Chain impacts, including soil contamination and damage will affect the road environs. Soil contamination
will possibly result in vegetation loss and therefore resistance to erosion. Construction of the road may
result in deforestation which, in turn, will lead to erosion of bare slopes, the re-channelling of rivers and
streams, possibly minor landslides and changes in the microclimatic conditions. The roadworks themselves
will temporarily increase waterborne material because rainfall erodes the surface of temporary or new
surfaces before they are stabilised.

5.5.1 Identifying and assessing potential erosion problems

The initial project survey will possibly indicate the range of problems that may be encountered, and the
design should include measures to mitigate the problems. The following should be considered:
ƒ Previous or similar construction projects. These can be useful indicators, and evidence of erosion
problems can be obtained from the local population.
ƒ Desk study as part of the pre-feasibility study. More detailed information can often be obtained
from a desk study, from maps (geological hydrological, and topographic) and aerial photographs
if available.
ƒ Historic evidence. Signs of erosion or soil instability and evidence of major floods and local
agricultural practices should be sought.
ƒ Drainage design. Consider how water flows will be concentrated by the construction of the road.
ƒ Cleared areas. Review the areas that will no longer be vegetated after the construction.
ƒ Cut or fill slopes. Review the slopes that will be at greater angles than previous natural slopes.

5.5.2 Mitigation measures to control erosion and scour

There are a wide range of methods and techniques that may be employed to prevent erosion and allow
for the construction of a road in an environment with little or no erosion impact. The simple technique of
replanting cleared areas will be effective generally, while the more difficult cases may be addressed with
measures such as retaining walls.

The simplest ways of controlling erosion of soil in road projects is by avoidance. This can be achieved by:
ƒ Reducing the area of ground that is to be cleared;
ƒ Quickly replanting cleared areas, maintaining the planted areas and specific bio-engineering
measures;
ƒ Avoiding erosion sensitive alignments;
ƒ Controlling the rate and volume of water flows in the area.

Replanting
An important method for reducing erosion and stability problems is by replanting cleared areas. It is
suggested that this procedure should be carried out as early as possible during the construction process,
and before the erosion becomes too advanced. It is important to select the correct vegetation that will
address the specific engineering function required for stabilisation.

The engineering function of vegetation in erosion protection measures are:

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 100

ƒ Retaining material from moving over the soil surface;


ƒ Armour plating the surface against erosion and abrasion;
ƒ Supporting the slope by stabilising it from the base;
ƒ Reinforcing the soil by increasing its effective shear strength;
ƒ Drain the soil profile by taking water into the roots.

Slope protection
Avoiding erosion by stabilising slopes requires good engineering design of the slope form and drainage.
This topic is dealt with in more detail in the Chapter 3 on geotechnical issues but is summarised here for
convenience.

Slope retaining techniques are necessary when:


ƒ The slopes are too high or steep;
ƒ There is a risk of internal erosion or localised rupture due to drainage problems;
ƒ It is necessary to decrease the amount of earthwork because the road width is limited.

Well-established techniques for slope protection against erosion are:


ƒ Intercepting ditches at the top and bottom of slopes. Gutters and spillways are methods used to
control the flow of water down the slope.
ƒ Stepped or terraced slopes to reduce the height of the slope.
ƒ Riprap or rock facing material embedded in a slope face, sometimes with planted vegetation.
ƒ Retaining structures such as gabion cribs.
ƒ Retaining walls.
ƒ Reinforced earth, where the embankment walls build up as the earth fill is placed, within anchors
compacted into the fill material.
ƒ Shotcreting and geotextiles; techniques that are usually expensive and should therefore be carefully
considered before being used for specific applications.

Table D.5.9 is a summary of the various mitigation measures, their effectiveness and comparative costs.

Table D.5.9: Comparison of various slope erosion mitigation measures

Measures Effectiveness Approximate comparative costs


Raises volume of earthworks
Stepped slopes High depending on distance to borrow and
spoil
Normally high but depends on source
Riprap High for embankment protection
of suitable material
Grass seeding or Only surface effective, avoids Inexpensive. May require short term
Turfing start of erosion watering to establish.
High, even in depth after several
Shrubs 2 to 3 times cost of grass
years of growth
Crib walls Good One quarter the cost of a retaining wall
High; good mechanical and
Geotextiles 10 to 20 times the cost of vegetation
chemical resistance
Gridwork and
Fairly good 5 times the cost of vegetation
wooden barricades

Riprap
The size of riprap needed to protect the stream bank and not move is related to the speed of flow as
shown in Figure D.5.13. The flow along a long tangent section of stream, or the flow parallel (VP) to the
stream, is assumed to be about 2/3, or 67%, of the average velocity (VAVE). The flow in a curved section
of stream, with an impinging flow, has an assumed impinging velocity (VI) equal to about 4/3, or 133%, of
the average velocity, VAVE. Thus, riprap in an area with relatively fast flow, such as a bend in the channel,

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 101

will have higher stresses and require larger rock than the size needed in a straight part of the channel.
Note that most of the rock should be as large as, or larger than, the size indicated in Figure D.5.13.
The Isbash Curve indicates the maximum size rock that might be considered in a critical application. If
suitably large rock is not available then the use of cement grouted rock, masonry, or gabions should be
considered.

Riprap installation details are shown in the Note in Figure D.5.11. Figures D.5.12 and D.5.13 illustrate
the use of rip rap. Ideally riprap should be placed upon a stable foundation and upon a filter layer made
either of coarse sand, gravel, or a geotextile. The riprap itself should be graded to have a range of sizes
that will minimize the voids and form a dense layer. The riprap should be placed in a layer with a thickness
that is at least 1.5 times the size (diameter) of the largest specified stone, with the thickest zone at the
base of the rock. In a stream channel, the riprap layer should cover the entire wetted channel sides, with
some freeboard, and it should be placed to a depth equal or greater than the depth of expected scour.

Figure D.5.11: Size of stone that will resist displacement for


various velocities of water flow and side slopes

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 102

Figure D.5.12: Use of riprap 1

Figure D.5.13: Use of riprap 2

Filters
A filter serves as a transitional layer of small gravel or geotextile placed between a structure, such as
riprap, and the underlying soil. Its purpose is to prevent the movement of soil behind riprap, gabions or
into under-drains, and allow groundwater to drain from the soil without building up pressure.

Traditionally, coarse sand or well-graded, free draining gravel have been used for filter materials. A sand
or gravel filter layer is typically about 150 to 300mm thick. In some applications, two filter layers may
be needed between fine soil and very large rock. Filter criteria have been developed to determine the

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 103

particle size and gradation relationships needed between the fine soil, a filter material, and coarse rock
such as riprap and are well documented elsewhere.

Today, geotextiles are commonly used to provide filter zones between materials of different size and
gradation because they are economical if manufactured locally, easy to install, and perform well with
a wide range of soils. When using geotextiles the fabric should be pulled tight across the soil area
to be protected before the rock is placed. The geotextile can be a woven monofilament or a needle
punched non-woven geotextile, but it must be permeable. The geotextile needs to have an apparent
opening size of 0.25 to 0.5 mm. In the absence of other information, a 200 g/m2 needle-punched non-
woven geotextile is commonly used for many soil filtration and separation applications. Other common
geotextile or geosynthetic material applications on roads include subgrade reinforcement to reduce
the thickness of needed aggregate over very weak soils; separation of aggregate from soft subgrade
soils; reinforcement of soils in structures such as retaining walls and reinforced fills; and entrapment of
sediment with silt fences.

If knowledgeable engineers are not available, then geotextile distributors or manufacturers should be
consulted regarding the function and appropriate types of geotextile to use in various engineering
applications. Alternatively, information is available on the requirements of different geotextiles for filter
applications in the references.

Erosion and scour protection of the road itself


Water produces harmful effects on road shoulders, slopes, drainage ditches and all the other road
structures featured in the design. Spectacular failures can occur when cuttings collapse or embankments
and bridges are carried away by flood-water. High water velocities can result in erosion which, if severe,
can lead to the road being destroyed. For each type of structure used in the road, specific erosion
problems can occur. Section 5.4.4 describes methods of minimising and controlling erosion in side
drains and other forms of drainage ditch. The specific problems associated with fords, drifts and bridge
foundations are discussed in Part E of this Manual.

It should also be noted that low flow velocities can result in silt being deposited which will, in turn, block
the various drainage structures. The blockage could result in the drainage structure being overtopped at
the next flood event. The diverted water then forges a new un-planned route which results in erosion and
possible washout in a new area. It is for this reason that the initial step in road drainage design is that of
a hydrological assessment. This will provide design discharges from all major drainage structures and for
rivers and streams adjacent to the road alignment.

5.6 Particular drainage problems in severe terrain


Much of what has been said already in previous sections is directly applicable in mountainous areas.
However, the situations in such terrain are usually much more severe. Drainage structures have to resist
scour and transmit debris every year and failure to do so will damage or destroy them more effectively
than a hydraulic surcharge. It is apparent from experience in mountainous terrain that culverts and bridges
are rarely subjected to surcharge or overtopping before the foundations are undermined by scour or the
waterway is blocked by debris, both of which can occur during the course of a single flood.

5.6.1 Drainage of hairpin stacks

The disposal of water from hairpin stacks on escarpments and steep hill sides is unquestionably a major
hazard, both for the road and for the surrounding hill slopes. It is imperative to ensure that water is
discharged into channels that are fully protected. As much effort should go into the protection of these
channels as into the protection of the road itself.

Side drain runoff at hairpin bends is often conveyed via contour drains to discharge into an adjacent
catchment, frequently causing severe erosion problems on the slopes and to the channels below. It is
preferable, therefore, to contain all water within the stack system itself, thus avoiding the construction
of drainage structures remote from the road whose inspection and maintenance might otherwise be
overlooked. This approach presents two alternatives for design. One is to construct a large reinforced
side drain around the outside of each hairpin bend, the other is to install a relief culvert beneath the

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 104

carriageway at each bend to take runoff into the inside side drain below. The main problem with the
former option is the fact that drain failure will lead to erosion of the hill slope below and eventual
undermining of the hairpin bend. Although the latter may be the preferred option, its main drawback is
one of awkward geometry. Side drain runoff is forced to make two ninety degree turns, and the reduced
gradient between the inlet and outfall restricts the size of culvert that can be utilized. There is also the
cost of these extra culverts and the required greater side drain capacity to be considered.

Where there is no choice but to discharge water onto a hill slope, the following sites should be sought in
order of preference:
ƒ Gently sloping or terraced ground;
ƒ A slope formed in strong bedrock;
ƒ A concave soil slope to assist in energy dissipation.

A standard detail for side drain turnout flow dissipation and erosion protection is illustrated in Figure
D.5.14. Most turnouts curve in plan, which throws high flows to one side, thus concentrating discharge
and increasing erosive power. The flow can be more evenly distributed by providing a flared and baffled
outlet.

Figure D.5.14: Suggested apron details for side drain turnout

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 105

5.6.2 Road construction along valley floors

Where there is a choice it is usually preferable, on hydrological and stability grounds, to adopt a hillside
rather than a valley floor alignment. However, the choice of corridor will depend upon the length and
practicality of hillside and valley floor options, and the degree of hazard posed by slope failure and
flooding along each. Furthermore, some valley floors are significantly more hazardous than others, and
it will be necessary to carefully evaluate the risk implications of these hazards before an alignment is
chosen.

Valley floor and lower valley side alignments can encounter some or all of the following landforms and
hazards:
ƒ Broad rivers that may rise and fall rapidly by several metres on a regular basis.
ƒ Rivers which are actively meandering and changing their plan-form, which could subsequently
encroach on the alignment.
ƒ Active river flood plains that are likely to flow full at least once a year. The erosive power against
the banks of a river in flood is very great.
ƒ Vigorous tributary streams that are usually highly erosive and capable of transporting large volumes
of sediment.
ƒ Fans from tributary valleys that are either eroding rapidly or building up by accumulating debris.
ƒ Flood plain terraces that may be susceptible to river scour on numerous occasions during the wet
season, and inundation once every 2-3 years.
ƒ Higher level terraces that may be subject to scour on a regular basis where they protrude onto the
active flood plain.
ƒ Rock spurs or promontories that project into the flood plain, forming obstacles to river flow and
road alignment.
ƒ Steep, and often eroded, rock slopes on the outside of valley meander bends.
ƒ Slope instability on the lower valley sides in general.

These conditions are most common on youthful valley floors, and especially those with gradients steeper
than 1 in 20. The rivers that occupy these valley floors drain steep and frequently unstable catchments.
Their flood plains will be either so confined and erosive that the development of terrace sequences has
not been possible, or will be subject to cyclic erosion and side slope instability over engineering time-
scales to an extent that any preserved terrace surfaces cannot be regarded as safe for road alignment.
In such situations, valley floor road alignments should be avoided altogether, otherwise frequent loss of
significant sections of road will be inevitable.

Where a valley floor is comparatively mature, and ancient high level terraces are well preserved, then
a road alignment located at the back of these terraces, combined with intervening rock cut, may
prove satisfactory. If valley side rock mass conditions are not especially adverse to stability, it is usually
preferable to construct a road in full cut, or a combination of cut and retained fill through these rocky
areas, with a freeboard above the highest anticipated flood level. Where valley side stability conditions
are unfavourable, or where river flooding could cause erosion and slope failure to extend far enough
upslope to undermine road foundations, it is advisable to examine the practicalities and costs of an
alternative alignment altogether.

The cost of constructing roads in major river valleys in mountainous terrain is high and largely independent
of traffic levels. The subject is dealt with in detail in ERA’s 2011 manual series.

The various design considerations associated with road construction in valley floor locations are discussed
below.

5.6.3 Freeboard

It is usual to provide the road surface and associated structures with a freeboard of 2m above design
flood level to accommodate surface waves and to provide some leeway in the estimation of flood level.
The freeboard can be reduced to 1 - 1.5m in cases where the hydraulic analysis is more reliable. However,
the calculation of the design flood is a particularly difficult task when rainfall and flow gauging data

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 106

are limited or non-existent; and where a catchment run-off regime is subject to short term fluctuations
brought about by road construction, land use change, extreme rainstorms and cycles of slope instability,
channel incision and sedimentation. Although widely appreciated, it is also important to remember that
flood levels can be substantially higher on the outside of meander bends than anywhere else along a
given reach.

5.6.4 Flood plain scour and embankment protection.

Flood plain scour, flood plain deposition and valley side instability usually occur at predictable locations.
However, external influences, such as tributary fan incursions onto the flood plain, temporary landslide
dams and engineering structures, can cause significant short-term modifications to flood plain processes
and flow patterns. These should be identified and monitored during the course of construction and
maintenance, with appropriate steps taken to protect or locally realign affected sections of road.

It can be assumed that maximum velocities around the concave (outside) banks of river bends and in
valley constrictions are between 1.5 - 2 times greater than average or calculated velocities. On highly
active flood plains with mobile bed material, predicted and actual scour depths can frequently exceed
5m, and occasionally 10m. Foundation excavations for road retaining walls and other structures are often
impracticable at these depths, given the nature of the bed material and the requirements for dewatering
the excavation.

Mortared masonry walls are more durable than gabion walls in abrasive riverside locations. and they have
the potential to arch over small areas of scour, where gabion walls are more likely to deform. Even when
heavy duty selvedge wire is used, gabion boxes are easily broken open by debris-laden water flowing at
velocities greater than 4m/s, which is not unusual.

Where there is no choice but to construct a retaining wall within the zone of highly erosive floodwaters,
it is worthwhile extending foundation excavations deeper than the depth required for bearing capacity
considerations alone, in the expectation that bedrock will be encountered, to obtain a stable foundation
for a masonry wall. Alternatively, where the foundation is composed of a significant proportion (usually
50% or greater) of large boulders, the softer materials can be excavated and replaced by concrete to
provide a stable foundation for a masonry wall.

However, it is frequently the case that neither of these foundation conditions are achievable within
practicable excavations depths, and especially on the outside of river bends where scoured bedrock
and boulders have been replaced by finer-grained materials. The potential for foundation scour in these
situations will usually dictate that a flexible gabion structure is adopted in preference to a more rigid
masonry one, and combined with whatever scour protection works are feasible under the circumstances.
Foundation stability can be improved by constructing the retaining wall on a concrete raft, thus reducing
differential settlements. Sacrificial walls, double thicknesses of gabion mesh, gabion mattresses and
stone rip-rap are likely to prove effective during small and medium-sized flood events only, and will
require regular repair or replacement. Reinforced concrete rip-rap can be fabricated in situ if sufficiently
large local stone rip-rap is unavailable or cannot be transported to the site, as is often the case. However,
the cost of fabricating rip-rap to the required dimension (3m in some cases) is usually prohibitive, and it
is usual to adopt a compromise solution under conditions of extreme scour potential.

5.6.5 Cross drainage and tributary fan crossings

Where alignments are located on the lower slopes of steep valley sides, cut slopes can truncate drainage
channels with the result that, during heavy rain, sediment and water may overshoot culvert inlets and
discharge directly onto the road surface. This is usually remedied by constructing a large catch-wall
between the culvert inlet and the road edge, or by providing a dished (concave) concrete causeway. On
occasions, concentrated run-off and slope failures will erode new gullies that will require some form of
culvert or other drainage provision. A causeway is likely to be the only practicable remedy.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 107

Tributary fans present a range of problems for road alignments. It is useful to differentiate between:
ƒ Mature and stable high level fans;
ƒ Equilibrium fans;
ƒ Immature and unstable fans.

Mature and stable fan surfaces are usually preserved as old, high level landforms that have become
incised by rejuvenated stream channels. The principal problem for road construction on high level fans,
other than alignment constraints, is the choice of a suitable site to cross the incised channel, bearing in
mind that its banks will be composed of unconsolidated and erodible materials.

In the case of equilibrium fan surfaces, all sediment supplied to the fan is transported out of the catchment
by one or a number of well-defined channels. Equilibrium fans, and their wide flood plains, will usually
comprise a number of distributary flow channels with only one or two of them occupied during normal
flow conditions. Other channels may become occupied every 2-3 years or so, in response to floods or
landslide (generated debris flows from further upstream). These fans are usually associated with terrace
sequences on the adjacent valley floor and, therefore, represent a stage of drainage development
between mature high level fans and immature, active fans on flood plains (described below).

A thorough understanding of the flow patterns across equilibrium fan surfaces is required before a road
alignment and bridging structures are designed. Artificially increased channelisation and bank protection
of the normal flow channel may increase its definition and capacity in order to allow the design of a road
crossing that consists of:
ƒ A relatively short span bridge;
ƒ Approach embankments;
ƒ Vented causeways and drifts (multi-culverted embankments to cater for other distributary channels.

The above combination will usually be cheaper to construct and represents a more convenient solution
than a multi-span bridge or a lengthy detour into the tributary to find a suitable shorter span bridging
site with stable abutment and pier foundations. However, river training and erosion protection of bridge
abutments, piers and approach embankments will prove costly and difficult to maintain in the long term.
In addition, vented causeways should only be considered where sediment loads are likely to be low.

Immature and unstable fan surfaces are usually characterised by cyclical regimes of erosion and
deposition across the whole fan surface during the course of individual storms, and a general process of
fan aggradation from one year to the next. The crossing of these fans presents severe problems to valley
floor alignments. Sediments will accumulate wherever flow velocity decreases as a result of an abrupt
concavity in the channel profile or a sudden increase in channel width. Ideally, valley floor alignments
should cross fans immediately upstream of these concavities and, if possible, at the fan apex.

Bridge clearance of at least 5 - 7m should be provided above the existing fan surface level wherever a
fan is actively aggrading and its tributary catchment is unstable. However, it is not always possible to
conform with this recommendation due to alignment constraints, unstable valley sides adjacent to the
fan apex, or the fact that the apex is poorly defined and the fan itself extends upstream into the tributary
valley. Under these circumstances, a combination of the following may be the only viable option:

ƒ Gabion check-dams in the stream channel above the fan and erosion control in the catchment
above, to control the stream bed level
ƒ River training and scour protection works upstream and downstream of the bridge, to control the
stream course
ƒ A commitment to regular maintenance and waterway clearing operations, to keep flow within the
channel and to provide room for the accumulation of debris during fan-building episodes.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 108

Design strategies for crossing unstable fan surfaces are summarised below.

1. Where rates of fan deposition are low and where the flow path across the fan is reasonably
consistent, a road can be formed on a causeway, preferably constructed in reinforced concrete
or gabion. If gabion construction is used, no wire baskets should be left exposed to abrasion by
passing rocks. They should be protected with a mortar rendering or equivalent durable surface.

2. Cross the fan via a track that is re-cut after every aggrading or eroding storm flow. This approach
will require the following considerations:
• Vehicular access must be prohibited during and for a few hours after each flood.
• As the fan surface builds up over time, temporary access will have to be cut deeper into the
fan surface and may eventually become waterlogged and impassable during the entire wet
season.
• An alternative to the above is an ever-enlarging detour downstream across the fan, eventually
coming to an end when the detour reached the flood plain at the base of the fan.
• Flood flows will tend to run down either side of the fan and erode the road on the fan
approaches.

3. Select a relatively narrow channel across or, preferably to one side of the fan surface (depending
on drainage pattern). Use river training gabions and excavation to concentrate flow through this
channel. Construct a bridge over the entire width of the fan with at least 7m clearance. This
approach will require the following considerations:
• River training gabions will tend to be scoured and undermined towards the fan apex.
• Deposition of fine-grained material towards the end of each storm may bury the river
training gabions to the extent that during the next storm a new channel will be formed and
the existing gabions may be outflanked or destroyed.
• If the bridge does not extend the full width of the fan, there is a risk that it may be outflanked
by changes in flow pattern across the fan surface leading to bridge redundancy and erosion
of the approach embankments.

4. Extending the concept of river training further, construct a continuous masonry or concrete spillway
from the fan apex to a point downstream of the bridge. The slope and cross-section of this
structure must be such that flow velocities are sufficient to transport bedload from the apex of the
fan to the flood plain downstream. This approach will require the following considerations:
• Flow along the external margins of the structure, leading to undermining, is likely to occur
unless its inlet is adequately keyed into both banks of the upstream channel.
• During peak run-off the floor of the spillway may be scoured and eventually destroyed by
passing boulders.
• During the later stages of storm run-off, the channel might still become blocked by fine-
grained sediment which will require clearing. The shape of the channel should be made to
facilitate this and access to it provided for machinery.

5. Install check dams in the gulley upstream of the fan apex to retain sediment. This approach will
require the following considerations:
• The checkdams may be destroyed during the first few storms in a channel where aggradation
of the fan itself is rapid.
• The volume of sediment that can be trapped behind a check dam system is usually
insignificant in comparison with the volume transported and transportable material.
• Artificially raising the channel bed upstream of the fan apex could easily lead to increased
rates of aggradation.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 5 - 109

5.7 Potential references and/or bibliography


American Association of State Highway and Transportation Officials (1986). Guide for the design of
pavement structures. Washington, DC.

Gourley C S and P A K Greening (1999). Performance of low volume sealed roads: results and
recommendations from studies in southern Africa. TRL Ltd, Crowthorne, UK.

Transport Research Laboratory (1993). A guide to the structural design of bitumen-surfaced roads in
tropical and subtropical countries. Overseas Road Note 31. 4th edition. TRL, Crowthorne, UK.

Transport Research Laboratory (1997). Principles of low cost engineering in mountain regions. Overseas
Road Note 16. TRL, Crowthorne, UK.

ARRB (2000). Unsealed Roads Manual, Guideline to Good Practice. Giummarra G, editor

Austroads (1989). Rural Road Design – Guide to the Geometric Design of Rural Roads,
Robinson R and B Thagesen, (editors) (1996). Highway and Traffic Engineering in Developing Countries.
E & FN Spon.Netherlands.

Cook J R and Petts R, 2005. Rural Road Gravel Assessment Programme. SEACAP 4 Module 4 Report.

Transport and Research Laboratory (1988). Overseas Road Note 5: A Guide to Project Appraisal. TRL,
Crowthorne.

Transport and Research Laboratory (1988). Overseas Road Note 6: A Guide to Geometric Design. TRL,
Crowthorne.

Thagesen B and R Robinson (editors) (1996). Highway and Traffic Engineering in Developing Countries.
E & FN Spon.Netherlands.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 110

6. MATERIALS AND PAVEMENT DESIGN

6.1 Introduction
The main objective of pavement design is to provide the best structural and economical combination of
pavement and surfacing materials, types, layer thicknesses and configurations to carry traffic satisfactorily
(ie to a pre-determined level of service with minimal maintenance) in a given climatic environment for the
design life adopted.

The load carrying capacity of the pavement is a function of both the thickness and stiffness of the materials
used in the pavement layers and the support provided by the subgrade. Consequently, a good knowledge
of the mechanical properties of the materials comprising the pavement layers and subgrade is important
for designing the structure. Climatic conditions as well as both internal and external drainage factors also
critically affect the performance of the pavement structure and must be given due consideration in the
design process.

The outcome of the design process, in terms of the type and thickness of structure chosen, is influenced
by the preceding planning phase and, in turn, determines many aspects of the subsequent construction
and maintenance phases of road provision and management. Thus, in order to achieve a successful
outcome, there is a need to ensure that the LVR design process is undertaken in a holistic manner with a
clear recognition that pavements perform as part of an overall system and must therefore be designed
as part of that system. Figure D.6.1 shows the pavement design system used as the basis of this chapter.

6.2 Underlying principles

6.2.1 Approach to low volume road pavement design

The general approach to the pavement design of LVRs differs in a number of respects from that of
HVRs. For example, conventional pavement designs are generally directed at relatively high levels of
service requiring numerous layers of selected materials. However, significant reductions in the cost of the
pavement for LVRs can be achieved by reducing the number of pavement layers and/or layer thicknesses,
by using local materials more extensively as well as at lower cost, and more appropriate surfacing options
and construction techniques.

Research has also indicated that the road deterioration mechanisms of LVRs are significantly different to
those of HVRs. One of the implications of this is that appropriate pavement design options need to be
fully responsive to a range of factors that may collectively be referred to as the road environment. This
approach pays justifiable attention to the reality of the varying environmental context of the road as
summarised in Section D.6.2.4 and discussed in detail in Sections D.6.3 onwards.

The adoption of appropriate designs for LVRs does not necessarily mean an increased risk of failure but,
rather, requires a greater degree of pavement engineering knowledge, experience and judgement and
the careful application of fundamental principles of pavement and material behaviour derived from local
or regional research.

Ultimately, the challenge of good pavement design for LVRs is to provide a pavement that is appropriate
to the road environment in which it operates and fulfils its function at minimum life cycle cost at an
optimal level of service. However, positive action in the form of timely and appropriate maintenance will
be necessary to ensure that the assumptions of the design phase hold true over the design life.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 6 - 111

Environment
(Figure D.6.3)

Subgrade Pavement
Traffic
Evaluation Materials
(Section D.6.6)
(Section D.6.5) (Section D.6.7)

Earth Gravel Paved


(Section D.6.15) (Section D.6.16) (Section D.6.17)

Surfaces
(Section D.7)

Practical Considerations
Drainage and Shoulders (Section D.6.18)
Problem Soils (Section D.6.19)
Construction Issues (Section D.6.20)

Cost Analysis

Implementation

Figure D.6.1: Pavement design system

6.2.2 Pavement structure and function

Road pavements have three primary components, namely, the wearing surface, pavement structure and
subgrade, with each serving a specific function. For paved roads the function of the surfacing is to keep
the pavement dry and waterproof. The function of the pavement structure for all road types is to support
the wheel load on the surface and to transfer and spread that load to the natural underlying subgrade
without exceeding either the strength of the subgrade or the internal strength of the pavement itself.
This implies that the pavement materials themselves should not deteriorate to such an extent as to affect
the riding quality and functionality of the pavement. These goals must be achieved throughout a specific
design period.

The function of the surfacing is slightly different for gravel and for earth roads where the wearing surface
is often permeable and actually wears away under the action of traffic and rainfall. However, the stresses
on the subgrade must be reduced to safe levels hence the gravel needs to be replaced regularly and
considerably more maintenance is necessary if gravel and earth roads are to perform satisfactorily.

Figure D.6.2 shows a wheel load, W, being transmitted to the pavement surface through the tyre at an
approximately uniform vertical pressure, P0. The pavement then spreads the wheel load to the subgrade

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 112

so that the maximum pressure on the subgrade is only P1. By proper selection of pavement materials and
with adequate pavement thickness, P1 will be small enough to be easily supported by the subgrade.

Because of the different functions of the surfacing and pavement structure, these basic components
of a road are often independent of one another and a large number of combinations are possible.
However, in terms of the design of the overall road, some surfacings (eg bituminous surface treatment
such as surface dressing) do not contribute to the overall structural strength of the road, while others
(eg penetration macadam) do. In the case of earth or gravel roads, the natural soil or gravel is the main
structural component.

Figure D.6.2: Wheel load transfer through pavement structure

6.2.3 Pavement and surfacing options

There is a wide range of pavement and surfacing options, both bituminous and non-bituminous, that
can be used in various combinations and are well suited for incorporation in LVR pavements (see Section
D 6.7). These options allow maximum use to be made of locally available materials and least use to be
made of more expensive high quality pavement materials, especially where they have to be processed
or hauled long distances.

Many of the pavement and surfacing options also allow the use of a high level of local labour, both skilled
and unskilled, and a low requirement for imported equipment. This provides the flexibility to use small
and medium scale enterprises (SMEs) with the accompanying benefits of higher local employment and
savings in costs and foreign exchange.

6.2.4 Road environmental factors

The pavement design process must be fully responsive to the Ethiopian road environment. The various
road environment factors that should be considered in the design of LVR pavements are illustrated in
Figure D.6.3 and described in the following sections.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 113

Road Safety
Climate
Regime

Maintenance Surface/sub-
Regime Surface Hydrology

Construction The Road


Regime Environment Subgrade

“Green” Terrain
Environment

Controllable Construction Uncontrollable


Traffic Materials
Factors Factors

Figure D.6.3: Road environment factors

6.3 Climate
The main features of climate that are of importance in the design of LVRs are:
ƒ Temperature and solar radiation;
ƒ Rainfall;
ƒ Water;
ƒ Winds.

Each of the above features can have a profound effect on the design, construction and maintenance
aspects of LVR roads and are considered below.

6.3.1 Temperature and solar radiation

Ethiopia’s diverse temperature patterns are largely the result of its location in the tropical zone of Africa
and its varied topography. In general, the climate comprises dry and hot zones in the eastern lowland
regions, a large moderate zone and several wet and cool zones mainly at high altitudes. Temperatures are
very varied, ranging from cool to very cold in the highlands to very hot in the lowland areas such as the
Dallol Depression. The climate is broadly divided into five zones as illustrated in Table D.6.1 and shown
in Figure D.6.5.

Table D.6.1: Ethiopia’s climatic zones

Zone Altitude (m) Temperature 0C Rainfall (mm)


Berha <500 >28 <400
Kola 500-1500 20-28 600-1000
Weina Dega 1500-2500 16-20 About 1200
Dega 2500-3200 10-16 1000-2000
Wurch >3200 <10 <800

Temperature and solar radiation both have significant implications on aspects of the design of LVR
pavements such as the availability of water for compaction purposes or the choice of binder used in
bituminous surfacings. These issues are discussed in Sections D.6.3.3, D.6.20.1 and Chapter D.7
respectively.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 114

6.3.2 Rainfall

Rainfall in Ethiopia is variable both in amount and frequency but, as indicated in Figure D.6.4, the general
pattern is for rainfall to increase with the increase in distance in a westerly direction. Average annual
rainfall is least in the east and south east of the country with a range from about 90 – 500mm and most in
the western highlands with a range of 1500 – 2000mm.

Figure D.6.4: Rainfall pattern in Ethiopia

The direct result of the relatively low precipitation in the eastern and south-eastern areas of the country is
scarcity of water. Other than in localised areas where there may be impeded drainage due to impervious,
low-lying strata, surface water and shallow water tables are rare. Ground water is usually very deep and
soaking of the subgrade does not normally occur other than in localised areas, a factor that needs to be
taken into account in correctly assessing the moisture content for assessing the strength of the pavement
layers.

In the western, higher rainfall areas of the country, surface run-off may be high leading to erosion
of shoulders and side slopes (see Plate D.6.1), increased soil erosion, flash flooding and siltation of
waterways from the disturbance of soil. Appropriate design measures must therefore be taken to combat
the potential erosive impacts of high/intense rainfall on road performance as discussed in Chapter D.5:
Drainage.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 115

Plate D.6.1: Severe erosion of road side slopes in high rainfall area

6.3.3 Moisture in the road pavement

The moisture environment in which a LVR pavement must operate has a particularly significant impact
on its performance due to the use of locally occurring, unprocessed materials which tend to be relatively
moisture sensitive. Of the sources of moisture entry into a pavement, a number can be controlled through
appropriate internal and external design measures which are discussed respectively in Section D.6.18
and Chapter D.5: Drainage.

Evaporation: The relatively low rainfall that occurs in the eastern and south-eastern areas of Ethiopia
coincides with long periods of intense sunshine that cause high evaporation rates and a net moisture
deficiency. This can affect road construction operations in the summer season in that much more water is
required for compaction purposes in a moisture-deficient environment. Various measures should therefore
be considered that minimise the evaporation of water during certain construction operations in order to
minimise loss by evaporation. Techniques are available for reducing compaction water requirements and
are discussed in Section D 6.20.1.

Climatic factor: The climatic descriptor which is used for the pavement design catalogues in this manual
is the Weinert ‘N’ value (Weinert, 1974). This index is calculated as follows:

N = 12Ej/Pa Equation 5

where:
Ej = evaporation for the warmest month
Pa = total annual precipitation

N-values less than 4 apply to a climate that is seasonally tropical and wet (the Kolla, Woina Dega, Dega
and Wurch regions of Ethiopia), whereas N-values greater than 4 apply to a climate that is arid, semi-
arid or dry (the Bereha region of Ethiopia). A map of equivalent N-values for Ethiopia is shown in Figure
D.6.5 and provides the means of placing a road in the appropriate climatic zone for design purposes as
discussed in Section D 6.17.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 116

Figure D.6.5: Climatic N-value map for Ethiopia

The climatic zones demarcated by the N-values are macro-climates and it should be kept in mind that
different micro-climates may occur within these regions. This is particularly important where such local
micro-climates can play a significant role in determining the in-situ moisture content of the various
pavement layers; a factor which needs to be carefully considered in the choice of N-Value or the subgrade
class used for design purposes.

6.3.4 Winds

Wind is a primary erosion and transportation agent in Ethiopia. Strong, hot summer winds are a common
occurrence in the eastern and south-eastern areas of the country and have the effect of shifting the
finer, generally non-plastic sands, which can cause the build-up of drifts against the edge of the road
pavement. Unprotected sand embankments can also migrate under the action of wind for which various
design counter-measure are available.

6.4 Surface/Sub-surface hydrology


The surface and sub-surface hydrological features of Ethiopia are characterised by the country’s diverse
terrain which is responsible for wide regional variations in water flow, whether it be on escarpments, in
water courses or in low-lying areas.

In the central and western areas of the country where there can be aggressive water flow within and
adjacent to road structures, appropriate internal and external drainage measures are needed to control
such water flow as discussed further in Sections D.6.18, D.6.19.4 and Chapter D.5: Drainage.

In the eastern and south-eastern areas of the country there is generally an absence of permanent surface
water, although it may accumulate in pans for a short time after rains. Groundwater is found at appreciable
depth below the ground surface and is relatively expensive to exploit. In many cases, the water tends to

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 117

be saline and may be unsuitable for use in road construction – an issue that is discussed further in Section
D.6.19.5.

6.5 Subgrade
Subgrades in Ethiopia are inherently variable and reflect the country’s diverse geology, topography, soil
type, climate and drainage conditions. As the foundation layer for the pavement, the assessment of the
subgrade condition in terms of the level of support provided to the pavement structure is one of the most
important factors, in addition to traffic loading, in determining pavement thickness design, composition
and performance. This level of support, as characterised by subgrade strength or stiffness, is dependent
on the soil type, density and moisture conditions at construction and during service. Hence, the selection
of a subgrade support value requires careful consideration of the quantity and quality of subgrade data
available to the design engineer and the variability of subgrade support within a particular project section.
The purpose of subgrade evaluation is therefore to estimate the support that the subgrade will provide
to the pavement during its design life.

This section also focuses on the classification of the subgrade in terms of the California Bearing Ratio
(CBR) to represent realistic conditions for design. In practice this means determining the CBR strength
for the wettest moisture condition likely to occur during the design life at the density expected to be
achieved in the field.

6.5.1 Subgrade Classification

Subgrades are classified on the basis of the laboratory soaked CBR tests on samples compacted to 97%
AASHTO T180 compaction. Samples are soaked for four days or until zero swell is recorded. On this
basis, the soaked CBR is used to assign a design subgrade class.

The structural catalogue given in this manual requires that the subgrade strength for design be assigned
to one of five strength classes reflecting the sensitivity of thickness design to subgrade strength. The
classes are defined in Table D.6.2.

Table D.6.2: Subgrade classes

Subgrade Class

Design CBR S2 S3 S4 S5 S6
Range % 3-4 5-8 9 - 14 15 - 29 30+

The following points should be noted with the subgrade classes defined in Table D.6.2:
ƒ No allowance for CBRs below 3% has been made because, from both a technical and economic
perspective, it would normally be inappropriate to lay a pavement on soils of such poor bearing
capacity. Moreover, the measurement of the bearing strength of such soft soils is generally most
uncertain and CBRs below 2% are of little significance. For such materials, special treatment is
required (see Section 6.19.7).
ƒ The use of Class S2 soils as direct support for the pavement should be avoided as much as possible.
Wherever practicable, such relatively poor soils should be excavated and replaced, or covered with
an improved subgrade.
ƒ Class S6 covers all subgrade materials having a soaked CBR greater than 30 and which comply with
the plasticity requirements for natural sub-base. In such cases, no sub-base is required.

Since the combination of density and moisture content wholly governs the CBR for a given material, it is
clear that changes in moisture content will alter the effective CBR in the field, and it is therefore also clear
that particular effort must be made to define the design subgrade condition.

The result of incorrect subgrade classification can have significant effects, particularly for poorer subgrade
materials with CBR values of 5 per cent and less. If the subgrade strength is seriously overestimated

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 118

(ie the support is actually weaker than assumed), there is a high likelihood of local premature failures
and unsatisfactory performance. Conversely, if the subgrade strength is underestimated (ie the support
is stronger than assumed), then the pavement structure selected will be thicker, stronger and more
expensive than necessary.

6.5.2 Specifying the design subgrade class

The CBR results obtained from the subgrade soils testing are used to determine which subgrade class
should be specified for design purposes in accordance with Table D.6.2. In some cases a variation in
results may make selection unclear. In such cases it is recommended that, firstly, the laboratory test process
is checked to ensure uniformity (to minimise inherent variation arising from, for example, inconsistent
drying out of specimens). Secondly, more samples should be tested to build up a more reliable basis for
selection.

Plotting these results as a cumulative distribution curve (S-curve), in which the y-axis is the percentage
of samples less than a given CBR value (x-axis), provides a method of determining a design CBR value
(Figure D.6.6).

The actual subgrade CBR values used for design depend on the traffic class as shown in Table D.6.3.
Thus, as indicated in the Table, for a design traffic class of LV5, the design CBR value should be the lower
10th percentile (ie the value exceeded by 90% of the CBR measurements).

Figure D.6.6: Illustration of CBR strength cumulative distribution

Table D.6.3: Design CBR values related to Traffic Classification

Traffic class Design CBR


LV5 (0.5-1.0 Mesa) Lower 10-percentile
LV3 and LV4 (0.1-0.5 Mesa) Lower 15-percentile
LV1 and LV2 (<0.1 Mesa) 30th percentile

6.5.3 Material depth

It is critical that the nominal subgrade strength is available to a reasonable depth in order that the
pavement structure performs satisfactorily. The concept of “material depth” is used to denote the depth
below the finished level of the road to which soil characteristics have a significant effect on pavement

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 119

behaviour. In addition, the moisture regime may need to be controlled by, for example, the provision of
adequate subsurface drainage and/or surface drainage. Below this depth the strength and density of
the soils are assumed to have a negligible effect on the pavement. The depth approximates the cover
required for a soil of less than 3% soaked CBR (ie less than Subgrade Class S2). However, this depth may
be insufficient in certain special cases where “problem” soils occur (See Section D.6.19).

Figure D.6.7 shows the material depth in relation to the main structural components of the road pavement,
while Table D.6.4 specifies typical material depths used for determining the design CBR of the subgrade
for the subgrade classes given in Table D.6.2.

Figure D.6.7: Material depth

Table D.6.4: Typical material depth by road category

Road Category Material Depth (mm)


DC 7 and DC 8 1,000 – 1,200
High volume roads
DC 5 and DC 6 800 – 1,000
DC 3 and DC 4 800
Low volume roads
DC 1 and DC 2 700

It should be clearly understood that the minimum depths indicated in Table D.6.4 are not depths to which
re-compaction and reworking is necessarily required. Rather, they are the depths to which the Engineer
should confirm that the nominal subgrade strength is available. In general, unnecessary working of the
subgrade should be avoided and limited to rolling prior to constructing overlying layers.

For the stronger subgrades, especially Class S4 and higher (CBR 9-14% and more) the depth check is to
ensure that there is no underlying weaker material which could lead to detrimental performance.

It is recommended that the Dynamic Cone Penetrometer (DCP) be used during construction to monitor
the uniformity of subgrade support to the recommended minimum depths given in Table D.6.4.

6.5.4 Dealing with poor subgrade soils

The cost of a road is integrally linked with subgrade conditions. The poorer and more problematic the
conditions, the greater the thickness required to support the design loads. Sometimes certain special
problems may arise in the subgrade below the material depth which requires individual treatment. Some
of the common problems which need to be considered include:
ƒ The excessive volume changes that occur in some soils as a result of moisture change (ie expansive
soils and soils with a collapsible structure);
ƒ The non-uniform support that results from wide variations in soil types over the road length;
ƒ The presence of soluble salts which, under unfavourable conditions, may migrate upwards leading
to several problems, including cracking of the surfacing;

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 120

ƒ The excessive deflection and rebound of highly resilient soils during and after the passage of a
load (eg micaceous soils).
Measures for dealing with these “problem” soils are addressed in Section D.6.19.

6.5.5 Improved subgrade layers

There are many advantages to improving the CBR strength of the in-situ subgrade to a minimum of 15%
(Subgrade Class S5) by constructing one or more improved layers where necessary. In principle, where a
sufficient thickness of improved subgrade is placed, the overall subgrade bearing strength is increased to
that of a higher class and the sub-base thickness may be reduced accordingly. This is often an economic
advantage as sub-base quality materials are generally more expensive than fill materials.

The use of improved subgrade layers also provides a number of other advantages, including:
ƒ Provision of uniform subgrade strength;
ƒ Protection of underlying earthworks;
ƒ Improved compaction of layers above subgrade level;
ƒ Provision of a more balanced pavement structure;
ƒ Provision of a running surface for the traffic during construction;
ƒ Provision of a gravel wearing course in the case of stage construction for future upgrading to a
paved road;
ƒ More economical use of pavement materials (thinner layers).

An improved subgrade placed on soils of any particular class must obviously be made of a material of a
higher class (up to Class S5, since Class S6 is of sub-base quality). The decision whether or not to consider
the use of an improved subgrade layer(s) will generally depend on the respective costs of sub-base and
improved subgrade materials.

6.6 Traffic
Determination of the amount and type of traffic is one of the most important factors in the design of
LVR pavements and will influence not only the type of surface needed (earth, gravel or bituminous) but
also the pavement thickness. The types of traffic using LVRs in Ethiopia vary significantly and include
both motorised and non-motorised traffic involving a wide spectrum of road users from pedestrians
to 3-wheeled rickshaws, motor cycle taxis to large commercial vehicles. Appropriate traffic surveys are
required to provide the information necessary for both geometric and pavement design as discussed in
Part B.

The deterioration of paved and unpaved roads caused by traffic results from both the magnitude of the
individual wheel loads and the number of times these loads are applied. It is necessary to consider not
only the total number of vehicles that will use the road but also the axle loads of these vehicles. Traffic
classes are defined for both paved and unpaved roads by ranges of the cumulative number of equivalent
standard axles (esas) (see Section D.6.6.2 below) and indicate the pavement structure requirements.

6.6.1 Estimating design traffic loading

The process by which the design traffic loading is estimated is illustrated in Figure D.6.8.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 121

Estimate Mean equivalent


Select Design Estimate initial Traffice Volume
Axle Load (esa) per Class
Period (initial AADT) per Vehicle Class
of Vehicle

Estimate Traffic Estimate Cumulative esas Over


Growth Design Period (in one direction)

Determine Cumulative Traffic Select Appropriate Traffic Class


Volumes (based on esas) for Pavemenet
over the Design Period Design

Figure D.6.8: Procedure to determine design traffic loading

In order to determine the cumulative number of vehicles over the structural design period of the road,
the following steps are required:
ƒ Determine the initial traffic volume (AADT0) using the results of a traffic survey and any other recent
traffic count information that is available. For paved roads, details of the AADT in terms of car, bus,
types of truck, and truck-trailers are required (see Table D.6.6).
ƒ Estimate the annual growth rate “i” expressed as a decimal fraction, and the anticipated number
of years “x” between the traffic survey and the opening of the road.
ƒ Determine AADT1 the traffic volume in both directions in the year of the road opening by:

AADT1 = AADT0 (1+i)x Equation 6

For paved roads, also determine the corresponding daily one-directional traffic volume for each type of
vehicle.
ƒ The cumulative number of vehicles, T over the chosen design period N (in years) is obtained by:

T=365.AADT1 [(1+i)N–1]/(i) Equation 7

For paved roads, conduct a similar calculation to determine the cumulative volume in each direction for
each type of vehicle.

Construction traffic can also be a significant proportion of total traffic on LVRs (sometimes 20 – 40% of
total traffic) and should be taken into account in the design of the pavement.

For very low volume roads (traffic <25 vpd), an elaborate traffic analysis is seldom warranted as
environmental rather than traffic loading factors generally determine the performance of roads.

6.6.2 Equivalent standard axles per vehicle class

The damage that vehicles impart on paved or unpaved road is highly dependent on the magnitude of the
axle loads of the vehicles and the number of times they are applied. Axle load data for design purposes
should preferably be obtained from surveys of commercial vehicles using the existing road or, in the case
of new roads on new alignments, from existing roads carrying similar traffic. Where this is not possible,
recourse may be made to historical information.

The damaging power of axles is related to a “standard” axle of 8.16 metric tons using empirical
equivalency factors (EFs). The number of standard axles for each vehicle is the sum of the standard axles
for each axle of the vehicle.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 122

In order to determine the cumulative axle load damage that a pavement will sustain during its design life,
it is necessary to express the total number of heavy vehicles that will use the road over this period in terms
of the cumulative number of equivalent standard axles (esas) for all the vehicles.

The relationship between an axle’s equivalency factor, EF, and the axle loading is:

EF = (P/8160)n (for loads in kg) Equation 8

or EF = (P/80)n (for loads in kN) Equation 9

Where:
EF = load equivalency factor in esas
P = axle load (in kg or kN)
n = relative damage exponent

The value of 4 or 4.5 for the exponent n is often used in line with early findings and the commonly
cited “fourth-power damage effects” of heavy axle loads. It is now clear that the value is influenced by
various factors, with the most significant being the type of materials used in the pavement structure (eg
granular/granular, granular/cemented, bituminous/cemented) and the thickness of the pavement. For
LVRs, which will normally comprise granular materials in both the base and subbase, the recommended
relative damage exponent ‘n’ is 4 (Table D.6.5).

General guidance on the likely average total equivalency factors for different vehicle types derived from
historical data in Ethiopia is given in Table D.6.6. However, data from any recent axle load surveys on the
road in question, or a similar road in the vicinity, is better than using countrywide averages.

The cumulative esas over the design period (N years) are calculated as the products of the cumulative
one-directional traffic volume (T) for each class of vehicle and the mean equivalency factor for that class
of vehicle. These are added together for all the vehicle classes for each direction. In some cases there will
be distinct differences in each direction and separate vehicle damage factors for each direction should
be derived. The higher of the two directional values should be used for design.

On narrow roads the traffic tends to be more channelised than on wider two lane roads. In such cases, the
effective traffic loading has been shown to be greater than that for a wider road. Table D.6.7 illustrates the
effect. The actual design traffic loading (esas) is calculated from Table D.6.7 using the design carriageway
widths and type of road to finalise the design values.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 123

Table D.6.5: Load equivalency factors for different axle load groups (esas)

Axle loads measured in kg Axle loads measured in kN


Axle load range (kg) EF Axle load range (kN) EF
Less than 1500 - Less than 15 -
1500 - 2499 - 15 - 24 -
2500 - 3499 0.02 25 - 34 .02
3500 - 4499 0.06 35 - 44 .06
4500 - 5499 0.15 45 - 54 .15
5500 - 6499 0.30 55 - 64 .32
6500 - 7499 0.56 65 - 74 .58
7500 - 8499 0.95 75 - 84 .99
8500 - 9499 1.5 85 - 94 1.6
9500 - 10499 2.3 95 - 104 2.4
10500 - 11499 3.3 105 - 114 3.6
11500 - 12499 4.7 115 - 124 5.0
12500 - 13499 6.5 125 - 134 6.9
13500 - 14499 8.7 135 - 144 9.3
14500 - 15499 11.5 145 - 154 12
15500 - 16499 15 155 - 164 16
16500 - 17499 19 165 - 174 20
17500 - 18499 24 175 - 184 25
18500 - 19499 30 185 - 194 32
19500 - 20499 36 195 - 204 39

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 124

Table D.6.6: Average equivalency factors for different vehicle types

No of Average esa per Average esa per


Class Type
axles vehicle - all loaded vehicle - half loaded(1)
1 Car 2 - -
2 4-wheel drive 2 - -
3 Minibus taxi 2 0.3 0.15
4 Bus/coach 2 2.0 1.0
5 Small truck/small bus 2 1.5 0.7
6 Medium truck 2 5 2.5
7 Large 2-axled truck 2 10 5
8 3-axled truck 3 12 3.5
9 4-axled truck 4 15 7.5
10 5-axled truck 5 17 8.5
11 6-axled truck 6 17 8.5
12 3-axled trailer 3 10 5
13 4-axled trailer 4 12 5
Notes:
1. It is common to find that vehicles have no back load hence half the vehicles are likely to be empty, or nearly so.

Table D.6.7: Factors for design traffic loading

Corrected design
Cross Section Paved width Explanatory notes
traffic loading (esa)
The driving pattern on
Double the sum of esas in
< 3.5m this cross-section is very
both directions
channelized.

Single Min. 3.5m but less The sum of esas in both Traffic in both directions
carriageway than 4.5m directions uses the same lane
Min. 4.5m but less 80% of the sum of esas in To allow for overlap in the
than 6m both directions centre section of the road
Total esas in the heaviest Minimal traffic overlap in the
6m or wider
loaded direction centre section of the road.
More than
one lane in 90% of the total esas in the The majority of vehicles use
each studied direction one lane in each direction.
direction

6.6.3 Design traffic classes

All survey data are subject to errors. Traffic data, in particular, can be very inaccurate and predictions about
traffic growth are also prone to large errors. Accurate calculations of cumulative traffic are therefore very
difficult to make. To minimize these errors there is no substitute for carrying out specific traffic surveys for
each project. Methods of carrying out classified traffic counts are described in ERAs 2011 Manual series.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 125

The design method recommended in this manual provides fixed structures for ranges of traffic as shown
in Table D.6.8. Therefore, as long as the estimate of cumulative equivalent standard axles is not near one
of the traffic class boundaries, traffic errors are unlikely to affect the choice of pavement design. However,
if estimates of cumulative traffic are close to the boundaries of a traffic class, then the basic traffic data
and forecasts should be re-evaluated and sensitivity analyses carried out to ensure that the choice of
traffic class is appropriate. If there is any doubt about the accuracy of the traffic estimates, it is prudent
to select the next higher traffic class for design.

Table D.6.8: Traffic classes for pavement design

Traffic class
Traffic LV1 LV2 LV3 LV4 LV5/T2(1)
range
(ESA x 106) < 0.01 0.01 – 0.1 0.1 – 0.3 0.3 – 0.5 0.5 – 1.0
Note:
1. LV5/T2 is the transition traffic zone between low-volume/high-volume roads with the former traffic class (LV5) applying
to the lower boundary of the traffic range and the latter traffic class (T2) applying to the upper boundary.

6.7 Construction Materials


The maximum use of naturally occurring unprocessed materials is a central pillar of the LVR design
philosophy. Current specifications tend to exclude the use of many naturally occurring, unprocessed
materials (natural soils, gravel-soil mixtures and gravels) in pavement layers in favour of more expensive
crushed rock, because they often do not comply with traditional (HVR-orientated) requirements. However,
recent research work has shown quite clearly that so-called “non-standard” materials2 can often be used
successfully and cost-effectively in LVR pavements provided appropriate precautions are observed as
discussed in this chapter.

The adoption of this approach provides the scope to consider a reduction in specification standard when
considering particular material types within defined environments. Recognising the material’s “fitness for
purpose” is central to assessing the appropriate use of non-standard materials. However, the use of such
materials requires a sound knowledge of their properties and behaviour in the prevailing environment.

Ethiopia’s diverse geology provides a variety of igneous, sedimentary and metamorphic rocks as well
as transported and residual soils, many of which are potentially suitable for use as road construction
materials. The selection of these materials for incorporation in a road pavement is generally based on
a combination of such factors as availability, structural requirements, environmental considerations,
method of construction, economics, previous experience and, of course, quality. These factors need to
be evaluated during the pavement design process in order to select the materials that are most suitable
for the prevailing conditions. An assessment of the durability of these materials is a key factor in the
selection process (See ERA Pavement Design Manual - 2011).

6.7.1 Performance characteristics of the pavement materials

Some understanding of the characteristics of the pavement material is necessary prior to any discussion
about the design of LVRs. Table D.6.9 summarises the typical characteristics of unbound and bound
materials that critically affect the way in which they can be incorporated into a pavement.

Category 1: materials are highly dependent on soil suction and cohesive forces for development of shear
resistance. The typical deficiency in hard, durable particles prevents reliance on inter-particle friction.
Thus, even modest levels of moisture, typically approaching 60% saturation, may be enough to reduce
confining forces sufficiently to cause distress and failure.

1 PIARC has defined non-standard and non-traditional materials as: “..any material not wholly in accordance with the
specification in use in a country or region for normal road materials but which can be used successfully either in special
conditions, made possible because of climatic characteristics or recent progress in road techniques or after having been
subject to a particular treatment.” (Brunschwig, 1989).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 126

Category 2: materials have a moderate dependency on all forms of shear resistance (ie friction, suction
forces and cohesion). These materials also have rather limited strength potential and therefore levels of
moisture, typically 60-80% of saturation, may be sufficient to reduce the strength contribution from suction
or cohesion, leading to premature distress and failure. This occurs at moisture contents lower than those
necessary to generate pore pressures.

Category 3: materials have only minor dependency on suction and cohesion forces but have a much greater
reliance on internal friction which is maximised when the aggregate is hard, durable and well graded. Very
high levels of saturation, typically 80-100% will be necessary to cause distress and this will usually result from
pore pressure effects.

Category 4: materials rely principally on physio-chemical forces which are not directly affected by water.
However, the presence of water can lead to distress under repetitive load conditions through layer
separation, erosion, pumping and breakdown.

More than anything else, the management of moisture during the construction and operational phases of
a pavement affects its eventual performance, especially when unbound, unprocessed materials are used.
Table D.6.10 shows the variation of laboratory CBR with moisture content expressed as the ratio of the field
(or in-situ) moisture content (FMC) and the optimum moisture content for compaction (OMC) (Emery, 1985).
The Table illustrates the significant effect of drying out of materials of varying quality on their strength (CBR).

As clearly indicated, if such materials can be maintained in a relatively dry state in service, then they can be
expected to perform satisfactorily at this “elevated” strength provided appropriate precautions are taken
to avoid their wetting up. Such precautions are discussed in Section D.6.18.

Table D.6.9: Pavement material categories and relative characteristics

Pavement Type
Unbound Bound
Parameter
Moderately Highly Very highly
Unprocessed
Processed processed processed
Category 1 Category 2 Category 3 Category 4
Material types
As-dug gravel Screened gravel Crushed rock Stabilised gravel
Variability High Decreases Low
Plastic Modulus High Decreases Low
Cohesion, Particle interlock
Development of Cohesion and
suction & some Particle interlock. & chemical
shear strength suction.
particle interlock. bonding.
Susceptibility to
High Decreases Low
moisture
Material strength Selection criteria reduces Material strength
Design
maintained only volume of moisture sensitive, soft and maintained even
philosophy
in a dry state. poorly graded gravels in wetter state.
Low traffic High traffic
Traffic loading increases,
Appropriate use loading in very loading in wetter
environment becomes wetter
dry environment. environments.
Cost Low Increases High High
Maintenance
High Decreases Low
requirement

Of particular significance for LVRs

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 127

Table D.6.10: Variation of CBR with moisture content

Laboratory Unsoaked CBR (%)


Laboratory at Varying FMC/OMC Ratios
Material Class
Soaked CBR (%)
1.0 0.75 0.50
G80 80 105 150 200
G65 65 95 135 185
G55 55 90 125 175
G45 45 80 115 165
G30 30 65 95 140
G25 35 60 90 135
G15 15 45 70 110
G10 10 35 60 100
G7 7 30 50 85

6.7.2 Pavement material types

The material code and outline characteristics of the material types for both paved and unpaved LVRs that
are used in the Catalogue of Designs adopted in the manual are described in Table D.6.11.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 128

Table D.6.11: Pavement material types and abbreviated nominal


specifications used in the paved and unpaved catalogue of designs

Code Material Abbreviated Specifications

Min. CBR: 80% @ 98/100% AASHTO T180 and 4 days soaking


Max. Swell: 0.2%
G80 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified.
PI: < 6 or as otherwise specified (material specific).
Min. CBR: 65% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G65 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 55% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G55 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 45% @ 98/100% AASHTO T180 and 4 days soaking
Max. Swell: 0.2%
G45 Natural gravel
Max. Size and grading: Max size 37.5mm, grading as specified
PI: < 6 or as otherwise specified (material specific)
Min. CBR: 30% @ 95/97% AASHTO T180 & highest anticipated moisture
content
G30 Natural gravel Max. Swell: 1.0% 1.5% @ 100% AASHTO T180
Max. Size and grading: Max size 63mm or 2/3 layer thickness
PI: < 12 or as otherwise specified (material specific)
Min. CBR: 30% @ 95/97% AASHTO T180 & highest anticipated moisture
content
G25 Natural gravel Max. Swell: 1.0% @ 100% AASHTO T180
Max. Size and grading: Max sixe 63mm or 2/3 layer thickness.
PI: <12 or as otherwise specified (material specific)
Min. CBR: 15% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G15 Gravel/soil Max. Swell: 1.5% @ 100% AASHTO T180
Max. Size: 2/3 of layer thickness
PI: < 12 or 3GM + 10 or as otherwise specified (material specific)
Min. CBR: 7% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G7 Gravel/soil Max. Swell: 1.5% @ 100% AASHTO T180
Max. Size: 2/3 layer thickness
PI: < 12 or 3GM + 10 or as otherwise specified (material specific)
Min. CBR: 3% @ 93/95% AASHTO T180 & highest anticipated moisture
content
G3 Gravel/soil
Max. Swell: N/A
Max. Size: 2/3 layer thickness
Note:
Two alternative minimum levels of compaction are specified. Where the higher densities can be realistically attained in
the field (from field measurements on similar materials or other established information) they should be specified by the
Engineer.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 129

6.7.3 Materials requirements for road base

A wide range of materials including lateritic, calcareous and quartzitic gravels, river gravels and other
transported and residual gravels, or granular materials resulting from weathering of rocks can be used
successfully as road base materials.

Particle size distribution: The grading envelopes to be used for road base are shown in Table D.6.12.
Envelope A varies depending whether the nominal maximum particle size is 37.5mm, 20mm or 10mm. A
requirement of five to ten per cent retained on successive sieves may be specified at higher traffic (>0.3
Mesa) to prevent excessive loss in stability. Envelope C extends the upper limit of envelope B to allow the
use of sandy materials, but its use is not permitted in wet climates.

Envelope D is similar to a gravel wearing course specification and is used for very low traffic volumes.
The grading is specified only in terms of the grading modulus (GM) and can be used in both wet and dry
climates.

Table D.6.12: Particle size distribution for natural gravel base

Per cent by mass of total aggregate passing test sieve


Test
Envelope A
Sieve
Nominal maximum particle size Envelope B Envelope C
size
37.5mm 20mm 10mm
50mm 100 100
37.5mm 80-100 100 80-100
20mm 55-95 80-100 100 55-100
10mm 40-80 55-85 60-100 40-100
5mm 30-65 30-65 45-80 30-80
2.36mm 20-50 20-50 35-75 20-70 20-100
1.18mm - - - - -
425µm 8-30 12-30 12-45 8-45 8-80
300µm - - - - -
75µm 5-20 5-20 5-20 5-20 5-30
Envelope D
1.65 < GM < 2.65

Strength and plasticity: The strength requirement varies depending on the traffic level and climate, as
outlined in the Catalogue of Structures in Part B (Figures B.5.3, B.5.4 and B.5.5. The soaked CBR test is
used to specify the minimum strength of road base material.

The plasticity requirement also varies depending on the traffic level and climate as shown in Tables
D.6.13 and D.6.14. A maximum plasticity index of 6 has been retained for higher traffic levels, where the
design chart merges to standard design documents, and also on weaker subgrades. For designs in dry
environments the plasticity modulus for each traffic and subgrade class can be increased depending on
the crown height and whether unsealed or sealed shoulders are used as described in Section D.6.17.2
and Figure D.6.22.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 130

Table D.6.13: Plasticity requirements for natural gravel road base materials

Traffic class (mesas)

Property of base
LV1 LV2 LV3 LV4 LV5

Subgrade
class4
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

Ip <12 <9 <6 <6 <6


S2 PM <400 <150 <120 <90 <90
Grading B B A(5) A(5) A(5)
Ip <15 <12 <9 <6 <6
S3 PM <550 <250 <180 <90 <90
Grading C1 B B A(5) A(5)
Ip Note(2) <12 <12 <9 <9
S4 PM <800 <320 <300 <200 <90
Grading D(3) B B B A(5)
Ip Note(2) <15 <12 <12 <9
S5 PM - <400 <350 <250 <150
Grading D(3) B B B A(5)
Ip Note(2) <15 <15 <12 <9
S6 PM - <550 <500 <300 <180
Grading D(3) C(1) B B A(5)
Notes:
1. Grading ‘C’ is not permitted in wet environments or climates (N<4); grading ‘B’ is the minimum requirement
2. Maximum Ip = 8 x GM
3. Grading ‘D’ is based on the grading modulus 1.65 < GM < 2.65
4. All base materials are natural gravels; Subgrades are non-expansive
5. Envelope A varies depending on whether the nominal maximum particle size is 37.5, 20 or 10mm

Lateritic road base gravels: Lateritic gravels occur in the west and northwestern parts of Ethiopia in the
Guraghe, Jimma, Wollega and Gojam highlands. A large number of factors control how a particular type
of laterite is developed and the material tends to exhibit both vertical and lateral variability within a deep
and irregular weathering profile.

The behaviour of lateritic materials in pavement structures depends mainly on their particle size
characteristics, the nature and strength of the gravel sized particles, the degree of compaction as well
as traffic and environmental conditions. The most important requirements for a laterite to show good
field performance are that the material is well graded with a high content of hard, or quartz particles
with adequate fines content. However, when judging the gradation of a lateritic gravel, it is important
to assess its composition to decide if separate specific gravity determinations of the fines and coarse
fractions should be made. For example, for nodular laterites, the coarse fraction is iron-rich whilst the fine
fraction is kaolinite. Thus, if there is a significant difference in the specific gravities of the coarse and fine
fractions, the grading should be calculated by use of both volume and mass proportions.

The requirements for selection and use of lateritic gravels for bases are slightly different to those given for
other natural gravels. These are presented in Table D.6.14. The maximum plasticity index of the lateritic
road base is also relaxed in comparison to Table D.6.13. A maximum plasticity index of 9 has been
specified for higher traffic levels and weak subgrades. For design traffic levels greater than 0.3 Mesa, a
requirement is set that the liquid limit should be less than 30. Below this traffic level, this requirement is
relaxed to a liquid limit of less than 35. Where sealed shoulders over one metre wide are specified in the
design, the maximum plasticity modulus may be increased by 40 per cent. A minimum field compacted
dry density of 2.0 mg/m3 is required for these materials.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 131

Table D.6.14: Guidelines for the selection of lateritic gravel road base materials

Traffic class (mesas)


LV1 LV2 LV3 LV4 LV5

Subgrade

Property
class
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

Ip <15 <12 <9 <9 <6


S2 PM <400 <150 <150 <120 <90
Grading B B A A A
Ip <18 <15 <12 <9 <6
S3 PM <550 <250 <180 <120 <90
Grading C(1) B B A A
Ip <201 <15 <15 <9 <9
S4 PM <800 <320 <300 <200 <90
Grading GM 1.6-2.6 B B B A
Ip <25(1) <18 <15 <12 <9
S5 PM - <400 <350 <250 <150
Grading GM 1.6-2.6 B B B B
Ip <25(1) <20 <18 <15 <12
S6 PM - <550 <400 <300 <180
Grading GM 1.6-2.6 B B B A
Notes:
1. Maximum Ip = 8 x GM
2. Unsealed shoulders are assumed. Further modification to the limits can be made if the shoulders are sealed.
3. The compaction requirement for the soaked CBR test to define the subgrade classes is 100% Mod. AASHTO with a
minimum soaking time of 4 days or until zero swell is recorded. This is a relaxation of the soaked CBR requirement for
natural gravel base materials given in the catalogues.

Basic igneous rock (including basaltic and doleritic gravels): These materials occur extensively in
Ethiopia and their more wide-spread use could result in significant savings provided the characteristics of
the material are good enough to serve as a road base material. However, more research work is required
before these materials can be used with confidence. Nonetheless, the following indicative limits can
contribute to successful use of the material in road bases:
ƒ Maximum secondary mineral content of 20 per cent (determined from petrographic analysis);
ƒ Maximum loss of 12 or 20 per cent after 5 cycles in the sodium or magnesium sulphate soundness
tests, respectively;
ƒ Clay index of less than 3 in the dye absorption test;
ƒ Increase in modified glycol-soaked aggregate impact value (AIV) from wet modified AIV should be
<4% units;
ƒ Durability mill index of less than 125.

In drier climatic areas (N>4), the materials can be used unmodified up to a maximum plasticity index
of 10. However, it is suggested that the materials should not be used in wet areas unless chemically
modified. The risk of using the material can be minimised if consideration is given to:
ƒ The variability of the material deposit, with good selection and control procedures in place for the
operation of the pit and on site;
ƒ The provision of good drainage conditions (these materials are particularly sensitive to moisture);
ƒ The adequacy of the pavement design (the use of Pavement Catalogue 2 with sealed shoulders is
suggested);
ƒ The use of double surface treatments or similar.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 132

Engineers need to use considerable judgement, experience and information from other roads in the area
to utilise these materials successfully. Risks must be identified and controlled.

Further information on the recommended durability specification limits for different types or road base
aggregates are given by the South African Roads Board report PR 88/032:1102 of 1990.

Cinder gravels: Volcanic cinder gravels are pyroclastic materials associated with recent volcanic activity.
Such materials occur extensively in Ethiopia, characteristically as straight-sided cone-shaped hills which
frequently have large concave depressions in their tops or sides where mixtures of solids and gases were
released during the formation of the cone. Cinders vary in colour often within the same cone and may
be red, brown, grey or black. The cinder particles also vary in size from large irregularly shaped lumps 50
cm in size, to sand silt sizes. In some cases, however, particles may be more uniform with the largest size
not exceeding 3 cm in diameter. Other characteristic features of cinders are their light weight, their rough
vesicular surface and their high porosity.

An advantage of cinders as a road construction material is the relative ease with which they can be dug
from the quarry; a mechanical shovel or hand tools are usually adequate for their extraction although
occasionally a bulldozer may be required. However, despite their extensive occurrence in Ethiopia,
cinders have been used for road construction only to a limited extent. Fortunately, however, this trend
has been reversed by research that has been carried out into the use of cinders (TRL, 1987). From full-
scale trials constructed in the late 1970s using cinders sourced from Modjo and Bekojo, it was concluded
that such gravels are capable of carrying in excess of 0.44 mesa when sealed with a surface dressing and
designed according to Road Note 31. Thus, with careful selection, cinder gravels typical of those used in
the full-scale trials can be used for road bases for LVRs with design traffic loadings up to about 0.5 Mesa.

The results of the full-scale trials also revealed that for gravel surfaced roads, mechanically stabilized
cinder gravels perform better than as-dug cinders which lack plastic binder even after the breakdown of
the cinder gravels. Recommended gradings for materials which are more resistant to corrugations are
presented in the TRL report (Newill et al, 1987).

6.7.4 Material requirements for sub-base

Strength requirements: A minimum CBR of 30% is required at the highest anticipated moisture content
when compacted to the specified field density, usually a minimum of 95% (preferably 97% where
practicable) AASHTO T180 compaction.

Under conditions of good drainage and when the water table is not near the ground surface, the field
moisture content under a sealed pavement will be equal to or less than the optimum moisture content
in the AASHTO T180 compaction test. In such conditions, the sub-base material should be tested in the
laboratory in an unsaturated state.

If the road base allows water to drain into the lower layers, as may occur with unsealed shoulders and
under conditions of poor surface maintenance where the road base is pervious, saturation of the sub-
base is likely. In these circumstances the bearing capacity should be determined on samples soaked in
water for a period of four days. The test should be conducted on samples prepared at the density and
moisture content likely to be achieved in the field.

Particle size distribution and plasticity requirements: In order to achieve the required bearing capacity,
and for uniform support to be provided to the upper pavement, limits on soil plasticity and particle size
distribution may be required. Materials which meet the recommendations of Tables D.6.15 and D.6.16
will usually be found to have adequate bearing capacity.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 133

Table D.6.15: Typical particle size distribution for sub-bases

Per cent by mass of


Sieve Size (mm) total aggregate passing
test sieve
50 100
37.5 80 – 100
20 60 – 100
5 30 – 100
1.18 17 – 75
0.3 9 – 50
0.075 5 - 25

Table D.6.16: Plasticity characteristics for granular sub-bases

Climate Liquid Limit Plasticity Index Linear Shrinkage


Moist tropical and wet tropical (N<4) < 35 <6 <3

Seasonally wet tropical (N<4) < 45 < 12 <6

Arid and semi-arid (N>4) <55 < 20 <10

6.7.5 Material requirements for gravel wearing course

Ideally, the wearing course material should be durable and of consistent quality to ensure it wears evenly.
The desirable characteristics of such a material are:
ƒ Good skid resistance;
ƒ Smooth riding characteristics;
ƒ Cohesive properties;
ƒ Resistance to ravelling and scouring;
ƒ Wet and dry stability;
ƒ Low permeability;
ƒ Load spreading ability.

For ease of construction and maintenance, a wearing course material should also be easy to grade and
compact. The material properties having the greatest influence on these characteristics are the particle
size distribution and the properties of the coarse particles.

Performance-related specifications: Performance related specifications for wearing course materials


have been developed for southern Africa based on extensive sampling, testing and monitoring of a large
number of test sections (Paige-Green, 1989). These specifications have been successfully implemented
in a number of African countries and are considered to be generally applicable to the Ethiopian
environment. The specifications identify the most suitable materials in terms of two basic soil parameters
– Shrinkage Product and Grading Coefficient – which are determined from particle size distribution and
linear shrinkage tests as shown in Figure D.6.9 and defined in Table D.6.18.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 134

Figure D.6.9: Material quality zones

The material quality zones define material quality in relation to their anticipated in-service performance.
The combination of grading coefficient and shrinkage product of each material determines which material
quality zone it falls into. The characteristics of materials in each zone are as follows:
A: Materials in this area generally perform satisfactorily but are finely graded and particularly prone
to erosion. They should be avoided if possible, especially on steep grades and sections with steep
cross-falls and super-elevations. Roads constructed from these materials require frequent periodic
labour intensive maintenance over short lengths and have high gravel losses due to erosion.

B: These materials generally lack cohesion and are highly susceptible to the formation of loose
material (ravelling) and corrugations. Regular maintenance is necessary if these materials are used
and the road roughness is to be restricted to reasonable levels.

C: Materials in this zone generally comprise fine, gap-graded gravels lacking adequate cohesion,
resulting in ravelling and the production of loose material.

D: Materials with a shrinkage product in excess of 365 tend to be slippery when wet.

E: Materials in this zone perform well in general, provided the oversize material is restricted to the
recommended limits.

Gravel loss: Gravel loss is the single most important reason why gravel roads are expensive in whole
life cost terms and often unsustainable, especially when traffic levels increase. Reducing gravel loss by
selecting better quality gravels or modifying the properties of poorer quality materials is one way of
reducing long term costs.

Based on research work carried out in Ethiopia (TRL, 2008), standardised gravel losses (gravel loss in mm/
year/100vpd) were determined in relation to the quality of the gravel wearing course (Table D.6.17).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 135

Table D.6.17: Typical standardised gravel loss

Typical gravel loss


Material Quality Zone(1) Material Quality
(mm/yr/100vpd)
Zone A Satisfactory 20
Zone B Poor 45
Zone C Poor 45
Zone D Marginal 30
Zone E Good 10
Note:
1. See Figure D.6.9.

The gravel losses shown in Table D.6.17 probably hold only for the first phase of the deterioration cycle
lasting possibly two or three years. Beyond that period, as the wearing course is reduced in thickness,
other developments, such as the formation of ruts, will also affect the loss of gravel material. However,
the rates of gravel loss given above can be used as an aid to the planning for regravelling in the future.
A more accurate indication of gravel loss for a particular section of road can be obtained from periodic
measurement of the gravel layer thickness.

Material requirements for gravel roads in rural areas: The following specifications for materials for
gravel roads in rural areas are recommended in Table D.6.18.

Table D.6.18: Recommended material specifications(1,3) for unsealed rural roads

Maximum size (mm) 37.5


Oversize index (Io)a ≤5%
Shrinkage product (Sp)b (2) 100 - 365 (max. of 240 preferable)
Grading coefficient (Gc)c (2) 16 - 34
Soaked CBR (at 95 per cent Mod AASHTO) ≥ 15 %
Treton impact value (%) (4) 20 – 65
a Io = Oversize index (percent retained on 37.5 mm sieve
b Sp = Linear shrinkage x percent passing 0.425 mm sieve
c Gc = (Percentage passing 26.5 mm - percentage passing 2.0 mm) x percentage passing
4.75 mm)/100
Notes:
1. Specifications should be applicable after placement and compaction
2. The Grading Coefficient and Shrinkage Product must be based on a conventional particle size distribution determination
which must be normalised for 100% passing the 37.5 mm screen.
3. Only representative material samples are to be tested.
4. The Treton Impact Value (TIV) limits exclude those materials that are too hard to be broken with a grid roller (TIV < 20%)
or too soft to resist excessive crushing under traffic (TIV > 65%).

Material requirements for gravel roads in ‘urban’ areas: The specifications in Table D.6.19 are
recommended for gravel roads in urban areas where there is a significant number of dwellings and local
businesses. In comparison with the limits for gravel roads in rural areas, the limits for the oversize index
have been reduced to eliminate stones whilst the shrinkage product has been reduced to a maximum
of 240 to reduce the dust as far as practically possible. This lower limit reduces the probability of having
unacceptable dust from about 70% to 40%. However, for such areas a sealed road surface is preferred.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 136

Table D.6.19: Recommended material specifications for unsealed ‘urban’ roads

Maximum size (mm) 37.5


Oversize index (Io) 0
Shrinkage product (Sp) 100 - 240
Grading coefficient (Gc) 16 - 34
Soaked CBR (at 95 per cent Mod AASHTO) ≥ 15 %
Treton impact value (%) 20 – 65

6.7.6 Material Improvement

Obtaining materials that comply with the necessary grading (particle size distribution-PSD) and plasticity
specifications for a gravel wearing course in Ethiopia can be difficult. Many of the natural gravels tend to
be coarsely graded and relatively non plastic and the use of such materials results in very high roughness
levels and high rates of gravel loss in service and, in the final analysis, very high life-cycle costs.

In order to achieve suitable wearing course properties a suitable PSD can be obtained by breaking
down oversized material to a maximum size of 50 mm or smaller. Atterberg limits may be modified by
granular/mechanical stabilisation (blending) with other materials. These material improvement measures
are discussed briefly below:

Reducing Oversize: There are various measures for reducing oversize including the use of labour, mobile
crushers, grid rollers or rock crushers. The choice of method will depend on the type of project and
material to be broken down:
ƒ Hand labour: This is quite feasible, especially on relatively small, labour-based projects where
material can either be hand screened and/or broken down to various sizes and stockpiled in
advance of construction.
ƒ Mobile crushers: The crushing of borrow pit materials may be achieved with a single stage crushing
unit or, in the other extreme, stage crushing and screening plant.
ƒ Grid rollers: These are manufactured as a heavy mesh drum designed to produce a high contact
pressure and then to allow the smaller particles resulting from the breakdown to fall clear of the
contact zone (see Plate D.6.2).
ƒ Rock crusher: The “Rockbuster” is a patented plant item which is basically a tractor-towed
hammermill. The hammermill action of the Rockbuster will act on the material that it passes over,
breaking down both large and small sizes. There is the potential to “over-crush” a material and
create too many fines in the product. It may be necessary to rill out only the larger particles in a
material and process these with the Rockbuster, with the crushed material then blended back into
the original product (see Plate D.6.3)

Plate D.6.2: Grid roller

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 137

Plate D.6.3: Rock Buster

Granular/Mechanical Stabilisation: Where materials with a suitable grading and/or plasticity are
unavailable locally, granular mechanical stabilisation may be possible by undertaking the following:
ƒ Mixing of materials from various parts of a deposit at the source of supply;
ƒ Mixing of selected, imported material with in-situ materials;
ƒ Mixing two or more selected imported natural gravels, soils and/or quarry products on-site or in a
mixing plant.

Such stabilisation can achieve the following:


ƒ Correction of grading generally associated with gap graded or high fines content gravels;
ƒ Correction of grading and increasing plasticity of dune or river-deposited sands which are often
single sized;
ƒ Correction of grading and/or plasticity in crushed quarry products;

The following methodology, using a ternary diagram (Figure D.6.10), has been developed for determining
the optimal mix ratio for blending two or more materials to meet the required specification:
ƒ Identify potential material sources that can be used to improve the available material;
ƒ Determine the particle size distribution of the available material and that considered for addition
or blending (wet sieve analysis recalculated with 100 per cent passing the 37.5 mm sieve);
ƒ Determine the percentages of silt and clay (<0.075 mm), sand (0.075 - 2.0 mm) and gravel (2.0 -
37.5 mm) for each source;
ƒ Plot the material properties on the ternary diagram as points a and b respectively (see example in
Figure D.6.10);
ƒ Connect the points. When the two points are connected, any point on the portion of the line in
the shaded area indicates a feasible mixture of the two materials. The optimum mixture should be
at point c in the centre of the shaded area;
ƒ The mix proportions are then the ratio of the line ac:bc. This can be equated to truck loads and
dump spacing;
ƒ Once the mix proportions have been established, the Atterberg Limits of the mixture should be
determined to check that the shrinkage product is within the desirable range (100 – 365 (or 240 if
necessary)). The quantity of binder added should be adjusted until the required shrinkage product
is obtained, but ensuring that the mix quantities remain within the acceptable zone;
ƒ If the line does not intersect the shaded area at any point, the two materials cannot be successfully
blended and alternative sources will have to be located, or a third source used for blending.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 138

Figure D.6.10: Ternary diagram for blending unsealed road materials

Example

Source material 1 - Grading coefficient of 20 and a shrinkage product of zero. This material plots in Zone
B of the specification and is therefore likely to corrugate and ravel.

Source material 2 - Grading coefficient of 4 and shrinkage product of 470. This material plots in Zone D
of the specification and would typically be dusty when dry and slippery when wet.

The particle size distributions and other relevant data of each material are provided below:

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 139

Material
Parameter
A B
% passing screen size (mm)
37.5 100 100
26.5 85 100
4.75 42 97
2.0 38 96
0.425 20 94
0.075 7 92
Linear shrinkage NP 5
Shrinkage product 0 470
Grading coefficient 20 4
% silt/clay (P075) 7 92
% sand (P2 - P075) 31 4
% gravel (100 - P2) 62 4

The relative proportions for each material are plotted onto the ternary diagram as points a and b which
are then connected (Figure D.6.10). The midpoint of the line within the shaded area is located at point
c. The mix proportions are thus the ratio of the line ac:ab. In this instance, the ratio is approximately 1:4,
which indicates that one part of Material B should be mixed with four parts of Material A (ie, one truck
load of Material B for every four truck loads of Material A). After blending, the grading coefficient and
shrinkage product are 18 and 138 respectively, which fall within Zone E of the specification.

6.8 Terrain
The Ethiopian topography is characterised by a blend of massive highlands, highly rugged terrain, valleys
surrounded by low-lands, steppes and semi-desert. The great diversity of terrain reflects the country’s
geological and geomorphological history which influences the types and occurrences of soils and
aggregates throughout the country. There is therefore a wide range of materials and resources available
for road construction in terms of location, type, suitability and variability.

Since much of the eastern and south-eastern areas of the country are relatively flat, most roads will tend
to be constructed on shallow embankments. This emphasises the importance of appropriate longitudinal
and cross drainage measures which are discussed in Chapter D.5: Drainage.

In the central and western regions of the country where the terrain is more mountainous and steep, the
physical environment presents a range of challenges that require careful consideration at all stages of
LVR projects, but particularly during feasibility stage of the project cycle which are considered further in
Chapter D.2: Site Investigations and Route Selection.

6.9 Construction regime


The construction regime governs whether or not the road design is applied in an appropriate manner.
Key elements include:
ƒ Appropriate labour, plant and equipment use;
ƒ Selection and placement of materials;
ƒ Quality assurance;

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 140

ƒ Compliance with the specification;


ƒ Technical supervision.

A construction strategy should be adopted that is appropriate to the prevailing social, economic, cultural
and other needs of Ethiopia. Such a strategy should be aimed at making maximum use of the relatively
abundant resource of labour through the use of labour-based technology. This approach involves using
a combination of labour and light equipment rather than heavy plant, without compromising the quality
of the end product.

6.10 Maintenance regime


All roads, designed and constructed, will require regular maintenance to ensure that the design life
is reached. Achieving this will depend on the maintenance strategies adopted, the timeliness of the
interventions, the local capacity and available funding to carry out the necessary works. Unless adequate
maintenance is provided, the anticipated pavement design life will not be attained and, indeed, the LVR
design philosophy promoted in this manual will be severely compromised.

Some types of surfacing require considerably less maintenance-than others and should hold sway in the
selection of appropriate surfacing for LVRs (See Chapter D.7: Surfacings).

6.11 Road safety regime


Some aspects of pavement structural design are affected by the road environment and can have an
impact on the safety aspects of the roads. For example, the generation of dust from some wearing course
gravels, the slipperiness of some earth surfaces and the skid resistance of bituminous surfacings . These
factors can be addressed through a surface selection procedure as discussed in Chapter D.7: Surfacings
and Part C: Complementary Interventions. However, many aspects of road safety are concerned with
geometric alignment and vehicle speeds and are addressed in Chapter D.4: Geometric Design.

6.12 The “green” environment


Any road construction and on-going road use and maintenance will have an impact on Ethiopia’s natural
or bio-physical environment including flora, fauna, hydrology, slope stability, health and safety. These
impacts have to be assessed through an Environmental Impact Assessment and mitigated as much as
possible by appropriate design and construction procedures.

6.13 Environmentally optimised design (EOD)


Invariably, the road environment factors encountered along a section of road will vary. In such a situation,
in order to be as cost-effective as possible, it is necessary to ensure that the use of materials and pavement
designs are matched to the road environment at a local level as illustrated in Figure D.6.11 .

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 141

Figure D.6.11: Application of the principle of environmentally optimised design

This can be achieved by selecting different pavement options in response to different impacting factors
along the road alignment ie by adopting what is called an environmentally optimised design (EOD)
approach. This approach economises on deployed resources to achieve an acceptable level of service
and results in the use of a spectrum of solutions incorporating spot improvement through to whole road
link, in the process using appropriate surfaces ranging from engineered natural surfaces (ENS), gravel
and, where appropriate, durable paved surfaces.

6.14 Structural design

6.14.1 Reliability and terminal condition

There are many reasons why it is impossible to predict exactly how a road will deteriorate and at what
rate. Roads built to apparently identical specifications and quality show an extremely wide range of
performances as illustrated rather emphatically in the definitive AASHO Road Test carried out in the USA
in 1960. In this road test, individual roads of similar design and quality reached their terminal level of
deterioration after anything between 30% and 200% of the average life for those sections. Thus a section
of road could last for only one third of the average life or could last twice as long. This variability is a
consequence of using relatively unprocessed and inherently variable materials and is an entirely natural
effect. Roads built to lower quality standards will exhibit even greater variability in performance.

Thus in any pavement design strategy it is necessary to specify the level of reliability required or, in other
words, the safety factor to be applied to the design. For roads carrying high levels of traffic, the safety
factor is normally set high because the poor performance of such roads has a large effect on travel times,
road roughness, vehicle operating costs and, ultimately, on the economic efficiency of the country. Thus
such roads are designed so that there is only a very small probability of not performing to the desired
standard for the desired length of time. This is defined as a high level of reliability, typically 95% or 98%.
This is equivalent to setting the planned terminal level of serviceability to a high value (ie the amount of
deterioration that is considered acceptable before the road needs improving is relatively small).

On the other hand roads carrying low levels of traffic can be allowed to deteriorate a little more before
they are deemed to need repair. They are therefore designed with a lower safety factor. The level of
reliability is set lower, typically 80% (or 50% for the lowest traffic levels). This is equivalent to designing
them with a lower level of planned terminal serviceability as shown in Figure D.6.12.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 142

The level of reliability is a statistical level of deterioration representing the probable behaviour of the
worst performing examples of road. The majority of roads will perform better and will show a level of
serviceability higher than the designed value at the end of their design period.

Although a pavement may have reached its terminal surface condition at the end of the design period, it
will still be able to carry traffic. However if, at this time, rehabilitation is not carried out, the level of service
provided will decrease quite quickly and will soon became unacceptable.

Figure D.6.12: Design reliability in relation to road category and terminal surface condition

6.14.2 Design and analysis period

The structural design period, or design life, is the period during which the road is expected to carry traffic
at a satisfactory level of service, linked to a specified level of risk or design reliability (see Section D.6.14.1),
without requiring major rehabilitation or repair work. It is implicit, however, that certain maintenance work
will be carried out throughout this period in order to achieve the expected design life.

In a country such as Ethiopia with a low density of roads and an expanding economy, many of today’s
LVRs will be tomorrow’s secondary or principal roads. Furthermore, LVRs are highly dependent on an
adequate maintenance regime if they are to perform as desired. For these reasons it is anticipated that
changes will be inevitable in a relatively short time hence the period of time for which LVRs are designed
is relatively short. In this manual a period of 10 years is used for earth and gravel roads and a period of
15 years for paved roads.

6.14.3 Upgrading Strategy

The decision as to when a road should be upgraded to a higher, more expensive structural standard
is often not a simple choice between a paved and an unpaved road. In practice, a spot improvement
strategy should be adopted as described in Section D.6.13. Thus, over a period of time, a road will
undergo a number of such improvements and will eventually consist of a mixture of structural designs.
Figure D.6.13 illustrates the various types of structure that are likely to be found on such a road. When

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 143

the traffic level is high enough a decision may be made to upgrade the entire road, or at least substantial
sections of it, to a specific structural standard. Such a decision will depend on a consideration of whole
life (or life-cycle) costs.

Increasing Cost, Maintenance Demand,


Sealed
Road

Skill Level and Equipment


Major Gravel
Road

Minor Gravel
Road

Formed
Natural Soil

Unformed
Natural Soil

Increasing demand, Traffic and Level of Service

Figure D.6.13: Components of a typical EOD designed LVR

6.15 Design of earth roads


Earth roads provide the cheapest, most basic form of access to rural communities for both non-motorised
and motorised traffic. Such roads are normally the first stage in the construction of a more durable road
and may be either “unformed” or “formed” as generally defined below:

Unformed roads: These are “non-engineered” roads that typically consist of a track that is cleared of
vegetation (see Plate D.6.4) but no significant earthworks are carried out. They are often not all weather
roads and can carry only very light traffic and then only in the dry weather or where the in-situ soils are
good (eg sand-clay or sand-silt-clay). Usually, minimum drainage is provided. Where poor soils are used,
the roads will generally be impassable in wet weather.

Formed roads: These are “engineered” roads that typically consist of the excavated in-situ material
(subgrade) in the vicinity of the alignment which is shaped to form a camber that is generally raised above
existing ground level and includes side drainage (see Plate D.6.5). When constructed with adequate
quality materials, provided with a proper camber (5-8%), adequately drained and properly maintained,
the performance is enhanced and they will normally carry higher volumes of traffic than unformed roads.
The term “engineered natural surface” (ENS) as used in this manual can also be used to describe such
roads.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 144

6.15.1 Rationale for ENS

The traffic-bearing ability of earth roads depends heavily on the type of soil forming the running surface,
and on the prevailing moisture conditions. In all but arid areas, the aim at every stage of development
should be to keep the road and its environs as dry as possible. When saturated, most soils are too
weak to carry any significant volume of traffic. However, attention to basic construction and maintenance
practice can greatly assist in extending the periods of the year during which these roads can carry traffic
satisfactorily. In all situations except arid conditions, ENSs, with the additional investment in camber and
drainage features, will usually have far superior performance to unformed roads.

Plate D.6.4: Typical unformed earth road

Plate D.6.5: Typical formed earth road

Experience shows that a well cambered and drained ENS will usually quickly dry out after rain so that
bearing capacity is rapidly restored. This suggests that Wereda and Kebele engineered earth LVRs can be
viable if communities and road users are aware of the implications and try to avoid significant trafficking
when wet. They also need to be aware of the importance of a well-maintained camber and drainage.

6.15.2 Design criteria

Although ENS roads are relatively simple structures, their analysis is quite complex because they are
subjected to rather severe environmental conditions and because the variability in their characteristics
is also very high. Their performance depends on the same factors as for all roads but the sensitivity of
performance is much greater:
ƒ Soil properties;
ƒ Rainfall (amount and intensity);

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 145

ƒ Traffic (type, volume and tyre pressure);


ƒ Longitudinal gradient, width;
ƒ Quality of drainage;
ƒ Level of water table.

As a consequence, there are no formal methods of design per se. Rather; the main focus of attention is
on the manner of their construction and maintenance which is intrinsically linked to their performance.

6.15.3 Estimating traffic capacity

Estimating the likely performance of earth roads requires an assessment of the traffic carrying capacity
of the soils under varying environmental conditions. Research undertaken on the trafficability of soils in
the United States of America (Alvin and Hammitt, 1975) provide some guidance on the traffic carrying
capacity of ENS roads from a knowledge of the bearing capacity (CBR) of the soil, the equivalent single
wheel load of the vehicles and the tyre pressures (Figure D.6.14). Thus, if the strength of the earth road
is known (in terms of its in-situ CBR), the nomograph permits predictions to be made of the load carrying
ability of the road. [The definition of the terminal condition in this study was when the rut depth in the
soil exceeded 75 mm].

Figure D.6.14: Carrying capacity of soils

As illustrated in the nomograph (Figure D.6.14), an ENS road with an in-situ CBR of 10% can be expected
to provide approximately 2,000 coverages of vehicles with a single wheel load of 10 kips (4.54 tonnes)
and a tyre pressure of 70 psi (482 kPa) before serious deformation is likely to occur. Since the wheel loads
will not be concentrated on exactly the same path, but will wander slightly across the width of a road, one
complete coverage is equivalent to the passage of 2.7 vehicles. Thus, 2,000 coverages is equivalent to
5,400 vehicles with the characteristics indicated above.

For a single lane road, the wheel loads will be restricted to narrower channels, as described in Table
D.6.7 and therefore the coverages will be different. For example, for a narrow single lane road and
using Table D.6.7, the number of vehicles that the earth road can accommodate before failure decreases

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 146

to approximately 1350 vehicles (5400/4). For a route carrying 50 vpd and assuming 15% of them are
relatively heavy (4.54 tonne wheels), this translates into a need to maintain, re-grade or reshape the
surface about every 4 to 6 months. For soils with higher CBR this will be longer. It is important for both
designers and road managers to appreciate that ENS are low initial cost but that they require an ongoing
commitment to regularly reshape/regrade the surface to keep it in a serviceable condition.

Although ENS roads can be constructed from in-situ soils with an in-service CBR of less than 15%, the
high maintenance requirements, costs, logistics and risk related to the lower strength soils mean that a
soil of CBR 15% should normally be used as the minimum target in-situ soil strength for a viable ENS. This
will require a significant proportion of sand and gravel in the natural soil. It should also be noted that in
dry weather, and when surface water can run-off quickly, the in-situ CBR is likely to be considerably higher
and the capacity of the ENS increases rapidly. Conversely, in the saturated state its capacity will be very
low.

6.15.4 Construction

ENS roads are normally constructed using the in-situ materials excavated from areas adjacent to the road
or from the side drains after any topsoil or vegetation matter has been removed from the ground surface.
The most commonly used technique is to excavate the material to form the side drains and use this to
build up the road cross section into a camber to allow rainwater to be drained off the surface into side
drains, or down the embankment slopes, and away from the road without causing erosion. This can best
be achieved using labour, motor or tractor towed graders. Side drains, turnout drains, cross drainage and
erosion control should be provided as discussed in Chapter D.5 as for all road surface types. The ditch-to-
camber earthworks technique usually provides sufficient material for an adequate cross section without
the need for expensive longitudinal haul or double handling of material. For LVR the soil can be cast in
one operation by labour from the ditch to the camber (see Plate D.6.6).

Plate D.6.6: Manual excavation of side ditch material to form


ENS camber (prior to spreading and compaction)

Guidance on the preferred method of construction is provided in Figures D.6.15, D.6.16 and D.6.17
(Howe and Hathway, 1996). In each case the road surface is 30 cm above the level of any shallow water
which may be standing alongside the road (ie on the ground in Figure D.6.15 and in the side drains in
Figure D.6.16). This 30 cm “freeboard: is necessary to enable the road to dry out a few centimetres below
the surface.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 147

Figure D.6.15: Road built above ground level – incorrect

Figure D.6.15 shows an ENS road built completely above ground level by importing soil from a borrow
site. This is expensive because of the cost of transporting the soil.

Figure D.6.16: The “high level method” - correct

Figure D.6.16 shows the “high-level” method of construction. Soil for the road is dug and thrown from
the side drain until the camber is high enough. A drain 15cm deep and a compacted road 15cm high
can provide the required 30cm free-board. In order to provide enough material, the width of the side
drain should be increased as the width of the road itself increases. Care is required to make sure enough
material is provided. Failure to do so will inevitably mean that the either the camber or crown height
(freeboard) or both will be inadequate Since the drain and camber are formed in one operation, the cost
of construction is minimised.

Figure D.6.17: Road built at ground level - incorrect

Figure D.6.17 shows what would happen if the road were built at ground level. Soil has to be excavated
to form the camber and side drains and dumped uselessly in the bush. Because the side drains are well
below ground level, it is difficult or impossible to run water from them to the surrounding ground.

Standard cross-sections for earth roads are shown in Chapter D.4: Geometric Design.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 148

6.15.5 Maintenance

Unformed roads: Particular attention must be paid to the problem of erosion in the wheel tracks and
side drains (see Figure D.6.18(a) and D.6.18(b)) for which efficient control and disposal of run-off water
is critical. Where the topography allows, wide, shallow longitudinal drains are preferred. They minimise
erosion, and will not block as easily as narrow ditches.

Figure D.6.18: (a) – Erosion in wheel tracks

Figure D.6.18: (b) – Erosion in side drains

Any unformed road will fail, probably sooner rather than later, if avoidable erosion due to water flow is
allowed to continue unabated. Each rainy season the level of the road will subside due to removal of soil.
The more it subsides, the more difficult it will be to drain the water off the road. Thus, erosion damage,
though imperceptible at first, is liable to increase rapidly within a few years. However, such avoidable
erosion may be easily prevented by adopting a variety of soil conservation and construction techniques
that are fully addressed in Earth Roads – Their Construction and Maintenance (Howe and Hathway, 1996).

ENS roads: Under traffic, ENS roads will become rutted and reshaping will be required. This should
consist of blading soil inwards from the outside of the road edges and serves to raise the road bed,
provide a cambered surface and initiate a drainage system (Figure D.6.19b). The simpler expedient of
digging out the rutted soil and throwing it to the edges of the road to expose a fresh soil surface will
result in a sunken road prone to water logging and being impassable to traffic in the rainy season. The
establishment of grass up to the road edges assists in preventing erosion. Simple turnout drains should
also be opened to discharge water collected from the road way.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 149

Figure D.6.19: (a) - Incorrectly maintained ENS road (Odier et al, 1971)

Figure D.6.19: (b) - Correctly maintained ENS road (Odier et al, 1971)

6.16 Design of gravel roads


A gravel road generally serves as the first stage in the making of an all-weather road designed to particular
standards of alignment and traffic carrying capacity. The gravel surface provides an improved and more
durable surface than earth.

The performance of a gravel surfaced road depends on the quality of the materials, the location of the
road (terrain and rainfall), and the traffic volume using the road. Where good quality in-situ road building
materials occur, they can provide a strong enough pavement structure to carry the expected traffic for
many years with no additional structural layers being required, but suitable drainage must be provided
and they must be properly shaped and compacted.

Generally, a gravel road consists of a wearing course and a structural layer (base) which covers the in-situ
material (Figure D.6.20).

Figure D.6.20: Gravel road - Typical pavement layers

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 150

The purpose of the structural layer is to protect the sub-grade from excessive stresses imposed by traffic
and thereby avoid unacceptable deformation within a specific design period. However, because the
stresses on the sub-grade increase as gravel is worn away, it is necessary to ensure that a minimum
thickness of the structural layer is maintained in service by providing a protective layer – the wearing
course – throughout the design life of the road.

To reduce the adverse impact of rainfall and water, the road must be constructed with an appropriate
camber (typically 4-6%) to effectively shed surface water. To achieve adequate external drainage, the
road must also be raised above the level of existing ground such that the crown of the road is maintained
at a minimum height (hmin) above the table drain inverts as shown in Figure D.6.21.

Figure D.6.21: Typical gravel road cross section in flat terrain.

The minimum height is dependent on the climate and road design class as shown in Table D.6.20. Unless
the existing road is well below existing ground level, this can usually be achieved by proper “forming”,
ie shaping of the road bed to ensure adequate road levels, coupled with the cutting of table drains to an
appropriate depth below existing ground level. Where necessary, additional fill will have to be imported
or obtained from shallow cuttings to achieve the required hmin.

Table D.6.20: Required minimum height (hmin) between road


crown and invert level of drain in relation to climate

Climate
Road Class Wet (N < 4) Dry (N > 4)
hmin (mm) hmin (mm)
DC-1 350 250
DC-2 400 300
DC-3 450 350
DC-4 500 400

6.16.1 Design method

Several methods have been developed internationally for designing gravel roads but most are based
solely on traffic volume and do not take into consideration the load characteristics of traffic. In Ethiopia
the heavy vehicle component on the higher categories of gravel roads can often be in excess of 20 per
cent. For this reason the design method adopted in this manual is based on the South African TRH20
manual “Unsealed roads: Design, construction and maintenance”. Ver. 1.4 (November 2008)’ which takes
both traffic volume and loading into account in the design process.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 151

The design catalogue in the TRH20 manual was developed from the DCP design method (Kleyn and van
Zyl, 1988) and was calibrated with Heavy Vehicle Simulator (HVS) testing carried out in South Africa. It
assumes a failure criterion of 25mm deformation of the subgrade.

Gravel roads are divided into two broad categories for design purposes namely ‘major’ and ‘minor’ gravel
roads:

6.16.2 Major and minor gravel roads.

A major gravel road is one that is very likely to be upgraded to a higher standard in the foreseeable
future with a surfacing that does not wear away as quickly as gravel, for example, a bituminous surfacing.
For major gravel roads, specific engineering features should be included at the time the gravel road is
first constructed. A major gravel road will normally have a design AADT between 75 and 300 and will
therefore fall into road class DC-3 or DC-4.

Ideally, the same engineered design approach can also be applied to gravel roads of classes DC-1 and
DC-2 but, for these classes, referred to as “minor” gravel roads, a simplified approach is presented in
Section 6.16.8.

The approach to the design of class DC-3 and DC-4 roads is as follows:
ƒ The sub-grade should be prepared in the same way as for a low volume sealed road to comply with
the requirements described in Section D.6.5.
ƒ It is assumed that the wearing course will be replaced at intervals related to the expected annual
gravel loss and before the structural layer is exposed to traffic and itself begins to wear away;
ƒ The geometry and drainage are upgraded to acceptable minimum levels during construction.
This may require the introduction of a fill layer between the compacted in-situ sub-grade and the
wearing course.
ƒ It should be noted that gravel roads in classes DC3 and DC4 are likely to incur extremely high
maintenance costs in some circumstances namely;
ƒ When the quality of the gravel is relatively poor (Section D.6.5).
ƒ Where no sources of gravel are available within a reasonable haul distance.
ƒ On road gradients greater than about 6%.
ƒ In areas of high and intense rainfall.

In these circumstances spot improvements will almost certainly be justified, as outlined in Section D.6.16
and, in some cases, it may prove to be more economical to build a fully paved road at the outset.

6.16.3 The structural design procedure

1. The first step is to determine the traffic volume and traffic loading. This step is similar for all roads and
is described in detail in Section D.6.6.
2. The second step is to determine the strength of the sub-grade at the appropriate moisture condition.
This is slightly different for an unsurfaced road and is described below.
3. Step 3 requires measurements of the quality of the gravel that is to be used. If only very poor gravel
is available, blending with another gravel or soil to improve its properties may be an option (Section
D.6.7.6).
4. Step 4 requires the thickness of gravel base that is necessary to avoid excessive compressive stresses
in the sub-grade to be determined. This depends on the information obtained in steps 1 to 3 but
the necessary calculations have already been carried out based on research in South Africa and the
results are presented in Tables D.6.21 (a), (b) and (c).
5. Finally the thickness of the wearing course needs to be calculated based on the expected rate of
gravel loss and a realistic choice of the frequency of re-gravelling. Estimating the annual gravel loss
is discussed in Section 6.16.7. If the annual gravel loss is expected to be GL and the road is likely to
be re-gravelled every R years, the gravel loss that occurs between re-gravelling operations will be R
x GL and therefore this depth of gravel needs to be provided for the wearing course.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 152

6.16.4 Moisture regime and determination of material strength

The determination of the material strength in the in-situ sub-grade and sub-base is related to the climate,
local moisture conditions and the elevation of the road above natural ground level.

In an arid, semi-arid or dry climate (the Bereha and Kolla regions of Ethiopia where the Weinert N-Value is
greater than 4) and under conditions of good drainage (Table D.6.20) with the water table not near to the
ground surface, both the in-situ sub-grade and sub-base layer strengths should assessed at the optimum
moisture content (OMC) for compaction at the appropriate compaction level.

In contrast, in a seasonally tropical or wet climate (eg the Weina Dega, Dega and Wurch regions in
Ethiopia, where the Weinert N-Value is less than 4) the in-situ sub-grade strength should be assessed in
the soaked condition. The design CBR should be based on the 50th percentile (See Table D.6.3).

6.16.5 Material classification

Test pits should be excavated as a part of a conventional centre-line survey to a depth of at least 500
mm to determine the existing pavement profile. Sampling and indicator testing (Atterberg limits, grading
and CBR) of selected or combined samples should be undertaken in each soil horizon. The upper 150
– 200 mm layer should be classified for the in-situ sub-grade in accordance with the sub-grade classes
presented in Table D.6.2. The gravel materials should be classified in accordance with the materials
classes presented in Table D.6.11.

6.16.6 Selection of appropriate pavement structure

Based on the information obtained in Steps 1 to 4, the road base thickness is selected from Tables
D.6.21a, D.6.2.1b or D.6.2.1c.

Table D.6.21a: Catalogue for major gravel roads – strong gravel (G45)

Subgrade Strength Traffic Classes (esa x 106)


Class
CBR (%) <0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

S2 (3-4) 175 225 250 300 350


S3 (5-7) 150 200 225 250 300
S4 (8-14) 100 150 200 200 250
S5 (15-29) 100 125 150 175 200

Table D.6.21b: Catalogue for major gravel roads – medium gravel (G30)

Subgrade Strength Traffic Classes (esa x 106)


Class
CBR (%) <0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

S2 (3-4) 175 250 290 325 370


S3 (5-7) 150 200 250 275 325
S4 (8-14) 125 175 200 220 275
S5 (15-29) 100 100 150 175 200

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 153

Table D.6.21c: Catalogue for major gravel roads – weak gravel (G15)

Subgrade Strength Traffic Classes (esa x 106)


Class
CBR (%) <0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.0

S2 (3-4) 225 325 375 NA NA


S3 (5-7) 200 250 325 350 NA
S3 (5-7) 200 250 325 350 NA
S5 (15-29) 150(1) 150(1) 200(1) 200(1) NA
Note:
1. This is the additional depth of compacted sub-grade material

The following points should be noted in connection with the selection of gravel for the structural base
layer:
ƒ The thicknesses required increases considerably if the gravel is weak hence stronger gravels should
generally be used if they are available at reasonable cost.
ƒ On relatively weak subgrades (S2 and S3), the use of strong gravels (G45) should be avoided
because of the poor “balance” of such pavements. Instead, the use of an improved subgrade layer
should be considered for the advantages provided (Section D.6.5.5).
ƒ Where the available gravel is not homogeneous, it will be necessary to substitute a particular
class of gravel with one or more different classes of gravel of appropriate thickness. The following
conversion factors may be used for this purpose (Emery, 1985).

G45 = 1.5 x G15


G30 = 1.2 x G15

Thus, a 200mm layer of G45 material could be substituted with a 300mm layer of G 15 material.
ƒ (4) For effective compaction of the gravel layer, it is necessary to restrict the loose thickness of
gravel to a maximum lift of about 200mm. Thus, any of the gravel layers that require a compacted
thickness of more than 150mm will have to be compacted in more than one 200mm lift.

6.16.7 Determination of wearing course thickness

The desired wearing course thickness depends on the annual gravel loss and the number of years between
re-gravelling operations.

Predicted gravel loss (GL).


The interaction between traffic and rainfall contributes significantly to the loss of material from a gravel-
surfaced road. Based on research work carried out in Ethiopia (TRL, 2008), standardised gravel losses
(gravel loss in mm/year/100vpd) were determined in relation to the quality of the gravel wearing course
as shown in Table D.6.22:

Table D.6.22: Typical standardised gravel loss

Material Quality Zone1 Description of Material Quality Typical gravel loss (mm/yr/100vpd)
Zone A Satisfactory 20
Zone B Poor 45
Zone C Poor 45
Zone D Marginal 30
Zone E Good 10
Note:
1. See Figure D.6.9.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 154

These rates of gravel loss increase quite significantly on gradients greater than about 6% and also in areas
of high and intense rainfall. On gradients, the increase could be greater than 50% depending on the
steepness of the gradient and material quality and therefore spot improvements should be considered.

The rates of gravel loss given here can be used as an aid to the planning for re-gravelling in the future.
A more accurate indication of gravel loss for a particular section of road can be obtained from periodic
measurement of the gravel layer thickness.

Re-gravelling frequency
Re-gravelling should take place before the sub-base is exposed in order to avoid significant deformation
which will necessitate reconstruction and loss of the strength that has been built up in the sub-grade by
traffic moulding over time. Where the initial gravel road has been properly designed and constructed
with appropriate quality materials, and also has adequate drainage, the re-gravelling frequency, R, will
be approximately as assumed (see Section 6.16.3) which is typically in the range 5 – 8 years. This can
decrease considerably if poor quality gravels have to be used. For example, if the gravel loss rate is
45mm per year per 100vpd, a class DC4 gravel road carrying 200vpd will lose 90mm per year and require
re-gravelling every two years at the most. It would be surprising if an economic analysis did not show that
such a road should be fully paved.

Wearing course thickness


This is determined from the product of the annual gravel loss, GL, and the re-gravelling frequency, R, as
discussed above.

6.16.8 Minor gravel roads

A minor gravel road is one which is unlikely, in the foreseeable future, to be upgraded to a bituminous
standard. This applies to roads which have a design AADT typically less than 50 and will fall into classes
DC-1 or DC-2. Where budgetary constraints or other reasons do not allow the construction of these
roads to an engineered gravel standard as described above, lesser gravel sub-base and wearing course
thicknesses may be used. However, a relatively lower level of service should be expected, coupled with
a greater risk of shear failures in the sub-grade.

The approach to the design of these categories of roads is as follows:


ƒ The catalogue is based on the AADT of the road and assumes approximately 30% commercial
vehicles;
ƒ The subgrade materials should not necessarily comply with the requirements of a low volume
sealed road;
ƒ A nominal wearing course thickness of 150mm of G15 is assumed for all road classes and sub-
grade conditions with the sub-base thickness being influenced by the sub-grade class;
ƒ Drainage, but not necessarily geometry, is upgraded to acceptable minimum levels during
construction. As for Class DC-3 and DC-4 roads, this can be achieved by building up the formation
to an appropriate height to achieve the hmin requirements given in Table D.6.20.

Based on the above approach, the recommended sub-base thicknesses and wearing course material
strengths for different sub-grade and traffic conditions are presented in the design chart in Table D.6.23.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 155

Table D.6.23: Typical standardised gravel loss

Traffic Classes (AADT)


Subgrade Strength
Class CBR (%) DC1/DC2(1)
(< 75)
150 WC
S2 (3-4)
200 G15(2)
S3 & S4 (5-14) 150 WC
S5 (15-29) Earth Road
Notes:
1. If more than 10 heavy vehicles per day, design as a major gravel road
2. If a G30 material is available the thickness can be reduced to 150 mm

6.17 Structural design of paved roads


The structure of a paved road consists typically of one or more layers of material with different strength
characteristics (Figure D.6.22), each layer serving the purpose of distributing the load it receives at the
top over a wider area at the bottom. The layers in the upper part of the structure are subjected to higher
stress levels than those lower down and therefore need to be constructed from stronger material. The
surfacing may be either structural or non-structural in terms of its contribution to the overall strength of
the road pavement.

Figure D.6.22: Paved road - Typical pavement layers

6.17.1 Design methods

There are a number of methods that have been developed for the design of flexible paved roads ranging
from the simple to the complex and based on both mechanistic/analytical and empirical methods. The
purely empirical design methods are limited in their application to conditions similar to those for which
they were developed whilst the mechanistic/analytical methods require a considerable amount of material
testing and computational effort and their application to highly variable, naturally occurring materials
which make up the bulk of LVR pavements is questionable.

The pavement design method used here is an empirically-based design method.

The DCP method is useful where a basic or more developed pavement structure is already in place and
needs to be enhanced or upgraded.

6.17.2 Design method for Bituminous surfaced roads

Design charts or catalogue methods are the easiest to use because all the practical and theoretical works
have been carried out and different structures are presented in chart form for various combinations of
traffic, environmental effects, pavement materials and design options.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 156

The design method presented in this manual is based on research undertaken in a number of countries
in southern Africa (TRL, 1999). It differs from the traditionally accepted design criteria applied to the
design of heavily trafficked roads in that it recognises the controlling influence of the road environment
on the deterioration of lighter pavement structures. By incorporating a recognised climatic variable, the
N-value (Section D.6.3.3), the geographical transferability of the research findings can be undertaken with
confidence in Ethiopia.

The LVR design process for bituminous surfaced roads is outlined in the flow chart presented in Figure
D.6.23. This process indicates the sequence of steps that are required to produce a pavement design that
is appropriate and adequate for an individual road.

Subgrade Evaluation
Section D.6.5

Traffic
Section D.6.6

Pavement materials
Section D.6.7

Geo-climatic Factor Geo-climatic Factor


N<4 N4

Sealed Width 8m (or Sealed Width 8m (or


Sealed Width Sealed Width
7m on embankments 7m on embankments
<7m <7m
>1.2m in height) >1.2m in height)

Design Chart 1 Design Chart 2 Design Chart 1 Design Chart 2

Materials relaxation Materials relaxation Materials relaxation


Materials relaxation Increase limits on
Increase limit on Increase limit on
None allowed PM by 40%
PM by 20% PM by 40%
PI by 3 units

Figure D.6.23: Flow chart for bituminous surfaced road pavement design process

Climate: The design method utilises two design charts each applicable to a different climatic zone
characterised by the Weinert N-values (Section D.6.3.3) and the shoulder and drainage design adopted.

An N-value map for Ethiopia is shown in Section 6.3.3 and Figure D.6.5 and provides the means of placing
the road in the appropriate climatic zone for design purposes. The two design charts are presented in
Part B: Design Standards, and offer a total of sixty different pavement structures depending on traffic and
subgrade class.

N-values less than 4 imply a climate that is seasonally tropical and wet (the Weina Dega, Dega and Wurch
regions in Ethiopia), whereas N-values of greater than 4 imply a climate that is arid, semi-arid or dry (the
Bereha and Kolla regions of Ethiopia).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 157

Traffic and environmental effects: For a correctly constructed pavement carrying low levels of traffic,
there is a low risk of a pavement failure being induced by traffic, and deterioration is controlled mainly
by environmental factors. However, as the traffic levels increase, the specification for road bases should
approach those of traffic design charts for high volume roads presented in ORN 31 (TRL, 1997). Experience
suggests that the transition from low-volume to high-volume roads is typically in the 1.0 Mesa range (see
Figure D.6.24).

Figure D.6.24: Traffic loading versus dominant mechanism of pavement distress (Schematic only)

Sealed width: When the total sealed width is 7 metres or less, the outer wheel-track is within one metre
of the edge of the seal. This affects pavement performance adversely because of seasonal moisture
ingress. Therefore, relatively stronger pavements are necessary in these situations. If the road width is
sufficient for the outer wheel to be more than 1.5 metres from the pavement edge, and good drainage
is ensured by maintaining the crown height at least 750mm above the ditch invert, an improvement in
performance occurs.

This is reflected in the catalogues where different sealed surface widths are treated separately. Thus a
wider sealed cross-section in climatic zones where N<4 (a relatively wet environment) allows a shift from
Catalogue 1 (N<4) to Catalogue 2 (N>4). This allows the use of thinner pavement layers and a relaxation
of the quality requirements for the base.

When a road is on an embankment of more than 1.2 m in height, the material in the road base and sub-
base stays relatively dry, even in the wet season. In this case, the design category can be relaxed, and a
pavement with a 7 m total sealed width can be designed to the same criteria as for an 8 m seal.

6.17.3 Use of the design charts

For guidance, the following design options are used in the catalogues related to the design traffic class
shown in Table D.6.24:

Table D.6.24: Design traffic classes

Design traffic classes


Traffic Classes
<0.01 0.01-0.1 0.1-0.3 0.3-0.5 0.5-1.
(ESA x 106)

Climatic zones N < 4


(a) Where the total sealed surface is 8 m or less, use Pavement Design Chart 1 (Table B.6.6). No road
base materials adjustments are allowed.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 158

(b) Where the total sealed surface is 8 m or more, use Pavement Design Chart 2 (Table B.6.7). The limit
on the plasticity modulus of the road base may be increased by 20 per cent.
(c) Where the total sealed surface is less than 8m but the pavement is on an embankment in excess of
1.2 m in height, use Pavement Design Chart 2 (Table B.6.7). The limit on the plasticity modulus of the
road base may be increased by 20 per cent.
(d) If the engineer deems that other risk factors (eg poor maintenance and/or construction quality) are
too high, then Pavement Design Chart 1 should be used.

Climatic zones N > 4


Use Pavement Design Chart 2 (Table B.6.7).
(a) Where the total sealed surface is less than 8 metres, the limit on the plasticity modulus of the road
base may be increased by 40%.
(b) Where the total sealed surface is over 8 metres and when the pavement is on an embankment in
excess of 1.2 metres in height, the plasticity modulus of the road base may be increased by up to 40%
and the plasticity index by 3 units.

The design flow chart in Figure D.6.23 should be used iteratively depending on conditions on the
individual project as in the following example:

Once the quality of the available materials and haul distances are known, the flow chart and the design
charts can be used to review the most economical cross-section and pavement; this involves assessment
of design traffic class, design period, cross-section and other environmental and design considerations. It
may be more economical to use a wider cross-section in the seasonal tropical and wet climate zone, and
then shift to Design Chart 2 than to design a narrow cross-section and a pavement using Design Chart
1, however the minimum width of carriageway and shoulders is controlled by the geometric standards
adopted and this depends on traffic volume and composition.

Reducing Risks in Special Cases: When the project is located close to the border between the two
climatic zones, the lower N-value should be used to reduce risks. When close to the borderline between
two traffic design classes, and in the absence of more eliable data, the next highest design class should
be used.

6.17.4 Design method for non-bituminous surfaced road

The design charts for non bituminous surfaced roads are given in Part B, Section 6.4.2, and as with the
design charts for bituminous surfaced roads, are empirically-based.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 159

6.18 Drainage and Shoulders


Moisture is the single most important factor affecting pavement performance and long-term maintenance
costs. Thus one of the significant challenges faced by the designer is to provide a pavement structure in
which the detrimental effects of moisture are contained to acceptable limits in relation to the traffic loading,
nature of the materials being used, construction/maintenance provisions and degree of acceptable risk.
This challenge is accentuated by the fact that most low volume roads will be constructed from natural,
often unprocessed, materials which tend to be moisture sensitive. This places extra emphasis on drainage
and moisture control for achieving satisfactory pavement life.

Two inter-related aspects of drainage need to be considered during road design, namely internal and
external drainage. This section focuses on internal drainage only which is concerned with water that
enters the road structure directly from above the road pavement or directly from below and the measures
that can be adopted to avoid trapping water within the pavement structure. External drainage which
seeks to control water before it enters the pavement structure is discussed in Chapter D.5: Drainage.

6.18.1 Sources of Moisture Entry into a Pavement

The various causes of water ingress to, and egress from, a pavement are listed in Table D.6.25 and
discussed in this Section.

Table D.6.25: Typical causes of water ingress to, and egress from a road pavement

Means of Water Ingress Causes

Through the pavement through cracks due to pavement failure


surface penetration through intact layers
artesian head in the subgrade
From the subgrade pumping action at formation level
capillary action in the subbase
seepage from higher ground, particularly in cuttings
reverse falls at formation level
From the road margins
lateral/median drain surcharging
capillary action in the subbase
through an unsealed shoulder collecting pavement and ground run-off
Through hydrogenesis condensation and collection of water from vapour phase onto underside
(aerial well effect) of an impermeable surface
Means of Water Egress Causes
Through the pavement
through cracks under pumping action through the intact surfacing
surface
soakaway action
Into the subgrade
subgrade suction
into lateral/median drains under gravitational flow in the subbase
To the road margins
into positive drains through cross-drains acting as collectors

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 160

6.18.2 Permeability

Permeability is a measure of the ease with which water passes through a material and is one of the key
material parameters affecting drainage. Moisture ingress to, or egress from, a pavement will be influenced
by the permeability of the pavement, subgrade and surrounding materials. The relative permeability
of adjacent materials may also govern moisture conditions. A significant decrease in permeability with
depth or across boundaries between materials (ie permeability inversion) can lead to saturation of the
materials in the vicinity of the inversion. Typical permeability values for saturated soils are presented in
Table D.6.26.
Table D.6.26: Typical material permeabilities (Lay, 1998)

Material Permeability Description


Gap-graded crushed rock > 30 mm/s
Gravel > 10 mm/s Free draining
Coarse sand > 1 mm/s
Medium sand 1 mm/s
Permeable
Fine sand 10 µm/s
Sandy loam 1 µm/s
Practically impermeable
Silt 100 nm/s
Clay 10 nm/s
Impermeable
Bituminous surfacing(1) 1 nm/s
Note:
(1) Applies to well-maintained double chip seal. Thicker asphalt layers can exhibit significant permeability as a result of a
linking of air voids. Permeability increases as the void content of the mix increases, with typical values ranging from 300
µm/s at 2% air voids to 30 µm/s at 12% air voids. Typically, a 1% increase in air voids content will result in a three-fold
increase in permeability (Waters, 1982).

6.18.3 Achieving effective internal drainage

The following guidance is provided for achieving effective internal drainage of the road structure.

Side drainage and crown height above drain invert. Side drainage is one of the most significant factors
affecting pavement performance. The “drainage factor” is the product of the height of the crown of the
road above the bottom of the ditch (h) and the horizontal distance from the centreline of the road to the
bottom of the ditch (d) and can be used to classify the type of drainage prevailing at the road site. This
classification of road drainage is shown in Table D.6.27.

Implied in the Table is the critical nature of the crown height which correlates well with the actual service
life of pavements constructed from natural gravels. A minimum value, h, of 0.75m is recommended as
illustrated in Plate D.6.7.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 161

Table D.6.27: Classification of road drainage

Drainage Factor
Classification
DF = d x h
<2.5 Very poor
2.6 – 5.0 Poor
5.1 – 7.5 Moderate
> 7.5 or free draining Good
Note:
1. Classification can move up one class if longitudinal gradient >1%

Irrespective of climatic region, if the site has effective side drains and adequate crown height, then the
in-situ subgrade strength stays above the design value. If the drainage is poor, the in-situ strengths will
fall to below the design value.

Plate D.6.7: Example of a well-drained pavement where the drainage is


classified as “good”, ie the drainage factor DF (d x h) >7.5

Drainage within pavement layers: Drainage within the pavement layers themselves is an essential
element of structural design because the strength of the subgrade in service depends critically on the
moisture content during the most likely adverse conditions. Since it is impossible to guarantee that road
surfaces will remain waterproof throughout their lives, it is critical to ensure that water is able to drain
away quickly from within the pavement. This can be achieved by a number of measures as follows:

Avoiding permeability inversion: A permeability inversion exists when the permeability of the pavement
and subgrade layers decreases with depth. Under infiltration of rainwater, there is potential for moisture
accumulation at the interface of the layers. The creation of a perched water table could lead to shoulder
saturation and rapid lateral wetting under the seal may occur. This may lead to base or sub-base saturation
in the outer wheeltrack and result in catastrophic failure of the base layer when trafficked. A permeability
inversion often occurs at the interface between sub-base and subgrade since many subgrades are
cohesive fine-grained materials. Under these circumstances, a more conservative design approach is
required that specifically caters for these conditions.

In view of the above, it is desirable for good internal drainage that permeability inversion does not occur.
This is achieved by ensuring that the permeability of the pavement and subgrade layers are at least equal
or are increasing with depth. For example, the permeability of the base must be less than or equal to the
permeability of the sub-base in a three layered system.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 162

Where permeability inversion is unavoidable, the road shoulder should be sealed to an appropriate width
to ensure that the lateral wetting front does not extend under the outer wheeltrack of the pavement.

Ensuring proper shoulder design: When permeable roadbase materials are used, particular attention
must be given to the drainage of this layer. Ideally, the roadbase and subbase should extend right across
the shoulders to the drainage ditches. In addition, proper crossfall is needed to assist the shedding of
water into the side drains. A suitable value for paved roads is about 2.5 to 3% for the carriageway, with
a slope of about 4-6% for the shoulders. Increased crossfalls, typically about 2-3% more, are required for
unsurfaced roads.

Lateral drainage can also be encouraged by constructing the pavement layers with an exaggerated
crossfall, especially where a permeability inversion occurs. This can be achieved by constructing the top
of the sub-base with a crossfall of 3-4% and the top of the subgrade with a crossfall of 4-5%. Although
this is not an efficient way to drain the pavement it is inexpensive and therefore worthwhile, particularly as
full under pavement drainage is rarely likely to be economically justified for LVRs. Figure D.6.25 illustrates
the recommended drainage arrangements for a LVR.

If it is too costly to extend the roadbase and subbase material across the shoulder, drainage channels at
3m to 5m intervals should be cut through the shoulder to a depth of 50mm below subbase level. These
channels should be back-filled with material of roadbase quality but which is more permeable than the
roadbase itself, and should be given a fall of 1 in 10 to the side ditch. Alternatively, a preferable option
would be to provide a continuous layer of pervious material of 75mm to 100mm thickness laid under the
shoulder such that the bottom of the drainage layer is at the level of the top of the subbase.

Figure D.6.25: Recommended drainage arrangements

Sealing of shoulders: It is now generally recommended that, wherever possible, shoulders should be
sealed, for the following reasons:
ƒ They provide better support and moisture protection for the pavement layers and also reduces
erosion of the shoulders (especially on steep gradients);
ƒ They improve pavement performance by ensuring that the zone of seasonal moisture variation
does not penetrate to under the outer wheel track (see Figure D.6.26);
ƒ They reduce maintenance costs by avoiding the need for regravelling at regular intervals;
ƒ They reduce the risk of road accidents, especially where the edge drop between the shoulder and
the pavement is significant or the shoulders are relatively soft.

For the above reasons, it is generally the case that if it is economically justifiable to pave a road then it
is very likely that it will also be economically justifiable to provide paved rather than unpaved shoulders.
This should be undertaken as part of the design of the pavement cross-section.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 163

Unsealed shoulders: A common problem associated with the use of unsealed shoulders is water
infiltration into the base and subbase for a number of reasons, which are illustrated in Figure D.6.26 and
include:
ƒ Rutting adjacent to the sealed surface;
ƒ Build-up of deposits of grass and debris;
ƒ Poor joint between the base and shoulder (more common when a paved shoulder has been added
after initial construction).

Figure D.6.26: Typical drainage deficiencies associated with pavement


shoulder construction (adapted from Birgisson and Ruth, 2003)

Avoiding trench construction: Under no circumstances should the trench (or boxed in) type of cross-
section be used in which the pavement layers are confined between continuous impervious shoulders.
This type of construction has the undesirable feature of trapping water at the pavement/shoulder interface
and inhibiting flow into drainage ditches which, in turn, facilitates damage to the shoulders under even
light trafficking.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 164

Plate D.6.8: Infiltration of water through a permeable surfacing


and subsequent outflow to an impermeable shoulder

This “boxed” construction is a common cause of road failure due to the reduction in strength and stiffness
of the pavement material and the subgrade below that required to sustain the traffic loading.

Adopting an appropriate pavement cross-section: In terms of pavement cross-section, the two moisture
zones in the pavement which are of critical significance are the equilibrium zone and the zone of seasonal
moisture variation (see Figure D.6.27: Right with a sealed shoulder; left with an unsealed shoulder).

Figure D.6.27: Moisture zones in a typical LVR

From extensive research work carried out in a number of tropical regions of the world (eg O’Reilly, 1968;
Morris and Gray, 1976; Gourley and Greening, 1999), it has been found that:
ƒ In sealed pavements over a deep water table, moisture contents in the equilibrium zone normally
reach an equilibrium value after about two years from construction and remain sensibly constant
thereafter.
ƒ In the zone of seasonal variation, the pavement moisture does not reach an equilibrium and
fluctuates with variation in rainfall. Generally, this zone is wetter than the equilibrium zone in the
rainy season and it is drier in the dry season. Thus, the edge of the pavement is of extreme
importance to ultimate pavement performance, with or without paved shoulders, and is the most
failure-prone region of a pavement when moisture conditions are relatively severe.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 165

In order to ensure that the moisture and strength conditions under the outer wheel track will remain
fairly stable and largely independent of seasonal variations, the shoulders should be sealed to a width of
between about 1.0 and 1.2 m from the edge of the sealed area (Figure D.6.28).

Adopting a holistic and integrated approach: The drainage measures highlighted above are all aimed
at:
ƒ Preventing water from entering the pavement in the first place;
ƒ Facilitating its outflow as quickly as is reasonable, given the cost implications;
ƒ Ensuring that the presence of water in the road for an extended period of time does not cause
failures.

It should be appreciated, however, that the adoption of any single measure on its own is unlikely to be
as effective as the adoption of a judicious mixture of a number of complementary measures applied
simultaneously. Such an approach forms part of the philosophy of minimising the risks associated with
using locally occurring natural materials in the pavements of LVRs.

6.19 Problem Soils


By virtue of their unfavourable properties, a number of subgrade materials fall into the category of
“Problem Soils” and, when encountered, would normally require special treatment before acceptance in
the pavement foundation. This category of soils includes:
ƒ Expansive clays;
ƒ Collapsible sands;
ƒ Dispersive soils;
ƒ Saline soils;
ƒ Micaceous soils; and
ƒ Low-strength soils.

This section focuses on typical measures that may be considered when dealing with problem soils that
occur in subgrades along the alignments of LVRs. The investigation and testing of such soils to determine
their engineering properties are not dealt with in this section but, rather, in the Site Investigation and
Route Selection chapter in Section D.2.6.7 and in the Site Investigation Manual - 2011.

6.19.1 Performance risk

In assessing the appropriateness of the measures available for dealing with problem soils, a careful
balance has to be struck between the cost of the measures and the benefits to be derived. This would
require that a life-cycle analysis be carried out to determine whether the costs of the measures would
be at least off-set by the benefits. Bearing in mind the relatively small user benefits generated by LVRs
when compared with higher trafficked roads, it is unlikely that the more extensive and costly measures
would be justified. This is particularly the case for unpaved roads where the consequences of rectifying
the problem as and when it arises are likely to be relatively small in comparison to the initial cost of
implementing the measures.

6.19.2 Expansive soils

Expansive soils are those which exhibit particularly large volumetric changes (swell and shrinkage)
following variations in moisture contents. The mechanism of expansion illustrated in Figure D.6.28 is
that of seasonal wetting and drying, with consequent movement of the water table. Soils at the edge
of the road wet up and dry out at a different rate than do those under a paved surface, thus bringing
about differential movement. It is this movement rather than the low soil strength (most expansive soils
are often relatively strong at their equilibrium moisture content) which brings about failure. Such failure
typically takes the form of associated longitudinal crack development, occurring first in the shoulder area
and developing subsequently in the carriageway, as well as general unevenness of the pavement surface,
arcuate cracking and settlement near trees and transverse humps and cracks at culvert sites (Plates D6.9
and D.6.10).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 166

Figure D.6.28: Moisture movements in expansive soils under a paved road

Generally, all of the following conditions must be satisfied before significant movement can take place:
ƒ The soil must be active; and
ƒ The changes in moisture content must be sufficiently great; and
ƒ The confining stresses must be sufficiently low.

When dry, some expansive soils present a sand-like texture and are prone to erosion to a much greater
extent than what would be normally expected from their plasticity and clay content.

Plate D.6.9: Expansive “black cotton” soil exhibiting wide-spaced shrinkage cracks

Plate D.6.10: Typical longitudinal cracking and pavement


deformation caused by an expansive soil subgrade

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 167

Countermeasures for dealing with expansive soils: Expansive soils are often thick and laterally
widespread which makes the implementation of countermeasures costly, particularly for LVRs. Any such
measures for dealing with such soils need to strike a balance between the costs involved and the benefits
to be derived over the design life of the road. Traditional countermeasures include the following:
ƒ Placing an uncompacted pioneer layer(s) of sand, gravel or rock fill over the clay and wetting up,
either naturally by precipitation or by irrigation;
ƒ Pre-wetting (2-3 months) to induce attainment of the equilibrium moisture content before
constructing the pavement;
ƒ Partially or completely removing the expansive soil and replacement with inert material;
ƒ Modifying or stabilizing the expansive soil with lime to change its properties;
ƒ Increasing the height of the fill (surcharge) to suppress heave;
ƒ Minimizing or preventing moisture change using waterproofing membranes (Weston, 1980) and/
or vertical moisture barriers (Evans and McManus, 1999).

Many of the above countermeasures would be prohibitively expensive and difficult to economically justify
for application to LVRs. Nonetheless, there are a number of relatively low-cost measures that can be
adopted based on practical experience in a number of African countries. The LVR countermeasures shown
in Table D.6.28 provide a guide for the selection of appropriate countermeasures that could be considered
for LVRs based on the degree of expansiveness, as defined in Chapter D.2 – Site Investigations and Route
Selection. The final choice should be based on a life-cycle cost analysis of the options presented.

Table D.6.28: Countermeasures for dealing with expansive soils

Alternative design and construction measures


Expansiveness of soil over expansive soils
Design Traffic < 100,000 esa Design Traffic > 100,000 esa
Low Countermeasure A Countermeasure A
Medium Countermeasure A Countermeasure A
High Countermeasure A Countermeasure B
Very high Countermeasure B Countermeasure C1 or C2

Countermeasure A: General good construction practice for all roads on expansive soils adds little, if any,
additional cost to construction works. Where possible:
ƒ Remove vegetation during the dry season as long as possible in advance of construction.
ƒ Construct any cuttings necessary, however shallow.
ƒ Undertake construction when the in-situ material is at equilibrium moisture content (ie at the end
of the rainy season). If construction takes place in the dry season, the roadbed should be watered
to saturation immediately prior to the placing of the backfill material.
ƒ Extend side slopes of the embankment to 1:4 for Design Traffic 1 and 1: 6 for Design Traffic 2 (there
is no design traffic 1 or 2 in the table above). Utilise excavated material to flatten the side slopes
of the embankment.
ƒ No side drains unless necessary due to site conditions in which case locate them from the toe of
the embankment for a distance of 4m for design traffic < 100,000 esa and 6m for design traffic 2
> 100,000 esa (see Figure D.6.29)
ƒ Seal shoulders.
ƒ Remove/do not plant trees within a distance of 1.5 times their mature height from the edge of the
seal.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 168

min. 6m (4m)

side drain

Figure D.6.29: Location of side drains in expansive soils

Countermeasure B: Use of a ‘pioneer” layer (Figure D.6.30) as follows:


ƒ Adopt countermeasures listed for Measure A.
ƒ Place a loose layer “pioneer” layer (about 100-200mm in thickness) of permeable sand, gravel
or rock fill over the clay to cover the full width of construction. It is essential that this layer should
remain loose and permeable and must therefore not be compacted or trafficked.
ƒ Allow the “pioneer” layer to stand through one full rainy season in order to pre-wet the roadbed as
much as possible by the elimination of evapotranspiration, and the collection of rainwater. Prevent
localised ponding of water.
ƒ Compact the “pioneer” layer in advance of construction during the following dry season and utilize
it as the first layer of fill.
ƒ Ensure minimum earthworks cover of 0.6m for Design Traffic 1 and 1m for Design Traffic 2.
ƒ Do not use active clay as fill.
ƒ Replace clay under culverts to a depth equivalent to the reduction of surcharge caused by the
culvert.
ƒ Waterproof culvert joints.
ƒ Prevent ponding of water at culvert inlets and outfalls and adjacent to road.

It important to note that when it is not possible to apply the “pioneer” layer technique, the vegetation
should be removed as far in advance of construction as is feasible. If the roadbed is to stand open during
the rainy season, it will be advantageous to plough or scarify it to a depth of about 150mm to promote
the collection and ingress of rainwater.

Figure D.6.30: Construction on expansive soils (use of pioneer layer)

Countermeasure C 1: Partial excavation (for embankments < 2m in height)


ƒ Adopt countermeasures listed for Measure A;
ƒ Excavate expansive soil over width to toe 1:2 side slope and to depth of 0.6m;
ƒ Stockpile excavated at sides for eventual grading on to shoulder slopes;
ƒ Backfill excavation with non-expansive fill. Ensure minimum earthworks cover of 0.6m for Design
Traffic 1 and 1m for Design Traffic 2;
ƒ Fill above ground level to be constructed with 1:2 side slopes;
ƒ Grade and spread excavated expansive soil on fill side slopes to lengthen their slope to 1:6 or
flatter, thereby extending the distance of the road over which transpiration will be reduced;
ƒ Expansive material must not be used for the shoulder slope to the pavement – these must be
constructed as wedges of permeable material as shown in Figure D.6.31.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 169

Figure D.6.31: Construction on expansive soils (embankment height <2m)

Countermeasure C 2: Partial excavation (for embankments >2m in height)


ƒ Adopt same countermeasures as for Measure (C.1) except for the following:
ƒ Excavate expansive soil under the width of the 1:2 side slopes (see Figure D.6.32).

Figure D.6.32: Construction on expansive soils (embankment height >2m)

The other countermeasures mentioned above, including stabilisation, surcharging and use of waterproofing
membranes would normally be ruled out on cost grounds for LVRs.

6.19.3 Collapsible soils

Collapsible soils occur mostly in the arid and semi-arid regions of eastern and south eastern Ethiopia.
They exhibit a weakly cemented soil fabric which, under certain circumstances, may be induced to rapid
settlement. A characteristic of these soils is that they are all unsaturated; generally have a low dry density;
and a low clay content. At the in-situ moisture content they can withstand relatively large imposed stresses,
well in excess of the overburden pressure, with little or no settlement. However, without any change in the
applied stress, but an increase in moisture content, additional settlement will occur as shown in Figure
D.6.33 and illustrated in Plate D.6.11. The rate of settlement will depend of the permeability of soil.

Countermeasures for dealing with collapsible soils: Methods for dealing with collapsible soils depend on
the degree of collapse potential as discussed Chapter D.2: Site Investigations and Route Selection. The
countermeasures include:
ƒ Excavation of material to the specified depth below ground level; break down collapsible structure;
replace in the excavation; and re-compact with conventional rollers in lifts typically not exceeding
250mm;
ƒ Ripping of the road bed, inundation with water and compaction with heavy vibrating rollers;
ƒ Use of high energy impact compactors from the surface of the subgrade, with or without the use
of water.

The risk of collapse occurring on LVRs, particularly in arid or semi-arid areas, is small. Thus, other than
exceptional circumstances, the above measures are unlikely to be economically justified for application
to LVRs.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 170

Plate D.6.11: Collapse settlement in excess of 150mm following impact compaction

Figure D.6.33: Manner of additional settlement due to collapse of soil fabric

6.19.4 Dispersive/erodible soils

Dispersive and erodible soils are prevalent over many areas of Ethiopia (Plates D.6.12 and D.6.13).
Although these soil types are similar in their field appearance (highly eroded, gullied and channelled
exposures), they differ significantly in the mechanisms of their actions and are differentiated as follows:
ƒ Dispersive soils are those soils that, when placed in water, have repulsive forces between the clay
particles that exceed the attractive forces. This results in the colloidal fraction going into suspension
and in still water staying in suspension. In moving water, the dispersed particles are carried away.
ƒ Erodible soils are those soils in which the cohesion (or surface shear strength when wet is insufficient
to resist the tractive forces of rain or runoff water flowing over them. Such soils tend to lose material
as a result of flowing water over the material exceeding the cohesive forces holding the material
together.

It is not normally important, or even easily possible, to quantify the actual potential loss of dispersive/
erodible material as the process is time related and given enough time, all of the colloidal material could
theoretically be dispersed and removed, leading to eventual loss of material on a large scale. However,
it is important to identify the presence of dispersive/erodible soils so that necessary precautions can be
taken if they affect the constructed pavement. Methods of identifying such soils are addressed in the Site
Investigation manual.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 171

Plates D.6.12 and D.6.13: Examples of severe erosion in erodible/dispersive soils in Ethiopia

Countermeasures for dealing with erodible/dispersive soils: Methods for dealing with dispersive soils
include (Paige-Green, 2008):
ƒ Avoiding the use of such soils in fills as far as possible and removing and replacing it in the subgrade;
ƒ Managing water flows and drainage in the area well;
ƒ Treating the soil with lime or gypsum to allow the calcium ions to replace the exchangeable sodium
cations and reduce the problem;
ƒ Compacting the soil at 2 to 3% above optimum moisture content to as high a density as possible
(Elges, 1985) .

Methods for dealing with erodible soils include (Paige-Green, 2008):


ƒ Ensuring that the drainage in the area is well controlled;
ƒ Covering the soils with non-erodible materials and vegetation;
ƒ Once erosion has occurred, back-filling the channels and gullies with less erodible material and
redirecting the water flows.

6.19.5 Saline soils

Saline soils occur mostly in the arid or semi-arid regions of Ethiopia where the dry climate combined with
the presence of saline materials and/or saline ground or surface water, create conditions that are conducive
to the occurrence of salt damage. The presence of soluble salts in the subgrade or pavement materials
can cause damage to the bituminous surfacings of roads. Such damage occurs when the dissolved salts
migrate to the road surface, mainly due to evaporation, become supersaturated and then crystallize with
associated volume change. This creates pressures which can lift and physically degrade the bituminous
surfacing and break the adhesion with the underlying pavement layer as illustrated in Plates D.6.14 and
D.6.15. Generally, the thinner the surfacing layer is, the more likely the damage, primes being the most
susceptible and thick, impermeable seals the least susceptible.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 172

Plate D.6.14: An example of severe distress to a road surfacing due to salt attack
resulting in damage within two years of its construction (Botswana)

Plate D.6.15: Salt damage may appear in the form of “blistering”,


“heaving” and “fluffing” of the prime surfacing.

As soluble salt problems arise from the accumulation and crystallization of the salts under the road
surfacing and the upper base layer, minimization of salts in the pavement layers can be achieved with the
use of impermeable surfacings. Such surfacings prevent water vapour passing through it, and, as a result,
crystallization will not occur beneath the surfacing (Netterberg, 1979). Construction should then proceed
as fast as possible to minimize the migration of salts through the layers. The addition of lime to increase
the pH to in excess of 10 will also suppress the solubility of the more soluble salts.

Guidelines for the prevention and repair of salt damage to roads and runways have been developed
based on research work carried out in the southern African region (Botswana Roads Department, 2003).
These guidelines provide guidance on methods of testing and measurement of salts as well as repair
methods where damage has already occurred.

6.19.6 Micaceous soils

Micaceous soils are those soils which contain large quantities of mica (muscovite) and occur in such
materials as weathered granite, gneiss, mica schist and phylite materials that occur in various areas
of Ethiopia. These soils often cause problems with compaction because of the “spring action” of the
muscovite materials which may prevent achievement of the intended density or, even if it is achieved
initially, can cause rutting in the compacted layer at a later stage.

Countermeasures for dealing with micaceous soils: Methods for dealing with erodible soils include:
ƒ Removing the micaceous soil layer to below the material depth in the subgrade;
ƒ Stabilizing the micaceous soil with lime or cement.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 173

For LVRs, the loss of shape associated with micaceous subgrades would generally have to be accepted
unless the overlying pavement warrants the expense of the countermeasures indicated above.

6.19.7 Low-strength soils

Soils with a soaked CBR of less than 3 per cent (< 2 per cent in dry climates) are described as Low-
Strength soils. These soils may be extremely soft in their natural state or become extremely soft on
soaking. They occur particularly in the low-lying, swampy areas of Ethiopia. They are easy to identify
either in-situ or during site inspections or laboratory testing of their soaked strengths. Typical treatment
measures for such soils include:
ƒ Removal and replacement with suitable material;
ƒ Stabilisation – chemical, modification with lime or mechanical;
ƒ Use of geo-synthetic products;
ƒ Raising of the vertical alignment to increase soil cover and thereby redefine the design depth
within the pavement structure.

Further details on the respective methods of treatment for low-strength soils need to be established
in the design stage at project level and the appropriate measure will depend on soil properties, site
conditions, available equipment, available materials, experience from other sites with similar conditions
and construction economy. The subgrade class will need to be re-defined according to the new subgrade
strength after treatment.

6.20 Construction Issues


One of the challenges of utilising natural gravels in LVR pavements is to maximise their strength, increase
their stiffness and bearing capacity, increase their resistance to permanent (plastic) deformation and
reduce their permeability (and, hence, susceptibility to moisture ingress). These attributes can be achieved
through effective compaction, as discussed below.

6.20.1 Subgrade compaction

Effective subgrade compaction is one of the most cost-effective means of improving the structural
capacity of pavements. A well compacted subgrade possesses enhanced strength, stiffness and bearing
capacity; is more resistant to moisture penetration; and less susceptible to differential settlement. The
higher the density, the stronger the subgrade support, the lesser the thickness of the overlying pavement
layers and the more economical the pavement structure. Thus, there is every benefit to achieving as high
a density and related strength as economically possible in the subgrade.

Maximising the strength potential of a subgrade soil can be achieved, not necessarily by compacting
to a pre-determined relative compaction level, as is traditionally done but, rather, by compacting to
the highest uniform level of density possible without significant strength degradation of the particles
(“compaction to near refusal”). In so doing, there is a significant, beneficial, gain in density, strength
and stiffness and reduction in permeability, the benefits of which generally outweigh the costs of the
additional passes of the roller.

Compaction to near refusal ensures that the soil has been compacted at an appropriate moisture content
to its near elastic state as shown in Figure D.6.34 at which point the air voids in the material are relatively
low (< 5%) with the significant benefit of reduced pavement deflection and increase in pavement life
as illustrated in Figure D.6.35. If, however, the volume of voids is high after construction, the pavement
will densify under traffic loading and rutting will appear in the wheel tracks. Further, if both the moisture
content is high in service and the air voids are also high, the pavement is potentially unstable and serious
deformation is likely to occur, particularly with heavy traffic using the road. These potentially adverse
situations emphasise the importance of ensuring that the subgrade compaction is carried out properly by
controlling both the air voids and moisture content at which the specified density is attained.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 174

Figure D.6.34: Illustration of concept “compaction to refusal”

Figure D.6.35: Deflection-life relationship and benefits of “compaction to refusal”

Compaction at low moisture content: In the arid or semi-arid, north-eastern and south-eastern
regions of Ethiopia where rainfall is less than 500mm per annum, water is often scarce and problems
arise when large quantities (up to 2,000 m3/km are needed for road construction. In these regions
qualified consideration can be given to “dry compaction” techniques for the compaction of the subgrade
and pavement layers. As illustrated in Figure D.6.36, high densities can be achieved at low moisture
contents using conventional compaction plant. However, as shown in Figure D.6.37, soils compacted
at low moisture contents will have high air voids. As indicated above, should the degree of saturation
increase in service, this may allow an ingress of water into permeable pavements even if the road surface
and shoulders are sealed, resulting in a loss in soil strength and resulting deformation of the pavement
structure. A life-cycle analysis will allow a determination to be made of the preferable option.

Impact compaction provides an alternative to conventional compaction plant for undertaking compaction
at low moisture contents. Impact Compactors are non-circular, relatively high-energy ‘rollers’, typically
three-(see Plate D.6.16), four- or five- sided. Large-wheeled tractors are used for pulling the compactors
at operational speeds of 12 – 15 km/hr producing a series of high amplitude/high impact blows delivered
to the soil at a relatively low frequency (90 – 130 blows per minute) with the energy per blow varying
between 10 and 25 kilojoules, depending on the mass and amplitude of the compactor.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 175

Figure D.6.36: Extended compaction curve for low moisture contents (Parsons, 1992)

Figure D.6.37: Air voids in dry and well compacted soil (Parsons, 1992)

Plate D.6.16: Three-sided impact compactor

Due to their very high energy density per blow, their main advantage over conventional compaction plant
is their depth effectiveness, typically of the order of one metre of fill or in-situ layers, thereby producing
deep, well-balanced, relatively stiff pavement layers. These rollers are well suited for densifying collapsible
soils. They have been successfully used in low-cost road systems and, when appropriately specified, offer
a cost-effective option for LVR construction.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 176

Minimum Compaction Requirements: Table D.6.29 gives the minimum compaction requirements for the
various layers in the pavement. For the reasons stated in Section D.6.20.1, where the higher densities can
be realistically attained in the field (compaction to refusal) from field measurements on similar materials
or other established information, they should be specified by the Engineer.

Table D.6.29: Minimum compaction requirements

Target Density
Pavement Layer Material Class
(Relative Compaction)
G80
G65
Roadbase 98% – 100% T180
G55
G45
Subbase G30 95% -97% T180
G15
Subgrade/Fill Wearing Course 93% -95% T180
G7
Sand 100% T180
Roadbed
Gravel 93% -95% T180

6.20.2 Quality Attainment

LVR design procedures assume that both the material properties and levels of density specified are
achieved in the field. However, in order to attain the specified densities, it is essential to ensure, as far
as practicable, the uniform application of water, the uniformity of mixing and uniformity of compaction
at or near OMC.

It is also important to note that layers below the one being compacted should be of sufficient density
and strength to facilitate effective compaction of the upper layer(s). Adherence to the compaction
recommendations given in Table D.6.24 should ensure this.

Granular materials which are well graded are easier to compact than poorly graded ones. It may therefore
be more economical to get the gradation right (eg by mechanical stabilisation) before wasting time and
energy with excessive rolling. Improved grading is also likely to improve the material strength (CBR) to an
extent where a subbase quality material could become eminently suitable for road base.

Whilst it is necessary for natural gravels to be brought to OMC for efficient compaction, it is necessary to
ensure that premature sealing does not lock in construction moisture. This can be achieved by allowing a
significant amount of drying out to occur before sealing takes place, particularly for materials that rely on
soil suction forces for strength gain and improved stability.

The variability of natural gravels is a significant factor in the reliability of performance of the pavement.
However, various measures can be taken during construction to reduce such variability. These include:

ƒ Careful selection during the winning process. Physical properties of natural gravels in most deposits
tend to change with depth and location. Careful selection of the material during the winning
process, coupled with appropriate testing on a grid pattern (eg use of the linear shrinkage test) will
often facilitate uniform stockpiling of the material.
ƒ Processing of stockpiled material: Power screens have been proved effective in screening out and
blending in to overcome deficiencies and can be particularly useful in attaining the requirements
for gravel wearing curse materials.
ƒ Quality control and assurance: Quality attainment and control are paramount when using
unprocessed materials for LVR construction. Quality assurance procedures and the use of statistical
control methods are recommended. Such measures will eliminate the costly ramifications flowing
from arbitrary decisions to include or exclude the use of certain readily available materials.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 177

6.21 References:
Weinert H. H. 1974. A Climatic Index of Weathering and its Application to Road Construction.
Geotechnique, Vol. 24, No. 4, pp 475-488.

Emery, S.J. 1985. Prediction of Moisture Content for Use in Pavement Design. PhD Thesis, Univ. of
Witwatersrand, Johannesburg.

Alvin, R.G. and Hamitt II, G.M. 1970. Load-supporting Capability of Low-volume Roads. Transportation
Research Board. National Academy of Sciences, Washington DC. Special Report 160. 1975.

Transport Research Laboratory (TRL). 1997. Overseas Road Note 31. A Guide to the Structural Design
of Bitumen-Surfaced Roads in Tropical and Sub-Tropical Climates. Crowthorne. HMSO, London.

Howe, J.D.G.F and G. Hathway. 1983. Earth Roads: Their Construction and Maintenance. Intermediate
Technology Transport Ltd, IT Publications, London.

Odier, L., Millard, R.S., Pimental dos Santos and S.R. Mehra. 1971. Low Cost Roads: Design, Construction
and Maintenance. UNESCO. London, Butterworths.

Transport Research Laboratory (TRL). 2008. Performance Criteria and Life-Cycle Costing for Unpaved
Roads Ethiopia Country Component. Unpublished Project Report UUPR/III/028/08.

Paige-Green, P. 1989. New performance-related specifications for unpaved roads. Proc. Annual
Transport Convention, Pretoria, South Africa, paper 3A-12, August 1989.

Gourley, C.S. and P.A.K. Greening. 1999. Performance of Low-volume Sealed Roads: Results and
Recommendations from Studies in Southern Africa. TRL Published Report PR/OSC/167/99. Crowthorne.

Newill, D., Robinson, R. and K.Aklilu. 1987. Experimental use of cinder gravels on roads in Ethiopia. 9th
Regional Conference for Africa on Soil Mechanics and Foundation Engineering. Lagos. A.A. Balkema/
Rotterdam/Boston/1987.

Lay, M.G. 1998. Handbook of Road Technology: Volume 1 – Planning and Pavements (3rd Ed). Gordon
and Breach Science Publishers, Australia.

Morris, P.O. and W.J. Gray. 1976. Moisture conditions under roads in the Australian environment.
Australian Road Research Board. Research Report, ARR No. 69.

Gourley, C.S. and P.A.K. Greening. 1999. Performance of Low-volume Sealed Roads: Results and
Recommendations from Studies in Southern Africa. TRL Published Report PR/OSC/167/99. Crowthorne.
Weston, D.J. 1980. Expansive soil treatment for southern Africa. Proc. Fourth Int. Conf. on Expansive
Soils, Denver.

Evans, R.P. and K.J. McManus. 1999. Construction of Vertical Moisture Barriers to Reduce Expansive
Soil Subgrade Movement. In Transportation Research Record 1652, TRB, National research Council,
Washington, DC. pp 108-112.

Paige-Green, P. 2008. Dispersive and Erodible Soils. Proceedings: Problem Soils in South Africa, 304
November 2008. South African Institute for Engineering and Environmental Geologists.

Elges, H.F.W.K. 1985. Dispersive soils. The Civil Engineer in South Africa. 27(7):347-353.

Obika, B. R.J.freer-Hewish, M. Woodbridge and D. Newill (1995). Prevention of Salt Damage to Thin
Bituminous Surfacings: Design Guidelines. Proc. Sixth Int. Conf. on Low-volume Roads, Minneapolis,
Minnesota, June 25-29, 1995.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 6 - 178

Botswana Roads Department, Ministry of Works, Transport and Communications. 2003. Guideline No.
6: The Prevention and Repair of Salt Damage to Roads and Runways. 2001.

Parsons, AW. 1992. Compaction of Soils and granular materials: a review of research performed at the
Transport Research Laboratory. Department of Transport (1992). London: HMSO.

Kleyn, E.G, Maree, J.H. and P.F.Savage 1982. The Application of a Portable Pavement Dynamic Cone
Penetrometer to Determine In-situ Bearing Properties of Road Pavement layers and Subgrades in South
Africa. European Symposium on Penetration Testing, Amsterdam.

Kleyn, Eg and G.D. van Zyl. 1988. Application of the DCP to light Pavement Design. First Int. Symposium
on Penetration Testing, Orlando, Florida.

South African Roads Board, 1990. Recommended durability tests and specification limits for base
course aggregates for road construction report PR88/032:1102.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 179

7. SURFACING

7.1 Introduction
There are a large number of bituminous and non-bituminous surfacing options available for use on LVR
pavements. These surfacings fulfil a variety of functions which, collectively, preserve the integrity of the
underlying pavement layers and improve the functionality of the road in service. The basic local materials
of natural soils/gravels, stone, fired clay brick can be used with or without a range of binders/sealers to
offer a range of attributes which need to be matched to such factors as expected traffic levels and loading,
locally available materials and skills, construction and maintenance regimes and the environment. Careful
consideration should therefore be given to all these factors in order to make a judicious, cost-effective
choice of surfacing to provide satisfactory performance and minimise life cycle costs.

This chapter provides an overview of the various types of surfacings available and appropriate for use in
Ethiopia in relation to a range of local factors. The chapter also provides information on the constituents
and performance characteristics of the surfacings, the factors affecting their choice and the general
approach to their design.

7.2 Types of Surfacings


Road surfacings may be grouped according to their main constituents as follows:

Basic
S-01: Engineered Natural Surface (ENS)
S-02: Natural gravel

Stone Paving
S-03: Waterbound/Drybound Macadam (WBM - DBM)
S-04: Hand Packed Stone (HPS)
S-05: Stone Setts or Pavé (SSP and MSSP)
S-06: Mortared Stone (MS)
S-07: Dressed stone/cobble stone (DS, CS, MDS, MCS)

Fired Clay Brick


S-08: Unmortared/mortared joints (CB, MCB)

Bituminous
S-09: Sand Seal
S-10: Slurry Seal
S-11: Chip Seal
S-12: Cape Seal
S-13: Otta Seal

Concrete
S-14: Non-reinforced concrete (NRC)
S-15: Ultra-thin reinforced concrete pavement (UTRCP)

An outline description of the above surfacing types is presented below while their relative advantages
and disadvantages are summarised in Annex D.3.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 180

7.2.1 Basic Surfacings

S-01: Engineered Natural Surface (ENS)


An Engineered Natural Surface (Plate D.7.1)using the compacted soil at the road location to form a
basic surface for traffic. Essential provisions are a compacted camber (3-6%), side drains and an effective
drainage system. Typically soils with an in service CBR of a minimum of about 15 or more can provide a year
round running surface for light motor traffic. Route sections with steep gradients, or weak or problematic
soils can be improved in situ by upgrading to higher standard surface under a spot improvement or EOD
strategy to improve their traffic carrying capacity throughout the year.

Plate D.7.1: An example of an Engineered Natural Surface (ENS)

S-02: Natural Gravel


One or more layers of natural gravel (Plate D.7.2) placed directly on the existing shaped earth formation
and compacted with an appropriate surface camber (typically 3-6%). The layers could be mechanically
stabilized or blended with other material to improve the properties.

Plate D.7.2: An example of a Natural Gravel Surfacing (Cinder gravel)

7.2.2 Stone Paving

S-03: Waterbound/Drybound Macadam


A Macadam layer essentially consists of a stone skeleton of single sized coarse aggregate in which
the voids are filled with finer material. The stone skeleton, because of its single size large material will
contain considerable voids, but will have the potential for high shear strength, if confined properly. The
stone skeleton forms the “backbone” of the macadam and is largely responsible for the strength of the
constructed layer. The material used to fill the voids provides lateral stability to the stone skeleton but
adds little bearing capacity.

In Waterbound Macadam (WBM) the aggregate fines are washed or slushed into the coarse skeleton with
water. Dry-bound macadam is a similar technique to the original WBM, however instead of water and
deadweight compaction being used in the consolidation of fine material, a vibrating roller is used. The

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 181

development of small vibrating rollers has made the use of this technique attractive for rural road works
in some locations.

WBM or DBM are commonly used as layers within a sealed flexible pavement, but in the appropriate
circumstances may be used as an unsealed option with a suitably cohesive material being used as the
fines component. The WBM or DBM may be constructed as a low cost, initial surface to be later sealed
and upgraded in a ‘stage construction’ strategy.

Plate D.7.3: An example of a Waterbound/Drybound Macadam

S-04: Handpacked Stone


Hand Packed Stone surfacing consists of a layer (typically 150 – 300 mm) of large broken stones pieces,
tightly packed together and wedged in place with smaller stone chips rammed by hand into the joints
using hammers and steel rods (Plate D.7.4). The remaining voids are filled with sand. The Hand Packed
Stone is normally bedded on a thin layer of sand gravel. For use by heavy traffic, the layer should be
compacted with a vibrating or heavy non-vibrating roller. An edge restraint or kerb constructed of large
or mortar jointed stones improves durability and lateral stability.

Plate D.7.4: Hand-packed Stone

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 182

S-05: Stone Setts or Pavé


Stone sett surfacing or Pavé (See Plates D.7.5 and D.7.6) is an historically well-established technique
that has been adapted successfully as a robust option on low volume rural roads where there is a good
local supply of suitable stone. It consists of a layer of roughly cubic (100mm) stone setts laid on a bed of
sand or fine aggregate within mortared stone or concrete edge restraints. The individual stones should
have at least one face that is fairly smooth, to be the upper or surface face when placed. Each stone sett
is adjusted with a small (mason’s) hammer and then tapped into position to the level of the surrounding
stones. Sand or fine aggregates is brushed into the spaces between the stones and the layer then
compacted with a roller.

Plate D.7.5: Stone Setts (Linear pattern) Plate D.7.6: Stone Setts (Radial pattern)

S-06: Mortared Stone


Mortared Stone Paving (Plate D.7.7) consists of a layer of natural selected stones, laid on a bed of loose
sand or fine aggregate with the joints filled with sand–cement mortar. The stones do not need to be
dressed to a regular shape. The individual stones should have at least one face that is fairly smooth
and even, to be the upper or surface face when placed. Stone size is typically from 100 – 300mm. The
bedding sand around each stone is adjusted with a small hammer and the stone is then tapped into
position and to the final level of the surrounding stones. Sand–cement mortar and small stones are used
to fill the joints between the individual stones. When the mortar has set the layer should be covered
in sand or other moisture retaining material and kept wet for a few days to aid curing. Mortared Stone
paving should not be trafficked until 7 days after laying.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 183

Plate D.7.7: Mortared Stone

S-07: Dressed Stone/Cobble Stone Paving


Dressed or Cobble Stone Paving (Plate D.7.8) has been used for centuries as a strong, durable road
surface. The technique is similar to Stone Setts or Pavé, however the individual stones are larger, normally
of size 100 – 300mm. They are cut from suitable hard rock and ‘dressed’ manually to a cubic shape with
a smooth, flat finish on at least one face using hammers and chisels. The dressed stones are laid on a
bedding sand layer (20 – 70mm) and tapped into final position with a hammer. Sand is brushed into the
joints between the stones. Covering with loose sand and compacting with a heavy roller can improve
durability. An edge restraint or kerb constructed (for example) of large or mortared stones is required for
durability. Sand-cement mortar joints and bedding can be used to improve durability and prevent water
penetrating to moisture susceptible foundation layers and weakening them.

Plate D.7.8: Dressed Stone/Cobble Stone

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 184

7.2.3 Fired Clay Brick

S-08: Unmortared or mortared joints


Bricks suitable for road surfacing can be produced by firing clay in large or small scale kilns using coal,
wood or some agricultural wastes as a fuel. The bricks must achieve certain strength, shape and durability
requirements. The fired bricks are generally laid on edge to form a layer of typical 100mm thickness
on sand or sand-cement bedding layer and jointed similarly (Plate D.7.9). Kerbs or edge restraints are
necessary and can be provided by sand-cement mortared fired bricks. The fired bricks are normally laid
in a herring bone or other approved pattern to enhance load spreading characteristics. Un-mortared brick
paving is compacted with a plate compactor and the jointing sand is topped up if necessary. For mortar
bedded and jointed fired clay brick paving, no compaction is required. When the mortar has set the layer
should be covered in sand or other moisture retaining material and kept wet for a few days to aid curing.
If mortared bedding and jointing are used the surface should not be trafficked until 7 days after laying.

Plate D.7.9: Fired clay brick

7.2.4 Bituminous surfacings

Bituminous surfacings or surface treatments generally comprise an admixture of different proportions


of stone or sand and bitumen. The bitumen may be a penetration grade, cutback or emulsion (the last
being particularly suitable for labour based methods of construction as heating is avoided; or small scale
works). The bituminous surfacings usually require good quality, screened or crushed stone or sand, but
lower quality aggregate may be used for some types of seals (eg Otta Seal).

An effective bond between the surface treatment and the surface of the roadbase is essential for good
performance. This can be achieved through the use of an appropriate grade of bitumen (the prime or
prime coat) before the start of construction of the surface treatment.

Types of Surface Treatments


Some typical types of surface treatment are shown in Figure D.7.1.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 185

Figure D.7.1: Examples of typical surface treatments

S-09: Sand Seal


This seal consists of a spray of binder followed by the application of a coarse, clean sand or crusher dust
as aggregate. This surfacing is used on low-volume roads, especially in drier regions, but can also be
used for maintenance resealing, or for temporary by-passes. For new construction two layers are usually
specified as single layers tend to be not durable. There is an extended curing period (typically 8 – 12
weeks) between the first and second seal applications to ensure complete loss of volatiles from the first
seal and thus prevent bleeding.

S-10: Slurry Seal


A Slurry Seal consists of a homogeneous mixture of pre-mixed materials comprising fine aggregate,
stable-mix grade emulsion (anionic or cationic) or a modified emulsion, water and filler (cement or lime).
The production of a slurry can be undertaken in simple concrete mixers and laid by hand, or more
sophisticated purpose-designed machines which mix and spread the slurry.

Slurry Seals can be used for treating various defects on an existing road surface carrying relatively low
traffic for which the following are typical applications:
ƒ Arrest loss of chippings;
ƒ Restore surface texture;
ƒ Reduce unevenness because of bumps, slacks and/or ruts;
ƒ Rectify low activity surface cracking;
ƒ New construction as a grout seal following a single Chip Seal or in multiple layers directly on the
base course of low traffic roads;
ƒ A component of a Cape Seal.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 186

S-11: Chip Seal


This seal (single or double) consists of a spray(s) of bituminous binder followed by the application of
a layer(s) of aggregate (stone chippings). The binder acts as a waterproofing seal preventing entry of
surface water into the road structure while the chippings protect this film from damage by vehicle tyres.

Chip Seals can be used for a number of purposes, including:


ƒ New construction (normally double surface dressings only);
ƒ Temporary by-passes (normally single surface dressings);
ƒ Maintenance resealing (normally single surface dressing);
ƒ First layer of a Cape Seal.

S-12: Cape Seal


A Cape Seal consists of a single 13mm or 19mm aggregate, penetrated with a binder and covered with
a slurry seal. If 19mm aggregate is used, the slurry is applied in two layers. The function of the slurry is to
provide a dense void filler to enhance the stability of the single-sized coarse aggregate layer. The coarse
aggregate is left proud to provide the macro texture for skid resistance.

S-13: Otta Seal


An Otta Seal is a sprayed bituminous surfacing comprising a mixture of graded aggregates ranging from
natural gravel to crushed rock with relatively soft (low viscosity) binder, with or without a sand cover seal.
This type of seal contrasts with the single sized crushed aggregate and relatively hard (high viscosity)
binders used in Chip seals. The following are the main types of Otta Seals:
ƒ Single / Double Otta Seal:
• Pen / medium / dense graded;
• Sand Seal / no Sand Seal cover.

Otta Seals can be used for a variety of purposes, including:


ƒ New construction (single or double Otta Seals with/without sand seal;
ƒ Temporary seal (normally single Otta Seal - diversions, haul roads, temporary accesses, etc);
ƒ Maintenance reseal (normally single Otta Seal).

Performance characteristics
The mechanism of performance of surface treatments varies in relation to the composition of their
constituents as illustrated in Figure D.7.2 and described below.

Type A: (eg Sand seal and Otta Seal):


These seal types, like hot-mix asphalt, rely to varying extents on a combination of mechanical particle
interlock and the binding effect of bitumen for their strength. Early trafficking and/or heavy rolling is
necessary to develop the relatively thick bitumen film coating around the particles. Under trafficking, the
seal acts as a stress-dispersing mat comprised of a bitumen/aggregate admixture.

Type B: (eg Chip seal, Cape Seal):


These seal types rely on the binder to “glue” the aggregate particles to the base. Where shoulder-to-
shoulder contact between the stones occurs, some mechanical interlock is mobilized. Under trafficking,
the aggregate is in direct contact with the tyre and requires relatively high resistance to crushing and
abrasion to disperse the stresses without distress.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 187

Figure D.7.2: Differing mechanisms of performance of surface treatments

Typical service lives


The life of a bituminous surfacing treatments can vary widely in relation to a number of factors as indicated
below:
ƒ Climate: Very high temperatures cause rapid binder hardening through accelerated loss of volatiles,
while low temperatures can lead to brittleness of the binder leading to cracking or aggregate loss
resulting in reduced surfacing life.
ƒ Pavement strength: Lack of underlying pavement stiffness will lead to fatigue cracking and
reduced surfacing life.
ƒ Base materials: Unsatisfactory base performance and absorption of binder into certain base
materials (eg pedogenic materials) will lead to reduced surfacing life.
ƒ Binder durability: The lower the durability of the binder, the higher the rate of its hardening, and
the shorter the surfacing life.
ƒ Design and construction of surfacing: Improper design and poor construction techniques (eg
inadequate prime, uneven rate of binder application or ‘dirty’ aggregates) will lead to reduced
surfacing life.
ƒ Traffic: The higher the volume of heavy traffic the shorter the surfacing life.
ƒ Stone polishing: The faster the polishing of the stone, the earlier the requirement for resurfacing.
ƒ Aggregate size: The larger the aggregate size, the shorter the service life, all other factors being
equal.

Typical service lives of bituminous surface treatments are given in Table D.7.1.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 188

Table D.7.1: Typical bituminous surfacing service lives1

Typical service
Type of surfacing
life (years)
Single Sand seal 2–3
Double sand seal 3-6
Slurry seal 2-4
Single chip seal 3-5
Double chip seal 7 - 10
Single Otta seal 6 - 10
Single Otta seal plus sand seal 8 - 12
Cape Seal (13mm + single slurry) 6 - 10
Cape seal (19mm + double slurry) 8 - 14
Double Otta seal 12 – 16
Note:
1. Assumes that timeous routine and periodic maintenance is carried out.

Factors affecting choice of bituminous surface treatments


The various factors affecting the choice of surface treatments in relation to the operational requirements
is indicated in Table D.7.2.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 189

Table D.7.2: Factors affecting choice of bituminous surface treatments

Type of Surfacing
Parameter Degree SOS+
SS SlS SCS DCS CS DOS
SS
Short
Service life
Medium
Required1
Long
Light
Traffic level Medium
Heavy
Low
Impact of traffic
Medium
turning action
High
Mild
Gradient Moderate
Steep
Poor
Material quality Moderate
Good
Poor
Pavement and
Moderate
base quality
Good
Suitability for labour-based methods
Contractor Low
experience/ Moderate
capability High
Low
Maintenance
Moderate
capability
High
Key:
SS = sand seal, SlS = Slurry Seal, SCS = single chip seal, DCS = double chip seal, CS = Cape seal,
SOS+SS = Single Otta seal + sand seal, DOS = double Otta seal

Suitable/Preferred Less suitable/not preferred Not suitable/not applicable

Note:
1. Short < 5 years; Medium 5 – 10 years, Long > 10 years

The final choice of a surface treatment should be based on the Surfacing Decision Management System
(SDMS) described in Section D.7.3.3 of this Chapter and a life-cycle cost analysis in which the various
factors discussed above, as well as the service life of the treatment, should all be taken into account.

Design of bituminous surfacings


The design of a particular type of surface treatment is usually project specific and related to such factors as
traffic volume, climatic conditions, available type and quality materials. Various methods of design have
been developed by various authorities for the design of surface treatments. The approach to the design
of surface treatments given in this section is generic, with the objective of presenting typical binder and
aggregate application rates for planning or tendering purposes only. Where applicable, reference has
been made to the source document for the design of the particular surface treatment which should be
consulted for detailed design purposes.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 190

Prime Coat
An effective bond between the surface treatment and the existing road surface is essential for good
performance of a bituminous surfacing. This generally requires that the non-bituminous road surface must
be primed with an appropriate grade of bitumen before the start of construction of the surface treatment.
Typical primes are:
ƒ Bitumen primes: Low viscosity, medium curing cutback bitumens such as MC-30, MC-70, or in rare
circumstances, MC-250, can be used for prime coats.
ƒ Emulsion primes: Bitumen emulsion primes are not suitable for priming stabilized bases as they
tend to form a skin on the road surface and to not penetrate this surface.
ƒ Tar primes: Low-viscosity tar primes such as 3/12 EVT are suitable for priming road surfaces but are
no longer in common use because of their carcinogenic properties which are potentially harmful
to humans and the environment.

The choice of prime depends primarily on the texture and density of the surface being primed. Low
viscosity primes are necessary for dense cement or lime stabilized surfaces while higher viscosity primes
are used for untreated, coarse-textured surfaces. Emulsion primes are not recommended for saline base
courses.

The grade of prime and the nominal rates of application to be used on the various types of pavements
are given in Table D.7.3.

Table D.7.3: Typical prime application rates in relation to pavement surface type

Prime
Pavement surface
Grade Rate of application (l/m2)
Tightly bonded
MC-70 0.6 – 0.7
(light primer)
Medium porosity
MC-30/MC-70 0.7 – 0.8
(medium primer)
Porous
MC-30 0.85 – 1.1
(heavy primer)

Sand Seal (S-09)


Design: There are no formal methods for the design of sand seals with the binder and aggregate
application rates being based on local experience.

Materials: Typical constituents for sand seals are:


ƒ Binder: The following grades of binder are typically used:
• MC-800 cut-back bitumen;
• MC-3000 cut-back bitumen;
• Spray-grade emulsion (65% or 70% of net bitumen);
• 150/200 penetration grade bitumen.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 191

ƒ Aggregate: The grading of the sand may vary to a fair degree, but the conditions of Table D.7.4
must be met:
Table D.7.4: Grading of sand for use in sand seal

Percentage by mass
Sieve size (mm)
passing through sieve
6.7 100
0.300 0-15
0.150 0-2
Sand equivalent (%): 35 Min

Application rates: For planning or tender purposes, typical binder and aggregate application rates for
sand seals are given in Table D.7.5.

Table D.7.5: Binder and aggregate application rates for sand seals

Hot Spray Rates of


Aggregate Application Rate
Application MC3000 cut-back
(m³/m²)
bitumen (l/m²)
Double sand seal used as a
1.2 – 1.4 per layer 0.010 – 0.012 per layer
permanent seal
Single sand seal used as a cover
seal over an Otta Seal or Surface 0.8 – 1.0 0.010 – 0.012
Dressing
Single seal used as a
maintenance remedy on an 1.0 – 1.2 0.010 – 0.012
existing surfaced road

Slurry Seal (S-10)


Design: The design of a Slurry Seal surfacing is based on semi-empirical methods or experience with the
exact proportions of the mix being determined by trial mixes for which the following guidelines may be
used:

Materials: The typical composition of a slurry is as follows:


ƒ Filler should be between 1% and 2% of the mass of fine aggregate.
ƒ Undiluted Bitumen Emulsion should be approximately 20% by weight of fine aggregate.

Application rates: For planning or tender purposes, the typical composition of the slurry may be based
on the mass proportions indicated in Table D.7.6.

Table D.7.6: Gives a nominal slurry seal mix.

Material Proportion (Parts)


Fine aggregate (dry) 100
Cement (or lime) 1.0 - 1.5
60% Stable grade emulsion 20
Water +/- 15

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 192

Chip Seal (S-11)


Design: The design methods for both single and double chip seals are presented in Overseas Road Note
3 (2nd edition): A guide to surface dressing in tropical and sub-tropical countries. In essence, the design
is based the concept of partially filling the voids in the covering aggregate and that the volume of these
voids is controlled by the Average Least Dimension (ALD) of the sealing chips. Corrections to the spray
rate need to be subsequently carried out to take account of site conditions as described in the guide.

Materials: Typical constituents for chip seals are:


ƒ Binder: The bituminous binder can consist of any of the following:
• 80/100 or 150/200 penetration grade bitumen;
• MC 3000 grade cutback bitumen;
• spray grade anionic (60) or cationic (65 or 70);
• Modified binders (polymer modified and bitumen rubber);
• Foamed bitumen.

ƒ Aggregate: The aggregate for a Chip Seal shall be durable and free from organic matter or any
other contamination. Typical grading requirements for Chip Seals are given in Table D.7.7.

Table D7.7: Aggregate requirements for Chip Seals

Nominal Aggregate Size (mm)


Sieve Size
19.0 13.2 9.5 6.7
(mm)
Grading (% passing)
26.5 100
19.0 85-100 100
13.2 0-30 85-100 100
9.5 0-5 0-30 85-100 100
6.7 - - 0-5 0-40
4.75 - - 0-5 0-40
2.36 - - - 0-5
0.425 (fines) <0.5 <0.5 <0.5 <2.0
0.075 (dust) <0.5 <0.5 <0.5 <1.0
Materials Properties
Flakiness Index Max 20 Max 25 Max 25 Max 30
10% FACT (dry) AADT > 1000 vpd: Min 160 kN; AADT < 1000 vpd: 120 kN
10% (wet) Min 75% of corresponding 10% FACT dry

Application rates: For planning purposes, typical binder and aggregate application rates for single Chip
Seals are given in Table D.7.8.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 193

Table D.7.8: Binder and application rates for Chip Seals

Double Chip Seal Single Chip Seal (Reseal)


Item 2nd 9.5 mm 2nd 6.7mm
13.2 mm 9.5 mm
1st 19.0 mm 1st 13.2 mm
Aggregate spread rates (m³/m²)
2nd layer 0.09 0.007
1st layer 0.015 0.011 0.012 0.010
Hot spray rates of 80/100 pen grade bitumen (l/m²)
Traffic
3.0 (total) 2.3 (total) 1.6 1.3
AADT < 200
Traffic
2.5 (total) 1.9 (total) 1.3 1.0
AADT 200-1000

Conversions from hot spray rates in volume (litres) to tonnes for payment purposes must be made for the
bitumen density at a spraying temperature of 180oC. For planning purposes, a hot density of 0.90 kg/l
should be used until reliable data for the particular bitumen is available.

Adhesion agents: The success of a bituminous seal depends not only upon the strength of the two main
constituents – the binder and the aggregate – but also upon the attainment of adhesion between these
materials - a condition that is sometimes not achieved in practice. In such a case a proprietary adhesion
agent could be used to facilitate the attainment of a strong and continuing bond between the binder and
the aggregate. The agent can be used in the aggregate pre-coating material (see below), in the binder
or in both.

Precoating agents: Surfacing aggregates are often contaminated with dust on construction sites and,
in that condition, the dust tends to prevent actual contact between the aggregate and the binder.
This prevents or retards the setting action of the binder which results in poor adhesion between the
constituents. This problem can be overcome by sprinkling the aggregate with water or, alternatively, by
using an appropriate pre-coating material which increases the ability of the binder to wet the aggregate
and improve adhesion between binder and aggregate.

A number of materials may be used for pre-coating aggregates including diesel fuel oil, cutback bitumen,
bitumen pre-coating emulsion and proprietary products.

Cape Seal (S-12)


Design: As a combination single seal + slurry seal, the design of a Cape Seal is similar to that for a Chip
Seal and Slurry Seal as described above.

Materials: Typical constituents for Cape Seals are:


ƒ Binder: As is the case with Chip Seals, a variety of binder types may be used for constructing a
Cape Seal.
ƒ Aggregate: The same requirements are required as for Chip Seals and Slurry Seals.

Application Rates: For planning purposes, typical binder and aggregate application rates for single Chip
seals are given in Table D.7.9.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 194

Table D.7.9: Nominal Application rates for single Chip Seals

Nominal ratesof application


For planning/tender purposes
Nominal size of aggregate (mm)
Binder (litres of net Aggreagte
bitumen cold per m2) (m³/m²)
13.2 0.6 110
19.0 1.1 75

Otta Seal (S-13)


General Design Principles: The design of the Otta Seal relies on an empirical approach in terms of the
selection of both an appropriate type of binder and an aggregate application rate. Full details of the
design methods are given in the Botswana Guideline No. 1: The Design, Construction and maintenance
of Otta Seals (1999).

As a general guide, the choice of binder in relation traffic and aggregate grading is given in Table D.7.10.

Table D.7.10: Choice of binder in relation to traffic and grading

Type of Bitumen
AADT (vpd) at time
of construction Medium Dense
Open Grading
Grading Grading
MC 3000
> 1000 N/A 150/200 pen. grade MC 800 in cold
weather
MC 3000
150/200 pen. Grade
100 - 1000 150/200 pen. grade MC 800 in cold
in cold weather
weather
< 100 150/200 pen. grade MC 3000 MC 800

For design purposes, preferred grading in relation to traffic

Application Rates: The following Application rates for binder and aggregates are recommended:
ƒ Binder: As a general guide, Table D.7.11 gives the hot spray rates for primed base courses.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 195

Table D.7.11: Nominal binder application rates for Otta Seal

Grading Dense
Open Medium
Type of Otta seal AADT < 100 AADT>100
Double
1.7 1.8 1.8 1.7
1st Layer
1.6 1.4 2.0 1.9
2nd Layer
Single with Sand Cover Seal
1.7 1.8 2.0 1.9
1st Layer
0.8 0.7 - 0.9
Fine sand
0.9 0.8 - 0.7
Crusher Dust/Coarse River Sand
Single 1.8 1.9 2.1 2.0
Maintenance Reseal (Single) 1.7 1.8 2.0 1.8

The following points should be noted with regard to the binder application rates:
• Hot spray rates lower than 1.6 l/m2 should not be allowed.
• Binder for the sand seal cover seal shall be MC 3000 for crusher dust or coarse river sand
and MC 800 for fine sand.
• Where the aggregate has a water absorbency of more than 2%, the hot spray rate should be
increased by 0.3 l/m2.

ƒ Aggregate: As a general guide, Table D.7.12 gives the aggregate application rates for Otta Seals.

Table D.7.12: Nominal aggregate application rates

Aggregate Application Rates (m2/ m3)


Type of Seal
Open Grading Medium Grading Dense Grading
Otta Seals 63 - 77 63 - 77 50 - 63
Sand Cover Seals 83 - 100

The following points should be noted with regard to the aggregate application rates:
• Sufficient amounts of aggregate should be applied to ensure that there is some surplus
material during rolling (to prevent aggregate pick-up) and through the initial curing period
of the seal.
• Aggregate embedment will normally take about 3 – 6 weeks to be achieved where crushed
rock is used, after which any excess aggregate can be swept off. Where natural gravel is
used the initial curing period will be considerably longer (typically 6 – 10 weeks).

7.2.5 Concrete Surfacings

Non-reinforced Concrete slab surfacing


Non-reinforced or reinforced cement concrete slab pavements can be used to provide a high strength,
durable road surface with very low maintenance requirements (Plate D.7.10). Concrete of minimum 20Mpa
quality is required to be used. Joints are required to accommodate thermal expansion and contraction.
Particular attention is required for the design and construction of these joints. When the concrete has set
the layer should be covered in sand or other moisture retaining material and kept wet for a few days to
aid curing. Concrete surfaces should normally not be trafficked until 7 days after casting. It will be difficult
to justify normal reinforced concrete paving for LVRs, however Ultra Thin Reinforced Concrete Paving may
be an affordable option see 7.2.5(b).

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 196

Plate D.7.10: Concrete slab surfacing

Ultra Thin Reinforced Concrete Paving


An Ultra Thin Reinforced Concrete Pavement (UTRCP) option has been developed in South Africa as a
low maintenance surfacing suitable for LVRs not subjected to heavy axle loading. A thin (50-60mm) layer
of reinforced concrete is used in essence as a rigid “structural surfacing” over a good sub-base layer
comprising well compacted good quality material, the top 150mm of which should have an effective CBR
of 80%. In contrast to a more conventional NRCP the pavement layers below a UTRCP must contribute
significantly to the strength of the pavement as a whole.

It should be emphasised that the formal design approach to this option is still under development and
that its use within an Ethiopian LVR road environment should be undertaken with caution.

Areas where the use of UTRCP can be considered include:


ƒ Surfacing of a new road or the rehabilitation/upgrading of an existing road;
ƒ All traffic and road classes from low-volume urban streets to inlays, to “provincial” roads where
typical traffic volumes are below 2 000 vehicles per day with less than 5% heavy vehicles (at this
stage);
ƒ Areas of steep grades and stop/start heavy traffic;
ƒ Areas where maintenance is unlikely.

The concrete is only 50-60mm thick and therefore tolerances and quality control are critical and the
success of the UTRCP process is therefore dependant on attention to detail. This applies not only to
the concrete layer (concrete strength, thickness, placing, curing) but also to the placing, supporting and
joining of the steel mesh panels as well as the tolerances of the layer supporting the UTRCP. The need
for meticulous monitoring and control during construction cannot be over-emphasised. Competent site
staff must be intensively involved in all the processes associated with and control of all the construction
activities.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 197

Plate D.7.11: Ultra Thin Reinforced Concrete Paving

7.3 Choice of pavement and surfacing


The various factors that typically affect the choice of a surfacing can be grouped under the following
headings:
ƒ Available materials;
ƒ Operational environment;
ƒ Road task;
ƒ Natural environment.

These factors are illustrated in Figure D.7.3.

AVAILABLE MATERIALS
Local Materials
Surface/Paving Options
Specifications

OPERATIONAL
ENVIRONMENT
NATURAL Construction
ENVIRONMENT OPTIMUM OR Regime
Climate APPROPRIATE Maintenance
Hydrology DESIGN Regime
Terrain Policies
Subgrade Socio-economic
factors

ROAD TASK
Traffic
Axle Loads
Standards

Figure D.7.3: General road surface selection factors

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 198

More specifically, the following factors should also be considered in short-listing surfacing types for more
detailed consideration:
ƒ Existing base/surface conditions;
ƒ Design life;
ƒ Materials (type and quality);
ƒ Safety (skid resistance - surface texture, etc.);
ƒ Riding quality required;
ƒ Maintenance (capacity and reliability).

The final selection of surfacing should then be made on the basis of life-cycle costing.

7.3.1 Evaluation framework

A rational method is required for the selection of the most appropriate surface or paving structure for a
particular section of low volume rural or urban road. The Surfacing Decision Management System (SDMS)
provides such a procedure for assessing the various factors that influence the suitability of surface-paving
options for a specific section of rural road.

When ENS or natural gravel are considered to be unsuitable options, the separate Matrices of Surfacing and
Paving Options (Tables D.7.15 to D.7.18) will further guide the user to identify the most appropriate options.
The key objective is the elimination of unsuitable or high risk options using a series of road environment
related “screens” before proceeding to Final Engineering Design (FED) for the surfacing/paving and their
Whole Life Costing. Figure D.7.4 shows the basic steps in the SDMS procedure.

1. Road Task 2. Available 3. Road


Definition Whole Life Budget Environment Data

4. ENS Road
Option Viable ?
(Sheet 1)

6. Only with
Yes No Spot
Improvement

5. Gravel Road
Option Viable ?
(Sheet 2)

Yes No 7. Only with


Spot
Improvement

8. Screen for sealed 9. Screen for Spot


and other options Improvement
(Tables D.7.5 to D.7.8) options

Phase II Detailed Design

Figure D.7.4: Overview of the SDMS procedure

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 199

7.3.2 SDMS Procedure

Step 1 of the three-step SDMS procedure is illustrated in Figure D.7.5 while each of the explanatory
sheets (Sheets 1-3) supporting the sequential activities are presented in Figures D.7.6 to D.7.8.

STEP 1 - Consideration of ENS or Natural Gravel as a Road Surface Option

ENS ENGINEERING
Sheet 1 & OPERATIONAL Whole Life Cost
(Figure D.7.6) ASSESSMENT OK? ASSESSMENT

US* US* = Unsuitable

NATURAL GRAVEL
Sheet 2 ENGINEERING
(Figure D.7.7) ASSESSMENT

OPERATIONAL
Sheet 3
ASSESSMENT
(Figure D.7.8)

Sheet 3 POLICY ASSESSMENT


(Figure D.7.8)

DECISION ON Whole Life Cost


SUITABILITY OF GRAVEL OK? ASSESSMENT

STEP 2 - Consideration of other Surfacing - Paving options

Whole Life Cost


TABLES D.7.15 - D.7.18 ASSESSMENT

Figure D.7.5: Overview of SDMS procedure

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 200

SHEET 1 - Assessment of suitability and Engineered Natural Surface - ENS (Engineered Earth Road) as a
feasible option

No upgrading option will be


Routine sustainable, consider
will sufficient FUNDING Maintenance* No
maintenance support
be available for: of the road? initiative

Yes

Can ENS will likely not be viable


is MAINTENANCE on at least 50% of the No maintenance No
Road Manager's network? due to the maintenance
effective: capacity be liability to retain passability
made effective
Yes within 2 years?

Yes

will the in situ soil and camber crossfall of 3 - 8% can be


cambered surface have maintained and side+turnout drainage No Option Inappropriate
a wet season system will be provided and maintained?
compacted CBR of >15
Yes

< 1000 1000 - 2000 > 2000


is RAINFALL: No No Yes Option Inappropriate
mm/year? mm/year? mm/year?

Yes Yes
OPERATIONAL ASSESSMENT

is longitudinal > 6%? > 4%? Option Probably


ROAD GRADIENT: Yes Inappropriate

Option Probably
No Yes Inappropriate

No

is TRAFFIC: < 100 ADT# Option Probably


No
(see ADT# Note) / day? Inappropriate

Yes

can TRAFFIC:
by natural conditions or by No Option may be Inappropriate
be prevented from using
physical barriers
the road in wet weather

Yes

by over-topping more Option Inappropriate


is road FLOODED: Yes
than one day/year?

No

ENS is Technically a feasible option


Make a Whole Life Cost assessment

For Option Inappropriate outcome proceed to Sheet 2 for Natural Gravel assessment.

Figure D.7.6: Decision Flow Chart for the Preliminary Consideration


of LVR Surface Options for a road section – STEP 1

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 201

SHEET 2 - Engineering Assessment of Natural Gravel Surface Option

is gravel of in sufficient quantities within 50km haul Option probably Inappropriate


Specification QUALITY for the construction and 10 year's No But check blending options to
available: maintenance meet specifications

Yes

< 1500 1500 - 2500 > 2500 Option Inappropriate


is RAINFALL: No No Yes
mm/year? mm/year? mm/year?

Yes Yes

is longitudinal > 6%? > 4%? Yes Option Inappropriate


ROAD GRADIENT:

No Yes Option Inappropriate

No
ENGINEERING ASSESSMENT

is wet weather in- Consider Engineered In-situ Material


is TRAFFIC: < 100 ADT# Option (Engineered Natural Surface)
Yes situ material Yes
(see ADT# Note) / day? (Sheet 1)
>15CBR?

No No

> 200 ADT#


Yes Option Inappropriate
/ day?

No

by over-topping more Option Inappropriate


is road FLOODED: Yes
than one day/year?

No

Option probably
is gravel material Inappropriate: Check by
more than 10km? Yes
HAULAGE: Whole Life Costing analysis

No

Natural Gravel is Technically


a feasible option.
Proceed to Non-technical
Assessment (Sheet 3)

Figure D.7.7: Decision Flow Chart for the Preliminary Consideration of


LVR Surface Options for a road section – Step 1 continued

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 202

SHEET 3 - Operational, Socio-economic and Economic Assessment of Natural Gravel as a surface option.

KEY CONSIDERATIONS
Who will be responsible for funding/resourcing ROUTINE maintenance of the road? ........................
Who will be responsible for funding PERIODIC maintenance of the road? ........................
Who is responsible for managing the maintenance of the road? ..
What is the annual rate of gravel loss predicted, that must be replaced by Periodic Maintenance? ..mm/year

No upgrading option will be


will sufficient FUNDING Routine sustainable, consider
No
be available for: Maintenance* maintenance support
initiative
Yes

Can Gravel will likely not be viable


OPERATIONAL ASSESSMENT

is MAINTENANCE on at least 50% of the maintenance due to the high maintenance


No No
effective: Road Manager's network? capacity be liability and additional burden
made effective
Yes within 2 years?

Yes
Gravel will not be viable as
will sufficient FUNDING Periodic Will the road be material losses will not be
Maintenance** No No replaced & road will revert to
be available for: upgraded within
2 years? (Stage earth standard
Yes Construction)

Yes

available to test & ensure the Gravel will likely not be viable
will sufficient QUALITY constructed materials comply with No unless improved Quality
ASSURANCE be: specifications? Assurance is provided

Yes

Natural Gravel is Operationally


a feasible option. Proceed to Policy
Assessment (below)

applicable to the road that will


are there any local or prejudice the use of gravel on
Yes Option probably Inappropriate
POLICY ASSESSMENT

national POLICY the grounds of dust nuisance,


considerations: pollution, resource depletion
etc?

No

Natural Gravel complies with Policy


requirements & is an acceptable option.
Proceed to Economic Assessment (below)

KEY CONSIDERATIONS
Carry out a Whole Life Costing of infrastructure improvement & maintenance costs, and road user costs for feasible paving options.

is gravel the lowest of all the technically,


WHOLE LIFE COST
ECONOMIC ASSESSMENT

operationally and No Option probably Inappropriate


option: socio-economically
feasible options?

Yes

Note: In Whole Life Costing,


Natural Gravel is an acceptable include damage to haul routes
option on Technical, Operational, Socio- caused by initial and periodic
economic & Economic grounds maintenance regravelling vehicles.

If neither ENS nor Natural Gravel is an Appropriate Option proceed to alternative surface options assessment

Figure D.7.8: Decision Flow Chart for the Preliminary


Consideration of LVR Surface – Step 1 continued

Step 2 involves the consideration of surfacing/paving options (S-01 to S-15) as listed in Section D.7.2.

If the Step 1 assessment indicates that neither ENS nor Natural Gravel are viable options for a particular
road section, then the assessment should proceed to the ‘screening’ process (see Tables D.7.15 to D.7.18)
to select a shortlist of appropriate and viable surface and/or paving options based on the evaluation
criteria included in these tables.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 203

In the screening process the Tables D.7.13 and D.7.14 set out the evaluation criteria in terms of indicative
traffic regime and erosion potential.

Table D.7.13: Definition of Indicative Traffic Regime

Indicative
Traffic Description
Category

Mainly non-motorised, pedestrian and animal modes, motorbikes & less than
25 motor vehicles per day, with few medium/heavy vehicles. No access for
Light
overloaded vehicles. Typical of a Rural Road with individual axle loads up to 2.5
tonne.

Up to about 100 motor vehicles per day including up to 20 medium (10t) goods
Moderate vehicles, with no significant overloading. Typical of a Rural Road with individual
axle loads up to 6 tonne.

Between 100 and 300 motor vehicles per day. Accessible by all vehicle types
High including heavy and multi-axle (3 axle +) trucks, Construction & timber materials
haulage routes. Specific design methodology to be applied.

Table D.7.14: Definition of Erosion Potential

Annual Rainfall (mm)


Road Alignment
Longitudinal Gradient
< 1000 1000-2500 2500-4000 >4000

Flat (< 1%) A A B C

Moderate (1-3%) A B B C

High (3-6%) B C C D

Very High (>6%) C C D D

A = Low; B = Moderate; C High; D = Very High


Note:
Areas prone to regular flooding should be classed as “High Risk” irrespective of rainfall.

In the following Tables

√ Indicates suitable for evaluation Mortared

Note:
Cost ratings are indicative only and will depend on local factors.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 204

Table D.7.15: Preliminary engineering filter - surfacing

PAVING CATEGORY BASIC STONE BR BITUMEN CONC

Fired Clay Brick Pavement: Un-/mortared Joints


Waterbround/Drybound Macadam

Ultra-thin Reinforced Concrete


Dressed Stone/Cobble Stone
Engineered Natural Surface

Non-Reinforced Concrete
Bituminous Slurry Seal
Bituminous Sand Seal

Bituminous Chip Seal


Stone Setts or Pavé
Hand Packed Stone

Mortared Stone
Gravel Surface

Cape Seal

Ottaseal
Economically
available
S01

S02

S03

S04

S05

S06

S07

S08

S09

S10

S11

S12

S13

S14

S15
Materials
Crushed stone
aggregate
√ √ √ √ √ √ √

Stone pieces/blocks √ √ √ √
Natural gravel √ √
Colluvial/alluvial
gravel
√ √

Weathered rock √
Fired clay bricks √
Clay soil √
Sand √ √ √ √ √ √ √ √
Cement √ √ √
Lime

Bitumen √ √ √ √
Bitumen Emulsion √ √ √ √
Reinforcement steel √

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
D - Chapter 7 - 205

Table D.7.16: Primary engineering filter - pavement layers / shoulders

SHOULDERS
SUB-BASES
CATEGORY
PAVING

BASES

Emulsion s tabilised soil

Emulsion stabilised soil


Waterbound macadam

Waterbound macadam
Cement stabilised soil

Cement stabilised soil

Cement stabilised soil


Drybound macadam

Drybound macadam
Lime stabilisede soil

Lime stabilisede soil

Lime stabilisede soil


Armoured gravel

Stone macadam
Natural gravel

Natural gravel

Natural gravel
Economically
available
Materials
Crushed stone
aggregate
√ √ √ √ √ √

Stone pieces/
blocks
Natural gravel √ √ √ √
Colluvial/alluvial
gravel
√ √ √ √

Weathered rock √ √ √ √
Fired clay bricks
Clay soil √ √ √
Sand √ √ √ √ √
Cement √ √ √
Lime √ √ √
Bitumen
Bitumen Emulsion √ √
Reinforcement
steel

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 7 - 206

Table D.7.17: Primary engineering filters (continued) - surfacing

PAVING CATEGORY BASIC STONE BR BITUMEN CONC

Fired Clay Brick Pavement: Un-/mortared Joints


Waterbround/Drybound Macadam

Ultra-thin Reinforced Concrete


Bituminous Chip Seal (double)
Bituminous Chip Seal (Single)
Dressed Stone/Cobble Stone
Engineered Natural Surface

Non-Reinforced Concrete
Bituminous Slurry Seal
Bituminous Sand Seal
Stone Setts or Pavé
Hand Packed Stone

Ottaseal (Double)
Ottaseal (Single)
Mortared Stone
Gravel Surface

Cape Seal
Traffic Regime:
S01
S02
S03

S04

S05
S06

S07

S08

S09

S10

S11

S11

S12

S13

S13
S14
S15
See Table D.7.3
Light traffic √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √
Moderate traffic √ √ √ √ √ √ √ √ √ √ √ √ √
Heavy traffic
(overload risk)
√ √ √ √ √
Construction Regime

High labour content √ √ √ √ √ √ √ √ √ √ √ √ √


Intermediate machinery √ √ √ √ √ √ √ √ √ √ √
Low cost √ √ √ √ √ √
Moderate cost √ √ √ √ √ √ √
High cost √ √ √ √
Maintenance Requirement

Low √ √ √ √ √ √
Moderate √ √ √ √ √ √ √ √
High √ √ √
Erosion Regime (See Table D.7.4)
A low erosion regime √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √
B Moderate
erosion regime
√ √ √ √ √ √ √ √ √ √ √

C High erosion regime √ √ √ √ √ √


D Very high
erosion regime
√ √ √ √ √ √

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


D - Chapter 7 - 207

Table D.7.18: Secondary engineering filters - pavement layers / shoulders

BASES SUB-BASES SHOULDERS

Emulsion s tabilised soil

Emulsion stabilised soil


Waterbound macadam

Waterbound macadam
Cement stabilised soil

Cement stabilised soil

Cement stabilised soil


Drybound macadam

Drybound macadam
Lime stabilisede soil

Lime stabilisede soil

Lime stabilisede soil


Armoured gravel

Stone macadam
Natural gravel

Natural gravel

Natural gravel

Sealed
Traffic Regime:
See Table D.7.3
Light
traffic
√ √ √ √ √ √ √ √ √ √ √ √ √ √

Moderate
traffic
√ √ √ √ √ √ √ √ √ √ √ √

Heavy traffic
(overload risk)
√ √ √ √ √ √

Construction Regime
High labour
content
Intermediate
machinery
√ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √ √

Low cost √ √ √ √
Moderate cost √ √ √ √ √ √ √ √ √ √ √ √
High cost √ √
Maintenance Requirement

Low √ √ √ √ √ √ √ √ √ √ √ √ √
Moderate √ √ √ √
High √
Erosion Regime (See Table D.7.4)
A Low erosion
regime
√ √ √ √ √

B Moderate
erosion regime
√ √

C High erosion
regime

D Very high
erosion regime

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 208

APPENDIX D.1

ASIST Information Service


Technical Brief No 9

Material selection and


quality assurance
for labour-based unsealed road
projects

International Labour Organisation


ADVISORY SUPPORT, INFORMATION SERVICES, AND T RAINING (ASIST)
Nairobi, Kenya

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.1 - 209

The Employment-Intensive Programme (EIP) is a sub-programme within the


Development Policies Department (POLDEV) of the ILO. Its objective is to promote
the use of local resource based technologies in infrastructure works in developing
countries and to strengthen their capacity to apply such technologies.

ASIST is a sub-regional programme under the EIP, one of whose objectives is to


achieve an improved effectiveness of road construction, rehabilitation and
maintenance in Sub-Saharan Africa and thereby promote employment and income
generation in the rural and urban areas.

The aim of ASIST Technical Briefs is to spread knowledge about labour-based


technology and management amongst policy makers, planners, designers,
implementers and trainers.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 210

Material selection and


quality assurance
for labour-based unsealed road
projects
First edition

This publication was developed by the


ASIST technical team in Harare, Zimbabwe,
and Nairobi, Kenya.

Written by Dr P Paige-Green of the CSIR


Division of Roads and Transport
Technology, Pretoria, South Africa (Contract
Report CR-97/047).
Editing and layout by David Mason.

International Labour Organisation


Advisory Support, Information Services, and Training (ASIST)
Nairobi, Kenya

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 211

Copyright ” International Labour Organisation 1998

Publications of the International Labour Organisation enjoy copyright under Protocol 2


of the Universal Copyright Convention. Nevertheless, short excerpts from them may be
reproduced without authorisation, on condition that the source is indicated. For rights
of reproduction, adaptation or translation, application should be made to the
ILO/ASIST Information Service, PO Box 60598, Nairobi, Kenya, or to ILO Publications
Branch (Rights and Permissions), International Labour Office, CH-1211 Geneva 22,
Switzerland. Both ASIST and the International Labour Office welcome such
applications.

Technical Brief No. 9: Material selection and quality assurance for labour-based
unsealed road projects

First published 1998

Produced by ASIST Information Service


with financial support from the
Swiss Agency for Development and Cooperation (SDC)

Set in Arial and Century Schoolbook typefaces


on an Hewlett-Packard LaserJet 4000TN printer
using Microsoft Word for Windows 8.0

ASIST publications can be obtained direct from ASIST Information Service,


PO Box 60598, Nairobi, Kenya, Tel +254-2-572555, Fax +254-2-566234,
email: iloasist@arcc.or.ke, Website: http://iloasist.csir.co.za

The designations employed in ILO publications, which are in conformity with United
Nations practice, and the presentation of material therein, do not imply the expression
of any opinion whatsoever on the part of the International Labour Office concerning the
legal status of any country, area or territory, or of its authorities, or concerning the
delimitation of its frontiers. The responsibility for opinions expressed in signed
articles, studies and other contributions rests solely with their authors, and
publication does not constitute an endorsement by the International Labour Office of
the opinions expressed in them. Reference to names of firms and commercial products
and processes does not imply their endorsement by the International Labour Office,
and any failure to mention a particular firm, commercial product or process is not a
sign of disapproval.

This is not an official ILO document

2 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 212

Contents
Acknowledgements.............................................................................4
List of abbreviations and definitions ........................................5
1 Background .................................................................................7
2 Material specifications ...........................................................8
2.1 RECOMMENDED SPECIFICATIONS....................................................................................8
3 Borrow material testing .......................................................10
3.1 GENERAL ............................................................................................................................10
3.2 TEST REQUIREMENTS.......................................................................................................10
3.2.1 Grading..............................................................................................10
3.2.2 Shrinkage...........................................................................................11
3.2.3 Aggregate hardness ............................................................................11
3.2.4 Material strength................................................................................11
3.3 FREQUENCY OF LABORATORY TESTING........................................................................13
3.4 FULL TEST METHODS.......................................................................................................14
3.4.1 Sieve analysis for grading coefficient...................................................14
3.4.2 Determination of the linear shrinkage of soils......................................15
3.4.3 The determination of the Treton impact value of aggregate..................18
4 Construction quality assurance testing .......................21
4.1 MATERIAL TESTING AND CONTROL...............................................................................21
4.2 CONSTRUCTION QUALITY ASSURANCE..........................................................................21
4.2.1 Moisture content ................................................................................21
4.2.2 Thickness ...........................................................................................23
4.2.3 Compaction ........................................................................................23
4.2.4 Visual inspection ................................................................................24
5 Thickness design .....................................................................25
6 Conclusions and recommendations ................................27
6.1 CONCLUSIONS ...................................................................................................................27
6.2 RECOMMENDATIONS ........................................................................................................27
7 References ..................................................................................28
Annex A: Guideline document .....................................................29
Annex B: Contents of CSIR testing kit ....................................47

LIST OF T ABLES
Table 2.1: Material specifications for labour-based road projects...... 9
Table 3.1: Penetration rates of DCP and RCCD ................................. 12
Table 5.1: Recommended thicknesses and material strengths......... 26

LIST OF FIGURES
Figure 3.1: DCP test apparatus............................................................. 13
Figure 3.2: Diagram of RCCD device ................................................... 13
Figure 3.3: Mould for bar linear shrinkage test.................................. 16
Figure 3.4: Preparation of material for shrinkage test...................... 17
Figure 3.5: Treton apparatus................................................................. 20
Figure 4.1: Thickness probe................................................................... 24

ILO/ASIST Technical Brief No. 9: Material Selection 3

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 213

Acknowledgements

This document is part of an occasional series of technical


briefs produced by ILO/ASIST to synthesise and summarise
technical information on important aspects of labour-based
technology.
The original work for this brief was undertaken under
contract by Dr P Paige-Green of the Division of Roads and
Transport Technology (Transportek) of the CSIR in Pretoria.
He produced a report, which was subsequently edited to
produce this brief.
Transportek have also put together a Gravel Road Test Kit
(see Annex B) for use with this brief. This kit was
demonstrated by them at the Sixth Regional Seminar for
Labour-based Practitioners, held in Jinja, Uganda, in 1997.
ASIST plans to evaluate the kit during 1999.

4 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 214

List of abbreviations and definitions

AASHTO American Association of State Highway and


Transportation Officials. This association adopted a test
proposed by the US War Department in 1943, to cater for
larger earth moving and compaction equipment. It is now
referred to as the AASHTO test.
ASIST Advisory Support, Information Services and
Training for labour-based technology.
Atterberg limits Atterberg limits are measured for soil
materials passing the No. 40 sieve: the shrinkage limit (SL) is
the maximum water content at which a reduction in water
content will not cause a decrease in the volume of the soil
mass. This defines the arbitrary limit between the solid and
semisolid states. The plastic limit (PL) is the water content
corresponding to an arbitrary limit between the plastic and
semisolid states of consistency of a soil. The liquid limit (LL)
is the water content corresponding to the arbitrary limit
between the liquid and plastic states of consistency of a soil.
BLS bar linear shrinkage.
BS British Standard. BS 1377 defines the British Standard
compaction test, introduced by R. R. Proctor in 1933. It used
a compactive effort which roughly corresponded to that
available in the field at the time.
CBR California Bearing Ratio. A measure of soil strength,
determined from the load required to penetrate the surface
of the compacted soil, expressed as a percentage of a
standard value.
Clegg Hammer A simple device utilising a
decelerometer, installed in a modified Proctor compaction
hammer, to evaluate the stiffness of a material by measuring
the deceleration encountered when the falling hammer
meets the material.
DCP Dynamic Cone Penetrometer. Apparatus for
estimating the in situ shear strength of a material by
dynamically driving a standard cone through the material.
Grading Coefficient (G c) A measure of the potential for
particle interlock defined by the product of the gravel
component of the material (the percentage retained between
the 26.5 and 2 mm sieves) and the percentage passing the
4.75 mm sieve.
Maximum dry density (MDD) The maximum dry
density which can be achieved under a specified compaction
effort at the optimum moisture content.

ILO/ASIST Technical Brief No. 9: Material Selection 5

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 215

Optimum moisture content (OMC) The moisture


content at which the maximum dry density for any
combination of material and compaction effort is obtained.
The importance of this is particularly relevant to labour-
based projects as the compaction effort using small
pedestrian rollers can seldom be equated to the traditional
AASHTO and BS compaction efforts. Higher OMCs will often
be necessary to achieve maximum density for these efforts.
Oversize index (I o) The stoniness as defined by the
percentage of material larger than 37.5 mm.
Proctor Mr R. R. Proctor was the author of the original BS
compaction standard. The compactive effort is supplied by a
2.5 kg hammer with a 50 mm diameter head falling freely
from 300 mm above the top of the soil sample.
Rapid Compaction Control Device (RCCD) A simple
impact penetrometer which injects a small cone into the
material to estimate the shear strength of the material.
Ravelling A process where the surface material of a road
is broken down by traffic to form loose material (e.g. gravel).
The process is likely to occur where there is a deficiency of
fine material, low cohesion between particles, poor particle
size distribution, and inadequate compaction.
Shrinkage Product (S p) A measure of the plasticity of
the soil defined by the product of the bar linear shrinkage
and the percentage passing the 0.425 mm sieve.
vpd Vehicles per day. That is, a count of the number of
vehicles passing along a road in one day.

6 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 216

1 Background

The implementation of labour-based construction techniques


for unsealed roads is beneficial in that it creates employment
opportunities and assists with the development of small
contractors whilst upgrading the transportation network in
developing countries. Improvement in the techniques
utilised during this type of construction project will result in
greater cost-effectiveness and better performance of the
completed product. The Advisory Support, Information
Services and Training (ASIST) programme of the
International Labour Organisation (ILO) has taken the lead
in this. With funding from the Swedish International
Development Cooperation Agency (Sida), ASIST is currently
involved in the implementation of innovative technologies on
various labour-based road projects in Zimbabwe and in other
countries of Sub-Saharan Africa.
During an earlier visit to labour-based projects in progress in
Zimbabwe, the Division of Roads and Transport Technology
(Transportek) of the CSIR in Pretoria was contracted by
ASIST to evaluate the procedures used regarding material
selection, testing and control. In the second phase of the
contract, the brief was to prepare a short guideline document
on the selection and control of borrow materials, and on
control of the construction process during labour-based
unsealed road projects. Recommendations on the thickness
design of the road are also provided.
The guidelines themselves are incorporated as an Annex to
this document. The background to the decisions as to what is
incorporated in the guidelines, with justification for these
decisions, makes up the main text of this Technical Brief.

ILO/ASIST Technical Brief No. 9: Material Selection 7

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 217

2 Material specifications
2.1 RECOMMENDED SPECIFICATIONS
The performance of any unsealed road is primarily a function
of the materials from which the road is constructed. It is
therefore essential that the best available materials which
comply with, or are as close as possible to, the appropriate
material requirements be used for construction. These
material requirements need, of necessity, to be simply and
rapidly determined at low cost to allow sufficient samples to
be tested prior to use on the road.
Numerous material specifications have been developed and
utilised over time in various countries, which take into
account the local material and environmental conditions. Most
specifications, however, have been derived from the original
AASHTO requirements which are primarily based on
theoretical considerations for maximum particle packing of
low plasticity materials (the dominant material derived from
glacial tills in the northern United States). Experience has
shown that materials with low plasticity lack adequate
cohesion to resist ravelling, or the formation of corrugations,
under traffic.
Regional specifications were subsequently adapted to allow
for slightly higher plasticities, but in very few cases was the
lower limit for plasticity specified. For this reason, variable
success was obtained using the available specifications.
Various projects to determine performance-related
specifications for unsealed road materials were therefore
carried out in South Africa and Namibia during the 1980s and
early 1990s (see References 1, 2, 3, and 4). These
specifications have recently been evaluated in a number of
regions and countries (including Zimbabwe) and have
generally been found to be more appropriate than those
previously used (and in many cases more appropriate than
even those currently used).
The traditional properties used in existing material
specifications for unsealed roads are particle size
distribution, Atterberg limits, remoulded strength, and
aggregate hardness. These are similar to those found by local
research to be necessary. All these parameters are critical to
the performance of materials in unsealed roads, but the
traditional methods of defining and evaluating them are
considered to be inappropriate for labour-based projects.
This is discussed further in this brief.
The material specifications recommended for the selection of
borrow materials for wearing courses for unsealed roads using
labour-based construction methods are given in Table 1. These
should be the desired specifications for a project. Testing of all
potential borrow materials for compliance with these should
optimally be carried out during the borrow pit or initial
materials evaluation. This should be done by a central
laboratory for a Public Authority, or by a commercial
laboratory, using traditional test methods, e.g. those contained
in TMH 1 (see References 6 and 7).

8 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 218

If no central laboratory is available, or if the results from the


laboratory are delayed, all the above testing except the
soaked CBR can be determined in a simple field ‘laboratory’
using minimal equipment. If full testing facilities do not exist
and if the road is likely to carry less than 50 vehicles per day
with less than 10 per cent heavy vehicles, the Shrinkage
Product and Grading Coefficient alone can be taken as the
preliminary acceptance criteria. The material strength (CBR)
can be evaluated during proof rolling trials as discussed in
Chapter 3. Full test methods are provided in Chapter 3.

Table 2.1: Material specifications for labour-based road projects


Maximum size (mm) 37.5
Oversize Index (Io) • 5%
Shrinkage product (S p) 100 – 365
Grading coefficient (Gc) 16 – 34
Soaked CBR (%) • 15 at 95 % Modified AASHTO density
Treton Impact value (%) 20 – 65
Io = Percentage retained on 37.5 mm sieve
Sp = Bar Linear shrinkage × per cent passing 0.425 mm sieve
Gc = (Per cent passing 26.5 mm − per cent passing 2.0 mm) × per cent
passing 4.75 mm/100
Treton Impact Value (see Section 3.4.3)

The relationship between the Shrinkage Product and the


Grading Coefficient is directly related to the performance as
shown in Figure 1, with zones E1 and E2 being the
recommended areas for best performance. This Figure shows
the predicted performance and the implications (potential
problems) of using material not falling within the specified
limits.

Figure 2.1: Relationship between Grading Coefficient,


Shrinkage Product, and performance

ILO/ASIST Technical Brief No. 9: Material Selection 9

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 219

3 Borrow material testing


3.1 GENERAL
It is essential that, during the initial proposal stage for any
project, suitable borrow pits are located, the materials are
adequately tested for compliance with the specifications
given in Chapter 2, and the suitable borrow areas are
carefully delineated in the field. In most cases, this should be
carried out by the regional soils laboratory, as far as possible
using traditional test methods and equipment as discussed in
Chapter 2. Problems have, however, been encountered in the
past, with the test results often only becoming available after
construction has commenced. The following methods are
proposed for control testing of materials during construction,
but could also be used to replace or complement the initial
borrow investigations where problems with obtaining results
in time are encountered.

3.2 T EST REQUIREMENTS


Traditional test techniques have been developed, based on
the assumption that various basic services and facilities are
available. On many labour-based projects, certain simple
assumptions, such as that electricity and running water will
be available, are invalid. The test techniques and methods
summarised in this chapter and in Annex A allow for these.
As far as possible, solar energy, local water (preferably
potable), and unsophisticated equipment are utilised. It is
assumed that everyday objects such as batteries are
available.
The specifications discussed in the previous chapter are
mostly based on simple tests which can be carried out rapidly
on site using minimal equipment. The following parameters
should be evaluated:
x Grading
x Shrinkage
x Aggregate hardness
x Material strength.
These tests are carried out as follows, with the complete
methods of non-standard tests being presented in Section 3.4.

3.2.1 Grading

The grading requirements for the characterisation of


material for unpaved roads are based on only five sieve sizes,
that is 37.5 mm, 26.5 mm, 4.75 mm, 2 mm, and 0.425 mm. For
the testing, the material needs first to be dried1, the mass
determined, and then the material sieved (manual shaking)
through the recommended sieves above with a soft brush

1 Air drying in direct sunlight is adequate for most materials which are
potentially suitable, although the use of a solar oven is recommended.
10 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 220

being used where necessary. The mass of each portion is


determined. The oversize index, grading coefficient and
percentage passing the 0.425 mm sieve can then be
determined. It should be noted that the influence of the
hygroscopic moisture content on the parameters determined
is negligible. The fraction passing the 0.425 mm sieve should
be retained for shrinkage testing.

3.2.2 Shrinkage

The bar linear shrinkage test is carried out on the fraction


passing the 0.425 mm sieve. The material should be
moistened until it is at or very near the liquid limit (this can
be checked with a simple fall-cone device (see Section 3.4)),
placed in the mould, and oven-dried at 105°C until all
shrinkage has stopped. The length of the sample is then
measured and the percentage shrinkage calculated. It is
recommended that the sample is dried for at least 12 hours
(overnight if not done in a solar oven), but experience has
shown that this can take as little as four or five hours,
depending on the soil. The length of time necessary can be
checked by drying to constant mass. However, preliminary
research has shown that air-drying of samples is not effective
for repeatable results.

3.2.3 Aggregate hardness

Aggregate hardness measurements are necessary to identify


those materials which will disintegrate under rolling or
traffic, as well as those which are excessively hard and will
result in a rough road if too much of this type of material is
included. The Treton test is used to determine this. The
Treton impact value is determined by means of a simple
impact hammer action on a single sized sample (obtained
during the sieve analysis). This test is unnecessary if the
road is unlikely to carry many buses or heavy vehicles (more
than two per day) or if the material lacks a significant
proportion of medium to coarse gravel (< 15 per cent
retained on a 16 mm sieve).

3.2.4 Material strength

Material strength is an indication of the capacity of the


material to support the wheel loads of the traffic using the
road. The traditional method for determining this property is
the soaked California Bearing Ratio (CBR) test. This test is
routinely carried out in a central or typical site laboratory
but is expensive to set up, requires a large amount of
equipment, and is relatively time consuming.
As an alternative, it is considered more practical to first
carefully compact a sample of the material, at the estimated
optimum moisture content, to the required thickness on a
subgrade prepared to the same standard as that which will
be used in construction. Then to measure the resistance to
penetration with a Dynamic Cone Penetrometer (DCP) (see
Figure 2), a Rapid Compaction Control Device (RCCD) (see
Figure 3) or a Clegg Hammer. The moisture content at the
time of testing (assumed to be at or about OMC) should be
ILO/ASIST Technical Brief No. 9: Material Selection 11

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 221

taken into account (See Section 4.2.1). Acceptable values of


penetration for the DCP and RCCD are given in Table 3.1.

Table 3.1: Penetration rates of DCP and RCCD


for equivalent soaked CBR values of 15 % (tested at OMC)
Apparatus Penetration Penetration Penetration
rate (3 blows) (20 blows)
(mm/blow) (mm) (mm)
DCP • 5 • 15 • 100
RCCD • 9 • 27 —

12 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 222

Figure 3.1: DCP test apparatus

Figure 3.2: Diagram of RCCD device

The RCCD is recommended for use since the test is simpler


and quicker. The apparatus is more robust (only periodic
calibration of the spring is necessary), but less bulky than
that required for the other two methods of control, and it has
less operator variability. More tests per job lot (day’s
production) can be carried out more economically with the
RCCD than with the other methods. However, the DCP
penetration rates given in Table 3.1 can also be used for
material characterisation and control purposes.
Selection of materials based on the specified Gc and Sp will in
most cases exclude those which are likely to have insufficient
CBR strength.

3.3 FREQUENCY OF LABORATORY TESTING


The frequency of testing of borrow pits needs to strike a
balance between cost and time and statistical validity of the
results. It is proposed that, even for labour-based projects,
the location of borrow materials and borrow-pit testing
should preferably be done according to traditional methods.
If full laboratory facilities are not available, the methods
described in this report can be substituted.
The frequency of testing will depend on the variability of the
material: the more homogeneous the material the less the
amount of testing necessary for statistical validity of the
results. Unless proper testing of the borrow materials is
carried out prior to commencement of the project, it is
usually not possible to quantify the variability in advance of
ILO/ASIST Technical Brief No. 9: Material Selection 13

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 223

construction. For projects of this nature it is thus necessary


to test samples from at least five locations per borrow pit
(covering the full depth of the layer to be used) in order to
quantify the variability. The sample locations should be
randomly selected within the pit. This variability is used as
an indication of the variation to be expected within the
borrow pit, and for a simple process control technique during
the construction operation.
It is recommended that at least ten RCCD or DCP tests (at
least two per square metre) be carried out at points selected
in a stratified random pattern when compaction is tested
during proof rolling.

3.4 FULL TEST METHODS

3.4.1 Sieve analysis for grading coefficient

SCOPE

In this method, a soil, sand or gravel sample is separated by


dry sieving for determination of the grading coefficient and
to prepare fine material for the bar linear shrinkage test.

APPARATUS
x Sheet of canvas 1 metre by 1 metre for coning and
quartering of the material
x The following test sieves: 37.5 mm, 26.5 mm, 4.75 mm,
2 mm and 0.425 mm with pan and cover
x Balance with pan, accurate to 1 g, to weigh up to 5 kg
x Various pans of 250 to 300 mm diameter and 20 mm deep
x Drying oven (Solar) to maintain a temperature between
105 and 110°C
x Various stiff brushes
x Thermometer (0 to 120°C)

METHOD

Size of sample The size of the test sample should be such


that at least 100 g of material passes the 0.425 mm sieve, but
not less than 2 kg in all. This should be prepared from a bulk
sample of at least 5 kg by coning and quartering on the
canvas sheet.
Preparation of the sample Air-dry the sample until it
is friable and particles separate with ease. If the sample is
still too wet, it should be dried in an oven at a temperature
not exceeding 50°C.
Dry sieving Dry sieve the material as follows: shake the
material through each sieve in turn, starting at the 37.5 mm
sieve, until further shaking results in minimal additional
material passing each sieve. The larger particles (> 4.75 mm)
should be brushed with a stiff bristle brush to remove all
fines adhering to them. Determine the mass of the soil fines
(< 0.425 mm) and transfer these to a marked paper bag. It is
14 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 224

recommended that they are dried in a solar oven prior to


being weighed but, if this will delay the testing, this step may
be omitted since the use of air-dried weights will not affect
the results unduly.
Determination of the particle size distribution The
masses of the individual fractions retained on each sieve
should be determined (preferably after being oven-dried but
after air-drying if necessary). The masses of these fractions
should be determined to the nearest 1 g. Record the masses
retained on each sieve and that of the material passing the
0.425 mm sieve.

CALCULATIONS

1 Calculate the total mass of material as the sum of the


masses retained on the individual sieves as well as of that
passing the 0.425 mm sieve.
2 Calculate the cumulative percentages passing each sieve
(by mass of the total dry sample) accurately to the nearest
1 per cent. All results should be normalised to 100 per cent
passing the 37.5 mm sieve by multiplying the percentage
passing each sieve by the percentage passing the 37.5 mm
sieve (P37) divided by 100. If 100 per cent passes the 37.5 mm
sieve, this step is not necessary.
3 Calculate the grading coefficient. This is the percentage
material passing the 26.5 mm sieve and retained on the 2 mm
sieve, multiplied by the percentage passing the 4.75 mm
sieve, as follows:
GC = (P26  P2) u P475/100

where P26 = cumulative percentage passing the 26.5 mm sieve


P2 = cumulative percentage passing the 2 mm sieve
P475 = cumulative percentage passing the 4.75 mm sieve

3.4.2 Determination of the linear shrinkage of soils

SCOPE

This method covers the determination of the linear


shrinkage of soil when it is dried from a moisture content
equivalent to the liquid limit to the oven-dry state.
Definition
The linear shrinkage of a soil, for the moisture content
equivalent to the liquid limit, is the decrease in one dimension,
expressed as a percentage of the original dimension of the soil
mass, when the moisture content is reduced from the liquid
limit to an oven-dry state.

APPARATUS
x A shrinkage mould made from 10 mm stainless steel bar
with internal dimensions of 150 mm ± 0.25 mm long u
10 mm ± 0.25 mm wide u 10 mm ± 0.25 mm deep, and open
on two sides (see Figure 3.3)
x A stainless steel plate to fit under the shrinkage mould

ILO/ASIST Technical Brief No. 9: Material Selection 15

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 225

x A small thick-bristle paint brush, about 5 mm wide


x Silicone lubricant spray (e.g. Q20 or WD 40)
x A spatula with a slightly flexible blade about 100 mm long
and 20 mm wide
x A solar drying oven
x A pair of dividers and a millimetre scale
x A standard cup, drop cone and guide-tube for estimating
the liquid limit
x A thermometer (0 to 120 °C).

Figure 3.3: Mould for bar linear shrinkage test

METHOD

Waxing the mould The interior of a clean, dry shrinkage


mould is sprayed evenly with the silicone lubricant
Filling the mould The moisture content at which the
test is carried out must be as close to the liquid limit as
practically possible. A simplified drop-cone device based on
the British Standard liquid limit method is used to ensure
that the moisture content is correct. Sufficient material to fill
the cup provided should be mixed up and placed evenly in
the cup to a level between 2 and 5 mm below the rim of the
cup. The cone should be placed in the guide tube on the
surface of the soil in the cup and allowed to penetrate for five
seconds. The cone should penetrate to a depth of 20 mm,
equivalent to the calibration mark on the cone. If the
penetration is below this mark, the material is too dry and
additional water is required. The material would then need
thorough re-mixing before the penetration test is repeated. If
the penetration is too high (i.e. the cone sinks into the
material to a depth above the calibration mark), the material
is too wet and needs to be dried out by mixing in sunlight
until repetition of the penetration test gives a result within
the defined limits.
The lubricated mould should be placed on the plate provided,
and one half should be filled with the moist soil by taking
16 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 226

small pieces of soil on the spatula and pressing the soil down
against one end of the mould. Then work along the mould
until the whole side is filled and the soil forms a diagonal
surface from the top of one side to the bottom of the opposite
side (see Figure 3.4(a)).
The mould is now turned round and the other portion is
filled in the same manner (see Figure 3.4(b)). The hollow
along the top of the soil in the mould is now filled so that the
soil is raised slightly above the sides of the mould (see Figure
3.4(c)). The excess material is removed by drawing the blade
of the spatula once only from one end of the mould to the
other. The index finger is pressed down on the blade so that
the blade moves along the sides of the mould (see Figure
3.4(d)). During this process the wet soil may pull away from
the end of the mould, in which case it should be pushed back
gently with the spatula.
NB The soil surface should on no account be smoothed or
finished off with a wet spatula.

Figure 3.4: Preparation of material for shrinkage test

Drying the wet material The mould with wet material


is now placed in the solar oven and dried at a temperature of
between 105 and 110°C (the lid may need to be partially
opened to maintain a reasonably constant temperature) until
no further shrinkage can be detected. As a rule, the material
is dried out for 12 hours, although three hours should be
ILO/ASIST Technical Brief No. 9: Material Selection 17

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 227

sufficient time in the oven. The mould (with the material) is


taken out of the oven and allowed to cool in the air.
Measuring the shrinkage It may be found that the ends
of the dry soil bar have a slight lip or projecting piece at the
top. These lips must be removed by abrading with a sharp,
narrow spatula, so that the end of the soil bar is parallel to
the end of the mould (see Figure 3.4 (e)). If the soil bar is
curved, it should be pressed back into the mould with the
fingertips so as to make the top surface as level as possible.
The loose dust and sand removed from the ends, as well as
any loose material between cracks, should be emptied out of
the mould by carefully inverting it whilst the material is held
in position with the fingers. The soil bar is then pressed
tightly against one end of the mould. It will be noticed that
the soil bar fits better at one end than at the other end. The
bar should be pressed tightly against the end at which there
is a better fit. The distance between the other end of the soil
bar and the respective end of the mould, is measured by
means of a good pair of dividers, measuring on a millimetre
scale, to the nearest 0.5 mm, and recorded.

CALCULATIONS

The bar linear shrinkage (BLS) is calculated as follows:


BLS = LS u 0.67 (%)

where LS = linear shrinkage in mm.

NOTES

After being tested, the soil bar should be examined to ensure


that the corners of the mould were filled properly and that
no air pockets were contained in the soil bar. If there are air
pockets, the test should be repeated.

3.4.3 The determination of the Treton impact value of aggregate

SCOPE

The Treton value is an indication of the resistance of


aggregate to impact. The aggregate is subjected to ten blows
of a falling hammer and the resulting disintegration is
measured in terms of the quantity passing the 2 mm sieve,
which is then expressed as a percentage of the mass of the
test sample. This is called the Treton value.

APPARATUS
x A Treton apparatus consisting of a base plate, anvil,
cylinder, and a hammer weighing 15 kg r 50 g (see Figure
3.5). The base plate should be placed on a firm concrete
block.
x The following test sieves, all 200 mm in diameter:
19.0 mm, 16.0 mm and 2 mm. The bigger sieves should be
made of perforated plate and the 2 mm sieve of wire
mesh.
x A balance to weigh up to 200 g, accurate to 1 g.
18 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 228

METHOD

From the field sample, screen out a sufficient quantity (at


least 200 g) of the fraction between 19 mm and 16 mm (see
Note (i)). Select a sample of 15 to 20 of the most cubical
pieces, so that their total mass (in grams) will be as close as
possible to 50 times the relative density of the aggregate in
grams (it is not necessary to determine the relative density.
An estimate will be satisfactory (2.65 for granitic and
sedimentary materials and 2.9 for dark basaltic and
metamorphic materials)). Weigh the sample accurately to 1 g,
and place the particles as evenly spaced as possible on the
anvil in such a manner that their tops are approximately in
the same horizontal plane.
Place the cylinder over the anvil and tighten the clamp
screws. Place the hammer in the cylinder so that the top of
the hammer is level with the top of the cylinder and let it
drop ten times from this position.
Remove the cylinder, and sieve all the aggregate on the anvil
and base plate thoroughly through a 2 mm sieve. Weigh the
aggregate retained on the sieve to the nearest 1 g, and record
the mass. The test should be carried out in triplicate (see
Note (ii)).

CALCULATIONS

Calculate the Treton value to the first decimal place as


follows and report to the nearest whole number:
Treton value = (A  B)/A u 100

where A = the mass of the stone particles before tamping (g)


B = the mass of the stone particles retained on the 2 mm
sieve after tamping (g).

NOTES

(i) If the aggregate is noticeably variable as regards type or


hardness, each type should be tested and reported
separately. In this case an estimate should be made of the
percentage of each type and a weighted average determined.
(ii) The Treton value, as reported, should be the average of
three determinations. If any individual result differs from
the others by more than five units, further tests should be
carried out.

ILO/ASIST Technical Brief No. 9: Material Selection 19

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 229

Figure 3.5: Treton apparatus

20 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 230

4 Construction quality assurance testing


4.1 M ATERIAL TESTING AND CONTROL
The properties of the material should be tested on a regular
basis during construction to ensure that they do not differ
from the accepted specification. For control testing purposes,
only the grading coefficient and shrinkage product need be
tested.
This testing shall be done daily to ensure that the material to
be used for the following job lot complies with the
specifications. It is recommended that samples of the
material to be used the next day (or for the next job lot if a
weekend follows) be taken during the morning and tested so
that the material can be approved first thing in the morning
before use. The test techniques are such that this is possible.
The individual results of the borrow pit testing should be
plotted on Figure 2.1 with the mean and standard deviations
of the two parameters which can be used to define a
rectangle. At least 90 per cent of the routine daily test
results should plot in this rectangle as work in the borrow
pit progresses. The test results should be plotted on this
figure on an ongoing basis. If there is a trend to move out of
the rectangle towards the limits of the E1/E2 block (in Figure
2.1), this would be indicative of a change in the material
properties. Additional testing should then be carried out to
determine the cause of this and to identify remedial action,
e.g. blending of different materials, redefinition of the
boundaries of the borrow pit, or adjustment of the depth of
excavation, etc.

4.2 CONSTRUCTION QUALITY ASSURANCE


A number of factors should be controlled during construction.
These include:
x Moisture content
x Thickness
x Compaction
x General finish.

4.2.1 Moisture content

One of the principal factors in the construction process, and


which affects the final compaction, is the moisture content.
In most soil materials the natural variation in optimum
moisture content (OMC) is wider than the limits around
OMC permitted for successful compaction. In addition, the
actual process of adding and mixing water to soil materials,
particularly in labour-based projects, often leads to
significant variation of the moisture content within the
material. In addition, most moisture content determinations
are slow (except for nuclear methods, but these are often
unreliable for moisture contents of natural gravels) and the

ILO/ASIST Technical Brief No. 9: Material Selection 21

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 231

results are frequently only available after compaction is


completed. For this reason, the manual control of the
moisture after laboratory calibration of the “feel” of the
material at various moisture contents at and around
optimum is considered the most practical and effective
solution.
In most cases, the test techniques for moisture content
render the results practically meaningless in the context of
labour-based construction. The process of moistening the
material, (whether this is done in the borrow-pit or on the
road) is not discussed in this report.
The control of moisture during construction should be
carried out visually by squeezing a sample of the material as
tightly as possible in the hand. The material should be moist
enough to stick together when squeezed without any visible
sign of free water on the surface. If the material
disintegrates, it is too dry for compaction. If free water is
ejected or if the soil sticks to the hand, it is too wet. If the
“sausage” formed by squeezing in the hand is squeezed
diametrically between the thumb and forefinger, it should
break with some crumbling. It should not break by
deformation under the finger pressure, nor should there be
excessive crumbling. It should be noted that non-cohesive
soils behave differently, but that all materials for wearing
courses should have some cohesion. The above technique is
considered most practical and suitable for the purpose. If
possible, this method should be practised in the laboratory
with material at various known moisture contents, and
correlated with the laboratory determined optimum
moisture content to “get the feel” prior to commencement of
compaction.
It is currently difficult to correct the field strength for any
deviation from the expected moisture content at the time of
testing. The following approximate model is based on the
combination of various parameters (soaked CBR (CBRs) and
optimum moisture content (OMC)) and models. It has been
developed to assist with evaluating whether the results are
in the right range for the DCP and RCCD penetration rates
(DNc and RCCDc) immediately after compaction at
compaction moisture content (CMC):

CMC
§ 1˜33OMC · 0˜787
DN c 0 ˜ 144 e
¨ ¸ CBRs0˜46
¨ ¸
© ¹

CMC
§ 1˜33OMC ·  0˜775
RCCDc 0 ˜ 0735 ¨¨ e ¸ CBRs0˜46
¸
© ¹

22 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 232

4.2.2 Thickness

It is important that the thickness of the material be closely


controlled. This should be controlled prior to compaction,
with allowance for the bulking factor. Spreading of the
material should be as consistent as possible to ensure that all
material is placed at a similar loose density. The bulking
factor is usually between 25 and 35 per cent, depending on
the gradation of the material and on the compaction effort,
but this should be determined accurately during the initial
proof rolling of the material.
Control of the thickness during construction is carried out by
inserting a calibrated probe (Figure 4.1) through the
uncompacted material to confirm that the thickness prior to
compaction is equivalent to the required final layer thickness
plus a correction for bulking. The bulking factor should be
determined during the proof rolling by measurement of the
thickness before and after rolling. It can also be estimated by
comparison of the mass of a known volume (best done in a
large measuring cylinder) with the maximum dry density of
the material determined in the laboratory. In most cases a
bulking value of 30 to 35 per cent may be assumed. One
advantage of knowing the bulking factor accurately is that
this can be used to ensure that adequate compaction has been
achieved by monitoring the initial and final thickness of the
layer.
Thickness before compaction = design thickness u (1 + BF (%) / 100)

where BF is the percentage bulking factor for the material.

4.2.3 Compaction

The compaction achieved in the field is arguably the most


important aspect of the construction process. It is neither
economically nor practically possible to determine sufficient
densities on labour-based projects for construction quality
control, to take into account the natural variability of the
material. It is thus recommended that a simple device such
as the Rapid Compaction Control Device (RCCD) or DCP be
used for this purpose. Both tests are quick and repeatable
and many tests can be done at little cost.
Should the test results show up areas which are
unacceptable, the reasons for this should be investigated.
Poor results can be attributed to the use of material that is
too wet, material that has not received adequate compaction
effort, or to the presence of a pocket of poor quality material.
The actual cause should be identified and corrective action
taken. This may involve scarifying and drying out prior to
recompaction if the material is too wet, additional
compaction if necessary, or replacement of poor material
where appropriate.
A simple method of sand replacement density determination
can be carried out, but it is difficult to relate this to a
standard for the evaluation of relative compaction in highly
variable natural materials. To determine relative
compaction, the density in the road should be compared with
ILO/ASIST Technical Brief No. 9: Material Selection 23

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 233

the actual laboratory maximum dry density (MDD) for that


material. This would require an MDD test to be done on the
identical material tested in situ. This is neither practical nor
economical.

4.2.4 Visual inspection

It is imperative that supervisors are trained to carry out a


comprehensive visual inspection of the completed layer prior
to excessive drying out of the material. This inspection
should be carried out during the latter part of the shift prior
to demobilisation of staff and plant for the day. It should be
extremely thorough and should cover the total job lot
completed in the shift. During this inspection, large stones,
excessively moist areas, poorly compacted areas, bumps and
depressions, areas of thin material, material segregation, etc.
should be located and the appropriate remedial action taken.

Figure 4.1: Thickness probe


24 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 234

5 Thickness design

Considerable debate has been held and continues to be held


regarding the need for a full structural design for unsealed
roads. The fact is that wearing courses seldom fail as a result
of punching into the subgrade (i.e. shear failure of either the
wearing course or subgrade material). Failure is typically
either the result of continued slippage of the vehicle tyre
against the soil when a high moisture content prevails (lack
of frictional resistance leading to plastic failure) resulting in
settlement of the tyre into the material; or shear failure of
the upper portion of the wearing course with lateral
displacement. The former situation is confirmed by the
fitting of chains to a vehicle wheel to increase this friction
which restores passability to the vehicle without excessive
additional deformation of the wearing course. The latter
situation is manifested by ruts in the wearing course with
lateral displacement of the wearing course material. Only
when the wearing course becomes excessively thin does
shearing of the underlying subgrade material become
possible when this is soaked.
The thickness design should take into account the fact that
most deformation which may occur in unpaved roads is
rectified during routine maintenance, and that with time, all
roads lose material through environmental factors and traffic
wear. It is thus necessary in terms of thickness design to
allow for these two aspects and for a minimum remaining
thickness to prevent the subgrade being exposed at the
surface.
The rate of gravel loss under traffic is typically greater than
that resulting from environmental influences and is mostly
related to traffic volume. Other factors such as climate,
material properties and, to a lesser extent, geometrics affect
the rate of gravel loss. As for most roads, the primary design
criterion should be the life before major maintenance
(regravelling) is required.
For roads carrying less than 100 vehicles per day (vpd) on a
subgrade material with a minimum soaked CBR of 3 per cent,
a wearing course 150 mm thick of material with a minimum
soaked CBR value of 15 will provide an adequate structure.
In areas where the subgrade (or wearing course) is likely to
become soaked (i.e. to be under standing water for more than
24 hours) resulting in the CBR of the subgrade and base
decreasing to less than 3 or 15 per cent respectively, the
application of two layers of wearing course quality material
(each 150 mm thick) is recommended. The standard thickness
of 150 mm of material as proposed in the specification has
been shown to be adequate to resist excessive deformation
under most conditions, even persistent rainfall, provided the
shape of the road is such that excessive ponding on the road
structure cannot occur. The necessity for proper
maintenance of unsealed roads cannot be overemphasised.
Extended periods of poor or no maintenance will always
result in significant deterioration of the road, loss of gravel,
ILO/ASIST Technical Brief No. 9: Material Selection 25

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 235

and in extreme difficulty in restoring the condition of the


road.
In addition to the subgrade being compacted to a high
density to provide a platform for compaction of the wearing
course, it is equally important that the subgrade density is
such that loss of wearing course by being punched into the
subgrade or by taking up significant rutting which may occur
in the subgrade, is minimised. This type of “gravel loss” is
typically made up during routine maintenance but results in
a premature need for regravelling.
Thinner layers (not less than 100 mm) can be placed to
reduce the cost of material and construction, but in the long
term this is generally not cost-effective. Once about 50 to
75 mm of material has been lost, the wearing course will
need to be replaced, resulting in a premature regravelling
operation. Layers thicker than 150 mm are not recommended
as it then becomes difficult to obtain adequate compaction,
particularly through the full depth, but also at the surface. It
should be remembered that, as the layer of uncompacted
material becomes thicker, the distance from the firm
platform required to compact against becomes greater and
more energy is absorbed by the loose material. This is
particularly relevant when light rollers are utilised.
If the traffic is likely to be higher than 100 vehicles per day,
including more than ten heavies, it may be necessary to
increase the thickness of the wearing course over those areas
with low subgrade strengths (soaked CBR of less than 5 per
cent); or else to include a 150 mm thick selected layer with a
minimum soaked CBR of 10 per cent. Recommended
thicknesses and wearing course material strengths are
summarised in Table 5.1.

Table 5.1: Recommended thicknesses and material strengths


for different subgrade and traffic conditions
Subgrade Traffic (vpd) Layers/thickness Min. soaked
CBR CBR
>3 < 100 150 wearing course 15

<3 < 100 150 wearing course 15


150 selected layer 5

<5 > 100 (> 10% 150 wearing course 15


heavy) 150 sub-base 10

It is possible to scientifically and mechanistically design the


layer thickness in which aspects, such as subgrade strength
and stiffness, wearing course material strength and stiffness,
annual gravel loss through traffic and environmental factors,
traffic compaction, etc., are taken into account. However,
these all require additional testing and environmental data,
much of which are not available in remote locations. In
general, the design proposed above will, with routine
maintenance, provide a road which will last between 5 to 10
years without requiring major maintenance other than
periodic grader blading.
26 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 236

6 Conclusions and recommendations


6.1 CONCLUSIONS
The construction of labour-based unsealed roads, if done with
the correct materials and appropriate quality assurance, will
result in roads which will perform as well as those built
conventionally using plant-based methods. Particular
attention should be paid to material selection and control, as
well as to construction control.

6.2 RECOMMENDATIONS
It is recommended that these guidelines be implemented and
augmented where necessary. Although every project will
have certain unique characteristics and problems, these
guidelines are seen as a generally applicable solution for
routine labour-based construction.
As a consequence of this assignment, Transportek is
currently producing a labour-based construction test kit
which will include all the necessary testing and analysis
equipment to carry out the testing recommended in this
Brief. Copies of the test methods to meet the requirements of
these guidelines will also be included (a summary of the
contents of the kit is provided in Annex B). It is
recommended that this be presented at appropriate venues
and seminars.
As these techniques are applied, any problems which may be
encountered in practice, and improvements in the methods
identified with experience, can be incorporated into the
Guidelines.

ILO/ASIST Technical Brief No. 9: Material Selection 27

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 237

7 References

1. Paige-Green, P. 1989. The influence of geotechnical


properties on the performance of gravel wearing course
materials. PhD Thesis, University of Pretoria, South
Africa.
2. Committee of State Road Authorities (CSRA). 1990. The
structural design, construction and maintenance of
unpaved roads. Technical Recommendations for
Highways. (Draft TRH 20). CSRA, Pretoria, South Africa.
3. Paige-Green, P, and Bam, A. 1991. Passability criteria for
unpaved roads. Research Report RR91/172, Department
of Transport, Pretoria, South Africa.
4. Paige-Green, P, and Bam, A. 1995. The hardness of gravel
aggregate as an indicator of performance in unpaved
roads. Research Report RR93/560, Department of
Transport, Pretoria, South Africa.
5. Secondary and feeder road development programme. Final
report. 1995. Swedish National Road Administration
(SweRoad), Solna, Sweden.
6. National Institute for Transport and Road Research.
1979. Standard methods of testing road construction
materials. Technical Methods for Highways (TMH 1), 1st
Edition, CSIR, Pretoria, South Africa.
7. National Institute for Transport and Road Research.
1986. Standard methods of testing road construction
materials. Technical Methods for Highways (TMH 1), 2nd
Edition, CSIR, Pretoria, South Africa.
8. Semmelink, CJ, Groenewald, M, and Du Plessis, EG.
1994. The optimisation of compaction specifications for
different pavement layers. Project Report PR 91/199,
Department of Transport, Pretoria, South Africa.
9. Cernica, JN. 1980. Proposed new method for the
determination of density of soil in place. Geotechnical
Testing Journal, Vol. 3,3, pp. 120-123; West
Conshohocken, PA, USA.

28 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 238

Annex A: Guideline document

WHERE THERE IS
NO SOILS LAB!
Material selection and tests for
labour-based gravel roads

ILO/ASIST Technical Brief No. 9: Material Selection 29

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 239

Contents

1 Introduction.............................................................................. 31
2 Material properties................................................................ 32
2.1 MATERIAL GRADING ........................................................................................................32
2.1.1 Determination of the grading coefficient and oversize index ................ 32
2.2 MATERIAL COHESION ......................................................................................................33
2.2.1 Determination of the linear shrinkage of soils ..................................... 34
2.3 MATERIAL STRENGTH......................................................................................................36
2.3.1 The determination of the compacted strength of material.................... 37
2.4 AGGREGATE STRENGTH..................................................................................................38
2.4.1 Determination of the Treton impact value of aggregate ....................... 38
2.5 APPLICATION.................................................................................................................40
3 Quality assurance during construction ....................... 42
3.1 MATERIAL TESTING AND CONTROL ..............................................................................42
3.2 CONSTRUCTION QUALITY ASSURANCE .........................................................................42
3.2.1 Moisture content................................................................................ 42
3.2.2 Thickness........................................................................................... 43
3.2.3 Compaction........................................................................................ 43
3.2.4 Visual inspection................................................................................ 43

LIST OF T ABLES
Table A1: Penetration rates of DCP and RCCD and equivalent
soaked CBR (tested at OMC)...........................................................37

LIST OF FIGURES
Figure A1: Preparation of material for the linear shrinkage test ....35
Figure A2: Treton apparatus .................................................................40
Figure A3: Relationship between Grading Coefficient,
Shrinkage Product and performance..............................................41
Figure A4: Thickness probe ...................................................................44

30 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 240

1 Introduction

These guidelines describe a step-by-step method of


evaluating borrow materials for use as the wearing course in
labour-based unsealed roads, and for ensuring that the
quality of the construction is appropriate.
The guidelines assume that all the required testing
apparatus has been provided for the project and that potable
water and normal day-to-day requirements such as batteries
are available. They also assume that supervisors have been
given adequate training in the use of these guidelines.
It will also be noted that the test methods given in this
section are a simplification of those described in the main
document, for ease of use in the field. The effect of this
simplification is considered to be insignificant in terms of the
repeatability and reproducibility of the tests and their
implications on the use of the materials.

ILO/ASIST Technical Brief No. 9: Material Selection 31

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 241

2 Material properties
2.1 M ATERIAL GRADING
In order to achieve good particle interlock from the wearing
course material, a good particle size distribution is
necessary. This is defined by the Grading Coefficient. In
addition, it is important to limit the quantity of large stones
in the material by specifying the Oversize Index. These
parameters are simply determined as described below.

2.1.1 Determination of the grading coefficient and oversize index

SCOPE

In this method, a soil, sand or gravel sample is separated by


dry sieving for determination of the grading coefficient and
to prepare fine material for the bar linear shrinkage test.

APPARATUS
x Sheet of canvas 1 metre by 1 metre for coning and
quartering material
x The following test sieves: 37.5 mm, 26.5 mm, 4.75 mm,
2 mm and 0.425 mm with pan and cover
x Balance with pan, accurate to 1 g, to weigh up to 5 kg
x Various pans of 250 to 300 mm diameter and 20 mm deep
x Drying oven (Solar) to maintain a temperature between
105 and 110°C
x Various stiff brushes
x Thermometer (0 to 120 °C)

METHOD

STEP 1 A test sample such that at least 100 g of material


passes the 0.425 mm sieve, but not less than 2 kg in total,
should be prepared from a bulk sample of at least 5 kg by
coning and quartering using a canvas sheet.
STEP 2 Air dry the sample until it is friable and the
particles can be separated with ease. If the sample is still too
wet, it should be dried in a solar oven at a temperature not
exceeding 50qC.
STEP 3 Dry sieve the material as follows: shake the
material through each sieve (37.5, 26.5, 4.75, 2.0 and
0.425 mm) in turn, starting at the 37.5 mm sieve, until further
shaking results in minimal additional material passing the
sieve. Brush the larger particles (> 4.75 mm) with a stiff
bristle brush to remove all fines adhering to these.
Determine the mass of the soil fines (< 0.425 mm) and
transfer them to a marked paper bag for subsequent testing.
(It is recommended that they be dried in a solar oven prior to
being weighed, but if this will delay testing, this step may be
32 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 242

omitted, since the use of air-dried weights will not affect the
results unduly.)
STEP 4 The masses of the individual fractions retained on
each sieve should be determined (preferably after being
oven-dried but after air-drying if necessary). The masses
should be determined to the nearest 1 g. Record the masses
retained on each sieve and that passing the 0.425 mm sieve.
A form for recording and calculating the results is included
at the end of this document.

RESULTS

Calculate the total mass of material as the sum of the


fractions retained on the individual sieves as well as that
passing the 0.425 mm sieve. Determine the percentage of
each of these as a percentage of the total mass of (dry)
material tested.
Calculate the cumulative percentages passing each sieve to
the nearest 1 per cent by summing the percentage passing
each sieve, starting from the finest sieve. All results should
first be corrected for the percentage retained on the 37.5 mm
sieve by multiplying the percentage passing each sieve by the
percentage passing the 37.5 mm sieve (P37) divided by 100. If
all the material passes the 37.5 mm sieve, this step is not
necessary.
Calculate the grading coefficient as the percentage
material passing the 26.5 mm sieve and retained on the 2 mm
sieve (P26 and P2 respectively) multiplied by the percentage
passing the 4.75 mm sieve (P475) using the following formula:
GC = (P26  P2) u P475/100

The oversize index is defined as the percentage of the


total material retained on the 37.5 mm sieve.

INTERPRETATION

The limits for the grading coefficient are shown in Figure A.3
and should be between 16 and 34. A maximum oversize index
of 5 per cent is permitted to retain a good riding quality over
time.

2.2 M ATERIAL COHESION


In order to minimise the loosening of the surfacing material
and the formation of corrugations, it is necessary that the
materials should have some plasticity. This is determined
using the bar linear shrinkage test described below by
measuring the linear shrinkage of a soil dried from a
moisture content equivalent to the liquid limit to the oven-
dry state.

ILO/ASIST Technical Brief No. 9: Material Selection 33

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 243

2.2.1 Determination of the linear shrinkage of soils

SCOPE

This method covers the determination of the linear


shrinkage of soil when it is dried from a moisture content
equivalent to the liquid limit to the oven-dry state.
Definition The linear shrinkage of a soil for the moisture
content equivalent to the liquid limit, is the decrease in one
dimension, expressed as a percentage of the original
dimension of the soil mass, when the moisture content is
reduced from the liquid limit to an oven-dry state.

APPARATUS
x A shrinkage mould made from 10 mm stainless steel bar
with internal dimensions of 150 mm ± 0.25 mm long u
10 mm ± 0.25 mm wide u 10 mm ± 0.25 mm deep, and open
on two sides (see Figure 3.3)
x A stainless steel plate to fit under the shrinkage mould
x A small thick-bristle paint brush, about 5 mm wide
x Silicone lubricant spray (e.g. Q20 or WD 40)
x A spatula with a slightly flexible blade about 100 mm long
and 20 mm wide
x A solar drying oven capable of maintaining a temperature
of 105 to 110qC
x A pair of dividers and a millimetre scale
x A standard cup, drop cone and guide-tube for estimating
the liquid limit
x A thermometer (0 to 120 qC).

METHOD

STEP 1 The interior of a clean, dry shrinkage mould is sprayed evenly with
the silicone lubricant.

STEP 2 The fines (i.e. the fraction passing the 0.425 mm sieve) saved
during the grading analysis are used for this test. Add water to the fines and
mix thoroughly until the consistency is at the liquid limit. The drop-cone
device is used to ensure that this initial moisture content is correct. Sufficient
material to fill the cup provided should be mixed up and placed evenly in the
cup to a level between 2 and 5 mm below the rim of the cup. The cone should
be placed in the guide tube on the surface of the soil in the cup and allowed to
penetrate for 5 seconds. The cone should penetrate to a depth of 20 mm,
equivalent to the calibration mark on the cone. If the penetration is below this
mark, the material is too dry and additional water is required. The material
would then need thorough re-mixing before the penetration test is repeated. If
the penetration is too high (i.e. the cone sinks into the material to a depth
above the calibration mark), the material is too wet and needs to be dried out
by additional mixing until repetition of the penetration test gives a result
within the defined limits.

STEP 3 The lubricated mould should be placed on the


plate provided and one half should be filled with the moist
soil by taking small pieces of soil on the spatula and pressing
34 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 244

the soil down against one end of the mould and working
along the mould until the whole side is filled and the soil
forms a diagonal surface from the top of one side to the
bottom of the opposite side (see Figure A1(a)).

Figure A1: Preparation of material for the linear shrinkage


test

The mould is now turned round and the other portion is


filled in the same manner (see Figure A1 (b)). The hollow
along the top of the soil in the mould is now filled so that the
soil is raised slightly above the sides of the mould (see Figure
A1 (c)). The excess material is removed by drawing the blade
of the spatula once only from one end of the mould to the
other. The index finger is pressed down on the blade so that
the blade moves along the sides of the mould (see Figure A1
(d)). During this process the wet soil may pull away from the
end of the mould, in which case it should be pushed back
gently with the spatula. On no account should the surface of
the soil be smoothed or finished off with a wet spatula.
STEP 4 The filled mould is now placed in the drying oven
and dried at a temperature of between 105 and 110qC until
no further shrinkage can be detected. As a rule, the material
is dried out overnight, although three hours in the oven
should be sufficient. The mould with the material is taken
out of the oven and allowed to cool.
STEP 5 It may be found that the ends of the dry soil bar
have a slight lip or projecting piece at the top. These lips
should be removed by abrading with a sharp, narrow spatula,
ILO/ASIST Technical Brief No. 9: Material Selection 35

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 245

so that the end of the soil bar is parallel to the end of the
mould (see Figure A1 (e)). If the soil bar is curved, it should
be pressed back into the mould with the fingertips so as to
make the top surface as level as possible.
The loose dust and sand, removed from the ends, as well as
any loose material between cracks, should be emptied out of
the mould by carefully inverting the mould whilst the
material is held in position with the fingers. The soil bar is
then pressed tightly against one end of the mould. It may be
noticed that the soil bar fits better at one end than at the
other end. The bar should be pressed tightly against the end
at which there is a better fit. The gap between the soil bar
and the end of the mould is measured by means of a good pair
of dividers, measuring on a millimetre scale, to the nearest
0.5 mm and recorded on the form included with this
document.

RESULTS

The bar linear shrinkage (BLS) is calculated from the


measured shrinkage LS (in mm) as follows:
BLS = LS u 0.67 (%)

NOTE

After the test, the soil bar should be examined to ensure that
the corners of the mould were filled properly and that no air
pockets were contained in the soil bar. If air pockets were
contained, the material should be tested again.

INTERPRETATION

A value for the shrinkage product in excess of 100 is required


but it should not exceed 365, otherwise slipperiness will
result when the material is wet.

2.3 M ATERIAL STRENGTH


In order to support the loads applied by vehicles, the
material should have an adequate strength at the density at
which it will perform in the field. The soaked California
Bearing Ratio is typically specified for this parameter but its
measurement requires bulky, expensive equipment. For
labour-based projects, the following simple but equivalent
method is, however, recommended. For this, a small section
of road equivalent to that of the full scale construction is
processed. This should be five metres long and one metre
wide (or the width of the roller if wider than one metre),
with one metre of material surrounding the central section to
be tested.

36 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 246

2.3.1 The determination of the compacted strength of material

SCOPE

The objective of this procedure is to ensure that the


compacted strength of the borrowed material as placed in the
field complies with the acceptance testing carried out on the
borrow material. Secondary objectives of this procedure are
to identify the limit to be used for control testing after
compaction and to identify the number of roller passes for
optimum compaction.

APPARATUS
x Compactor equivalent to that proposed for use
x DCP or RCCD apparatus.

METHOD

STEP 1 A sufficient quantity of the proposed material


(between 4.0 and 4.5 cubic metres) should be dumped on a
section of the proposed subgrade prepared to the same
standard as that of the proposed road. Water should be
added to bring this material to its estimated optimum
moisture content.
STEP 2 The moist material should be spread to a
thickness which will provide a compacted thickness
equivalent to the design thickness of the layer.
STEP 3 Using the compaction method and plant which
will be used during full construction, a complete roller pass
should be given to the layer. The strength of the layer is then
determined using a DCP or RCCD and the results recorded
on the field test data form provided.
STEP 4 Repeat Step 3, plotting the penetration rate
obtained from testing against the number of roller passes
until no further strengthening of the material occurs (i.e.
until the measured penetration rate reaches a minimum).
This identifies both the number of passes above which no
additional benefit from rolling is obtained, and the final
maximum strength of the material at compaction moisture
content.

INTERPRETATION

The maximum strength of the material should comply with


the requirements given in Table A1 below:

Table A1: Penetration rates of DCP and RCCD and


equivalent soaked CBR (tested at OMC)
Apparatus Penetration Penetration Penetration Equivalent
rate (3 blows) (20 blows) soaked CBR
(mm/blow) (mm) (mm) (%)
DCP • 5 • 15 • 100 15

ILO/ASIST Technical Brief No. 9: Material Selection 37

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 247

RCCD • 9 • 27 — 15

STEP 6

If the above requirements are not met, either the material


has inadequate strength to resist deformation or to avoid
becoming slippery when wet, or additional compaction
energy is required in order to achieve a higher density and
increase the strength of the material.
A minimum soaked equivalent CBR strength of 15 per cent is
required for acceptable passability and trafficability. This is
related to the DCP and RCCD results as indicated in the
table above.

2.4 AGGREGATE STRENGTH


In order to ensure that the aggregate particle strength is
sufficient to avoid this component of the material breaking
down excessively under rolling and traffic, a simple strength
test (Treton Impact Value test (Figure A2)) is recommended,
as described below. This test also identifies those materials
which are too hard to break down and which could result in
excessive stoniness of the road. A sample of the aggregate is
subjected to ten blows of a falling hammer and the resulting
disintegration is measured in terms of the quantity passing
the 2 mm sieve.

2.4.1 Determination of the Treton impact value of aggregate

SCOPE

The Treton value is an indication of the resistance of


aggregate to impact. The aggregate is subjected to ten blows
of a falling hammer and the resulting disintegration is
measured in terms of the quantity passing the 2 mm sieve,
which is then expressed as a percentage of the mass of the
test sample. This is called the Treton value.

APPARATUS
x A Treton apparatus consisting of a base plate, anvil,
cylinder and a hammer weighing 15 kg r 50 g (Figure A2).
The base plate should be placed on a firm concrete block.
x The following test sieves, 200 mm in diameter: 19.0 mm,
16.0 mm and 2 mm. The bigger sieves should be made of
perforated plate and the 2 mm sieve of wire mesh.
x A balance to weigh up to 200 g, accurate to 1 g.

METHOD

STEP 1 From the field sample, screen out a sufficient


quantity (at least 200 g) of the – 19.0 + 16.0 mm fraction. If
the aggregate is noticeably variable as regards type or
hardness, each type should be tested and reported

38 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 248

separately. In this case an estimate should be made of the


percentage of each type.

ILO/ASIST Technical Brief No. 9: Material Selection 39

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 249

Figure A2: Treton apparatus

STEP 2 Select 15 to 20 of the most cubical pieces. Weigh


the aggregate pieces to an accuracy of 1 g, and place them as
evenly spaced as possible on the anvil in such a manner that
their tops are approximately in the same horizontal plane.
STEP 3 Place the cylinder over the anvil and tighten the
clamp screws. Place the hammer in the cylinder so that the
top of the hammer is level with the top of the cylinder and let
it drop ten times from this position.
STEP 4 Remove the cylinder and sieve all the aggregate
on the anvil and base plate thoroughly through a 2 mm sieve.
Weigh the aggregate retained on the sieve to the nearest 1 g,
and record the mass. The test should be carried out in
triplicate. (If any individual result differs from the others by
more than five units, further tests should be carried out.)

RESULTS

Calculate the Treton value to the first decimal place as


follows:
Treton value = (A - B)/A u 100

where A = the total mass of the stone particles before tamping (g)
B = the total mass of the stone particles retained on the 2 mm
sieve after tamping (g).

Report the value to the nearest whole number.

INTERPRETATION

Recommended Treton impact values should lie between 20


and 65. Materials with values less than 20 will be too hard
and cause excessive roughness whilst those with values
higher than 65 will be too soft and break down under traffic.

2.5 APPLICATION
The results obtained from the grading and linear shrinkage
testing are evaluated using Figure A-3
For the best performance, the results should plot in zone E1
or E2 of the diagram. The potential problems associated with
materials plotting in the other zones are identified in the
zones. To reduce dust, materials should preferably plot in the
E2 zone.

40 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 250

Figure A3: Relationship between Grading Coefficient,


Shrinkage Product and performance

ILO/ASIST Technical Brief No. 9: Material Selection 41

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 251

3 Quality assurance during construction


3.1 M ATERIAL TESTING AND CONTROL
The quality of the borrow material should be controlled on a
regular basis during construction to ensure that its
properties do not differ from the accepted specification. For
this purpose, only the grading coefficient and shrinkage
product need be tested.
This testing is required on a routine basis to ensure that the
material to be used for the following job lots comply with the
specifications. It is recommended that samples of the
material to be used should be taken two or three days prior
to that material being used and tested so that the material
can be approved at least the day before it is processed in the
borrow pit, loaded and hauled. The test techniques described
in the section on Material Properties above are such that this
is possible.
The individual results of the borrow pit testing should be
plotted on Figure A3. The test results should be plotted on
this figure on an ongoing basis and if there is a trend to move
out of Zone E, it is indicative of a change in the material
properties. Additional testing should be carried out to
determine the cause of this and to identify remedial action,
e.g. blending of material, redefining of boundaries of the
borrow pit, adjustment of depth of excavation, etc.

3.2 CONSTRUCTION QUALITY ASSURANCE


A number of factors should be controlled during construction.
These are:
x Moisture content
x Thickness
x Compaction
x General finish.

3.2.1 Moisture content

It is very difficult to get a high degree of compaction if the


moisture content is not close to the Optimum Moisture
Content (OMC) for that material. As it is difficult to get an
accurate, usable determination of the moisture content
quickly, the visual determination of this in the field is
recommended.
The control of moisture during construction must be carried
out by squeezing a sample of the material as tightly as
possible in the hand.
The material should be moist enough to stick together when
squeezed, with no visible sign of free water on the surface.

42 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 252

If the material disintegrates, it is too dry for compaction. If


free water is ejected or if the soil sticks to the hand, it is too
wet.
If the “sausage” formed by squeezing in the hand is squeezed
diametrically between the thumb and forefinger, it should
break with some crumbling. It should not break by
deformation under the finger pressure, nor should there be
excessive crumbling.

3.2.2 Thickness

The thickness of the layer should be closely controlled prior


to compaction, with allowance for the bulking factor.
Spreading of the material should be as consistent as possible
to ensure that all material is placed at a similar loose
density.
To produce a 150 mm thick compacted layer, 190 to 200 mm
of loose material is typically required.
Control of the thickness during construction is carried out by
inserting a calibrated probe (Figure A4) through the
uncompacted material to confirm that the thickness prior to
compaction is within the required limits.
This should be checked in at least 25 locations and, where
the thickness is deficient, more material should be added.

3.2.3 Compaction

The compaction achieved in the field is the most important


aspect of the construction process. It is recommended that a
simple device such as the Rapid Compaction Control Device
(RCCD) or DCP is used for this purpose. Both tests are quick
and repeatable, and many tests can be done rapidly and at
little cost.
Not less than six tests should be done on any job lot and in no
case should the penetration rate exceed the specified
maximum permissible penetration rate.
Should the tests show up areas which are unacceptable, the
reasons for this should be investigated. Poor results can be
attributed to the material being too wet, material that has
not received adequate compaction effort, or to a pocket of
poor quality material.
The actual cause needs to be identified and corrective action
taken. This may involve scarifying and drying out prior to
recompaction if the material is too wet, additional
compaction if necessary, or replacement of poor material
where appropriate.

3.2.4 Visual inspection

Supervisors should be trained to carry out a comprehensive


visual inspection of the completed layer prior to excessive
drying out of the material. This inspection should be carried
out during the latter part of the shift prior to demobilisation
ILO/ASIST Technical Brief No. 9: Material Selection 43

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 253

of staff and plant for the day, should be extremely thorough


and should cover the total job lot completed in the shift.
During this inspection the following should be evaluated:
x the presence of large stones
x excessively moist areas
x poorly compacted areas
x bumps and depressions
x areas of thin material
x material segregation.
Appropriate action to rectify the problems should be taken
before the material has dried out.

Figure A4: Thickness probe

44 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 254

LABORATORY TEST RESULTS

PROJECT: ______________________________________________ DATE: _____________

SAMPLE NUMBER: _______________________________ OVEN DRIED: YES † NO †

SAMPLE LOCATION: _____________________________________________________________

GRADING ANALYSIS

Sieve size Mass % of total Cumulative % Normalised


mm retained retained passing cumulative %
(g) passing
37.5 D
26.5 A A
4.75 C C
2.0 B B
0.425
< 0.425

GC = (A – B) x C/100 = (........ – ........) x (......./100) =..................

IO = 100 – D = (100 – ........) = ..............

LINEAR SHRINKAGE

Mould Number Linear shrinkage Bar linear shrinkage Shrinkage


(LS) (BLS)* product (SP)**
mm mm

* BLS = LS x 0.67
** SP = Mean BLS x Percent passing 0.425 mm

TRETON IMPACT VALUE

Test Number Total Mass (A) Mass Retained on 2 Treton value*


g mm Sieve (B) g

* Treton value = ((A – B)/A) x 100

Operator: _______________________________________

ILO/ASIST Technical Brief No. 9: Material Selection 45

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 255

FIELD TEST RESULTS

PROJECT: ______________________________________________ DATE: ______________

TEST NUMBER: __________________________________________________________________

TEST LOCATION: ________________________________________________________________

COMPACTED STRENGTH

Pass number 1 2 3 4 5 6 7 8 9 10
RCCD reading (3 blows)
DCP penetration (mm/blow)

THICKNESS

(25 readings in mm from the probe per job lot)

Operator: _______________________________________

46 ILO/ASIST Technical Brief No. 9: Material Selection

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.1 - 256

Annex B: Contents of CSIR testing kit


The following are the basic contents of the CSIR field testing
kit. Certain items (e.g. balance and oven), which are
necessary in one test, are not repeated in the requirements
for subsequent tests.

Grading
x Sheets of canvas 1 metre by 1 metre for coning and
quartering material.
x Test sieves: 37.5 mm, 26.5 mm, 4.75 mm, 2 mm and
0.425 mm with pan and cover.
x Balance with pan, accurate to 1 g, to weigh up to 5 kg.
x Various pans of 250 to 300 mm diameter and 20 mm deep.
x Drying oven (Solar) to maintain a temperature between
105 and 110qC.
x Various stiff brushes.
x Electronic calculator (Solar powered).
x Wind shield for balance.
x Levelling platform for balance.
x Thermometer (0 to 120qC).

Linear shrinkage
x Shrinkage moulds with internal dimensions of 150 r
0.25 mm long u 10 r 0.25 mm wide u 10 r 0.25 mm deep
and made of 10 mm thick stainless steel bar, open on two
sides.
x A steel plate to fit underneath the shrinkage moulds.
x Silicone lubricant spray (e.g. Q20 or WD 40).
x A spatula with a slightly flexible blade about 100 mm long
and 20 mm wide.
x A pair of dividers and a millimetre scale.
x A standard drop cone and calibrated tube for estimating
the liquid limit.

Material strength
x RCCD or DCP test apparatus.

Aggregate strength (Treton)


x A Treton apparatus consisting of a base plate, anvil,
cylinder and a hammer weighing 15 kg r 50 g (Figure A2).
The baseplate should be placed on a firm concrete block.
x The following test sieves, 200 mm in diameter: 19.0 mm,
16.0 mm and 2 mm. The larger sieves should be made of
perforated plate and the 2 mm sieve of wire mesh.

Thickness
x Thickness probe.
ILO/ASIST Technical Brief No. 9: Material Selection 47

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.2 - 257

APPENDIX D.2

1. Rainfall analysis for hydrological design

The high variability associated with rainfall events requires a statistical approach to rainfall analysis. For
hydrological design it is the worst storm events in a reasonably long period of time (typically 10, 20, or 50
years) that the engineer is concerned with. Because these storms are relatively rare events the statistical
information about them lies near the edges of any rainfall-frequency distribution curve and it is at these
extremes that the shape of any distribution curve is least well defined. The technique used to analyse the
data is therefore called ‘extreme value statistical analysis’. In practice the rainfall-frequency distribution is
not a normal distribution (in the statistical meaning of the term) but usually a non-symmetrical distribution
with a long tail. The most common is the ‘gumbel’ distribution but a Log-normal distribution is sometimes
applicable. Such distributions can be applied to rainfall data as follows:

For a selected duration of rainfall, for example, the 24-hour, 2-hour, 1-hour, 30-minute rainfall, the highest
value is selected for each calendar year. This series is termed an ‘annual maximum series’. The annual
data are then listed in order of descending values and a rank assigned to each, rank 1 being the highest
rainfall level. Thus if there is 20 years of data the lowest rainfall will be rank 20. The cumulative frequency
distribution (or non-exceedence probability) and the recurrence interval i.e. the return period, are then
calculated.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.2 - 258

20 year historical data for the 1-hour storm

Recurrence Non exceedence


Rainfall
Rank interval T Probability Probability of storm in any year
(mm)
years F
95% chance of a storm not
90 1 21.0 0.95
exceeding this size in any year
or 10% chance of a storm
70 2 10.5 0.90
exceeding this size
63 3 7.0 0.86 etc
55 4 5.3 0.81
48 5 4.2 0.76
45 6 3.5 0.71
41 7 3.0 0.67
39 8 2.6 0.62
38 9 2.3 0.57
35 10 2.1 0.52
33 11 1.9 0.48
30 12 1.8 0.43
28 13 1.6 0.38
27 14 1.5 0.33
25 15 1.4 0.29
23 16 1.3 0.24
22 17 1.2 0.19
20 18 1.2 0.14
18 19 1.1 0.10
15 20 1.1 0.05

The recurrence interval T = (N+1)/n where N is the number of years of data and n is the rank

The non-exceedence probability F = 1 – 1/T

These data should, ideally, fit a Gumbel or Log normal distribution. The equations are:
Gumbel F = exp[-exp((-r-B)/A)] and
Log-normal F = [(ln r-B)/A] where is the standard normal distribution function

In these equations A and B are the parameters that need to be fitted to the data and r is the depth of
rainfall from the data above.

It should be noted that the same method can be applied to other data, for example a set of data
showing maximum rainfall depths for storms of other durations during each year or a set of data showing
maximum rainfall intensities in each year. The actual data can be fitted to these distributions using the
traditional least squared deviation approach to obtain the parameters A and B.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.2 - 259

Once the relationships have been established the rainfall depth for any recurrence interval can be
determined. Usually the values for longer recurrence intervals are needed because rainfall records
rarely go back far enough in time. Therefore it should be noted that however well the Gumbel or Log-
normal distribution appears to fit the data for the lower and more frequent storm intensities, this is an
extrapolation of very variable data and should be borne in mind when designing the drainage system; the
apparent accuracy of many of the hydraulic calculations is simply not justified or required.
It is unusual for adequate rainfall data to be available to develop a full set of rainfall intensity curves for
the location of interest. Normally only rainfall data for a fixed duration, most commonly for 24-hours, is
available and no information on the time duration within the 24-hour period exists. In such cases it may
be appropriate to use a generalized relationship between the rainfall falling within a time t hours and that
falling in 24 hours. Such a relationship might be of the following form:

Rt/R24 = (t/24)[(b+24)/(b+t)]n

Where Rt = rainfall in a duration t hours


R24 = rainfall in 24 hours
B, n = constants.

The value of b ranges from 0.2 to 0.5 and n ranges from 0.5 to 1.1 and they will need to be computed
by analyzing rainfall data from as many rain gauges as possible and including data about shorter time
duration storms.

Annex B
Culvert capacity (alternative method)
Maybe we don’t need this and the nomographs in the chapter. I will assess which is best

(This Annex provides an alternative method of estimating culvert sizes. The use of equations is preferred
by some readers to the use of the nomographs.

If the culvert is relatively short and the streambed slope is sufficient to avoid accumulation of water on
the downstream side of the culvert, inlet conditions are likely to prevail. In this case, the culvert will act
as an orifice and the capacity can be determined in a relatively simple manner on the basis of headwater
height and inlet geometry (barrel shape, cross-sectional area and the inlet edge). Barrel slope affects the
inlet control performance to a small degree but may be neglected.

These conditions, known as inlet control, are those that normally occur in most parts of Ethiopia. Estimates
of culvert capacity in cases where a high degree of accuracy is not required may be approximated by the
expressions in Table B1. The expressions are based on the following assumptions:
ƒ Inlet Control;
ƒ Wingwall Angle = 450;
ƒ Vertical Headwall.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.2 - 260

Table B1: Simplified Formulae for Calculation of Discharge Capacity

Discharge Capacity Q (m3/s)


Type (with inlet control)
Hw/D=1.00 Hw/D=1.25 Hw/D=1.50
Concrete Pipe 1.3 x D2.5 1.9 x D2.5 2.2 x D2.5
Corrugated Metal Pipe 1.1 x D2.5 1.6 x D2.5 1.8 x D2.5
Arch Culvert (semi-circular) 2.3 x H2.5 3.4 x H2.5 4.0 x H2.5
1.5 x B x
Box Culvert 2.1 x B x H1.5 2.5 x B x H1.5
H1.5
D : diameter of a pipe culvert (m) B : width of a box culvert (m)
Hw : headwater height (m) H : height of a box/arch culvert (m)

For major culverts, where an under-design could have serious consequences in terms of road failure or
damage caused by flooding of upstream areas, detailed calculation of expected culvert performance
should be carried out taking into account the geometry of the culvert and the characteristics of the
surrounding area.

The calculations should be carried out using recognised computer programs or nomographs, as presented,
for example, by the U.S. Department of Transportation, Federal Highway Administration (1985). Tables
for the hydraulic design of pipes sewers and channels Volumes I & II, 7th edition, published by HR
Wallingford (UK), may also be used where different conditions exist, or greater accuracy is needed.

Table B2: Simplified Formulae for Calculation of Discharge Capacity

Discharge Capacity Q (m3/s)


Type (with inlet control)
Hw/D=1.00 Hw/D=1.25 Hw/D=1.50
Concrete Pipe 1.3 x D2.5 1.9 x D2.5 2.2 x D2.5
Corrugated Metal Pipe 1.1 x D2.5 1.6 x D2.5 1.8 x D2.5
Arch Culvert (semi-circular) 2.3 x H2.5 3.4 x H2.5 4.0 x H2.5
1.5 x B x
Box Culvert 2.1 x B x H1.5 2.5 x B x H1.5
H1.5
D : diameter of a pipe culvert (m) B : width of a box culvert (m)
Hw : headwater height (m) H : height of a box/arch culvert (m)

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.3 - 261

APPENDIX D.3

1. Advantages and disadvantages of Surfacing/Paving Options

1.1 Basic
S-01: Engineered Natural Surface (ENS)
ƒ Advantages:
• Lowest cost option for basic access (no imported materials required);
• Easy to construct and maintain using labour-based methods or simple, low-cost grading
equipment.
ƒ Disadvantages:
• Suitable only for relatively light traffic (normally up to AADT 50) in gentle terrain (gradient <
6%) with relatively low rainfall (< 2000mm per annum) with moderately strong soils (soaked
CBR>15%);
• May be impassable in wet weather, particularly with plastic materials;
• Relatively high maintenance requirements;
• Susceptible to water erosion on gradients;
• Dust pollution in dry weather.

S-02: Natural Gravel


ƒ Advantages:
• Good performance when properly constructed with adequate quality material
• Suitable for light to medium traffic (normally AADT < 200)
• Usually lower initial cost than most other surfacing options
• Can be used as an intermediate surface in a planned and resourced “stage construction”
strategy (however surfacing gravel and road base gravel specification requirements differ)
ƒ Disadvantages:
• A diminishing, finite resource of often variable quality;
• Need for sustained maintenance programme;
• High maintenance costs, particularly re-gravelling;
• Dust pollution in dry weather with health and environmental concerns;
• Traffic, climatic and longitudinal gradient (typically < 6%) constraints on use relating to rate
of gavel loss.

1.2 Stone Paving


S-03: Waterbound/Drybound Macadam
ƒ Advantages:
• Proven performance in all climates
• Suitable for light and medium traffic
• Does not require expensive equipment
• Stone may be broken manually and/or mechanically crushed
• Can be later upgraded/overlaid with another surface type
• Can be used in water-scarce and water sensitive locations
ƒ Disadvantages:
• Hand crushed aggregate is usually single sized which creates poor interlock and strength.
Improving grading can be achieved by adding different size material but this requires careful
supervision and additional labour resources.
• Stone type and shape is important
• If non-vibrating equipment used for WBM it should be heavy
• WBM not suitable for use on weak subgrade which will be weakened by excessive use of
water.
• If used as road surface, will probably require medium levels of maintenance.
• WBM requires water to be available.

PART D: EXPLANATORY NOTES FOR LOW VOLUME ROAD DESIGN


Appendix D.3 - 262

• Smooth to medium roughness.

S-04: Handpacked Stone


ƒ Advantages:
• Proven performance in all climates
• Suitable for light and medium traffic
• Does not require expensive equipment to construct or maintain
• Suitable for construction by small contractors or communities in remote areas with access
problems for crushing equipment or heavy plant.
• Can be later upgraded/overlaid with another surface type
• Low maintenance, easily repairable
• Can be later upgraded by sealing in a stage construction strategy
ƒ Disadvantages:
• Requires hard stone to be available locally
• Stones must not be rounded in shape
• Smooth to high surface roughness, depending on skill of laying. Rough surface can be
uncomfortable for traffic, especially bicycles, motorcycles or carts.
• Surface is porous, so foundations should not be liable to severe weakening when wet
• Medium to high surface roughness

S-05: Stone Setts or Pavé


ƒ Advantages:
• Proven performance in all climates
• Suitable for heavy traffic
• Does not require expensive equipment to construct or maintain
• Suitable for construction by small contractors or communities in remote areas with access
problems for crushing equipment or heavy plant.
• Can be constructed at any gradient.
• Low maintenance, easily repairable.
• Can be later upgraded by sealing in a stage construction strategy
ƒ Disadvantages:
• Requires hard stone to be available locally
• Cobble stones must not be roughly cubical in shape
• Requires skill in laying to achieve a smooth finished surface
• If non-vibrating equipment used it should be heavy
• Surface is porous, so foundations should not be liable to severe weakening when wet
• Smooth to medium surface roughness
• Stones that polish by traffic, or are slippery when wet, must not be used.

S-06: Mortared Stone


ƒ Advantages:
• Proven performance in all climates.
• Suitable for light to heavy traffic.
• Does not require expensive equipment to construct or maintain.
• Built with unshaped stones and laid by hand. It is therefore suitable for construction by
small contractors or communities themselves, or in remote areas with access problems for
crushing equipment or heavy plant.
• Can be constructed at any gradient.
• Low maintenance, easily repairable.
• Light compaction equipment is only required for the foundation layers.
ƒ Disadvantages:
• Requires hard stone to be available locally.
• Stone requires to have at least one smooth, even face.
• Requires skill in laying to achieve a good bedding and smooth, even finished surface.
• Smooth to medium surface roughness.
• Stones that ‘polish’ by traffic, or are slippery when wet, must not be used.

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.3 - 263

• Cannot be used until the mortar joints have set and hardened sufficiently (usually about 5-7
days in hot/warm climate).

S-07: Dressed Stone/Cobble Stone Paving


ƒ Advantages:
• Proven performance in all climates.
• Suitable for light to heavy traffic.
• Does not require heavy compaction equipment or any other expensive equipment to
construct or maintain.
• It is suitable for construction by small contractors or communities themselves, or in remote
areas with access problems for crushing equipment or heavy plant.
• Erosion resistant, durable, not damaged by diesel/lubricant spillage.
• Can be constructed at any gradient.
• Minimal maintenance required, easily repairable.
• Surface easy to clean, suitable also for urban use.
• High residual value; the materials can be recycled into other types of paving, or be overlaid
with another surface.

ƒ Disadvantages:
• Requires hard stone to be available locally.
• Stone must be suitable for dressing by hand into a cubic shape.
• Requires skill in laying to achieve a smooth finished surface.
• Surface is porous (unless mortar jointed), so foundations should not be liable to severe
weakening when wet.
• Smooth to high surface roughness, depending on skill of dressing/laying. Rough surface can
be uncomfortable for traffic, especially bicycles, motorcycles or carts.
• Stones that ‘polish’ by traffic, or are slippery when wet, must not be used.

S-08: Fired Clay Brick – Unmortared or mortared joints


ƒ Advantages:
• Proven performance in all climates
• Options to produce the bricks in a sustainable or low-carbon-footprint way
• Bricks can be produced in small scale kiln close to the road site, thus reducing transport
costs and energy
• Suitable for heavy traffic
• Does not require expensive equipment to construct or maintain
• Suitable for construction by small contractors or communities in remote areas with clay and
energy sources but no/few hard stone resources
• Mortared option can be constructed at any gradient.
• Surface easy to clean, suitable also for urban use.
• Low maintenance, easily repairable.
ƒ Disadvantages:
• Requires suitable clay and energy sources to be available locally
• High quality control required on production and burning process
• Requires skill in laying to achieve a smooth finished surface
• Surface of un-mortared option is porous, so foundations should not be liable to severe
weakening when wet
• Smooth to medium surface roughness

1.3 Bituminous Surfacings


ƒ S-09: Surface Dressing (Chip Seal)
ƒ S-10: Sand Seal
ƒ S-11: Slurry Seal
ƒ S-12: Graded Aggregate (Otta) Seal
ƒ S-13: Cape Seal

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.3 - 264

ƒ Advantages:
• Seals and protects the base and provides strength at the road surface so that the latter can
resist the abrasive and disruptive forces of traffic.
• Protects the pavement from moisture ingress and, in so doing, prevents loss of pavement
strength thereby permitting the use of many materials that would otherwise not be
appropriate.
• Improves safety by providing a superior skid-resistant surface, free from corrugations, dust
and mud, often increasing light-reflecting characteristics and allowing the application of
pavement markings.
• Reduces vehicle operating and maintenance costs and extends vehicle life.
• Most bituminous seals (except Ottaseal) are suitable for labour based methods if emulsions
are used.

ƒ Disadvantages:
• Relatively costly to construct where bitumen is a high cost item and/or has to be hauled long
distances to the project site.
• Some types (eg Chip Seal, Cape Seal) require relatively high quality crushed aggregates
• Labour based operations require a high level of quality control to ensure, inter alia, correct
and consistent bitumen application rates

1.4 Concrete
Unreinforced:
ƒ Advantages:
• Proven experience in all climates
• Long expected life span
• Suitable for all traffic, including heavily loaded trucks
• Good load spreading properties and suitable for weak subgrades
• Does not require expensive equipment to construct or maintain
• Suitable for labour-based, small scale contractor or community construction
• Erosion resistant, durable, not damaged by diesel/lubricant spillage
• Can be constructed at any gradient
• Minimal maintenance required, easily repairable
• Suitable for roads that suffer flooding (providing foundation remains intact)
• Surface easy to clean, suitable for urban use

ƒ Disadvantages:
• Expensive to construct
• Requires to be cured and to gain strength before opening to traffic
• Cement and steel reinforcement are high components of the total cost, particularly if they
have to be imported and/or hauled for a long distance
• If frost conditions are expected, a higher quality concrete is required.

Ultra-thin reinforced concrete


ƒ Advantages:
• Proven experience for LVR
• Long expected life span
• Suitable for light and medium traffic
• Good load spreading properties and suitable for weak subgrades
‚ Does not require expensive equipment to construct or maintain
‚ Suitable for labour-based, small scale contractor or community construction
‚ Erosion resistant, durable, not damaged by diesel/lubricant spillage
‚ Can be constructed at any gradient
‚ Minimal maintenance required, easily repairable
‚ Surface easy to clean, suitable for urban use

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
Appendix D.3 - 265

ƒ Disadvantages:
• Relatively expensive to construct
• Requires high quality concrete and good quality control processes from sub-grade
preparation through to curing
• Requires to be cured and to gain strength before opening to traffic
• Cement and steel reinforcement are high components of the total cost, particularly if they
have to be imported and/or hauled for a long distance

PART D: EXPLANATORY
PART D: EXPLANATORY
NOTES FORNOTES
LOW VOLUME
FOR ROADS
ROAD DESIGN
)('(5$/'(02&5$7,&5(38%/,&2)
('(5$/'(02&5$7,&5(38%/,&2)

(7+,23,$152$'6$87+25,7<

,7 <
( 7+

25
,2

,$
15 87
+
3

2$'6 $

'(6,*10$18$/)25/2:92/80(52$'6
3$57(
),1$/'5$)7$35,/
Part E
EXPLANATORY NOTES AND DESIGN
STANDARDS FOR SMALL STRUCTURES

Project planning

Design criteria

Structural options

Site selection and appraisal

Part E
Explanatory notes and
design standards for Watercourse characteristics
small structures

Materials

Structure design

Construction

Maintenance
E - iii

E TABLE OF CONTENTS

E. TABLE OF CONTENTS ...................................................................................................... E.III


E. LIST OF TABLES ................................................................................................................E.IX
E. LIST OF FIGURES ..............................................................................................................E.XI
E. LIST OF PLATES ............................................................................................................. E.XIV
1. INTRODUCTION ................................................................................................................E.1
1.1 Definition of a small road structure ................................................................................. E.2
1.2 Scope of the guidance .................................................................................................... E.2
1.3 Types of structures ........................................................................................................... E.3
1.4 How to use Part E ............................................................................................................ E.5
2. PROJECT PLANNING ........................................................................................................E.7
2.1 Setting priorities .............................................................................................................. E.7
2.2 Assessment of the problem or need ............................................................................... E.8
2.3 Assessment of potential structures ................................................................................. E.8
2.3.1 Desk Study ........................................................................................................ E.9
2.3.2 Field study ........................................................................................................ E.9
2.4 Collection of initial design data....................................................................................... E.9
2.5 Field assessment practicalities ...................................................................................... E.12
3. DESIGN CRITERIA ...........................................................................................................E.14
3.1 Selecting design parameters ......................................................................................... E.14
3.2 Design life ..................................................................................................................... E.14
3.3 Design flood .................................................................................................................. E.14
3.4 Traffic categories and widths ......................................................................................... E.15
3.5 Design code .................................................................................................................. E.19
3.6 Serviceability ................................................................................................................. E.19
3.7 Drainage of the structure............................................................................................... E.19
3.8 Maintenance capability ................................................................................................. E.19
3.9 Safety ............................................................................................................................. E.19
3.10 Future changes in road use ........................................................................................... E.20
3.11 Funding ......................................................................................................................... E.20
4. STRUCTURAL OPTIONS ..................................................................................................E.21
4.1 Drifts .............................................................................................................................. E.21
4.1.1 Key features .................................................................................................... E.22
4.2 Culverts ......................................................................................................................... E.23
4.2.1 Key features ................................................................................................... E.24
4.3 Vented Fords and Causeways ....................................................................................... E.25
4.3.1 Key features ................................................................................................... E.26

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - iv

4.4 Large Bore Arch Culverts .............................................................................................. E.27


4.4.1 Key features ................................................................................................... E.28
4.5 Bridges (arch or simply supported deck) ....................................................................... E.29
4.5.1 Key features .................................................................................................... E.30
4.6 Structure selection ......................................................................................................... E.31
4.6.1 Costs ............................................................................................................... E.31
4.6.2 Amount of traffic per day / acceptable duration of traffic interruptions ........ E.31
4.6.3 Frequency of flooding .................................................................................... E.31
4.6.4 Emergency / principal route ........................................................................... E.31
4.6.5 Availability of alternative route ....................................................................... E.32
4.6.6 Damage to land or property ........................................................................... E.32
4.6.7 Uncertainties in flood prediction .................................................................... E.32
4.6.8 Bank elevation and bed material of the watercourse ..................................... E.32
4.6.9 Complexity of the structure ............................................................................ E.33
5. SITE SELECTION AND APPRAISAL ..................................................................................E.35
5.1 General requirements ................................................................................................... E.35
5.1.1 Road Alignment .............................................................................................. E.36
5.1.2 Location .......................................................................................................... E.36
5.1.3 Existing Structure Assessment ........................................................................ E.36
5.1.4 Site Investigation ............................................................................................ E.37
5.1.5 Bearing capacity ............................................................................................. E.37
5.2 Specific requirements .................................................................................................... E.40
5.2.1 Drifts ............................................................................................................... E.40
5.2.2 Culverts .......................................................................................................... E.40
5.2.3 Vented fords ................................................................................................... E.44
5.2.4 Large bore culverts ......................................................................................... E.45
5.2.5 Bridges ........................................................................................................... E.45
6. WATERCOURSE CHARACTERISTICS ................................................................................E.47
6.1 Maximum peak flow ...................................................................................................... E.47
6.1.1 Method 1 - Observation ................................................................................. E.49
6.1.2 Method 2 - Interviews ..................................................................................... E.49
6.1.3 Method 3 - Rational Method .......................................................................... E.50
6.1.4 Method 4 - Estimation ................................................................................... E.50
6.2 Flood return period ....................................................................................................... E.50
6.3 Duration of peak flow .................................................................................................... E.51
6.4 Flow velocity .................................................................................................................. E.51
6.4.1 Direct observation in flood conditions ........................................................... E.51
6.4.2 Mannings formula ........................................................................................... E.52
6.4.3 Flat terrain....................................................................................................... E.52

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E-v

6.5 Normal flow rate............................................................................................................ E.53


6.6 Perennial / seasonal flow ............................................................................................... E.53
6.7 Amount of debris in water ............................................................................................. E.53
6.8 Type of river (incised or alluvial) .................................................................................... E.53
6.8.1 Incised ............................................................................................................ E.53
6.8.2 Alluvial ............................................................................................................ E.53
6.9 Watercourse bank and bed characteristics .................................................................... E.53
6.10 Catchment area and shape ........................................................................................... E.54
6.11 Cross-sections at crossing point .................................................................................... E.54
6.12 Hydraulic gradient upstream and downstream ............................................................. E.55
6.13 Permeability of soil ........................................................................................................ E.55
6.14 Rainfall intensity............................................................................................................. E.55
7. MATERIALS ......................................................................................................................E.56
7.1 Stone and stone masonry .............................................................................................. E.56
7.1.1 Stone sources and extraction ......................................................................... E.58
7.1.2 Properties of stone ......................................................................................... E.60
7.1.3 Field Testing ................................................................................................... E.61
7.1.4 Mortars ........................................................................................................... E.62
7.1.5 Stone walls ...................................................................................................... E.63
7.1.6 Dry stone walling ............................................................................................ E.65
7.1.7 Hybrid walls .................................................................................................... E.66
7.1.8 Masonry culverts ............................................................................................. E.68
7.1.9 Gabion works .................................................................................................. E.69
7.2 Brick and block masonry ............................................................................................... E.71
7.2.1 Properties of bricks and blocks ....................................................................... E.73
7.2.2 Field testing .................................................................................................... E.75
7.2.3 Mortars for brick and block masonry .............................................................. E.76
7.3 Timber and organic materials ........................................................................................ E.77
7.3.1 Characteristics and utilisation of timber ......................................................... E.79
7.3.2 Properties of timber ........................................................................................ E.82
7.3.3 Field testing .................................................................................................... E.84
7.3.4 Uses of timber................................................................................................. E.86
7.4 Plain and reinforced concrete........................................................................................ E.87
7.4.1 Materials for concrete ..................................................................................... E.88
7.4.2 Production and placement ............................................................................. E.91
7.4.3 Properties of concrete .................................................................................... E.94
7.4.4 Field testing .................................................................................................... E.95
7.4.5 Uses of concrete ............................................................................................. E.98
8. STRUCTURE DESIGN .....................................................................................................E.100

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - vi

8.1 Scour ........................................................................................................................... E.100


8.1.1 Rule 1 - Provide minimum foundation or cut-off wall depths ....................... E.102
8.1.2 Rule 2 - Create a minimal constriction to the water flow .............................. E.102
8.1.3 Rule 3 - Avoid the use of piers...................................................................... E.105
8.2 Foundations................................................................................................................. E.106
8.3 Structural Slabs ............................................................................................................ E.107
8.3.1 Drifts ............................................................................................................. E.107
8.3.2 Concrete slab................................................................................................ E.109
8.3.3 Cement mortar bonded stone paving .......................................................... E.109
8.3.4 Hand pitched stone ...................................................................................... E.109
8.3.5 Gabions and gravel ...................................................................................... E.110
8.3.6 Slab construction (vented fords and large bore culverts) ............................. E.111
8.4 Cut-off walls ................................................................................................................. E.111
8.5 Pipes ............................................................................................................................ E.112
8.5.1 Vertical positioning of pipes ......................................................................... E.112
8.5.2 Pipe sizing ..................................................................................................... E.113
8.5.3 Pipe options.................................................................................................. E.121
8.5.4 Pipe inlets ..................................................................................................... E.127
8.5.5 Pipe bedding and cover arrangements ....................................................... E.128
8.5.6 Multiple culverts and vented fords .............................................................. E.130
8.6 Headwalls and Wingwalls ............................................................................................ E.131
8.6.1 Culverts ......................................................................................................... E.131
8.6.2 Wingwalls - Larger structures........................................................................ E.133
8.6.3 Stone, brick and blockwork walls.................................................................. E.134
8.6.4 Gabion baskets ............................................................................................. E.136
8.6.5 Timber walls .................................................................................................. E.137
8.7 Aprons ......................................................................................................................... E.137
8.7.1 Drift aprons ................................................................................................... E.137
8.7.2 Culvert aprons .............................................................................................. E.137
8.7.3 Vented ford aprons ....................................................................................... E.138
8.8 Approach ramps .......................................................................................................... E.138
8.9 Downstream Protection ............................................................................................... E.141
8.9.1 Rip-rap .......................................................................................................... E.144
8.9.2 Masonry slabs ............................................................................................... E.145
8.9.3 Gabions ........................................................................................................ E.145
8.9.4 Vegetation .................................................................................................... E.147
8.9.5 Steep channels ............................................................................................. E.148
8.9.6 Drain protection............................................................................................ E.149
8.10 Arches .......................................................................................................................... E.149

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - vii

8.10.1 Arch shape .................................................................................................... E.150


8.10.2 Formwork reuse ............................................................................................ E.150
8.10.3 Bridge/culvert layout .................................................................................... E.150
8.10.4 Construction sequence ................................................................................. E.151
8.10.5 Arch materials ............................................................................................... E.151
8.10.6 Fill options .................................................................................................... E.152
8.11 Bridge design .............................................................................................................. E.152
8.11.1 Choice of bridge site .................................................................................... E.153
8.11.2 Loading ......................................................................................................... E.153
8.11.3 Scour ............................................................................................................. E.154
8.11.4 Drainage ....................................................................................................... E.154
8.11.5 Maintenance ................................................................................................. E.154
8.11.6 Choice of structure ....................................................................................... E.155
8.11.7 Choice of materials and form of construction .............................................. E.155
8.11.8 Foundations .................................................................................................. E.156
8.11.9 Arch bridges ................................................................................................. E.156
8.11.10 Deck .............................................................................................................. E.159
8.11.11 Abutments .................................................................................................... E.163
8.11.12 Piers .............................................................................................................. E.164
8.11.13 Bearings and joints ...................................................................................... E.165
8.11.14 Parapets ........................................................................................................ E.166
8.12 Other design issues ..................................................................................................... E.167
8.12.1 Debris control ............................................................................................... E.167
8.12.2 Road signage ................................................................................................ E.167
8.12.3 Carbon footprint ........................................................................................... E.167
8.12.4 Climate change............................................................................................. E.167
9. CONSTRUCTION ...........................................................................................................E.168
9.1 Preparatory work ......................................................................................................... E.168
9.1.1 Planning of site works ................................................................................... E.169
9.2 Site works .................................................................................................................... E.172
9.2.1 Simple setting out techniques ...................................................................... E.173
9.2.2 Setting out culverts and drifts....................................................................... E.173
9.2.3 Setting out bridges and large structures ...................................................... E.179
9.2.4 Excavations ................................................................................................... E.179
9.2.5 Temporary works........................................................................................... E.181
9.2.6 Shuttering/formwork and steel reinforcement for concrete work ................. E.181
9.2.7 Steel reinforcement ...................................................................................... E.181
9.2.8 Concrete work .............................................................................................. E.183

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - viii

9.2.9 Precast concrete .......................................................................................... E.185


9.2.10 Timber stave culverts (Plate E.9.4) ................................................................ E.185
9.2.11 Masonry work ............................................................................................... E.186
9.2.12 Timber superstructure................................................................................... E.187
9.2.13 Earthworks/Backfilling................................................................................... E.189
9.2.14 Safety measures ............................................................................................ E.190
9.3 Site administration....................................................................................................... E.190
10. MAINTENANCE .............................................................................................................E.192
10.1 Managing the structure ............................................................................................... E.192
10.2 Maintaining the structure ............................................................................................ E.194
10.3 Common maintenance requirements .......................................................................... E.195
10.3.1 Drifts ............................................................................................................ E.195
10.3.2 Culverts ......................................................................................................... E.195
10.3.3 Vented drifts and large bore culverts............................................................ E.196
10.3.4 Bridge Maintenance ..................................................................................... E.197
11. REFERENCES .................................................................................................................E.199

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - ix

E LIST OF TABLES

Table E.2.1: Data Requirements ..........................................................................................................E.10


Table E.3.1: Design storm return period (years) ..................................................................................E.15
Table E.3.2: Typical loaded weights and dimensions of vehicles that may use low volume roads .....E.16
Table E.3.3: Safe Sight Stopping Distance (single lane) ......................................................................E.20
Table E.4.1: Drifts advantages and disadvantages ..............................................................................E.23
Table E.4.2: Culvert advantages and disadvantages ...........................................................................E.24
Table E.4.3: Vented fords/causeways advantages and disadvantages................................................E.27
Table E.4.4: Large bore culverts advantages and disadvantages........................................................E.29
Table E.4.5: Bridges advantages and disadvantages ..........................................................................E.30
Table E.4.6: Example comparison of timber and masonry bridge costs .............................................E.31
Table E.4.7: Closure Times ..................................................................................................................E.33
Table E.5.1: Trial Pits: Requirements and Locations ............................................................................E.38
Table E.5.2: Rock bearing capacity .....................................................................................................E.38
Table E.5.3: Clays and silts bearing capacity .......................................................................................E.39
Table E.5.4: Sands and gravels bearing capacity ................................................................................E.39
Table E.5.5: Minimum recommended relief culvert spacing ...............................................................E.41
Table E.6.1: Hydraulic and watercourse data required to undertake design ......................................E.48
Table E.6.2: Runoff coefficient: Humid catchment ..............................................................................E.50
Table E.6.3: Runoff coefficient: Semi-arid catchment ..........................................................................E.50
Table E.6.4: Flood Period Factors .......................................................................................................E.51
Table E.6.5: Mannings formula ............................................................................................................E.52
Table E.6.6: Roughness coefficient......................................................................................................E.52
Table E.7.1: Classes of rocks used for building ...................................................................................E.59
Table E.7.2: Durability Issues ...............................................................................................................E.61
Table E.7.3: Mortar Proportions by Volume ........................................................................................E.63
Table E.7.4: Heavy hardwoods ............................................................................................................E.79
Table E.7.5: Lighter hardwoods ...........................................................................................................E.79
Table E.7.6: Softwoods ........................................................................................................................E.80
Table E.7.7: Design stresses for the three principal timber groups .....................................................E.83
Table E.7.8: Limits of visible defects for structural timber from tropical hardwoods ...........................E.85
Table E.7.9: Concrete Types ................................................................................................................E.91
Table E.7.10: Concrete grades, strengths and batching strengths ......................................................E.91
Table E.7.11: Recommendations for good quality concrete .................................................................E.93
Table E.7.12: Maximum slump values for particular uses ......................................................................E.94
Table E.7.13: Agents of concrete deterioration ....................................................................................E.95
Table E.8.1: Guidance on design aspects .........................................................................................E.100

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E-x

Table E.8.2: Foundation depths ........................................................................................................E.102


Table E.8.3: Opening widths .............................................................................................................E.102
Table E.8.4: Scour depth adjustment ................................................................................................E.105
Table E.8.5: Bearing safety factor ......................................................................................................E.107
Table E.8.6: Cut-off wall locations .....................................................................................................E.111
Table E.8.7: Maximum flow rates.......................................................................................................E.114
Table E.8.8: Simplified Formulae for Calculation of Discharge Capacity ..........................................E.121
Table E.8.9: Masonry culverts ............................................................................................................E.123
Table E.8.10: Timber barrel culverts ....................................................................................................E.123
Table E.8.11: Timber log culverts ........................................................................................................E.125
Table E.8.12: Cast in-situ concrete culverts .........................................................................................E.125
Table E.8.13: Precast unreinforced concrete culverts ..........................................................................E.126
Table E.8.14: Steel culverts..................................................................................................................E.127
Table E.8.15: Pipe spacing and flow reduction factors........................................................................E.131
Table E.8.16: Wingwalls.......................................................................................................................E.133
Table E.8.17: Height of wingwall without surcharge ...........................................................................E.135
Table E.8.18: Height of wingwall with surcharge .................................................................................E.136
Table E.8.19: Height and width of gabion walls ..................................................................................E.137
Table E.8.20: Fill material in the approach way ...................................................................................E.139
Table E.8.21: Maximum water velocities .............................................................................................E.144
Table E.8.22: Stone sizes for rip-rap bed protection ...........................................................................E.145
Table E.8.23: Small Versus Large arches? ............................................................................................E.150
Table E.8.24: Minimum arch ring thickness .........................................................................................E.152
Table E.8.25: Sawn timber girder bridge deck for 6 ton vehicles .......................................................E.161
Table E.8.26: Nailing requirements .....................................................................................................E.162
Table E.8.27: Minimum deck clearances .............................................................................................E.167
Table E.9.1: Checklist of cost components for detailed costing of a structure .................................E.169
Table E.9.2: Checklist for planning site works ...................................................................................E.170
Table E.9.3: Checklist for preparing a construction programme ......................................................E.170
Table E.9.4: Recommended productivity standards ..........................................................................E.171
Table E.9.5: Checklist of handtools and site equipment ...................................................................E.172
Table E.9.6: Supervision check box - Excavation...............................................................................E.181
Table E.9.7: Supervision Check Box - Shuttering/formwork ..............................................................E.182
Table E.9.8: Supervision check box - Reinforcement steel fixing ......................................................E.183
Table E.9.9: Supervision check box - Concreting ..............................................................................E.185
Table E.9.11: Supervision check box - Backfilling................................................................................E.190
Table E.9.12: Checklist of site administration tasks .............................................................................E.190
Table E.10.1: Structure - Routine .........................................................................................................E.194
Table E.10.2: Structure - Periodic ..........................................................................................................E.19

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xi

E LIST OF FIGURES

Figure E.1.1: Flow diagram of the planning, design and construction process ...................................E.6
Figure E.2.1: Example of a Structural Survey Format .........................................................................E.13
Figure E.4.1: Key features of a stream drift ........................................................................................E.21
Figure E.4.2: Key features of a relief culvert .......................................................................................E.23
Figure E.4.3: A typical vented ford / causeway ..................................................................................E.25
Figure E.4.4: Key features of large bore culvert .................................................................................E.27
Figure E.4.5: Key features of a simply supported deck bridge ..........................................................E.29
Figure E.4.6: Large embankments required to prevent road flooding ...............................................E.32
Figure E.4.7: Complexity and Cost.....................................................................................................E.33
Figure E.4.8: ‘Route Map’ for the selection of a suitable structure ....................................................E.34
Figure E.5.1 Suitable crossing points for larger structures ....................................................................E.35
Figure E.5.2: Right angle crossings reduce the length and cost of structure required. .....................E.36
Figure E.5.3: Road alignment raised over culvert (Schematic) ...........................................................E.42
Figure E.5.4: Culvert drop inlets .........................................................................................................E.43
Figure E.5.5: Culvert alignment options .............................................................................................E.44
Figure E.5.6: Key design criteria for a vented ford .............................................................................E.45
Figure E.6.1: Catchment characteristics .............................................................................................E.54
Figure E.6.2: Stream crossing point cross-sections ............................................................................E.55
Figure E.7.1: Hammer test for stone ..................................................................................................E.62
Figure E.7.2: Rubble masonry wall .....................................................................................................E.64
Figure E.7.3: Squared masonry walling ..............................................................................................E.64
Figure E.7.4: Dry stone retaining wall.................................................................................................E.65
Figure E.7.5: Hybrid wall ....................................................................................................................E.67
Figure E.7.6: Gabion details ...............................................................................................................E.69
Figure E.7.7: Brick types .....................................................................................................................E.73
Figure E.7.8: Saw cuts ........................................................................................................................E.80
Figure E.7.9: Natural defects in timber .............................................................................................E.82
Figure E.7.10: Timber shape criteria ....................................................................................................E.85
Figure E.7.11: The constituents of concrete .........................................................................................E.88
Figure E.7.12: Cement stored in dry conditions ...................................................................................E.89
Figure E.7.13: Concrete mixes (guidance)............................................................................................E.92
Figure E.7.14: Field fines test ...............................................................................................................E.95
Figure E.7.15: Slump test mould ..........................................................................................................E.96
Figure E.7.16: Slump test .....................................................................................................................E.97
Figure E.8.1: Scour allowances .........................................................................................................E.103
Figure E.8.2: Scour – Fine silt ...........................................................................................................E.103

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xii

Figure E.8.3: Scour - Fine sand.........................................................................................................E.104


Figure E.8.4: Scour – Coarse sand....................................................................................................E.104
Figure E.8.5: Depth of scour around piers (from TRL ORN9) ...........................................................E.105
Table E.8.5: Bearing safety factor ...................................................................................................E.107
Figure E.8.6: Construction on sloping bedrock ................................................................................E.107
Figure E.8.7: Cut-off wall ..................................................................................................................E.112
Figure E.8.8: Culvert pipe requirements ..........................................................................................E.113
Figure E.8.9: Pipe design cases ........................................................................................................E.114
Figure E.8.10: Friction factor ..............................................................................................................E.116
Figure E.8.11: Flow Rate.....................................................................................................................E.116
Figure E.8.12: Restriction ..................................................................................................................E.117
Figure E.8.13: Headwater depth and capacity for corrugated metal pipe culverts with inlet control.
(Adapted from FHWA, 1998).......................................................................................E.118
Figure E.8.14: Headwater depth and capacity for concrete pipe culverts with inlet control.
(Adapted from FHWA, 1998).......................................................................................E.119
Figure E.8.15: Headwater depth and capacity for concrete box culverts with inlet control.
(Adapted from FHWA, 1998).......................................................................................E.120
Figure E.8.16: Transportation issues ...................................................................................................E.122
Figure E.8.17: Masonry Culverts.........................................................................................................E.122
Figure E.8.18: Timber log culvert details............................................................................................E.124
Figure E.8.19: Drop inlet on relief culvert ..........................................................................................E.127
Figure E.8.20: L-shaped inlet..............................................................................................................E.128
Figure E.8.21: Pipe granular bedding and cover ...............................................................................E.129
Figure E.8.22: Pipe arrangement with minimum cover ......................................................................E.130
Figure E.8.23: Possible culvert headwall positions .............................................................................E.131
Figure E.8.24: Position of culvert headwalls .......................................................................................E.132
Figure E.8.25: Headwall and wingwall arrangements ........................................................................E.133
Figure E.8.26: Stone, brick or blockwork wall with and without sloping backfill (surcharge) .............E.135
Figure E.8.27: Gabion baskets for walls ............................................................................................E.136
Figure E.8.28: Culvert apron ..............................................................................................................E.138
Figure E.8.29: Construction of approach ways ...................................................................................E.140
Figure E.8.30: Approach way cross section ........................................................................................E.140
Figure E.8.31: Energy dissipating apron ............................................................................................E.145
Figure E.8.32: Gabion protection on steep banks .............................................................................E.146
Figure E.8.33: Gabion protection on shallow banks...........................................................................E.146
Figure E.8.34: Gabion basket step waterfall ......................................................................................E.148
Figure E.8.35: Arch forces ..................................................................................................................E.150
Figure E.8.37: Two course arch...........................................................................................................E.151
Figure E.8.36: Use of tyre in formwork ...............................................................................................E.151

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xiii

Figure E.8.38: Corrugated metal sheet arch ......................................................................................E.152


Figure E.8.39: Arch bridge details ......................................................................................................E.157
Figure E.8.40: Abutment details .........................................................................................................E.163
Figure E.8.41: Gabion Abutment Log Deck Bridge ...........................................................................E.164
Figure E.8.42: Pier shapes (Plan view) ................................................................................................E.165
Figure E.9.1: Levelling with a water hose .........................................................................................E.173
Figure E.9.2: Setting out a right angle ............................................................................................E.173
Figure E.9.3: Use of batter boards ...................................................................................................E.173
Figure E.9.4: Culvert arrangement A (flat outfall) .............................................................................E.174
Figure E.9.5: Culvert arrangement B (intermediate outfall)..............................................................E.174
Figure E.9.6: Culvert arrangement C (steep outfall) .........................................................................E.174
Figure E.9.7: Procedure for setting out a culvert .............................................................................E.175
Figure E.9.8 (a) and (b): Setting out culvert profiles ............................................................................E.177
Figure E.9.9: Basic setting out techniques for structures .................................................................E.179
Figure E.9.10: Safety issues ................................................................................................................E.180
Figure E.9.11: Pumping arrangements ...............................................................................................E.180
Figure E.9.12: Formwork detail ..........................................................................................................E.181
Figure E.9.13: Wall Formwork detail ..................................................................................................E.182
Figure E.9.14: Reinforcement spacers ................................................................................................E.183
Figure E.9.15: Batch box ....................................................................................................................E.183
Figure E.9.16: Concrete hand mixing .................................................................................................E.184
Figure E.9.17: Masonry Templates .....................................................................................................E.186
Figure E.9.18: Timber beam launching ..............................................................................................E.187
Figure E.9.19: Log packing.................................................................................................................E.187
Figure E.9.20: Fixing deck timbers .....................................................................................................E.188
Figure E.9.21: Trimming running boards ............................................................................................E.188
Figure E.9.22: Kerb timber fixing .......................................................................................................E.189
Figure E.9.23: Backfilling ....................................................................................................................E.189

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xiv

E LIST OF PLATES

Plate E.1.1: Concrete arch bridge with masonry abutments and spandrels...................................... E.1
Plate E.1.2: Modular steel panel bridge ............................................................................................ E.4
Plate E.1.3: Steel cable pedestrian bridge ........................................................................................ E.4
Plate E.3.1: Overloading and risk of related structures failure are important considerations for some
routes ........................................................................................................................... E.16
Plate E.3.2: Robust width restrictions can be used in some instances to restrict heavy vehicles .... E.17
Plate E.3.3: Culvert head stone outside main running lanes........................................................... E.17
Plate E.3.4: Guide stones narrowing road width ............................................................................. E.18
Plate E.3.5: Pedestrian guard rails (in need of repair) ..................................................................... E.20
Plate E.4.1: A stream drift................................................................................................................ E.22
Plate E.4.2: A basic hand packed stone drift ................................................................................. E.22
Plate E.4.3: A three barrel corrugated steel culvert with stepped outfall apron ............................. E.25
Plate E.4.4: A vented ford ............................................................................................................... E.26
Plate E.4.5: Concrete vented causeway .......................................................................................... E.26
Plate E.4.6: Downstream protection to a vented ford ..................................................................... E.27
Plate E.4.7: Large bore arch culverts ............................................................................................... E.28
Plate E.4.8: Guide stones on large bore culvert structure ............................................................... E.28
Plate E.5.1: Collapse due to settlement .......................................................................................... E.37
Plate E.5.2: Flat arid area location of the low point ........................................................................ E.40
Plate E.5.3: Ponding at culvert outlet .............................................................................................. E.41
Plate E.5.4: Long culvert ................................................................................................................. E.43
Plate E.6.1: Flood debris caught on vegetation .............................................................................. E.49
Plate E.7.1: Dry jointed, multiple opening, cantilevered laterite stone culvert approximately
1,000 years old and still in service. .............................................................................. E.57
Plate E.7.2: Dry jointed, laterite, stone viaduct, approximately 1,000 years old, still in service on a
main road. .................................................................................................................... E.57
Plate E.7.3: Examples of established Ethiopian techniques of dry stone walling ........................... E.58
Plate E.7.4: Hand dressed stone ..................................................................................................... E.60
Plate E.7.5: Hand quarrying of stone .............................................................................................. E.60
Plate E.7.6: Hand quarried granite stone blocks ............................................................................. E.60
Plate E.7.7: Manual quarry operations can create local employment,
however good management is required to ensure a safe working environment......... E.60
Plate E.7.8: Rubble stone ............................................................................................................... E.63
Plate E.7.9: Dry stone wall supporting embankment ...................................................................... E.65
Plate E.7.10: Dressed dry stone wall ................................................................................................. E.66
Plate E.7.11: Hybrid retaining wall .................................................................................................... E.67
Plate E.7.12: Masonry culvert with steel former to be left insitu after construction .......................... E.68

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xv

Plates E.7.13-15:Dry stone masonry culvert extension works - Dry stone Conservancy, USA ............. E.68
Plate E.7.16: Gabions supporting a road on a steep slope............................................................... E.69
Plate E.7.17: Filling gabion baskets .................................................................................................. E.70
Plate E.7.18: Gabion retaining wall used as river bank protection.................................................... E.70
Plate E.7.19: Natural fibre matting inserted in gabion protection works to encourage
rapid growth of vegetation cover................................................................................ E.70
Plate E.7.20: Steel mesh gabions ...................................................................................................... E.70
Plate E.7.21: Gabion river crossing ................................................................................................... E.70
Plate E.7.22: Fired clay brick is a very versatile material. This elliptical brick arch bridge over the
River Thames at Maidenhead, UK was constructed in 1838 and has two main spans of
38 metres. It is still in use today (designer: Isambard Kingdom Brunel) photograph
by Jon Combe.............................................................................................................. E.71
Plate E.7.23: Fired clay brick structures exist in many countries and still provide service
for today’s traffic after centuries of use ........................................................................ E.72
Plate E.7.24: High quality burnt clay bricks produced in small ‘beehive’ kilns fired by rice husk.
Bricks can be fired using coal, wood or agricultural waste products ........................... E.72
Plate E.7.25: Simple burnt clay brick clamp ...................................................................................... E.72
Plate E.7.26: Bricks baked in a kiln .................................................................................................... E.73
Plate E.7.27: Bricks in a bridge pier .................................................................................................. E.77
Plate E.7.28: Steel beams being moved into position using labour methods,
without the use of cranes; ready for completion with a timber deck ........................... E.78
Plate E.7.29: Timber deck on recycled steel beams ......................................................................... E.78
Plate E.7.30: Completed timber deck ............................................................................................... E.78
Plate E.7.31: Timber bridge deck, pier and abutments .................................................................... E.78
Plate E.7.32: Timber poles for piles .................................................................................................. E.81
Plate E.7.33: Timber poles as deck beams of a twin span bridge..................................................... E.81
Plate E.7.34: Timber decked bridge ................................................................................................ E.86
Plate E.7.35: Log abutments & deck ................................................................................................. E.87
Plate E.7.36: Treated timber culvert .................................................................................................. E.87
Plate E.7.37: Aggregate crushed and screened by hand ................................................................. E.89
Plate E.7.38: Steel reinforcement cage being assembled................................................................. E.90
Plate E.7.39: Concrete culvert rings .................................................................................................. E.98
Plate E.7.40: Concrete drift under construction ................................................................................ E.98
Plate E.7.41: Reinforced concrete slab .............................................................................................. E.99
Plate E.7.42: Flow spreader structure at the outlet of a mitre drain on a steep fragile slope.
Bio-engineering planting should be established downhill of such structures .............. E.99
Plate E.7.43: A simple reinforced concrete walled box culvert ready to receive pre-cast top
slab units (timber or reinforced concrete) .................................................................... E.99
Plate E.8.1: Bridge damage due to scour of abutment................................................................. E.101
Plate E.8.2: Failure of structure due to a combination of constriction of the watercourse
(structure too small), scour and inadequate protection of abutments. ...................... E.106
Plate E.8.3: Guide stones at the edge of a drift ............................................................................ E.108

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xvi

Plate E.8.4: Concrete slab drift...................................................................................................... E.109


Plate E.8.5: Hand pitched stone drift ............................................................................................ E.110
Plate E.8.6: Gabion drift ................................................................................................................ E.110
Plate E.8.7: Natural fibre matting inserted in the top and face of the gabion .............................. E.111
Plate E.8.8: Ponding at culvert outfall ........................................................................................... E.115
Plate E.8.9: Ponding in outfall channel.......................................................................................... E.115
Plate E.8.10: Timber barrel culvert .................................................................................................. E.123
Plate E.8.11: Timber pole culvert during reconstruction ................................................................. E.124
Plate E.8.12: Reusable steel mould for cast in-situ culvert .............................................................. E.125
Plate E.8.13: Cast in-situ culvert ...................................................................................................... E.125
Plate E.8.14: Steel mould ................................................................................................................ E.126
Plate E.8.15: Precast concrete culvert ring ...................................................................................... E.126
Plate E.8.16: Steel culvert................................................................................................................ E.127
Plate E.8.17: Drop inlet on stream culvert....................................................................................... E.128
Plate E.8.18: Skew culvert crossing with splay wingwall to prevent water continuing in side drain E.128
Plate E.8.19: Poor quality jointing and bedding support for pipe culvert ...................................... E.129
Plate E.8.20: Wingwall cascade....................................................................................................... E.134
Plate E.8.21: Masonry side drains at the edge of an approach way ............................................... E.141
Plate E.8.22: Example of damage due to lack of downstream protection ..................................... E.141
Plate E.8.23: Minor erosion in watercourse upstream from culvert ................................................. E.142
Plate E.8.24: Serious erosion downstream from the same culvert due to concentration of flow and lack
of appropriate protection measures........................................................................... E.142
Plate E.8.25: Severe erosion downstream of a relief culvert due to inadequate protection
measures .................................................................................................................... E.143
Plate E.8.26: Severe erosion downstream of a cross culvert due to inadequate protection
measures .................................................................................................................... E.143
Plate E.8.27: Gabion erosion control structure downstream of a culvert ........................................ E.147
Plate E.8.28: Gabion basket step waterfall ..................................................................................... E.148
Plate E.8.30: Semi-circular Arch bridge ........................................................................................... E.149
Plate E.8.29: Masonry Arch ............................................................................................................. E.149
Plate E.8.31: Two span concrete deck bridge with masonry pier and abutments ......................... E.153
Plate E.8.32: Reinforced concrete deck on masonry abutments and pier .................................... E.155
Plate E.8.33: Masonry arch under construction with wooden formwork ......................................... E.158
Plate E.8.34: Completed masonry arch bridge with splayed wing walls on hard rock foundations E.159
Plate E.8.35: Timber deck with floor planking and edge beams in need of repair ......................... E.161
Plate E.8.36: Log stringers from underside of bridge deck............................................................. E.162
Plate E.8.37: Brick pier .................................................................................................................... E.165
Plate E.8.38: Raised kerbs on a vented causeway allow water to pass over the structure with minimum
disturbance, but provide protection from vehicles driving off the structure .............. E.166
Plate E.8.39: Timber bridge with handrails ..................................................................................... E.166

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - xvii

Plate E.8.39: Road narrows sign ...................................................................................................... E.167


Plate E.9.1: Construction of culvert inlet apron with guide pegs and strings ............................... E.178
Plate E.9.2: Completed culvert inlet apron ................................................................................... E.178
Plate E.9.3: Hand mixing mortar on a clean surface ..................................................................... E.184
Plate E.9.4: Timber stave culvert ................................................................................................... E.185
Plate E.9.5: Example of Masonry works ........................................................................................ E.186
Plate E.10.1 and Plate E.10.2: Culvert cleaning tool .......................................................................... E.195
Plate E.10.3: Box culvert partially blocked by vegetation ............................................................... E.196
Plate E.10.4: Cleaning the drop inlet and barrel of a large bore culvert ........................................ E.196
Plate E.10.5: Serious vehicle impact damage to steel panel bridge .............................................. E.198

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 1

1. INTRODUCTION

Part E of the LVR Design Manual - 2011 deals with small drainage and watercourse crossing structures,
typically up to 10 metres span (as shown in Plate E.1.1), and retaining structures. It provides detailed
guidance on the processes involved with the planning, design, construction and maintenance of small
drainage and other structures for low volume roads.

It is clear that road structures are an important aspect of road design and construction. Unfortunately it
is an aspect that is often given little or insufficient attention which is shown by the fact that when roads
become impassable it is usually where they cross a watercourse. Although the length of road structures
forms only a very small fraction of the total road length the time spent on their design must be a much
greater portion of the total planning and design process.

There are manuals for the design of structures on Ethiopia’s main roads1. The predominant construction
materials used are concrete and steel. However, little guidance has hitherto been available concerning
small structures, particularly with respect to the optimum use of resources such as labour, local skills (which
may include masonry and carpentry), local materials and small local enterprises, while still achieving
durable and adequate structures. Intelligent use of these resources will often produce the lowest cost
structures. This is particularly important in the limited resource environment expected to prevail in the
LVR sector in Ethiopia for some time. It is certainly not advisable to blindly apply standards, practices and
‘rules of thumb’ derived from rich economies for use in Ethiopia where the balance of influential factors
such as labour wage rates, availability and cost of standard materials and equipment, skills, access to
finance and the support environment can be very different.

Plate E.1.1: Concrete arch bridge with masonry abutments and spandrels

1 ERA Bridge Design Manual - 2011

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 2

This LVR Design Manual - 2011 Part E on Small Structures aims particularly to satisfy the need to assist
engineers and technicians in the planning and provision of road structures by:
ƒ providing concise and complete information in one document;
ƒ explaining the steps required in the design process;
ƒ providing different levels of information depending on the complexity of the structure;
ƒ providing guidance on costing, construction and maintenance of structures;
ƒ assisting in the approval and adoption of low cost structural designs.

The lack of access for designers and planners to design information and other resources requires this
Manual to provide all the basic information needed in the design of small structures up to about 10
metres in span. References are provided at the end of Part E for more complex structures or problems,
where these issues were considered outside of the scope and objectives of the Manual.

Part E of this Manual has been written as a design guide, to complement existing national design codes
and standards from the Ethiopian Roads Authority. It is also intended that this Part of the Manual will
assist in the process of establishing more comprehensive and appropriate planning, design, construction
and maintenance procedures and practices for small structures.

Investigations and fieldwork have shown that steps in the design process are often missed or neglected.
Therefore the steps that should be carried out and the reasons for undertaking them are explained along
with it the type and detail of data that are required and how they should be used in order to undertake
a design.

1.1 Definition of a small road structure


For the context of this Manual a road structure is a construction which provides support and/or drainage
to the road carriageway or associated road works. In practical terms this manual deals with structures of
span up to 10 metres. Roads form a barrier to the natural drainage of surface water from the surrounding
land into streams, lakes and rivers. In the absence of any control arrangements the water would find its
own way across the road, resulting in gullies and washouts along the road. An effective drainage system
is therefore the most important Part of a low volume road and should protect the road from damage due
to water. The most basic drainage provision is the camber of the road carriageway which directs water
off the road to each side. Water is then removed from the road by the side and mitre (turn-out) drains. In
some cases it may be necessary for water to be moved across the road at a low point in the alignment
or at a stream, for example. As quantities of runoff water build up in the side drains at low points in the
alignment or at watercourses, it will be necessary to allow water to cross from the high side of the road
to the lower. This Part of the Manual deals with the road structures required to manage the drainage of
water across a road. The other features of the drainage system are dealt with in Part D of the Manual with
the design standards shown in Part B.

1.2 Scope of the guidance


The scope of Small Structures guidance in the Manual includes:

Rural / Urban Roads


Although the Manual primarily discusses issues associated with the design and construction of structures
on low (traffic) volume rural roads, many of the ideas and design factors discussed are applicable to urban
and peri-urban roads. In these cases it will be necessary to consider pedestrian issues in more detail.
Existing built infrastructure and planned development can also influence options with regard to the siting,
type, size and ancillary works associated with structures design.

Paved / Unpaved Roads


The majority of low volume roads will be unpaved. However, many of the structures discussed in this
Manual will also be suitable for low volume paved roads. Roads may initially be built to earth or gravel
surface standard and then upgraded by spot improvements or comprehensive paving to partial or fully
sealed/paved roads at a later date. Road structures designed and constructed with reference to this

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 3

Manual will be suitable for paved roads provided that possible increased loadings and higher design
standards such as roadway widths are satisfied.

Structural assessment
Although this Manual is primarily a design guide, principally dealing with the design and construction of
road structures, it may also serve as a useful reference for the assessment and maintenance of existing
structures. As assessment is a check of an existing design, the Manual highlights structural aspects which
should be checked during an inspection and assessment under an appropriate asset management and
maintenance regime.

Reconstruction / Construction / Maintenance


The Manual primarily deals with new structures; however, the design principles are the same for
reconstruction, rehabilitation, extension and upgrading of existing structures. In these cases it may be
possible to make use of elements of existing structures, for example, using an old drift slab as downstream
protection for a new piped drift built adjacent to the existing structure. The Manual deals with the
construction aspects of structures which are of interest to supervisors and engineers. The construction
chapter concentrates on the management and supervision of the construction that must be undertaken
by the field officer/engineer overseeing the project.

Better use of local resources


Adoption of the recommendations in this Manual will increase the use of local material and labour
resources. This will help to relieve the constraints that road authorities face due to a shortage of funding
and may allow a foreign exchange saving as fewer materials may have to be imported. The increased
use of local labour will assist in stimulating the local economy and greatly reduce the mobilisation costs
of road construction. The maintainability of structures will also be improved as the skills required will be
established during the construction phase within the local community.

Unskilled and semi-skilled labour could be utilised for a range of tasks in the construction of road structures,
such as timber growing preparation and formwork, quarrying dressing and crushing stone, fired clay
brick production, local transport, masonry and brickwork in structures, retaining walls, ditch linings and
culverts, collection and preparation of river gravel for structural fill, and construction of components such
as gabion baskets. The creation of jobs in the area will not only provide socio-economic development but
will also allow the development of skills which will have three benefits. Firstly, there will be the capacity
in the local community and enterprises to undertake maintenance on the structures; secondly, there will
be an increase in employment opportunities in other construction sectors for the labourers employed on
the road works; and thirdly, studies have also shown the employment generation multiplier effect of jobs
created on rural infrastructure works.

1.3 Types of structures


The Manual covers a wide range of drainage structures from drifts to small bridges (Chapter 4 describes
the characteristics of these structures). These structures vary in complexity and are ranked in order of
increasing complexity as follows:
1. Drifts
2. Simple culverts
3. Vented fords
4. Large bore culverts
5. Small bridges

It is difficult to define the boundaries between the categories above: for example, when does a large
bore culvert become an arched bridge? The background information, site data and technical knowledge
and support required to undertake the design also vary significantly. This Manual therefore addresses
the information required for the more complex structures but also indicates the reduced level of survey
and technical knowledge required to design more simple structures. Other road structures which are not
covered in the Manual include large bridges and viaducts. Further information on these structures can be
found in the Bridge Design Manual - 2011.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 4

The Manual does not cover steel girder or lattice frame structures as these structures require specialist
design and erection expertise. Neither does the Manual cover modular panel steel bridges (eg Bailey) as
shown in Plate E.1.2. These bridges are intended as a temporary steel structure that can be erected at
short notice from panels that would normally be held in a store.

Plate E.1.2: Modular steel panel bridge

Modular panel bridges are suitable for short term measures where an unforeseen flood disrupts access
at a critical location. The steel panel bridges can also be readily dismantled and the units returned to
store once a permanent solution has been constructed on the access alignment. For such structures the
specialist manufacturers’ manuals should be used.

This Manual also excludes suspension and suspended steel cable bridges (Plate E.1.3) for pedestrian,
animal and light motor traffic. Design and construction of such bridges are covered in specialist
documentation developed for application in Ethiopia. (Ethiopian Suspended Trail Bridge Survey, Design
and Construction Manual)

Plate E.1.3: Steel cable pedestrian bridge

Chapters E.1 to E.7 cover planning and initial design assessment of structures. Chapters E.8 to E.10
focus on detailed design, construction and maintenance. The separate Volume of LVR Standard Drawings
contains some standard designs of appropriate structures in A3 format. It is intended that standardising
these designs which have been independently checked technically, will result in:
ƒ reduced design costs and economies of scale, leading to an improvement in cost and quality;
ƒ increased speed of construction, as labourers, supervisors and engineers will become more familiar
with the standardised design;
ƒ simplified approval procedures.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 5

For complex structures, for bridges with main spans of more than 10 metres or bridges expected to be
trafficked with trucks of gross weight more than 6 tonnes, the ERA Bridge Design Manual - 2011 should
be used.

1.4 How to use Part E


There is a logical sequence of work that must be undertaken in the selection and design of any road
structure. This Part of the Manual is laid out in sequence with each chapter covering one aspect of
the process shown in the diagram Figure E.1.1. The two initial tasks which should be carried out are
to identify the problem or task (Chapter E.2) and determine the design criteria (Chapter E.3) for the
structure. The initial design data may then be collected (Chapter E.2) to enable the preliminary design
to be carried out. Preliminary design, shaded yellow in the flow diagram, involves four different stages
which may be performed a number of times before a design solution is proposed. It is suggested that
a review of structural options (Chapter E.4) is initially undertaken followed by an appraisal of a potential
construction site (Chapter E.5). The water flow characteristics of the watercourse (Chapter E.6) should
then be considered before a selection of the most appropriate structure is made (Chapter E.4). It is likely
that the preliminary design loop will need to be followed a number of times to review different potential
structures and construction sites.

Following completion of the preliminary design the proposed design solution should be checked to
ensure that it complies with the design criteria. Detailed design of the structure can then be undertaken
(Chapter E.8) which will require further reference to be made to chapters covering site selection and
appraisal (Chapter E.5) and watercourse characteristics (Chapter E.6). It will also be necessary to review
the options for construction materials (Chapter E.7) that may be available.

A separate Volume of the LVR Design Manual - 2011 contains standard design drawings that may be
useful and contribute to the preparation of design drawings. During supervision of the construction work
(Chapter E.9) it will be necessary to ensure materials used in the structure meet and are used according to
the specifications. This may require additional reference to the materials chapter and the Specifications.
Chapter E.10 covers the maintenance requirements of structures after they have been built and highlights
the problems that may be encountered if maintenance is not carried out.

Depending on the complexity of the structure, the level of work and detail required at each stage will
vary. Although each stage of the design process shown in the Figure must be covered, it may be possible
to skip more detailed issues in each chapter for simple structures such as drifts or culverts. Throughout
the subsequent chapters there is guidance to indicate which sections may be ignored depending on the
type of structure to be built.

For complex structures, for those with main spans of more than 10 metres, and structures crossing other
roadways or railways, the ERA Bridge Design Manual - 2011 should be used.

A qualified civil, highway or structural engineer should certify all bridge and completed structures as fit
for purpose.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 1 - 6

A road structure
is required

Identify the problem Determine design criteria


Chapter E.2 Chapter E.3

Collect initial design data


Chapter E.2

Review structure options


Chapter E.4

Select preferred site for


Determine of structure
structure
Chapter E.4
Chapter E.5

Watercourse Considerations
Chapter E.6

Does the proposed solution


meet the design criteria?

Carry out detailed design Collect detailed design data


Chapter E.8 Chapter E.5, E.6

Prepare design drawings,


Materials Chapter E.7 BoQs and specification
Chapter E.8

Supervise construction work Maintain finished structure


Chapter E.9 Chapter E.10

Figure E.1.1 Flow diagram of the planning, design and construction process

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 7

2. PROJECT PLANNING

2.1 Setting priorities


The approach adopted assumes that a road network and the associated structures are the responsibility
of a road authority. From time to time there will be a requirement for new, rehabilitated or upgraded
structures. The approach is also applicable for a ‘one-off’ initiative to provide, replace or rehabilitate
a structure by an authority or community group. It may take many years to construct all the roads and
associated structures to all-weather standard required by a community due to the limited financial
resources and the capacity of the available equipment and labour. Priorities must therefore be set on the
order that work should be undertaken. It may be possible to build a high priority road in the short term,
but construct some of the structures at a later date. However, these roads may be seasonally impassable
until the structures have been completed. A more pragmatic strategy with limited resources may be to
initially provide all of the structures and durable surfacing on problem sections of the route (Basic Access
strategy), and provide an engineered earth surface to the remainder of the route until additional resources
are available to attain a more durable road surface throughout. This can be termed a stage construction,
spot improvement or differential upgrading strategy. In setting priorities the following factors should be
taken into account.

General
ƒ The first question to be answered is “will a low cost drift suffice until resources for a more expensive
structure can be mobilised?”;
ƒ Reconstruction of a damaged structure may have a higher priority over provision of a new structure
in a different location.

Wereda road network / location


ƒ The level of priority given to the road/structure within the road inventory;
ƒ The location of the road in relation to other structures/roads. For example, is there an alternative
route with an acceptable detour?
ƒ The requirements of access for construction. Is it necessary to construct a new road or upgrade an
existing alignment before work can commence on the structure?
ƒ Proximity to other work in order to avoid transportation of labour equipment and materials over
long distances.

For example, if there are 3 potential structures that are required and two are close together while the
other is a long distance away. It will be more efficient to construct the two structures that are close
together at the same time as labour and equipment can easily be transferred between the two sites. If
the programme requires the construction of two structures that are a long distance apart it would be less
efficient to move labour and equipment between the two sites as the construction demand varied.

Road category
ƒ The class of road and hence its strategic importance within the road network.
ƒ The design level of structure required on the road network which will determine the resources and
time required for construction.

Work status
Any work that has already commenced should be given the highest priority for funding in order to be
completed so that the benefits of the investment already made will be realised.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 8

Justification
A simple cost benefit analysis and assessment of social benefits can be useful, both to raise the finances
and compare the various options to utilise the available resources.
ƒ Need: An assessment of the number of people who will benefit from the construction of the road
or structure coupled with the availability of other access in their area. Improved access to important
services such as health centres should also be considered.
ƒ Costs: The cost of providing one road or structure should be compared against providing another.
For example, if a budget of 200,000 Birr is available would the best option be to construct one
structure which costs 200,000 Birr or provide 5 smaller structures around the road network, which
only cost 40,000 Birr each?

Resource availability
It will be necessary to make an assessment of the resources (equipment, labour, artisans, supervisors,
materials, enterprises) which may be available in the locality. Assessments must also be made for the
timeframe required to obtain equipment and materials from other areas. Labour may not be so freely
available in agricultural areas at certain times of the year. Specific skills may need to be trained or imported
into the locality. It would also be easier to manage if the labour resource requirements were steady rather
than increasing and decreasing throughout the year. For many authorities the expected timing of funding
availability from internal/external sources (and possible conditionality) is an important consideration.

Climatic factors
In regions which have a pronounced wet and dry season, or occasional flooding, it may only be possible
(or much more straightforward) to undertake construction in the dry season. Drifts constructed in seasonal
streams will not require the additional cost and time for diverting the water or providing cofferdams if
they are built during the dry season.

2.2 Assessment of the problem or need


In order to set priorities it will be necessary to assess the general condition of the road network, highlighting
which roads require improvement and where new or improved structures are required. This assessment
should allow the responsible engineers to:
ƒ Prioritise construction of structures;
ƒ Calculate the structures programme budget requirements;
ƒ Develop work programmes and construction timeframe;
ƒ Identify resource requirements.

This information can then be collated for senior planning engineers to co-ordinate the overall budget
and resource requirements for the whole road network. The basis of this work should be an inventory of
all structures (or required structures) on the road network. TRL Overseas Road Note 7 provides guidance
on the preparation of such inventories.

It is essential that detailed assessments are undertaken at each structure site as structures form a large
percentage of the overall cost of the road infrastructure. Assessments undertaken at sites of proposed
structure locations should be sufficiently detailed to ensure:
ƒ Enough time is spent identifying the best location for the structure. (If the road is already built and
the structure is being upgraded it may not be possible to identify new crossing sites);
ƒ The appropriate type of structure is chosen;
ƒ The structure is adequate for the purpose (traffic type and numbers, water flows and size etc.);
ƒ The design should not need to be significantly changed during construction, as this would result in
an increase in the cost of the structure.

2.3 Assessment of potential structures


The main issues to be decided during the assessment of new structures are:
ƒ Type of structure - Chapter E.4;
ƒ Location of the structure - Chapter E.5;
ƒ Size of structure - Chapter E.6.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 9

The assessment may be undertaken for either a new structure or the upgrading of an existing structure.
In either case the design work will be similar. There are two main stages to be undertaken in the planning
and assessment of potential structures, or Site Investigation.

2.3.1 Desk Study

The desk study should allow the designer to develop an initial idea of the size and possibly type of
structure required. The following information should be obtained and assessment made at this stage:
ƒ Obtain a map of the area. Ensure that it shows the important features (roads, villages, watercourses
and contours);
ƒ Mark the catchment areas on the map and calculate the catchment size for each structure location;
ƒ Review the topography of the area.

2.3.2 Field study

The following should be undertaken as Part of the field study:


ƒ Prepare a sketch map of potential site (s) - plan and x-section;
ƒ Field investigations of the soil conditions and strengths (See Part D, Chapter 5);
ƒ Surface exploration - identify soil types in the water course for potential erosion;
ƒ Sub surface exploration - trial pits;
ƒ Record results in tables or on maps;
ƒ Site survey, including water measurements (See Part D, Chapter 6);
ƒ Determine section and gradient near potential sites by surveying the watercourse;
ƒ Determine area of the waterway for normal and flood flows;
ƒ Check local resource availability;
ƒ Cross check information with local community members regarding flood levels, frequency and
duration.

The value of consulting with the local community should not be underestimated, particularly with regard
to levels and frequency of flooding and waterborne debris. Also in generally flat terrain their knowledge of
the extent and direction of flood flows is valuable. They should be able to inform regarding local materials
and labour/skills resources and any seasonal accessibility problems. The consultation opportunity should
also be used to listen to any concerns and allay fears regarding the potential effects or impact of any new
or rehabilitated structure.

2.4 Collection of initial design data


The collection of the initial design data will affect the primary choice of structure. This design data should
be gathered during the desk and field studies. Table E.2.1 shows the range of data that can be collected
in the assessment of the potential structure, which is discussed in more detail in the various chapters. It
will not be necessary to collect all the data for the more simple structures. Some of the data may also be
collected on a further visit during more detailed survey of the selected structure site.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 10

Table E.2.1: Data Requirements

Category Item Information required


ƒ Is there an availability of trade skills in the locality eg carpentry, stone
masons?
ƒ What is the standard of workmanship available?
ƒ Options of:

Labour
1. Specialist skills vs. training
local labour
2. Time/cost vs. skills transfer
and ongoing maintenance
potential
ƒ Labour wage rates
Local Resources

ƒ What is the availability of local materials (eg masonry stone (rough/


Materials (Part E.7)

dressed), timber, locally manufactured brick and blockwork)?


ƒ What is the strength, quality, durability and quantity of local materials?
ƒ Steel: what are the imported and delivery costs to site, delays, welding,
bending and fixing skills available?
ƒ Cement: what are the strengths achievable, delivery/ import delays,
types of concrete and experience, quality control and possible testing
arrangements?
ƒ What are the unit costs of materials?
Equipment

ƒ What basic specialist equipment is available / would be required for


construction AND maintenance (transport, production, loading unloading,
mixing, placing, craneage etc.?)
What are the costs of equipment (including transport and servicing costs)?

ƒ What is the reliability of the collected data?


General

ƒ Is a separate structure needed to allow work to commence further along


the road?
ƒ What will be the cost for construction AND maintenance?
ƒ Do pedestrians, animals or IMTs frequently travel along the road?
ƒ What is the class of road?
Are local standards established for structures on this category of road?
Design Criteria
(Chapter E.3)

ƒ What is the largest type of vehicle that uses the road?


Does vehicle and axle load data exist?
ƒ If funds are severely constrained, is a one lane, alternate traffic flow
option feasible?
Traffic

ƒ What is the traffic density, does it vary eg seasonally or on market days in


the local wereda or kebele?
ƒ Review standards used elsewhere & recommend appropriate ones. Will
the vehicle size or loading increase if the road or structure is improved
(new or re-routed traffic)?
ƒ Are any exceptional loads transported? - check for logging, quarries,
mining or other industries in the area. What are the possible traffic,
economic and safety implications?

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 11

Category Item Information required

New
ƒ Which types of structure would be acceptable?

Type of Structure ƒ What is the general condition of the structure?


(Chapter E.4) ƒ What was the original design life?

Existing Structure
ƒ Do as-built records exist?
ƒ Are there indications of maximum flood levels on structure?
ƒ Are there any signs of post construction settlement?
ƒ What are the main problems with the existing structure?
ƒ Are there failures in any of the structural elements?
ƒ What is the current level of scour around structure?
ƒ Indications of excessive loading or abuse?
ƒ Dimensions and any possibility of refurbishment or adaptation?

ƒ Is the depth to firm strata or rock known?


ƒ What type of material is available to build on for foundations?
Site Selection
(Chapter E.5)

ƒ What is the level of the water table?


General

ƒ What is the compressibility or strength of subsoil?


ƒ What is the best location of trial pits - to provide the most valuable
information?
ƒ Is the water / soil chemistry aggressive to building materials? (specialist
advice may be required)
ƒ Is the stream perennial or seasonal?
ƒ What is the type of watercourse? (meandering, straight, bends, presence
of weeds)
ƒ Is the watercourse and bed stable, eg in rock?
ƒ What is the low water level?
ƒ What are the minimum or normal flow levels?
ƒ What are the maximum flood levels (MFL)? (frequency of occurrence and
Watercourse Details

duration)
ƒ What are the watercourse cross sections at potential site?
ƒ What is the gradient of watercourse upstream and downstream of the
crossing point?
Water Parameters

ƒ Is there evidence of course/bank or level changes, erosion/deposition


(Chapter E.6)

at the site, upstream or downstream? Consult with old maps and the
community
ƒ Is there sometimes floating debris in the water?
ƒ What is the water velocity during floods?
ƒ What is the longitudinal section or profile along the watercourse? Is
the watercourse used for private or commercial traffic with headroom
requirements?
ƒ Size and amount of sediment supplied from catchment area.
ƒ Area of catchment?
Catchment Details

ƒ Are sudden floods encountered?


ƒ Shape of catchment?
ƒ Gradient of terrain?
ƒ Permeability of soil?
ƒ Vegetation coverage and type?
ƒ Rainfall intensity?
ƒ Is the vegetation coverage changing rapidly eg Deforestation?

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 2 - 12

The importance of collecting accurate information cannot be over emphasised. Although it may prove
difficult to collect the required data it is not good practice to make superficial or un-supported assumptions
as this will almost invariably result in higher costs due to either additional resources being required to
amend the design during construction or the structure being unfit for its purpose.

2.5 Field assessment practicalities


To undertake a survey of a new road or site for a structure it is usually necessary to have the following
equipment:
ƒ Vehicle - with an odometer;
ƒ Map of the road network;
ƒ Note book;
ƒ Tape measure;
ƒ Ranging rods;
ƒ Graduated line and weight for measuring water depths;
ƒ Hammer, nails, wooden stakes and paint for site survey marks;
ƒ Abney level (or simple survey level) and survey staff (only required for bridges);
ƒ Camera (optional - may be useful for recording potential sites for reference in design office);
ƒ Shovel and pickaxe/mattock for trial holes;
ƒ Materials sample bags;
ƒ Containers for water samples;
ƒ Dynamic Cone Penetrometer (DCP) for soil strength assessment (desirable);
ƒ Water craft for deep water sites;
ƒ GPS.

It is likely that more complex structures will require a second or even third site visit in order to collect the
necessary detailed information required. These visits will probably require additional survey equipment
to determine more accurate levels. Information about land use will also need to be collected from the
whole catchment area upstream of the potential construction site.

Always ask the question:


What is the contribution of the information to the design process?

Following initial field investigations along a potential route or rehabilitation/ improvement of an existing
route, the field engineer should compile a Table of the structural works which may be required. An
example of a Structural Survey form is shown in Table E.2.2. This Table can be used to assess the physical
resources and financial costs required to provide the structures. It can also be utilised in assessing
priorities and determining work plans for construction units.

The actual costs of structures will vary according to local resource costs and factors. The benefits of
keeping a database of actual and estimated construction costs cannot be overemphasised. Because of
the many factors that influence local costs and construction practices it is highly risky to transfer unit cost
knowledge from one location to another, and most certainly between regions and countries. There is no
substitute for careful consideration of all local cost components and variables.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


Date: Road From - To: Surveyed by:

Existing Structures & Reference Points Preliminary Survey Recommendations

Diameter/Height Culvert Diameter/Height


Culvert
Length*
Length* Culvert? Culvert/
No. Chainage or
Principal or Bridge Structural Bridge Proposed
Type Pipe Remarks bridge Pipe
Material bridge span Height Condition span Height solutions
Diam. width* Diam.
width* (m) (m) (m) (m)
(mm) (m) (m)
(m)

Remove soil and debris.


1 0+260 PC RC 6.00 0.60 1Ø600 0.60 Good Sitted Check + Survey outlet
conditions
Erosion base slab
2 0+477 PC RC 6.50 1.00 2Ø1000 1.00 Fair Repair Base Slab
at RHS
3 1+111 Bridge RC 5.50 4.60 - 2.30 Good No work required
Provide gabion
4 1+864 Drift Masonry 5.50 12.50 - - Fair Erosion on LHS
protection
Headwall
5 2+106 AC Brick 5.50 1.20 - 0.80 Good demolished by Rebuild brick headwall
vehicle LHS
Channel erosion
RHS downstream. Prelim assessment
Vented Masonry/
6 2+750 6.00 8.50 1Ø600 0.60 Good Substantial 2Ø1000? indicates enlargement
Drift Concrete
erosion around needed
structure
Culvert installed too
The Culvert is 2/3
low. Remove top slab
7 3+113 BC RC 6.50 1.20 - 0.60 Good filled by soil and 1.2
and increase opening
debris
to 1.2m
0.6 metre step at outlet.
Severe erosion at Requires new catchpit
8 3+367 PC RC 6.00 0.60 1Ø600 0.60 Fair
outlet RHS structure and repairs
downstream
Water crossing
Provide new 1Ø600
9 3+960 road but no 1Ø600
culvert
structure exists

Figure E.2.1: Example of a Structural Survey Format


In centre of village
10 4+335 Drift Masonry 5.50 6.40 Poor with steep eroded 6.5m ~6m New bridge
approaches
Cascade and catchpit
Erosion on cutting
11 4+900 4.5 required, reline cut-off
face RHS
ditch

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


LHS Parapet rail Replace one section of
12 5+215 Bridge RC 6.50 5.50 - 4.00 Good
damaged parapet rail.
Key
* Clear width across carriageway between headwalls or kerbs RC = Reinforced Concrete
PC = Pipe Culvert RHS = Right hand side in direction of increasing chainage
BC = Box Culvert LHS = Left hand side in direction of increasing chainage
E - Chapter 2 - 13

AC = Arch Culvert
E - Chapter 3 - 14

3. DESIGN CRITERIA

3.1 Selecting design parameters


Ethiopia has national standards established for the design of road structures on the primary road network.
However, these standards may be inappropriate for the size and level of traffic on low volume roads. For
example, vehicle loading is based on the largest long distance haulage trucks which rarely use some
minor roads in their fully loaded condition. Designs based on these standards would therefore usually
incur excessive construction costs. Unfortunately, heavy and overloaded trucks are commonplace on some
routes in Ethiopia due to factors of driver/operator discipline, economic pressures, or other local factors.
This can lead to vehicle and axle loading being experienced well in excess of those in accordance with
the national loading regulations. Such occurrences are usually related to haulage of particular products
such as bulk fuel, minerals, construction materials and timber. Therefore, when designers are selecting
design parameters for a particular structure they must ensure that they are appropriate for the conditions
that will be experienced on that particular road. Examples of the factors which designers should consider
are:
ƒ What is the nature and loading of traffic currently using the route? (Carry out loading surveys if
necessary). Are conditions likely to change substantially in the foreseeable future? (eg could new
quarrying operations start up?)
ƒ Are local design standards established for the relevant road category? Are these appropriate or
achievable?
ƒ If overloading is prevalent, are there realistic possibilities to physically restrict access?
ƒ What are the cost implications relating to the loading criteria or restrictions?

It is impossible to state definitive design criteria in this Manual as overall site conditions will vary between
locations and weredas. The information given below should be considered as a guide to designers, and
adapted according to specific conditions in the area or the structure being designed.

3.2 Design life


The design life of a structure is the length of time that the structure can be expected to carry traffic
without reconstruction or replacement of structural elements. It assumes that throughout the life of the
structure regular standard maintenance is carried out.

When determining the structure’s design life, the factors which must be taken into account are:
ƒ Expected life spans for different structure types and materials;
ƒ Expected initial and recurrent costs for the design life options;
ƒ Finance currently available and future maintenance / rehabilitation finance probability;
ƒ Future changes in the use of the road (eg increased traffic volumes or loadings);
ƒ Flood return periods (see below);
ƒ Consequences of structural failure;
ƒ Likely influence of climate change on future life of the structure, risk and consequences of failure.

The design life of the road itself (ie the length of time before the road will become obsolete or require
substantial improvement) should also be taken into account. After consideration of all of the relevant
local factors, it is probable that a design life of between 10 and 40+ years will be appropriate for an
individual structure. The selected design life should be clearly stated in the design dossier.

3.3 Design flood


One of the major design factors in the selection and size of road structures is the amount of storm water
that will flow past the structure. Each year there will usually be a few heavy storms which will result in
peaks in the water flow over or through the structure, but the largest of these peaks will vary in size each
year. If the flows are recorded over a number of years, a longer period of recording will result in a larger
maximum peak flow. The highest known flood that has ever occurred may be referred to as the high

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 15

flood. For minor structures on low volume roads the designer cannot be expected to propose a design
that is so large or wide that it could cope with a storm water flow of the high flood. Structures should
therefore be designed to have the capacity to cope with a smaller flood; for example, the largest flood
that occurs every 10 years. This flood is called the design flood and the time period between successive
design floods is called the return period. The design flood is the largest flood that is practical and/or
economic for design. Structures should withstand the design flood without any significant damage to
the structure or adjacent road and/or embankments. Structures will have a design life greater than the
return period between design floods. The designer should therefore consider the effects on a structure
of a flood that is larger than the design flood to ensure that significant or unacceptable damage will not
occur. Further information about return periods is given in Chapter E.6 of this Manual.

In addition to the practical and economic considerations, the choice of return period for a design should
be based on the risk of failure of the structure if a larger flow is encountered. It can be very difficult for the
designer to undertake this risk analysis with the limited data that may be available. Table E.3.1 therefore
shows suggested return periods for design flood flows for different types of structures.

Table E.3.1: Design storm return period (years)

Geometric design standard2


Structure type
DC4 DC3 DC2 DC1
Gutters and inlets (1) 2 2 2 1
Side ditches (1) 10 5 5 2
Ford (1) 10 5 5 2
Drift (1) 10 5 5 2
(1)
Culvert diameter <2m 15 10 10 5
Large culvert diameter 2 - 6m 25 15 10 5
Gabion abutment bridge 25 20 15 -
Short span bridge 6 - 10m 50 25 15 -
Note
1. These periods should be doubled if the alternative route in the event of a drainage failure is more than an additional
75km, or no alternative exists.
2. For further guidance see Table D.5.1 and D.5.2 in Part D.

Clearly drifts and vented drifts may be overtopped during or after any storm. In these cases the design
period would indicate a peak flow where it would be impossible for a vehicle to cross the structure
safely for an extended period. This period would be determined according to the road’s importance in
the network. The strategic importance of a structure should also be considered. For example, will it be
possible to use an alternative route if the structure is temporarily unusable or damaged? The selected
storm return period should be clearly stated in the design dossier.

3.4 Traffic categories and widths


Careful consideration must be given to the types of vehicles which may use the road, both at the present
time and in the future after road improvements have been made. For example, if the road is close to
quarries or a logging area, extremely heavy vehicles may travel down the road. While it may be possible
to establish a weight restriction on vehicles using the road due to the loads that particular structures can
carry, they are often ignored by drivers and operators. It may only take one overweight vehicle to destroy
a structure and make the road impassable. Engineers should therefore design structures to withstand the
load of any vehicle that could travel down the road. Typical loaded weights and vehicle dimensions are
shown in Table E.3.2.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 16

Table E.3.2: Typical loaded weights and dimensions of vehicles that may use low volume roads

Typical max. Length Width


Vehicle
weight (kg) (m) (m)
Bicycles 250 - -
Motorcycles 400 2 1
Carts 1500 - -
Car / pick up 2500 5 1.75
4WD pick up 3000 5 1.75
Minibuses 5000 7 2
Tractor & trailers 12 000 10 2
2 axle small/medium
17 000 8 2.5
trucks
Large buses 25 000 15 2.5
2/3 axle heavy trucks 30 000 10 2.5
5/6 axle heavy truck &
60 000 18 2.5
trailer combinations (1)
Note:
1. usually used for paved main road and urban routes only

Experience has shown that some locations are particularly prone to grossly overloaded vehicles. If vehicle
overloading is common practice the suggested vehicle weights may be up to twice some of the values
shown in Table E.3.2.

If a type of vehicle can physically travel down a road then one of these vehicles will almost
certainly pass down that road at some time in the life of the structure – therefore structures should
be designed to withstand the weight of the heaviest vehicle which can pass down the road.

Signage should be provided to clearly state the loading capacity of any structure if it is limited in any way.
Local road network managers and administrators should also be made aware of any load limitations and
the likely consequences of these being exceeded as illustrated in Plate E.3.1.

Plate E.3.1: Overloading and risk of related structures failure


are important considerations for some routes

With the resources available, if it is not possible to construct a crossing which will withstand the largest
vehicle that could travel down the road shown in Table E.3.2, it will be necessary to install a robust non-
removable barrier each side of the structure to prevent overloaded vehicles crossing (Plate E.3.2).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 17

Plate E.3.2: Robust width restrictions can be used in some instances to restrict heavy vehicles

When the structure is designed, the size of vehicle should also be taken into consideration to ensure that it
can safely cross the structure without damage to the vehicle or structure.

The scope of this Manual covers low volume roads generally carrying up to 300 motor vehicles per day
equivalent. However it is recognized that with double digit annual percentage increases in traffic typical
of some rural routes, the current flow volumes could at least triple even in a 10 year design period. The
width of a structure will substantially influence the initial construction cost, for bridges the cost is roughly
proportional to deck area and for culverts, roughly proportional to barrel length. In a severely constrained
resource environment a vital decision is therefore required with respect to whether one or two way traffic
flow will be accommodated over the structure. It is probable that two way traffic for bridges will only
be justifiable for some category DC4 roads and above; although local conditions may override this. The
secondary decision is with respect to the safe width for the predominant traffic type and driver behaviour.
These decisions become more important with the increasing size of the proposed structure.

For culverts, a typical provision rate for rolling terrain will be about two or three per km. In severe terrain
or in flat, floodable areas the frequency will be expected to be higher. However, it should be noted that
a culvert or other drainage structure is required in all low points in a road. The cost of their provision is
usually significant in the overall cost of the low volume road provision, particularly for unpaved roads. The
frequent occurrence of culvert headwalls and width narrowings, and the difficulty for drivers to see them in
advance particularly for travel at night without public lighting and hazard signing, raises important safety
issues. The provision of minimum two-lane width culverts can therefore often be justified in all except the
most constrained finance resource situations. Furthermore, culvert headwalls should not restrict the general
roadway width. They should be set back behind the carriageway and shoulder, and clearly marked or
have guide stones at each end of the culvert to prevent vehicles driving into the inlets, outfalls or ditches
when passing on-coming traffic (Plate E.3.3). These requirements may be relaxed to provide only clear
carriageway width in slow speed mountainous alignments.

Plate E.3.3: Culvert head stone outside main running lanes

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 18

The argument for restricting larger structures to one lane is more easily supported. At the very basic
level, bridges for loaded motorcycle and bicycle traffic on village access tracks can be provided with a
carriageway width from about 1.5 metres.

For single lane motor vehicle traffic the clear carriageway width (between kerbs or guide stones) is
recommended to be a minimum of 3.65 metres.

If the traffic is mostly light in nature (motorcycles, cars, carts or light goods vehicles) then a 4.6 metres
‘one and a half’ lane option may be appropriate to allow for the occasional safe passage of a heavy goods
vehicle.

Where justifiable, full two lane motor traffic provision should allow a minimum of 6.5 metres between
kerbs provided that vehicles are restricted to slow speed passage.

Where physical restrictions are necessary to prevent passage of heavy good vehicles these will need to
limit free passage to about 2.3 metres.

It is recommended that the carriageway width (between kerbs or guide stones) should be between 3.75
and 4.5 metres for larger structures such as drifts, vented drifts and bridges. This width should allow easy
single way traffic but restrict two vehicles from passing on the structure (see Plate E.3.4).

Plate E.3.4: Guide stones narrowing road width

It is likely that these width restrictions will result in a reduction in the general road width which will require
a clear indication that the roadway narrows (advance warning signs) as recommended by the national
standards for the category of road shown in Part B.

Although the widths given above should generally be followed, cross drainage structures are difficult to
widen at a later stage. Consideration must therefore be given at the planning stage regarding the future
use of the road and whether the traffic volumes are expected to increase significantly. It may prove more
cost effective to construct a structure wider than current requirements in order to avoid reconstruction at
a later date.

It is evident that close liaison is required with the road alignment designer in the selection of and decision
on structures width.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 19

3.5 Design code


Bridge decks and structural components should be designed according to the Design code set out in
ERA Bridge Design Manual - 2011 and selected vehicle loading.

3.6 Serviceability
Vehicle Impact: One of the most common causes of damage to structures is vehicle impact. It is therefore
important that reinforcement be placed in culvert headwalls and guide stones to prevent them being
demolished by traffic. Safety barriers should be installed in the situations of particular hazards, according to
the Ethiopian standards for the category of road.

Fatigue Deflections: The majority of codes in use limit deflections to prevent fatigue damage to structural
members by specifying permissible deflections as a function of length. Typically the permissible deflection
is 1/800 of the span length. It will be suitable to relax this requirement to a deflection of 1/100 of the span
for LVR small structures (ie a 6mm deflection on a 6m span bridge) if only one vehicle will be on the bridge
at one time and this level of deflection will not be noticed when compared to the ride from the approach
roads.

3.7 Drainage of the structure


There should be a camber or cross fall on any highway structure to ensure that water does not collect
and lay on the structure, increasing the rate of deterioration or acting as a safety hazard. A minimum
camber of 2.5% will normally be acceptable. Bridges should be constructed with adequate drainage
arrangements, such as pipes, which drain water off or through the deck away from abutments or piers.
Careful consideration should be given to water flow in the side drains along the road adjacent to the
structure to ensure that it does not erode a deep channel along the side of the structure.

3.8 Maintenance capability


When materials are chosen, consideration should be given to the predicted life of the material in relation
to the design life of the whole structure. The resources required and frequency of maintenance should
also be carefully reviewed.

3.9 Safety
Where there are a large number of pedestrians using the road, provision should be made for a 1.5m wide
segregated footway across or on the side of the structure. If the structure is over 20m long but the number
of pedestrians cannot economically justify a pedestrian footway it may be advisable to construct a limited
wider section in the middle of the structure (or regular refuges) where pedestrians can wait safely while
vehicles pass. In some cases it may be justifiable to construct a separate low cost, lightweight structure for
pedestrian passage.

Guard rails and kerbs can be provided to prevent vehicles or pedestrians from falling off the structures.
For structures which have pedestrians regularly crossing, it is highly advisable to construct some form of
guard rail to prevent pedestrian and child accidents. This guard rail will not normally be required to restrain
vehicles from falling off the structure (Plate E.3.4). The provision of guard rails or kerbs to prevent vehicle
accidents will depend on the level of vehicle traffic. It is unlikely that vehicle guard rails can be economically
justified where the vehicle flows are less than 50 vpd. If vehicle guard rails are not provided it is imperative
that clearly marked kerbs or kerb stones are provided to indicate the extent of the roadway lanes. Where
the structure is designed to be overtopped it is necessary to indicate the depth of water over the roadway
and whether it is safe to cross. As it will normally be safe to cross fast flowing water up to a depth of 200mm,
guide stones on overtopped structures should be made at least 200mm high. The stones will then remain
visible and mark the edge of the roadway when the structure is safe to cross and be submerged under the
water when it is unsafe to cross the structure. Guard rails should not be used on structures that are designed
to be overtopped as they will trap debris.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 3 - 20

Sight distances: Drivers should always be able to see the road far enough ahead to be able to stop safely
if required (eg when there is an obstruction on the road). It may not be possible to maintain a full sight
distance over a cross drainage structure. However, the distances in Table E.3.3 provide a guide to the
desirable minimum distance that should be provided.

Plate E.3.5: Pedestrian guard rails (in need of repair)

Table E.3.3: Safe Sight Stopping Distance (single lane)

Speed km/h 30 40 50 60
Distance m 50 70 100 130
(Source: TRL, 1984 Towards Safer Roads in Developing Countries - A Guide for Planners and Engineers)

3.10 Future changes in road use


During the initial design of the structure, careful consideration must be given to probable future changes
in road use. For example, the type of traffic and number of vehicles of each type which may affect the
requirements of the structure must be taken into account. The future changes should be reviewed for the
predicted life of the structure but consideration should be given to the financial costs of building a larger
structure if a smaller, simpler structure will be acceptable for the majority of the design life.

3.11 Funding
In selecting design parameters and ultimately the choice of structure, the economic benefits of different
types of structures should be taken into account. These economic considerations do not only include the
physical costs of the structure and measurable benefits of increased access, lower transport costs, time
savings and increased economic activity but also social benefits of increased access. For example, it may
be considered beneficial to provide a small bridge across a river which will provide constant access to
a health centre for a village on the opposite bank. A vented ford may be more suitable for the level of
traffic using the road, but high flood flows may prevent a patient receiving treatment in an emergency.

In many cases, engineers will not have all the financial resources that they need to satisfy all the structures
needs. If a structure is to be provided which does not fully meet the design requirements, the design
should enable the structure to be upgraded at a later date with minimal reconstruction if further resources
become available.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 21

4. STRUCTURAL OPTIONS

The greatest potential cost savings for water crossing options is in the choice of structure type. This
chapter considers different water crossing options, explaining the characteristics of each and highlighting
the conditions suitable for their use. The advantages and disadvantages associated with each structure
are also discussed.

At the most basic level, a ford can be created in a stable sandy bed of an occasional watercourse by
burying stones of 15 - 30 cm size just below the surface and re-covering them with sand. This substantially
improves bearing capacity for vehicles. The Chapter deals with improved crossing structures from drifts
to small bridges with spans of <10m.

4.1 Drifts
Drifts are the most basic structure and can be the lowest cost form of watercourse crossing construction.
There are two types of drift:
Relief drifts: relieve side drains of water where the road is on sloping ground and water cannot be
removed from the uphill side drain by mitre drains, or as an alternative to a relief culvert.
Small watercourse (or stream) drifts: where stream flows are very small (as five years flow 6m3 / sec)
or perennial, drifts may be used to allow the stream to cross the road (see Figure E.4.1 and Plate E.4.1).

Figure E.4.1: Key features of a stream drift

Drifts can also be referred to as Irish bridges, fords or splashes. The terms describe essentially the same
structure, however, it is generally accepted that a ford or splash is constructed from the existing riverbed
eg a sandy river bed or level rock. A drift is a ford or splash with an improved running surface constructed
from imported (or gathered) materials an example of which is shown in Plate E.4.2. A low water crossing
is the collective term used to describe all drifts, fords, splashes and vented fords.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 22

Plate E.4.1: A stream drift

Plate E.4.2: A basic hand packed stone drift

4.1.1 Key features

The key features of drifts are:


ƒ Stream drifts are structures which provide a firm place to cross a river or stream. Relief drifts transfer
water across a road without erosion of the road surface. Water flows permanently or intermittently
over a drift, therefore vehicles are required to drive through the water in times of flow.
ƒ Drifts are particularly useful in areas that are normally dry with occasional heavy rain causing short
periods of flood water flow.
ƒ Drifts provide a cost effective method for crossing wide rivers which are dry for the majority of the
year or have very slow or low permanent flows.
ƒ Alternative solutions may be preferable for small permanent watercourses to prevent vehicles
having to drive through the water.
ƒ Drifts are particularly suited to areas where material is difficult to excavate, thus making culverts
difficult to construct.
ƒ Drifts are also particularly suited in flat areas where culverts cannot be buried because of lack of
gradient.
ƒ The drift approaches must extend above the maximum design flood level flow to prevent erosion
of the road material.
ƒ If necessary guide stones should be provided on the downstream side of the drift and be visible
above the water when it is safe for vehicles to cross the drift.
ƒ Buried cut off walls are required upstream and downstream of the drift to prevent under cutting by
water flow or seepage.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 23

ƒ The approach road level will normally mean that approach ramps are required. Approach ramps
should be provided to the drift in the bottom of the watercourse with a maximum gradient of 10%
(7% for roads with large numbers of heavy trucks).
ƒ Drifts should not be located near or at a bend in the river.
ƒ Some form of protection is usually required downstream of a drift to prevent erosion.

The advantages and disadvantages of using drifts for water crossings are shown in Table E.4.1.

Table E.4.1: Drifts advantages and disadvantages

Advantages Disadvantages
ƒ Low cost: at the most basic level, can be ƒ Drifts require vehicles to slow down when
constructed and maintained entirely with local crossing
labour and materials ƒ The crossing can be impassable to traffic
ƒ Ease of maintenance and repair during flood periods
ƒ Volume of excavated material in most cases is ƒ Foot passage can be inconvenient or
minimal hazardous when water is flowing
ƒ Do not block with silt or other debris carried
by flood water.
ƒ Can accommodate much larger flows than
culverts
ƒ Easier to repair than culverts
ƒ Water flows over a wide area, resulting in less
water concentration and erosion downstream
than piped culverts

4.2 Culverts
Culverts are the next step upwards from drifts in terms of cost and complexity of structure. There are two
types of culvert:
Relief culverts: at low points in the road alignment or where there is no definable stream, but the
topography of the ground requires a significant amount of cross drainage, which cannot be accommodated
by side drains (See Figure E.4.2).
Stream culverts: which allow a watercourse to pass under the roadway.

Figure E.4.2: Key features of a relief culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 24

4.2.1 Key features

The key features of culverts are:


ƒ Culverts are the most commonly used structures on low volume roads. They can vary in number
from about one each km in dry and gently rolling terrain up to six or more for severe terrain with
high rainfall. In high rainfall, flat areas the frequency may also increase to allow water to cross the
road alignment in manageable quantities.
ƒ Culverts channel water under the road, avoiding the need for vehicles to drive through the
watercourse.
ƒ In addition to well defined water crossing points culverts should normally be located at low points
or dips in the road alignment.
ƒ Relief culverts may be required at intermediate points where a side drain carries water for more
than about 200 metres without a mitre drain or other outlet.
ƒ Culverts can be pipe, box, slab or arch type.
ƒ Headwalls are required at the inlet and outlet to direct the water in and out of the culvert and
prevent the road embankment sliding into the watercourse. Wingwalls at the ends of the headwall
may also be used to direct the water flow and retain material.
ƒ Aprons with buried cut off walls are also required at the inlet and outlet to prevent water seepage,
scouring and undercutting.
ƒ Culvert alignment should follow the watercourse both horizontally and vertically where possible.
ƒ Gradient of the culvert invert should be between 2 and 5%. Shallower gradients could results in
silting whereas steeper gradients result in scour.
ƒ Culvert invert levels should be approximately in line with the water flow in the stream bed, otherwise
drop inlet and/or long outfall excavations may be required.
ƒ Common culvert diameters are 600mm and 900mm.
ƒ Cross culverts smaller than 600mm in diameter should not be installed as they are very difficult to
clean.
ƒ Where foundation material is poor, culverts should be placed on a good foundation material to
prevent settlement and damage. On very soft ground, it may be necessary to consider concrete,
steel or timber piles to provide adequate foundations. This will require specialist design expertise
not covered by this manual.
ƒ It is necessary to protect the watercourse from erosion downstream from the structure.
ƒ Culverts can exist in pairs or in groups to enable larger stream flows to be accommodated using
standard unit designs. An example of a three-barrel corrugated steel culvert is shown in Plate E.4.3.
ƒ When silt supply is high, pipe culverts shall not be used.

The advantages and disadvantages of using culverts are shown in Table E.4.2.

Table E.4.2: Culvert advantages and disadvantages

Advantages Disadvantages
ƒ Culverts provide a relatively cheap and ƒ Regular maintenance is often required to
efficient way of transferring water across a road prevent the culvert silting up, or to remove
ƒ Can be constructed and maintained primarily debris blockage
with local labour and local materials ƒ Culverts act as a channel, forcing water flow
ƒ Culverts allow vehicle and foot passage at all to be concentrated, so there is a greater
times potential for downstream erosion compared
ƒ Culverts do not require traffic to slow down with drifts
when they are crossed ƒ Culverts are not suited to occasional high
ƒ Culverts allow water to cross the road at volume flows
various angles to the road direction for a
relatively small increase in costs

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 25

Plate E.4.3: A three barrel corrugated steel culvert with stepped outfall apron

4.3 Vented Fords and Causeways


These generally have higher capacity and construction costs than drifts or culverts. A typical vented ford/
causeway is illustrated in Figure E.4.3.

Figure E.4.3: A typical vented ford / causeway

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 26

4.3.1 Key features

The key features of vented fords and causeways are:


ƒ These structures are designed to pass the normal dry weather flow of the river through pipes below
the road (see Plate E.4.4 and E.4.5). Occasional larger floods pass through the pipes and over the
road, which may make the road impassable for short periods of time.
ƒ Vented causeways are the same concept as vented drifts but are longer with more pipes, to cross
wider watercourse beds.
ƒ The level of the road on the vented drift should be high enough to prevent overtopping except at
times of peak flows.
ƒ There should be sufficient pipes to accommodate standard flows. The location of pipes in the drift
will depend on the flow characteristics of the river.
ƒ Vented fords should be built across the whole width of the water- course.
ƒ A vented ford requires approach ramps, which must be surfaced with a non erodible material and
extend above the maximum flood level.
ƒ Watercourse bank protection will be required to prevent erosion and eventually damaging the entire
structure
ƒ The approach ramps should not have a steeper grade than 10% (7% where there is significant heavy
vehicle traffic).
ƒ The upstream and downstream faces of a vented drift require buried cut off walls (preferably down
to rock) to prevent water undercutting or seeping under the structure.
ƒ An apron downstream of the pipes and area of overtopping is required to prevent scour by the water
flowing out of the culvert pipes or over the structure.
ƒ There is also a requirement to protect the watercourse from erosion downstream from the structure
(see Plate E.4.6). There will be considerable turbulence immediately downstream of the structure in
flood conditions.
ƒ The road surface longitudinal alignment of the vented ford should be a slight sag curve to ensure
that, at the start and end of overtopping, water flows across the centre of the vented drift and not
along it.
ƒ There should be guide stones on each side of the structure to mark the edge of the carriageway and
indicate when the water is too deep for vehicles to cross safely.
ƒ Vented fords can also be known as piped drifts

Plate E.4.4: A vented ford Plate E.4.5: Concrete vented causeway

The advantages and disadvantages of using vented fords and causeways are shown in Table E.4.3.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 27

Table E.4.3: Vented fords/causeways advantages and disadvantages

Advantages Disadvantages
ƒ Vented fords can allow a large amount of ƒ Vented fords can be closed for short periods
water to pass without overtopping during periods of flooding and high flow
ƒ They are cheaper to construct and maintain ƒ Floating debris can lodge against the
than bridges upstream side of the structure and block pipes
ƒ Construction of vented fords is fairly ƒ Foot passage can be inconvenient or
straightforward compared with bridges hazardous when water is flowing
ƒ Vented fords are well suited to cope with short
high volume flows
ƒ Can be constructed and maintained primarily
with local labour and local materials

Plate E.4.6: Downstream protection to a vented ford

4.4 Large Bore Arch Culverts


Lareg bore arch culverts are illustrated in Figure E.4.4 with typical examples shown in Plate E.4.7 and
E.4.8.

Figure E.4.4: Key features of large bore culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 28

4.4.1 Key features

The key features of large bore arch culverts are:


ƒ Large diameter culverts typically have openings greater than 1 metre and are capable of passing
high flows, either through one large opening or a number of medium sized openings.
ƒ Very large bore arch culverts may also be called arch bridges.
ƒ Formwork is required to construct the openings. This formwork can be made from wood, stones
or metal sheeting and either incorporated into the structure or removed once construction is
complete.
ƒ Although these structures are not in generally designed to be overtopped, they can be designed
and constructed to cope with an occasional overtopping flood flow.
ƒ The road alignment needs to be a minimum of 2 metres above the bottom of the watercourse.
ƒ Approach embankments are required at each end of the structure.
ƒ Large bore culverts require solid foundations with a buried cut off wall on both upstream and
downstream sides to prevent water seepage erosion and scouring.
ƒ These structures require large amounts of internal fill material during construction.
ƒ Guide stones or kerbs should be placed at the edge of the carriageway to increase vehicle safety.
ƒ If the crossing is to be used by pedestrians, consideration should be given to installing guard rails
and central refuges for long crossings where pedestrians can move off the roadway to allow traffic
to pass.
ƒ Water from the road side drains should be carefully channelled into the watercourse away from the
structure to prevent erosion of the bank or scour of the culvert structure.

Plate E.4.7: Large bore arch culverts

Plate E.4.8: Guide stones on large bore culvert structure

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 29

The advantages and disadvantages of large bore arch culverts are shown in Table E.4.4.

Table E.4.4: Large bore culverts advantages and disadvantages

Advantages Disadvantages
ƒ Large bore culverts are usually easier and ƒ The water opening in large bore culverts is
cheaper to construct than bridges smaller than for a bridge of the same size,
ƒ They can accommodate flows significantly which reduces the potential flow rate past the
higher than smaller culverts and vented fords structure at peak flows
ƒ Can be constructed and maintained primarily ƒ Large bore culverts can require a significant
with local labour and materials, without the amount of internal fill material
need for craneage
ƒ They may easily be designed and constructed
for occasional overtopping
ƒ Central ‘piers’ are not so susceptible to
damage by scour and erosion when compared
with bridge piers
ƒ They generally require less maintenance than
conventional bridges

An alternative to a large or multi-bore culvert is a reinforced concrete box culvert. This type of structure
is not covered by the Manual. For guidance on such structures refer to the ERA Bridge Design Manual
- 2011 and publications such as TRL Overseas Road Note 9.

4.5 Bridges (arch or simply supported deck)


These are generally the highest cost structures to construct. This Manual does not cover multiple span
bridges, which may be simply supported or continuous over piers. For such structures and bridges with
spans more than 10 metres, refer to the ERA Bridge Design Manual - 2011. Simply supported bridge
decks are illustrated in Figure E.4.5.

Figure E.4.5: Key features of a simply supported deck bridge

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 30

4.5.1 Key features

Key features of the bridges covered in this Manual are:


ƒ The arch is the simplest form of bridge.
ƒ There are a number of different elements to a simply supported deck bridge, which will comprise
of a superstructure (deck, parapets, guide stones and other road furniture) and substructure
(abutments, wingwalls, foundations, piers and cut off walls).
ƒ Bridges are generally the most expensive type of road structure, requiring specialist engineering
advice and technically approved designs.
ƒ Bridges can be single span or multi span, with a number of openings for water flow and intermediate
piers to support the superstructure.
ƒ The main structure is always above flood level, so the road will always be passable.
ƒ Abutments support the superstructure and retain the soil of the approach embankments.
ƒ Wingwalls are needed to provide support and protect the road embankment from erosion.
ƒ Embankments must be carefully compacted behind the abutment to prevent soil settlement, which
would result in a step on the road surface at the end of the bridge.
ƒ Weep holes are needed in the abutment to allow water to drain out from the embankment, and
avoid a build up of ground water pressure behind the abutment.
ƒ Bridges should not significantly affect the flow of water (ie the openings must be large enough to
prevent water backing up and flooding or over topping the bridge).
ƒ The shape of the abutments and piers will affect the volume of flow through the structure and also
the amount of scouring.
ƒ Bridges require carefully designed foundations to ensure that the supports do not settle or become
eroded by the water flow. On softer ground this may require piled foundations which are not
covered in this Manual.
ƒ Water from the road side drains should be channelled into the watercourse to prevent erosion of
the bank or scour of the abutment structure.
ƒ Guide stones or kerbs should be placed at the edge of the carriageway to increase vehicle safety.
ƒ If the crossing is to be used by pedestrians, consideration should be given to installing guard rails
and a central refuge for long crossings where pedestrians can move off the roadway for passing
traffic.

Advantages and disadvantages of bridges with spans <10 m are shown in Table E.4.5.

Table E.4.5: Bridges advantages and disadvantages

Advantages Disadvantages
ƒ The road is always passable as the structure ƒ Bridges are normally significantly more
should not be overtopped expensive than other road structures
ƒ Simple arch bridges can be constructed ƒ They are more complex than other structures
primarily with local skills and local materials, and will require specialist engineering
without the need for craneage (however simply support for design and construction
supported spans are more complex) ƒ Additional height of and earthworks in
approach embankments
ƒ Bridges may require heavy duty lifting cranes
for the deck components
ƒ Although all structures should be inspected
for defects, bridges require regular detailed
checks
ƒ Bridges are likely to fail if flood flow
predictions are incorrect and they are over
topped
ƒ A small amount of scour and erosion can
often result in major damage to the structure

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 31

4.6 Structure selection


The objective in selecting a structure for a water crossing is to choose the most appropriate design for
each location. This selection should be based on the factors outlined in the following sections.

4.6.1 Costs

Assessments will have to be made of the initial cost of construction which should include materials,
transportation, equipment, labour, and supervision as well as overheads (and for a contractor, the profit
margin). An assessment will also have to be made of the on-going maintenance costs that will be required
for each structure.

The example in Table E.4.6 compares the costs of a timber bridge with a masonry vented ford. Initially it
may appear that the timber bridge is the cheaper option but even without inflation over the first 15 years,
the masonry culvert can be shown to be the cheaper when whole life costs are considered. Furthermore,
there may be risks that funding will not be available for maintenance, or that defects will not be identified
and repaired in a timely manner on a high maintenance structure.

Table E.4.6: Example comparison of timber and masonry bridge costs

Timber Bridge Masonry vented ford


Year
Work Undertaken Cost Work undertaken Cost
1 Construction 10000 Construction 15000
Inspection and replacement of Repair of downstream
4 1000 300
running boards protection
Inspection, replacement of
Replacement of
8 running boards and 2 deck 2000 700
downstream protection
members
Inspection and replacement of Repair of downstream
12 4000 300
decayed structural members protection
Inspection replacement of Replacement of
15 1000 700
running boards downstream protection
Total Cost 18 000 17 000

4.6.2 Amount of traffic per day / acceptable duration of traffic interruptions

The amount and type of traffic using the road each day will help determine carriageway width and the
length of time that the road could be closed due to overtopping during periods of peak flood. The
seasonality of traffic flows and relationships to likely flood periods should also be considered in terms of
the risk to local perishable goods for example.

4.6.3 Frequency of flooding

The frequency and size of peak flows will determine the level of the structure’s roadway to ensure that the
road remains open for all but the largest peak flows.

4.6.4 Emergency / principal route

Principal routes such as access roads to local markets or emergency routes to a nearby hospital will
require higher levels of access and shorter periods of closure due to high water levels.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 32

4.6.5 Availability of alternative route

The proximity and distance of an alternative route will also affect the choice of structure, as an alternative
secure route with a short acceptable detour will allow the road to be closed for longer periods.

4.6.6 Damage to land or property

Whenever watercourses are channelled through pipes, such as in culverts and vented fords or through
narrow openings in bridges, severe erosion can be caused to land and property downstream of the
structure. If agricultural land or buildings are close to the proposed structure careful consideration must
be given to erosion protection. Undersized structures can also cause water to back up causing flooding
upstream and possible property damage.

4.6.7 Uncertainties in flood prediction

The choice and design of the structure will depend on the maximum water flow during flood conditions.
If the maximum water flow is not known sufficiently accurately it may be necessary to provide a structure
that can be over-topped during periods of unpredicted water flow.

4.6.8 Bank elevation and bed material of the watercourse

The resistance of the watercourse banks and bed to erosion will dictate the type of foundation bank
protection and hence structure that can be built. For material which is easily erodible it will be necessary
to have deep foundations and possibly extensive bed and bank protection, or structures which are not
susceptible to damage. The steepness of the banks and difficulty in excavating soil material will also
determine the most convenient approach roads.

A major factor affecting the cost of building a structure is the amount of material which needs to be
imported to, or exported from, the site. Where the road alignment is at a similar level to the river bed
it may be difficult to construct a structure that will not be overtopped without large approach ramps/
embankments as illustrated in Figure E.4.6.

Figure E.4.6: Large embankments required to prevent road flooding

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 33

4.6.9 Complexity of the structure

There is a general progression in complexity, and hence cost, of structures with the cheapest structure
being a drift and the most expensive a bridge (see Figure E.4.7).

Figure E.4.7: Complexity and Cost

It may also be difficult to define the boundaries of different structures (for example, “when does a vented
ford become a multi bore culvert?”). In reality there are overlaps of suitability of each structure type so
that in a particular situation more than one structure type may be suitable.

For small watercourses and relief structures the choice of structure will, in general, be between a culvert
and drift, and for large watercourses between a vented ford and a large bore culvert, or possibly a
bridge. The choice of structure will be determined by all the factors discussed above, but particularly by
the predicted maximum water flow, its seasonal variations and the length of road closures that can be
tolerated.

The flow diagram in Figure E.4.8 shows in more detail the questions and decisions that should be made
when choosing a structure. Factors affecting the choice of structure are different for each location;
therefore a number of questions need to be addressed. It should also be noted that Figure E.4.8 only
highlights the key issues and it should only be used as a guide when determining the most appropriate
structure.

Figure E.4.8 also asks questions regarding the permissible closure time for a road during floods. Each
individual case will have to be assessed separately depending on its particularly circumstances. In the
absence of any local guidelines Table E.4.7, gives suggested upper and lower bounds for closure times.

Table E.4.7: Closure Times

Criteria Drift most favourable Drift least favourable


Average daily traffic (ADT) Less than 5 vehicles per day More than 200 vehicles per day
Average annual flooding Less than twice per year More than 10 times per year
Average duration of traffic
Less than 24 hours More than 3 days
interruption per occurrence
Extra travel time for detour Less than 1 hour More than 2 hours/no detour

When the problem is ‘beyond the scope of this Manual’, specialist bridge engineering skills should be
mobilised.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 4 - 34

START

Is water Are there YES


Is the road
NO NO
flowing short periods of an important link
all year? high water? in the network?
YES NO YES Can the road YES
be closed for the
duration of flood
water levels?
Problem is
NO
beyond the
scope of this
Manual Is another What is
secureroute YES the profile of
available with the channel at
an acceptable the proposed
detour? crossing?
YES

Does WIDE
Are flows
water flow in Provide a
greater than NARROW & FLAT
a well defined DRIFT
channel? NO approximately 30 WITH BANKS
m 3/s
NO
YES NO
How difficult Can peak
is it to excavate flows be handled
bank and bed EASY by simple culvert
Provide a long
material? (s)?
VENTED FORD
DIFFICULT YES

Can
water be
accommodated Provide a
by a single YES CULVERT
small bore*
culvert? Can a vented
NO ford cope with
Are there the standard flow
YES with acceptable YES
short peaks in
water overtopping for
Provide a
flow peak flow?
VENTED
NO NO FORD

Is another
secure route
Provide a
available with an YES
MULTI BORE
acceptable detour?
CULVERT
NO
Compare
Provide
Can a large or the advantages
a LARGE
multi bore culvert YES of different
BORE
cope with the peak culvert types
CULVERT
flows? based on cost
and materials &
skills availability Provide an
NO
ARCH
Can a small BRIDGE
Problem is bridge cope
NO
beyond the with the peak Provide a
scope of this flows without YES
SMALL
Manual overtopping? BRIDGE

*Small bore – diameters less than 900mm

Figure E.4.8: ‘Route Map’ for the selection of a suitable structure

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 35

5. SITE SELECTION AND APPRAISAL

5.1 General requirements


For minor structures such as drifts or culverts on existing routes there may be little choice available in site
selection. Changing the existing road alignment could incur substantial additional road works costs.

For relief drifts or culverts that are necessary to allow the build up of water in side drains to cross the road
alignment, there is usually some flexibility in location. Normally side drains will be relieved by a turn out
or cross structure after a maximum length of about 200 metres to avoid exceeding capacity or causing
erosion in the drain or in the outfall watercourse. Ideal outfall sites are at field boundaries or where there is
vegetation or stable ground to minimize the risk of damage or erosion downstream. If not, provide some
mitigation works (eg drop structures of gabions)

For larger structures and watercourses the selection of site location requires more attention (See Figure
E.5.1). Adjustment of the road alignment is often justifiable to minimize the cost of structures and risk of
damage or erosion.

Careful site selection is essential to ensure ease of construction and to minimise the whole life cost of the
structure. Poor site selection can result in a longer, wider or higher structure than is actually necessary. Poor
siting can also lead to excessively high maintenance costs and, in extreme cases, a high risk of destruction
of the structure. Regardless of the type of structure to be constructed the following criteria should ideally be
met when determining a site for a water crossing (other than at side drain relief, drift and culvert crossings):

ƒ The crossing should be located away from horizontal curves in the watercourse, as these areas are
unstable, with the line of the watercourse tending to move towards the outside of the bend with
time; If no option is available a new channel should be made (See Figure E.5.1)
ƒ The crossing should be at an area of uniform watercourse gradient. If the gradient is steepening
there is a greater possibility of scour and erosion, and if the gradient is reducing there is the potential
for silt and other debris to be deposited near or inside the structure;
ƒ The crossing should ideally be at an area of the channel with a non-erodible bed. These areas have a
reduced scour potential, reducing the amount of watercourse protection required;
ƒ The road should cross the watercourse at a point with well defined banks, where the stream will
generally be narrower;
ƒ The watercourse should not be prone to flooding at the crossing point;
ƒ The scew angle shall be <15o.

Figure E.5.1 Suitable crossing points for larger structures

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 36

5.1.1 Road Alignment

In addition to the watercourse requirements noted in Section E.5.1, the road should:
ƒ Cross the watercourse at 90 degrees as this minimises the span length of the bridge or pipe. A
comparison of length of culvert L1 with a culvert on a skew crossing L2 is shown in Figure E.5.2.

Figure E.5.2: Right angle crossings reduce the length and cost of structure required.

ƒ Cross on a straight length of road, rather than a curve, to reduce the width of a bridge or length of
a culvert. For bridges the minimum straight approach will be 6m.
ƒ Be fixed vertically at the minimum elevation necessary to pass above the design flood flow (this
is obviously not required for drifts and vented fords). If the road alignment is fixed too high,
unnecessary costs will be incurred in abutment/wingwall/ headwall construction and approach
embankments.
ƒ Be centred above the centre line of the substructure.

5.1.2 Location

In locating a structure, the following criteria should be considered:


ƒ A site with a natural narrow channel width rather than a wide one should be used.
ƒ The crossing should be constructed at a straight stretch of river or watercourse, rather than a
curved one where the stream is likely to cause erosion of the bank on the outside of the curve.
ƒ Alignment should be at right angles to the water flow to avoid additional scouring. A skew crossing
may channel the water towards one of the river banks. This channelling may erode the approach
way and/or the bank eventually resulting in the river flowing around the bridge rather than under it.
ƒ The approach roads should preferably be straight on each side to ensure sufficient sight distances
and prevent traffic hazards;
ƒ The site of the river crossing should be away from waterfalls and confluence zones.

It is very rare that all the criteria above can be satisfied for each crossing, therefore a balanced consideration
of the various factors is required. It is necessary to establish the most cost effective solution for each
structure depending on individual circumstances.

5.1.3 Existing Structure Assessment

Where existing roads are being improved, existing drainage sites should already have been provided
with an appropriate structure. However, it is possible that an inadequate structure has been provided or
the need for a structure had been overlooked. A common fault is that culverts have been installed at the
wrong level; too high often results in erosion downstream; and too low leads to repeated silting and a
maintenance problem. When the road is inspected the following conditions indicate that further drainage
work needs to be undertaken:

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 37

ƒ Small gullies exist on the road due to water flowing across the running surface;
ƒ Existing culverts are damaged due to:
• standing water softening the soil around the culvert;
• insufficient capacity.
ƒ Sand and silt has been deposited on the road in patches due to standing water;
ƒ Culverts, inlets or outlets are silted due to incorrect design or installation;
ƒ Evidence of erosion around the structure or culvert;
ƒ Debris trapped at inlet due to incorrect type, sizing or lack of protection.

5.1.4 Site Investigation

The objective of site investigation is to provide a, clear picture of the ground conditions, to enable
a suitable design to be carried out. The level of site investigation clearly depends on the type and
complexity of the proposed structure. When bridges are considered, it is advisable to refer to the Site
Investigation Manual (2011) for foundations investigations. A site investigation will involve taking samples
of the ground material to determine its bearing capacity. These samples can either be obtained through
digging trial pits or by using a hand auger.

5.1.5 Bearing capacity

The ground underneath a proposed structure should have an adequate bearing capacity to support the
load of the structure itself and the vehicles which pass over it. If the soil has insufficient strength it will
compress and the structure will subside, possibly causing failure (see Plate E.5.1).

The bearing capacity will depend on a range of different factors including; the proportions of sand
clay; organic and other material in the soil; the mineralogy of the clay materials; and the level of the
water table. As the type of soil may change with depth it is necessary to dig trial pits at the proposed
site to determine the bearing capacity at the proposed foundation level. By identifying and sampling
the material excavated from different depths of the trial pits the bearing capacity of the soil can be
determined. Bearing capacities are particularly important in the design of structures where large localised
loads are expected, (eg bridge abutments and piers) as the soil must have a high bearing capacity to
support these loads.

Plate E.5.1: Collapse due to settlement

The number of trial pits that should be dug will depend on the complexity of the structure and the
uniformity of the soil. Table E.5.1 gives a guide to the number and depth of trial pits that should be dug
for different structures. If the ground conditions are known to vary over the proposed site, or two trial pits
show markedly different results, then further trial pits should be dug as appropriate. The trial pit depth is
only given as a guideline figure. If the soil conditions are very poor it may be necessary to increase the
depth. Where bedrock exists close to the ground surface this offers the best foundation.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 38

Table E.5.1: Trial Pits: Requirements and Locations

Structure Number Location Depth


Drift Not required
Culvert 1 At outlet 1.5 metres
At each end of the vented section,
Vented 2 (only 1 required if ford is
preferably one on the upstream and 1.5 metres
ford shorter than 15 metres)
one on the downstream side
2.5 metres
Large bore 2 (additional pits at each (deeper in
At each abutment and each pier
culvert pier location if required) poor ground
conditions)
2 (additional pits at each To firm strata
Bridge At each abutment and each pier
pier location if required) (minimum of 3m)

The only accurate method for determining the bearing capacity of any soil is through detailed field and
laboratory investigations. If soils testing facilities are not available, Tables E.5.2, E.5.3 and E.5.4 may be
used for determining an approximate bearing capacity of the in situ rock and soil material. A Dynamic
Cone Penetrometer (DCP) is a low cost, portable device that can also provide an approximation for in
situ soil strength for some materials. However, care must be taken in interpreting results, particularly with
regard to possible variations of in-service moisture conditions.

The engineer should take samples of the soil from the trial pits and compare its properties with the
descriptions in the tables. As different materials have different strength criteria, the three tables are
applicable to rocks, clays and silts, and sands and gravels. The soils used for bearing capacity estimation
should be in the same condition and state as they would be found at the proposed site. As a general
guideline, the more complex and expensive the structure, the more extensive the soil and foundation
investigation should be to minimise initial and whole life costs and the risk of later damage or failure of
the structure.

Table E.5.2: Rock bearing capacity

Rock Allowable bearing Uniaxial compressive


Soil description
strength capacity (kN/m2) strength (mN/m2)
A hammer blow required to break
specimen, can be scratched with firm Strong 10 000 50 - 100+
pressure from knife
Easily broken with hammer, can be
Moderately
easily scratched with knife and pick end 2000 12.5 - 50
strong
indents approx. 5mm
Broken in hand by hitting with hammer, Moderately
1000 5.0 - 12.5
can be grooved 2mm deep with a knife weak
Broken by leaning on sample with a
hammer, can be grooved or gouged Weak 750 1.25 - 5.0
easily with a knife
Can be broken by hand and knife will
Very weak 250 0.6 - 1.25
penetrate approx. 5mm
Note:
1. The uniaxial compressive strength will normally be determined from laboratory tests. It has been included in this Table
for comparison against laboratory soil data where available.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 39

Table E.5.3: Clays and silts bearing capacity

Allowable bearing Undrained shear


Soil description Strength
capacity (kN/m2) strength
A thumb nail will not indent the
Hard 600 300+
soil
Indented by a thumb nail,
penetrated about 15mm with a Very stiff 300 150 - 300
knife
Indented by a thumb with effort,
Stiff 150 75 - 150
cannot be moulded by fingers
Penetrated by thumb with
pressure, moulded with strong Firm 75 40 - 75
finger pressure
25 (should not
Easily penetrated by thumb,
Soft be used as a 20 - 40
moulded by light finger pressure
foundation soil)
Extrudes between fingers when
Very soft 0 < 20
squeezed in hand
Note:
1. The undrained shear strength will normally be determined from laboratory tests. It has been included in this Table for
comparison against laboratory soil data where available. It is important to appreciate that clay soils in particular vary
enormously in strength with moisture content. Dry weather visual assessment is certainly no indication of likely wet
season performance.

Soft or very soft clay/silt soils at the level of proposed foundations will indicate the likely requirement
for special arrangements such as piling. This would require specialist expertise and designs for such
conditions are beyond the scope of this Manual.

Table E.5.4: Sands and gravels bearing capacity

Allowable bearing Standard


Soil sample description Strength capacity penetration test
(kN/ m2) N-Value
High resistance to repeated
Very dense 500 >50
blows with a pick
Requires pick for excavation, a
50mm diameter peg is hard to Dense 300 30 - 50
drive in
Considerable resistance to
Medium dense 100 10 - 30
penetration by sharp end of pick
Can be excavated by spade, a
50mm peg is easily driven, can Loose 50 5 - 10
be crushed between fingers
Crumbles very easily when
Very loose Negligible <5
scraped with a pick
Note:
1. The standard penetration test N-value will normally be determined from in situ tests. It has been included in this Table
for comparison against laboratory soil data where available.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 40

5.2 Specific requirements


In addition to the general site selection criteria given above, the following factors should be taken into
account for the different types of structures.

5.2.1 Drifts

The following site selection criteria should be considered when locating drifts:
ƒ Avoid areas with steep banks (greater than 1.5m) as these require a large amount of excavation
to achieve acceptable approach gradients, and can cause erosion/siltation problems at the banks.
ƒ The level of the drift should be as close as possible to the existing river bed level. This is most
important as it will affect the amount of water turbulence and erosion that may occur around the
drift.
ƒ The normal depth of water should be a maximum of 150mm and the maximum 5 years flow should
be 6m3/second on the drift to allow traffic to pass.
ƒ The watercourse should be clearly defined and stable at the crossing point to ensure that the water
will not alter its flow away from the drift slab.
ƒ In flat arid areas the exact location of the low point in the alignment or occasional watercourse may
not be possible to determine without a detailed level survey (Plate E.5.2).

Plate E.5.2: Flat arid area location of the low point

5.2.2 Culverts

Culverts (or drifts) are usually required at every low point in the road alignment when run-off should be
cross-drained or when it is not possible to be drained forward, backward or with turn of, in addition to
actual water crossing points. Exceptions to this are for an alignment along a hill or mountain ridge and
where drifts or other structures are more suitable for a particular location. The culverts also allow water
from the side drains to cross the road. In areas of sloping ground with little vegetation water will tend to
run down the surface of a hill side and collect in the side drains. In these areas further culverts may be
required to transfer this water across the road. On long continuous gradients without turnout possibilities
it may also be necessary to provide additional intermediate culverts to transfer water across the road to
avoid large quantities of water building up and causing erosion to the drain, road or land downhill of the
road. In such circumstances, these ‘relief’ culverts will normally be expected to be required at intervals of
no more than about 400m.

It is impossible to define the number of culverts required per km as this will vary according to the
topography, catchment and weather characteristics and must be determined by an investigation of the
proposed route. However, as a guide there will typically be 1-3 culverts per km in arid flat or undulating
land and up to 6 culverts per km in more severe terrain, high rainfall areas.

In hilly areas on long steep gradients relief culverts will be required at regular intervals to transfer water
from the uphill side drain to the downhill side of the road to prevent erosion from a large build up of water
in the side drain. Table E.5.5 suggests intervals between relief culverts on long grades. Culverts will also

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 41

be required at points where a stream or waterfall crosses the road. These culverts may also be used as
relief culverts to transfer the water across the road.

Table: E.5.5: Minimum recommended relief culvert spacing

Road gradient (%) Culvert intervals (m)


12 40
10 80
8 120
6 160
4 200
2 >200

The location of a culvert will be determined from the foregoing considerations. The next concern will be
the level at which the culvert should be installed. On rural roads there is often insufficient attention paid
to the alignment and forces related to the water flow, even when this is infrequent. This often causes
problems for the performance and maintenance of the culvert (see Plate E.5.3).

Plate E.5.3: Ponding at culvert outlet

Careful selection of the culvert alignment and size is important to:


ƒ Achieve good hydraulic performance;
ƒ Ensure stability of the stream bed;
ƒ Reduce risks for vehicles;
ƒ Minimise construction and maintenance costs.

It is important to design the culvert to be free from sediment deposits, which tend to occur on the inside
of stream bends, or where there is an abrupt change from the stream slope to a flatter grade in the culvert.
For reasons of economy, culverts should always be laid on a straight alignment that may be perpendicular
or skewed to the road centre line.

In rolling and mountainous terrain culverts usually operate as hydraulically short drainage structures under
conditions of inlet control. The slope of the culvert invert should be 2-5%. Typically, they are sized to
flow 75-90% full, with measures to reduce velocities at the outlet. In flat terrain the culvert slope should
be the same as that of the stream or water course but should never be less than 1% to prevent siltation.

For relief cross culverts, where sediment loads are low to moderate, the combination of a nominally 1m
deep catch-pit inlet, a moderately sloping culvert long-section, and sufficient energy dissipation and

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 42

erosion protection works at the outlet, is recommended. A culvert catch-pit inlet area should be designed
to be easily manually cleared of debris during maintenance operations. The catch-pit should have raised
side walls or wing walls to contain water splash. Where sediment loads are high, a chute inlet, a wide
culvert and greater erosion protection works at the outlet are usually required.

Typical examples of problems that could occur if attention is not given to appropriate horizontal and
vertical road alignments are:
ƒ In flat ground, the invert of the culvert outfall should be determined by the level of the surrounding
ground. Box culverts and arch culverts are preferable in these circumstances as the flat invert slabs
cause less disturbance to the flow of water. Barrel culvert inverts should be similarly determined,
however an outfall apron should be provided to ensure that the flow is stabilised and distributed
horizontally before it reaches the natural ground downstream. If the invert is placed too low then
the culvert outfall and opening will silt up. If the invert is fixed too high there will be ponding (Plate
E.5.3) or silting upstream of the structure and the risk of erosion as the water drops to its natural
vertical alignment downstream of the structure. This also results in the shifting of the stream line
and changes in the stream morphology. It follows that the alignment of the road should be raised
if necessary to provide the correct invert, adequate height for the structure and any necessary
protective cover.
ƒ Where the road is on ground sloping across the alignment, a frequently observed mistake is to
leave the road vertical alignment unchanged and ‘bury’ the culvert so that the outlet discharges
in a long trench with a flat grade. Not only does this ditch often encroach substantially on the
surrounding land, but it is also prone to silting and consequently to causing blockage of the culvert.
Furthermore, vegetation growth and bank erosion are common related problems. In essence a
maintenance problem is created. Localised raising of the road alignment can alleviate this potential
problem (Figure E.5.3). Long culvert (Plate E.5.4) outfall ditches should be avoided and their grade
should not be less than 2% under normal conditions.

Figure E.5.3: Road alignment raised over culvert (Schematic)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 43

Plate E.5.4: Long culvert

ƒ On steep sidelong ground a key consideration is to minimise the erosion risk. In these circumstances
there is usually more opportunity to ‘bury’ the culvert under the road and provide a short outlet
ditch. A ‘drop inlet” or ‘catchpit’ arrangement (See Plate E.5.5 and Figure E.5.4) is normally required
at the inlet to provide a controlled drop in the water flow. Particular attention may still need to be
paid to downstream erosion protection. Special arrangements such as energy dissipating cascades
or gabions may be required in extreme cases. The drop inlet arrangements also need consideration
to be made regarding risk of blockage and maintenance arrangements.

Plate E.5.5 and Figure E.5.4: Culvert drop inlets

In all situations the road alignment standards and structure protection cover requirements should be
complied with. Erosion, silting potential and maintenance implications should be seriously considered in
all cases. Further guidance on culvert setting out is provided in Chapter E.9.

Although it is desirable for culverts to cross roads at 90 degrees to minimise the length and hence the
cost of the culvert, it is not essential and various alignment options are shown in Figure E.5.5. However,
it is important to avoid abrupt changes in stream flow direction at the inlet or outlet of the culvert as this
will result in severe erosion risk for the channel (without suitable control arrangements such as a drop inlet
or erosion protection).

It is not possible to achieve this requirement for relief culverts which transfer water across the road from
the high side channel to the low side channel. These culverts will have an abrupt bend at the inlet and

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 44

require careful protection to ensure that erosion does not occur. The design of these inlets is discussed
in more detail in Chapter E.8.

Figure E.5.5: Culvert alignment options

5.2.3 Vented fords

As vented fords are designed to be overtopped during flood periods it is necessary for the watercourse to
be well defined both for normal flows and flood flows. During flood flows the watercourse will generally
be wider but should still have clearly defined banks to enable the position and size of the structure to be
identified (see Figure E.5.6).

A vented ford provides a constriction to the water flow, due to the solid fill between the pipes. The
proposed location should allow sufficient pipes to be constructed to prevent normal flows overtopping
the structure. In areas where the flow level regularly varies, it is desirable that there are sufficient pipes to
only cause overtopping for larger flood flows. The proposed site should require neither long approach
embankments, as these will increase the cost of the structure, nor steep approaches, which will make the
structure difficult for larger vehicles to cross.

Vented fords can be built on relatively weak ground as their dead weight is spread over the whole area
of the structure. However, the ground should not be susceptible to long term settlement under the dead
weight of the fill material, as this could result in damage to the structure. To minimise the cost of the
vented ford a suitable source for fill material should be available close to the proposed site.

If the volume of traffic using these structures cannot justify two way traffic, the proposed site should allow
drivers to see the opposite end of the crossing and have waiting areas at each end to allow vehicles to
pass each other safely. On road networks where there is a long detour to avoid the vented ford when

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 45

it is impassable, the proposed site should have a waiting area on both sides of the structure sufficiently
large enough for the expected number of waiting vehicles. This waiting area may consist of widening the
carriageway
g y or an area where vehicles can p
pull off the road.

Figure E.5.6: Key design criteria for a vented ford

5.2.4 Large bore culverts

Large bore culverts require the bed of the watercourse to be at least 2 metres below the proposed road
level, to allow sufficient cover over the culvert barrel. Proposed sites for these culverts should have
watercourse banks higher than two metres to prevent the need for long approach embankments which
increase the cost of the structure.

If the crossing site will require more than one arch there should be suitable ground conditions to construct
firm foundations for the piers as well as the abutments. Large arches can exert substantial forces on the
ground at each end, and therefore usually require firm ground on each side of the watercourse. If the
foundation strength is insufficient to support arch springing thrust blocks or pier foundations, it may be
necessary to consider provision of a foundation slab across the entire structure.

Large bore culverts are usually not expected to be overtopped. Consideration of the consequences of
the high flood and its potential to overtop the structure should be made for the proposed site.

As a substantial amount of fill material can be required for a large bore culvert, the total construction
costs can be reduced if suitable fill material is available near the crossing site.

5.2.5 Bridges

The site selection of bridges often involves detailed site investigations which are beyond the scope of
this Manual. For further guidance refer to the ERA Bridge Design Manual - 2011 or publications such as
Overseas Road Note 9. For bridges up to 10 metre spans the guidelines given below should be followed.

The most common cause of failure of bridges is scour of the abutments or piers. In addition to the
factors discussed for all structures above, a site which can avoid the use of piers and has firm ground for
abutment foundations is the overriding criteria in selecting a suitable site for a bridge crossing.

Additional factors which should be taken into account:


ƒ Artificial constriction of the watercourse due to the proposed position of the abutments should be
minimised to reduce the depth of scour.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 46

ƒ The stream velocity should be modest (ie the watercourse should be on a shallow gradient to
reduce the possibility of scour).
ƒ The proposed site should require a minimal amount of work to be carried out underwater. Where
work in the water is unavoidable, a site which reduces the amount of underwater work either by a
simple cofferdam or construction during a dry period, is preferable.
ƒ The bridge superstructure should be above the design flood level. Consideration should also be
given to the possible consequences of a high flood on the bridge superstructure.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 5 - 47

6. WATERCOURSE CHARACTERISTICS

Road structures are constructed not only to remove water from the road and side drains, but also to
transfer water from one side of the road to the other where the road crosses a natural watercourse. These
natural watercourses transfer rainfall from higher ground to lowlands and eventually, usually, into the sea.
The water flowing in a stream or river is called the runoff and will usually be expressed as mm per unit
area or a total volume in cubic metres for a stated period of time. There are many factors which will affect
the runoff, or amount of water in a watercourse, and hence the type of structure which will be required:
ƒ Rainfall
• Annual and seasonal variations
• Extent and duration of the rainfall
• Intensity and distribution
ƒ Geological features
• Type and permeability of the soil
• Natural water storage characteristics of the catchment area
• Size of the catchment
• Intense rainfall only occurs over a small area at any point in time so runoff is not proportional
to size of catchment
• Fan shaped catchment will give higher peak flows as runoff will generally arrive at the
confluence of all streams at the same time. Long and thin catchment areas have the discharge
spread over a longer period and have relatively smaller peak flows for a similar area.
ƒ Topography
• The relief of the ground
• Character of the area: smooth or rugged
ƒ Land use
• Natural drainage of the area
• Vegetation cover

A wide range of hydraulic information may be required for the design and construction of water crossing
structures. The amount of information required and its accuracy will depend on the type and complexity of
the proposed structure. Table E.6.1 indicates hydraulic data and other information about the watercourse
which is required for different structures, along with the potential inaccuracies that may be encountered.

The most important hydraulic factor for structures is the maximum peak flow (or runoff). Culverts and
bridges must be capable of accommodating the peak runoff, after heavy rain, without overtopping,
and vented fords or drifts must be able to pass the peak runoff without erosion or other damage to the
watercourse or roadway. In the case of drifts and vented fords the normal runoff or average flow will also
be important, to ensure that the drift will be passable or the pipes on the vented ford can accommodate
normal flows.

6.1 Maximum peak flow


The maximum peak flow is the most important information to be collected as it is used to determine the
size of the chosen structure. There are a number of methods which can be used to assess the maximum
peak flow. These methods vary in complexity of calculation and accuracy. The option chosen will depend
on the availability of topographical data and the accuracy required for the structure to be constructed.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


Table: E.6.1: Hydraulic and watercourse data required to undertake design
E - Chapter 6 - 48

Hydraulic Data Drift Culvert Vented Ford Large Bore Culvert Bridge
Maximum peak flow Use methods 1,2,4 Use methods 1,2,4 Use methods 1,2 & 4 Use all methods Use all methods
described below, described below, as a cross check for described below and described below and
but see flood return but see flood return method 3 described use worst case result. use worst case result.
period period below Accurate rainfall data is Accurate rainfall data is
required required
Duration of peak flow Required Not required Required Not required Not required
Flow velocity Desirable to know Required Required Required Required
Normal flow rate Required Not required Required May be required May be required
Perennial / seasonal flow Required Not required Required Not required Not required
Amount of debris in Not required Required Required Required Required
watercourse
Type of watercourse (alluvial/ Required Required Required Required Required
incised)
Watercourse bank and bed 25m above and 25m above and below 25m above & 50m 100m above & 200m 200m above and below
characteristics below crossing point crossing point below crossing point below crossing point crossing point
Catchment area & shape May be required to May be required to Required Required Required
calculate peak flow calculate peak flow
rate rate
Cross-sections at crossing Required Required Required Required Required
point
Cross-section 100m above Not required Required Required Required Required
crossing point
Cross-section 400m above Not required Not required Not required Not required Required
crossing point
Cross-section 150m below Not required Not required Not required Required Required
crossing point
Hydraulic gradient up- and 50m above & 25m 100m above and 50m 100m above and 50m 250m above and 100m 250m above and 100m
down-stream below crossing point below crossing point below crossing point below crossing point below crossing point

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


Permeability of soil Desirable for
determining peak flow
Rainfall intensity Desirable for
determining peak flow
E - Chapter 6 - 49

6.1.1 Method 1 - Observation

It may be possible to observe previous high water marks from existing structures, trees or other vegetation
near the watercourse. Small debris floating down the river will be caught on branches and twigs during
floods and indicate the water level during a flood (Plate E.6.1).

The highest flood is the most likely to be visible as it will often ‘rub off’ smaller flood tide marks. The
problem with this method is that there is often no indication of how old the flood level indicators are and
hence what the return periods will be. There may in the past have been higher floods but these marks
have been removed by natural weathering. This method will therefore give an indication of a recent high
flood level but will not be guaranteed to be the highest expected flood level. The information gathered
by observation may be supplemented by interviews with local residents.

Plate E.6.1: Flood debris caught on vegetation

6.1.2 Method 2 - Interviews

If there are people living near the proposed crossing point it will be possible to ask them how high the
water level has risen in previous floods, as these occurrences tend to intimately affect their activities. If
this method is adopted a number of people should be questioned as memories ‘fade’ over time and
floods may ‘get bigger’ each time the story is told. It may be possible to ask people individually how high
the biggest flood had been over the previous years and then take an average of the results obtained.
Validation may be improved if enquiries are made for each bank independently and for different locations
along the banks that provide information that can be correlated. Alternatively a group may be asked to
collectively agree the maximum height of the flood water. It will also be necessary to ask how often floods
of the maximum size occur in order to determine the return period.

Methods 1 and 2 can often form a good cross check between the data obtained for each method.

The interviews shall also acquire changes to the upstream line such as diversion, overtopping, floods from
adjacent streams, land use change and irrigation projects.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 50

6.1.3 Method 3 - Rational Method

This method is accurate for smaller catchments up to 10 km2. The rational method may be used for larger
catchments but the results obtained will tend to be larger than the actual floods encountered.

where
q = flood flow in m3/s
q = 0.278 c i A c = runoff coefficient (Table E.6.2 or E.6.3)
i = rainfall in mm/hr (rainfall intensity)
A = drainage area contributing to runoff in km2

Table: E.6.2: Runoff coefficient: Humid catchment

Soil Permeability
Average Ground very low
Slope low medium high
(rock & hard
(clay loam) (sandy loam) (sand & gravel)
clay)
Flat 0-1% 0.55 0.40 0.20 0.05
Gentle 1-4% 0.75 0.55 0.35 0.20
Rolling 4 - 10 % 0.85 0.65 0.45 0.30
Steep > 10 % 0.95 0.75 0.55 0.40

Table: E.6.3: Runoff coefficient: Semi-arid catchment

Soil Permeability
Average Ground very low
Slope low medium high
(rock & hard
(clay loam) (sandy loam) (sand & gravel)
clay)
Flat 0-1% 0.75 0.40 0.05 0.05
Gentle 1-4% 0.85 0.55 0.20 0.05
Rolling 4 - 10 % 0.95 0.70 0.30 0.05
Steep > 10 % 1.00 0.80 0.50 0.10
Note:
The soil permeability will also be affected by the type of cultivation; these values may be increased by 0.1 for cultivated land
and decreased by 0.1 for forested land.

6.1.4 Method 4 - Estimation

The main problem with the rational method is the requirement to have data available for the predicted
rainfall intensity. In many regions this data may not exist or be incomplete. For catchments up to 15km2
an approximate maximum flood flow can be calculated by assuming a discharge of 1 - 2 m3/s per 25
hectares of catchment area.

6.2 Flood return period


Each flood will have a different size and intensity. Over a period of years the maximum flood experienced
each year will vary. Chapter 3 indicated that the maximum design flood that a structure should be able
to accommodate will depend on the type of structure and be related to a number of years. A 100 year
flood (the largest flood expected in 100 years) will be much greater than a 5 year flood (the largest
flood expected in 5 years). Data may be available for the size of flood, or rainfall intensity, which will be
predicated over a particular period (12.5 years and 100 years are popular record assessment periods).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 51

However, designs for drifts and small culverts will suggest a shorter return period due to the less serious
(physical and financial) consequences of storm damage and hence a smaller storm basis may be justified
for their design. Table E.6.4 provides factors to adjust the size of a given storm period to another period
based on a 12.5 year flood.

Table E.6.4: Flood Period Factors

Flood period 2 3 5 7 10 12.5 15 20 25 50 100


Adjustment factor 0.15 0.3 0.5 0.65 0.9 1.0 1.1 1.3 1.5 2.0 2.5

Example 1
The 12.5 year flood has a rainfall intensity of 35 mm/hour. What will be the rainfall intensity of a 5
year flood ?

From Table E.6.4, the 5 year flood factor is 0.5. Therefore rainfall intensity = 35 x 0.5 = 18 mm/hour.

The Table can also be used to adjust flood flows for other return periods.

Example 2
The 25 year flood results in a flood flow of 12 m3/s. What will the 10 year flood flow be?

From the Table the 25 year factor is 1.5 and the 10 year factor is 0.9.
Therefore the 10 year flood flow = 12 x 0.9 / 1.5 = 7 m3/s.

6.3 Duration of peak flow


The duration of peak flow will not usually affect the design of the structure, but will determine how long
the crossing may be impassable. It is therefore necessary to collect this data if a drift or vented ford is
proposed. The duration of the peak flow depends on the factors which affect the rainfall runoff described
above, and may therefore be difficult to calculate. As this data is only required for simple structures it
will be acceptable to rely on information gathered from the local population and/or the following rule of
thumb.

For catchment areas less than 10km2 the designer can assume that the duration of peak flow will last no
longer than twice the length of rainfall periods.

6.4 Flow velocity


The velocity of the water flow during peak floods is important to determine as it affects the amount
of scour that can be expected around the structure, and hence the protective measures that may be
required. The velocity can be measured in two ways.

6.4.1 Direct observation in flood conditions

An object which floats, such as a stick or piece of fruit, may be thrown into the river upstream of the
potential crossing point. The time it takes to float downstream a known distance (about 100m is a suitable
distance) should be measured. The velocity can then be calculated by dividing the distance the floating
object has travelled by the time taken. This exercise should be repeated at least 3 times, but preferably 5
times, to get an accurate result. Tests where the floating object is caught on weed or other debris in the
water should be discarded. The opportunities for making such observations during flood conditions are
obviously very limited.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 52

6.4.2 Mannings formula

Table E.6.5 shows Mannings formula for calculating flow velocity.

Table E.6.5: Mannings formula

V = velocity in m/s
R = hydraulic depth (the area for the
stream flow divided by the wetted
perimeter)
S = hydraulic gradient ( the slope of the
river bed over a reasonable distance either
side of the crossing point)
n = roughness coefficient (see Table E.6.6) Definition of hydraulic depth

It is difficult to define an exact value of n in tables. It is necessary for the engineer to relate the characteristics
described above in relation to the watercourse being considered to interpret the value of ‘n’ to be used.

Table E.6.6: Roughness coefficient

Ranges of
Stream characteristics
values of n
ƒ Streams in upland areas
‚ Gravels, cobbles and boulders with no vegetation 0.030 - 0.050
‚ Cobbles and large boulders 0.040 - 0.070
ƒ Streams on plains
‚ Clean straight bank with no rifts or pools 0.025 - 0.033
‚ Same as 1 but with some weeds and stones 0.030 - 0.040
‚ Winding watercourse, some pools and shoals but clean banks 0.035 - 0.050
‚ As 3 but straighter river with less clearly defined banks 0.040 - 0.055
‚ As 3 but with some weeds and stones 0.035 - 0.045
‚ As 4 but with stony sections 0.045 - 0.060
‚ River reaches with weeds and deep pools 0.050 - 0.080
‚ Very weedy river reaches 0.080 - 0.150
‚ Stream out of channel flowing across grass 0.030 - 0.050
‚ Stream out of channel flowing through light bush 0.040 - 0.080

6.4.3 Flat terrain

In flat terrain where obtaining such a slope may not be possible and where water flow at the outlet may
be constrained by downstream flow restrictions, considerably more care is needed to ensure sufficient
flow to minimise siltation. Usually it is sufficient to make sure that the slope of the culvert is not less than
1% or, if it is greater, equal to the slope of the water course itself. However some engineering work may
also be required to ensure the downstream flow is not restricted.

In completely flat terrain that is liable to seasonal flooding, the road will usually be on an embankment and
culverts are required to allow cross flow when the flood water ebbs or flows. Under these circumstances
the flow can be relatively slow provided that enough culverts are available, but insufficient culverts can
lead to rapid flow along the side of the embankment and consequent scouring. The best method of
estimating this is by asking the local people how long the water usually takes to dissipate from peak flood
condition after the rain. Calculating the likely volume and required number and size of culverts necessary
to prevent the flow velocity exceeding the velocities shown in Table E.8.20 is then straightforward.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 53

6.5 Normal flow rate


The normal flow rate, like the peak flow rate, is linked to the total runoff and dependent on the rainfall in
the catchment area. As the design of structures is primarily based on the peak flow rate it is necessary to
know the normal flow rate for two reasons:
ƒ For the design of drifts and vented fords it is necessary to ensure that vehicles can cross the drift
during the normal flow or that in the case of a vented ford the water passes through the pipes and
the vented ford is not overtopped.
ƒ To check that there will be no long term damage to the structure due to erosion. The short period
of peak flows may not damage erodible parts of the structure. However it is necessary to ensure
that parts of the structure permanently in contact with the water flow are not damaged.

6.6 Perennial / seasonal flow


An investigation into the variation in seasonal water flows is required if the proposed structure will be
overtopped. It is necessary to determine the proportion of the year that higher flows will be experienced
to estimate the number of days the structure may not be passable. It may be necessary to raise the
running surface of the structure, such as a vented ford, to ensure that the structure is only overtopped
during particularly rainy months. Unless detailed rainfall data is available for the area it is likely that the
only suitable methods for collecting seasonal water levels and flows will be from the knowledge of the
local population.

6.7 Amount of debris in water


Any debris, such as tree branches, carried in the water could cause blockages in the structural openings.
A few small branches can quickly block the 600mm opening of a culvert pipe resulting in water backing up
in the watercourse and potential damage to the culvert or road. Investigations must therefore be carried
out to determine the amount of debris that is typically carried downstream during a flood, to determine
if protective measures are required and/or determine the frequency of maintenance required to the
completed structure to remove trapped vegetation. The funding, resources and likely responsiveness of
the maintenance authority responsible for the route should be investigated to assess the risk of debris
blockage and subsequent damage.

6.8 Type of river (incised or alluvial)


A stretch of river may be described as incised or alluvial. The upper sections of a river are classified
as incised, where the river is eroding the sides and bed of the watercourse. Incised water flows are,
in general, irregular with faster and slower flowing sections. The lower sections of a river are typically
alluvial, with the watercourse meandering across flat plains. Each of the river characteristics provides
challenges for the designer:

6.8.1 Incised

This section of the river is particularly prone to scour, especially around piers and abutments, which
requires careful consideration to protection measures.

6.8.2 Alluvial

The lower reaches of a river normally flow at a steadier rate. There is an equal amount of erosion and
deposition of material in the channel as the stream is already carrying a large amount of sediment.
Although scour will still occur around abutments and piers an additional problem for a designer is that
the watercourse is often unstable; changing its route. It may therefore be necessary to train the river to
ensure that it continues to flow through the structure rather than breaking through the road alignment at
an alternative point.

6.9 Watercourse bank and bed characteristics


Visual inspections of the watercourse bank and bed should be carried out to determine the type of soil
and depth to a firm stratum or rock. The ground conditions will determine the size, depth and shape

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 54

of the structures foundations. Watercourse characteristics will also determine the amount of protection
required to the river bank downstream of the structure. It is not necessary to dig trial pits in the actual river
bed unless piers are required in the watercourse. In this case a temporary cofferdam may be required to
enable investigations if the bed does not dry out in the dry season. Pits in the sides of the watercourse
around the site of the proposed structure foundations provide useful information.

6.10 Catchment area and shape


The size of the catchment will determine the maximum peak runoff that may be experienced after heavy
rain. It may be determined from topographical maps, if they are available, or from a simple field survey of
the area. The shape of the catchment area may also be of interest to the designer as it will affect the size
and duration of peak flows (see Figure E.6.1).

A long thin catchment area results in a lower but sustained A round or square catchment area will tend to have shorter but
peak flow as the rainfall has a range of distances to flow to higher peak flows when compared against a long thin catchment
the proposed structure location. area of the same area. The rainfall in this catchment area has a
similar distance to flow to the proposed crossing site.

Figure E.6.1: Catchment characteristics

6.11 Cross-sections at crossing point


The cross-section of the watercourse at the crossing point will affect the design of the structure (Figure
E.6.2). The watercourse should be surveyed and a section drawn with an exaggerated (5 or 10 times)
vertical scale for the design process. It is also useful to know the cross-section of the watercourse above
and below the crossing point, particularly in the case of incised rivers, as this information will give an
indication of the ‘movement’ of the watercourse and possible additional erosion and training measures
that may be required upstream of the structure.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 6 - 55

Figure E.6.2: Stream crossing point cross-sections

6.12 Hydraulic gradient upstream and downstream


The hydraulic gradient is the slope of the river bed and is normally expressed as a fraction. The hydraulic
gradient will determine how fast the water will flow and hence how much damage it can cause to the
structure and river bed. It will also help to determine how much downstream protection is required for
the watercourse. Some simple surveying is required to determine the slope of the watercourse around
the proposed crossing point. The extent of the survey will depend on the type of structure proposed.

6.13 Permeability of soil


The permeability of the soil in the whole catchment area will affect the peak water flow after heavy rain.
See the section on peak flow rate above. The permeability of the soil in the river banks at the proposed
structure site will also affect the bearing capacity of the soil and hence the design of the structural
foundations.

6.14 Rainfall intensity


The rainfall intensity in the whole catchment area will affect the peak water flow after heavy rain (see
Section E.6.1 on peak flow rate).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 56

7. MATERIALS

This chapter aims to provide sufficient information to enable road designers and builders:
ƒ to identify potentially suitable materials;
ƒ to know what materials’ properties will be required for various uses;
ƒ to know how testing for these properties can simply be conducted;
ƒ to know something about the range of uses appropriate for each material.

The chapter will also discuss potential causes of deterioration and damage and how these might be
avoided by good design.

A particular aim of this chapter is to enable road builders to make more extensive use of local materials
and existing craft skills in the construction of road structures. These are particularly suitable for use by
small local enterprises and community implementation. Where manufactured materials such as steel and
cement are used, initial costs tend to be high; special training in the correct procedures for construction
are required. This Manual focuses particularly on materials such as brick, stone masonry and timber,
which have until now been comparatively neglected both in guidelines for road construction and in the
training of the designers and builders of roads. These materials usually allow lower cost durable solutions
to structure requirements by making better use of locally available materials, skills and labour resources.
They can also have more favourable carbon footprint characteristics and reduce foreign exchange and
national importation requirements.

The remainder of this chapter on materials is organised into four sections:


ƒ Stone masonry;
ƒ Brick and block masonry;
ƒ Timber and organic materials;
ƒ Concrete and reinforced concrete.

It is recommended that consideration be given to all locally available construction materials. For all
materials meeting the specification requirements, a costing of the various options and consideration of
training, maintenance and other factors will enable a rational decision to be made regarding the final
choice of materials. National bridges and structures standards, which have often been ‘imported’ from
developed country conditions or are aimed at structures on main roads, often ignore the possibility of
using some of the materials covered by this Manual. This may deny the benefits of lower costs, use of
local resources, labour, skills and enterprises, and reduce the likelihood of maintenance being carried out
in a timely manner.

7.1 Stone and stone masonry


The density and durability of natural stone make it an ideal material for road structures, where it is available.
Fortunately, in hilly territory where road-building entails frequent retaining walls and river crossings, stone
of building quality can often be found relatively close by, or may even be generated in forming the
roadway. Even in lowland terrain, “field” stone of suitable quality may often be found; and because of
the inherent durability of most types of stone, relatively simple field tests are appropriate to assess its
suitability for building. In addition, suitable stone may be excavated, prepared and incorporated in the
works using hand tools and manual methods, or by mechanised means.

For these reasons, where good quality stone is available, it should be the first choice for retaining walls,
piers and wingwalls for river crossings, the formation of masonry culverts, and for low-level drifts. Stone
can also be used for masonry arches and other simple bridges. Examples of stone and stone masonry
structures are shown in Plates E.7.1 and E.7.2 with examples of established Ethiopian dry stone walling
expertise shown in Plate E.7.3

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 57

Plate E.7.1: Dry jointed, multiple opening, cantilevered laterite stone


culvert approximately 1,000 years old and still in service.

Plate E.7.2: Dry jointed, laterite, stone viaduct, approximately


1,000 years old, still in service on a main road.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 58

Plate E.7.3: Examples of established Ethiopian techniques of dry stone walling

7.1.1 Stone sources and extraction

The well-known classification of types of stone by their geological origin is valuable, because each class
has recognisable characteristics. Table E.7.1 show typical classes and types of stone that could be used
for these structures.

Although the classification of a stone is not essential for its successful use, knowing the origin and type of
stone does help to know what properties to expect. Stone from an existing quarry will probably already
be classified. Unless it is obvious, help in classifying stone from an unknown source should be sought from
professional geologists - samples sent to laboratories can usually be very quickly identified.

Methods of quarrying stone vary greatly from one quarry to another, and are developed to suit the
character of the particular stone being worked and resources available. There are basically two different
approaches. Where the stone is evenly bedded and valuable, a stone-by-stone approach may be used.
The stone is cut straight from the bed to the size required, largely with hand tools, and hand drills, ‘plugs
and feathers’, chisels, crowbars or explosives may be used to assist the cutting. The operation is labour-
intensive, but little waste is produced; in some quarries, the stone is even mined from underground.
Examples of this approach are shown in Plates E.7.4, E.7.5, E.7.6 and E.7.7.

Alternatively, large-scale blasting may be used, bringing down many tonnes at a time, including large
blocks of various shapes and sizes which can be further split down or removed by cranes for cutting.
There will be a large amount of waste which can be crushed for use as concrete aggregate; this may even
be the main product of the quarry. This method uses less labour and more mechanical equipment, and in
certain circumstances may be more economical.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 59

Table E.7.1: Classes of rocks used for building

Granite, Basalt, Pumice or Tuff. These are of volcanic origin and form as a
result of cooling of molten material either within the earth’s crust or at its
surface. Granites and basalts are hard, dense, strong and impermeable,
and can form excellent rubble building stone, but they require a lot of work
Igneous rocks to quarry and form to precise dimensions. Pumice and tuff are relatively soft
and porous materials formed by depositions of ash materials on the surface
or under water. Strength is very variable, but they can often be easily cut
and worked, and may be suitable for building road structures where they are
protected from water.
Sandstones, Limestones. These are formed by deposition (usually under
water) of particles from older rocks and organic materials, and chemical
precipitation. They show natural stratification with separate layers having
Sedimentary rocks different properties, and natural bedding planes. The stratification makes
quarrying and working to precise dimensions easier. Limestones and
sandstones often have colour and texture varieties which make them
attractive as well as durable building stones.
Slates, Quartzite, Marble. These are rocks which can be either igneous
or sedimentary in origin, but which have been subsequently altered due
to movements in the earth’s crust causing them to experience enormous
heat and pressure. As a result they are often hard and durable, and they
tend to have a foliated structure with layers of stratification. Slates are
Metamorphic rocks
metamorphasised clay and shale which quarry easily and are frequently
suitable for walling and roofing stones. Marbles are metamorphasised
limestones which have been crystallised by heat and pressure. They are hard
and durable, and suitable for sawing and carving and can often take a high
polish.
This is strictly a form of soil rather than a rock. It is the end product of the
intense tropical weathering of primary rocks, and it consists largely of the
oxides of iron and aluminium, but it has the useful property of hardening on
exposure to air. When soft it can easily be cut with a hoe, but on hardening
Laterite
it becomes weather resistant, and has a durability comparable with some
building stones. It is widely used as a building material in humid tropical
areas. However the quality can be very variable and care needs to be taken
in selecting suitable material.
Stone which is found away from quarries or other formal deposits, usually
transported by water or landslides, and may be of any of the geological
Field stone types described. It is usually of a durable type. Field stone can be a useful
source of stone for small road projects, but should be subjected to tests as
described below to determine its suitability.

In some locations suitable stone may be lying on the ground surface and may be collected by local
labour. This can even benefit the local land users by clearing fields to improve crop yields.

Cutting and finishing methods also range from very labour-intensive techniques using only hand tools to
highly mechanised operations.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 60

Plate E.7.4: Hand dressed stone Plate E.7.5: Hand quarrying of stone

Plate E.7.6: Hand quarried Plate E.7.7: Manual quarry operations


granite stone blocks can create local employment, however
good management is required to
ensure a safe working environment

7.1.2 Properties of stone

Size
The most important prerequisite of a good building stone is that the stone is available in pieces of a size
and shape suitable for the type of wall or structure to be built. Stones should also be small enough to
be lifted and placed by hand. For use in rubble walling, a range of sizes is needed. The individual stone
height may be up to 300 mm, the length should not exceed three times the height and the breadth on
base should not be less than 150 mm, or more than three-quarters of the wall thickness. A range of sizes
should be used, with larger stones being used for corners (quoins) and for through (bonding) stones.

Durability
Durability is the resistance of the stone to weathering or deterioration from other causes. The structure
of the stone is the most important aspect of its resistance to decay. Stone used for building should be
uniform in colour and texture, without soft seams or veins or other visible blemishes. The surface of a
freshly broken stone should be bright, clean and sharp without loose grains and be free from an earthy
appearance. Visual tests are sufficient to assess its durability characteristics. Other durability issues are
shown in Table E.7.2.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 61

Table E.7.2: Durability Issues

Some types of stone are seriously affected by frost, and in cold climates
must not be used in positions where they can become saturated. The
Frost Action remedy is to detail the wall to protect the stone from becoming saturated,
by means of a coping, and providing protection for the base from upward
percolating water by means of a damp-proof course.
Soluble salts can disfigure and ultimately cause deterioration of some
sedimentary stones. Soluble salts may occur in the sands used for mortar, in
Soluble salts the water behind retaining walls, or in road salts. The remedy is not to use a
stone which is liable to react poorly to soluble salts in circumstances where
it will be exposed to them.
Some small variations in the dimensions of stones always occur as a result
Thermal and moisture of changes in temperature and moisture. These are rarely sufficient to cause
movement any cracking problems, but it is a good precaution to insert movement
joints in mortared masonry walls at intervals of approximately 15 metres.

Compressive strength
There are significant problems of strength testing of stone in rural areas. The compressive strength of
dense stone is generally greatly in excess of that required in any small road structures. A few porous
stones, like pumice or tuff, or soft stones like laterite, may require some testing to establish that they
have a suitable compressive strength. In other cases, the compressive strength can be assumed to be
adequate for the small road structures described in this Manual based on evidence of established local
use. However, for stones subject to abrasive conditions or just use in arches it is advisable to confirm the
compression strength is a minimum of 15MPa unless otherwise specified.

Seasoning
Certain stones, notably soft limestones, sandstones and laterites, increase significantly in strength and
durability after quarrying. The appropriate time for seasoning depends on the quarry, and local knowledge
is needed to decide on the correct seasoning time. It is a potential problem mostly where there is frost.

Porosity
Porosity is not in itself a disadvantage in most cases, but some stones are capable of absorbing substantial
amounts of water; this can reduce the strength and also, in cold climates, freezing can cause disintegration
of the stone. A good building stone should not absorb more than 5% of its weight in water.

7.1.3 Field Testing

In many cases the best test of the suitability of a stone from a local quarry or other source is its previously
successful use in structures in the area which have been subjected to the local climate for a long period
of time. Enquiries to local builders and contractors may result in knowledge gained regarding the best
sources of building stone, and any local characteristics. This information can be supplemented by
additional tests as required.

Structure test
The structure of a stone from sedimentary rock sources can be tested by immersing small pieces in clear
water in a glass jar for about an hour and then shaking them vigorously. If the water discolours, the stone
is not well cemented and should not be used.

Water absorption
The water absorption of a stone is a measure of its porosity and of its liability to frost damage. The water
absorption of a stone can be assessed by:
ƒ Weighing it when dry (stored in a dry environment for at least 5 days);
ƒ Immersing it in water for 24 hours at ambient temperature;
ƒ Weighing it again after removing excess surface moisture.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 62

The difference in weight should not exceed 5% of the initial weight.

Soundness test
The soundness (freedom from cracks or weaknesses) of a stone can be tested by means of the hammer
test (see Figure E.7.1).

Figure E.7.1: Hammer test for stone

Acid test for weathering potential


A small sample is immersed into a 1% solution of hydrochloric acid for seven days, during which time it is
frequently agitated. If the sample has retained the sharpness of its edges and corners, it will weather well.

Compressive strength
There is no adequate field test for compressive strength. This is not normally an important consideration
except with blocks made from rather weak stones such as tuff. Where needed, testing should be entrusted
to a competent laboratory.

Hardness
The surface hardness can be tested by scratching with a penknife. All types of stone will be marked by a
knife blade under firm pressure; but stone in which a penknife blade can make a groove exceeding 2 mm
is likely to be moderately weak in compression, and compression testing may be needed.

7.1.4 Mortars

Unless dry stone walling skills are available or can be introduced, stone masonry usually involves mortar
jointing. The principal function of mortar in masonry is to provide an even bed to distribute the load over
the whole bearing area of the units, and to bond the masonry units together.

Good mortars should:


ƒ Be cohesive, spread easily and retain water so that they remain plastic while the masonry units are
positioned and adjusted;
ƒ Set and develop strength rapidly after the units are in place;
ƒ Have a final strength adequate to carry the load without cracking the masonry;
ƒ Be impermeable to moisture movement, and resistant to weathering.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 63

Mortars are composed of clean sand and a binding agent (usually Portland cement) and often some
additive (either lime or plasticizer) to improve plasticity and workability. Sand should be soft building sand
free of organic particles and clay. Lime should be bagged dry hydrated lime or lime putty. A plasticizer is
an admixture to the mortar used in small quantities to improve the workability of the mix or to achieve the
same workability with less water, thus improving both strength and durability. Plasticizers are proprietary
materials and should be used according to manufacturers’ instructions.

It is important that the strength should not be greater than that of the units being joined so that movement
cracking will be dispersed through the mortar joints and not lead to a few wide cracks which could affect
strength and weather resistance.

Table E.7.3 below shows typical mortar mixes using cement-sand or cement-lime-sand.

Table E.7.3: Mortar Proportions by Volume

Type of mortar
Cement:lime:sand Cement:sand
Higher strength for structural
1 : 0.5 : 4 1:4
use or contact with water
Lower strength for general use 1:1:6 1:6

Commonly used mixes are 1:4 cement:sand for structural use, or where there is contact with water, and
1:6 in other cases. For a good quality mortar, water content should be low (typically 0.4 water/cement
ratio). The quantity mixed in any one batch should not be more than can be used in about one hour;
during that time unused mix should be covered to protect it from excessive evaporation.

7.1.5 Stone walls

Random stone masonry is constructed from stones


as they came from the quarry or source with minimal
dressing. The laying skill is selecting individual stones
so that they create a reasonable joint with the adjacent
stone without the excessive use of jointing mortar.
Stone should be bonded both longitudinally (along the
wall) and transversely (across the thickness of the wall).
Longitudinal bond is achieved by placing each joint
more than one-quarter of a stone’s length away from the
joint below. Transverse bond is obtained by the use of
bonders (at least one per m2 of wall), extending about
two thirds to three quarters across the width of the wall
or right through the wall if water penetration is not a
problem.

Random stone walls may be constructed without any


courses, or brought to level courses for example every
600 or 900mm.

Plate E.7.8: Rubble stone


large bore culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 64

Figure E.7.2: Rubble masonry wall

Dressed stone masonry is built with stones which are dressed to approximately rectangular shape, usually
before leaving the quarry. It is built in courses which may vary in height from 100mm to 300mm, often
with thicker courses lower in the wall. All stones in any course are squared to roughly the same height.
Bond stones are laid in each course at about 1.5m spacing. Examples of rubble masonry walls are shown
in Plate E.7.8 and in Figures E.7.2 and E.7.3.

Figure E.7.3: Squared masonry walling

Some general requirements for stone walls are:


ƒ The minimum thickness of a stone masonry wall should be 400mm;
ƒ The height of a free-standing wall should not be more than six times its width at the base, and may
be tapered over its height;
ƒ Mortar joints should be between 10 and 40mm thick, and have a minimum overlap of one quarter
of the length of the smaller stone;
ƒ Mortar joints should be pointed on the face of the masonry;
ƒ No stone should touch another, but should be laid into mortar;
ƒ Mortar should be made of cement and sand using volumetric proportions shown in Table E.7.3
above. A common mortar is 1:4 cement:sand.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 65

Some additional requirements for retaining walls are:


ƒ The thickness of a gravity retaining wall at any point should be at least one third of the retained
height above that point;
ƒ Retaining walls should be provided with regular weep holes just above ground level on the outer
face. Weep holes should be of 75mm diameter and spaced at 1.5m centres. A filter of loose stone
or lean concrete should be placed at the back of the weep holes to permit free drainage of water,
but not allow material to be washed through.

7.1.6 Dry stone walling

Dry stone walling is a form of stone walling built without mortar. A dry stone wall costs substantially less
than a mortared wall, but calls for considerable skill in choosing and laying stones if instability and rapid
deterioration is to be avoided. The face stones are usually roughly dressed and laid on a firm natural soil
bed; the core is formed from earth or smaller stones. Typical arrangements are shown in Figure E.7.4.
Undressed stone may be suitable for retaining walls (Plate E.7.9) where stone is available locally and the
ground beneath is firm. Walls should not exceed 5m in height.

Figure E.7.4: Dry stone retaining wall

Plate E.7.9: Dry stone wall supporting embankment

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 66

Some guidelines for the construction of dry masonry walls are as follows:
ƒ All front-side inclinations should be in the range of 3:1 to 4:1 (vertical:horizontal);
ƒ All back-side inclinations should be 1:10;
ƒ The top width of the wall should be 1m;
ƒ The tilt of the bottom of the foundation should be at right angle to the front inclination of the wall
and hence fixed in the range of 1:3 to 1:4;
ƒ The stones used should be generally of a large size and care must be taken in dressing and placing
them;
ƒ The joints between the stones must be arranged in a staggered fashion as in brick masonry (see
Figure E.7.4 and Plate E.7.10);
ƒ All other construction details can be considered analogous to the rules applicable to gabion wall
construction.

Plate E.7.10: Dressed dry stone wall

7.1.7 Hybrid walls

Hybrid walls are walls made of bands of mortared stone masonry reinforcing areas of dry stone masonry.
The construction of a hybrid wall is recommended when the height of a dry stone wall is greater than 5m.
The construction technique is the same as that of a dry wall, except for providing 0.6m bands of mortared
cement masonry at intervals of 2m–3m both horizontally and vertically. This type of wall is recommended
for heights between 5m and 12m both in valleys and on hillsides. Examples are shown in Figure E.7.5
and Plate E.7.11

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 67

Figure E.7.5: Hybrid wall

Plate E.7.11: Hybrid retaining wall

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 68

7.1.8 Masonry culverts

Masonry arch culverts may be more economical than pipe culverts where
stone is locally available. Some general requirements are:
ƒ Culverts are usually up to 2 metres in span;
ƒ Strip foundations of concrete or stone masonry should be laid on
firm ground: the foundation walls are brought up to the level of the
arch springing;
ƒ Arch formwork may be made from corrugated steel roof sheets (Plate
E.7.12), timber or reusable steel formwork (Plates E.7.13,14 and 15).
Simple compacted earth fill can also be used and excavated after
the masonry has been constructed;
ƒ The arch should have a minimum thickness of 400-500mm with all
stones having the same dimensions as the arch thickness;
ƒ The ground seepage cut-off, invert (base) slab, headwalls, wingwalls,
drop inlets and aprons may also be constructed of masonry as Plate E.7.12:
required. Masonry culvert with
steel former to be left
The masonry is normally mortar jointed using skills commonly available insitu after construction
with local building contractors. Dressed stone skills are required for dry stone masonry culverts.

Plates E.7.13-15: Dry stone masonry culvert


extension works - Dry stone Conservancy, USA

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 69

7.1.9 Gabion works

Gabions are wire mesh boxes filled with stones and tied together
to form basic structures. Their principal uses are for retaining walls,
drifts and erosion protection (Plate E.7.16). Gabion boxes may be
made from purpose made gabion cages, welded steel mesh sheets
or galvanised chain link fencing.

Gabions are used as a substitute for concrete or masonry, and


gabion structures should be built with the same principles of good
foundation, stability and quality control. The advantages of gabions
are their simplicity of construction (requiring low levels of skill),
use of local materials (stones), ability to let moisture pass through,
avoiding the build up of water pressure, and flexibility (should
minor settlement occur). Flat gabions are also referred to as gabion
mattresses. Plate E.7.16:
Gabions supporting a road
The process of gabion construction (illustrated in Figure E.7.6) is as on a steep slope
follows:
ƒ Foundations should be excavated
level and cleaned as for a conventional
structure, with any unsuitable material
removed and replaced with good soil,
stone or gravel, and compacted;
ƒ The baskets should be erected in their
final position;
ƒ Cages should be woven together using
3mm binding wire securing all edges
every 150mm with a double loop. The
binding wire should be drawn tight with a
pair of heavy duty pliers and secured with
multiple twists;
ƒ The connected baskets should be stretched
and staked with wires and pegs to achieve
the required shape;
ƒ Filling should only be carried out by hand
using hard durable stones not larger than
250mm and not smaller than the size of the
mesh. The best size range is 125 - 200mm.
The stones should be tightly packed from
the edges inwards with a minimum of
voids;
ƒ Boxes of 1 metre height should be filled to
one-third height. Horizontal bracing wires
should then be fitted and tensioned with
a windlass to keep the vertical faces even
and free of bulges;
ƒ Further bracing should be fixed after filling
to two-thirds height. 500mm height boxes
should be braced at mid height only. 250
Figure E.7.6: Gabion details
- 330mm gabions do not require internal
bracing;
ƒ Where water falls directly onto the top of the gabion, vertical bracing wire should also be fitted to
secure the gabion lid when closed;
ƒ The stones should be carefully packed to about 3 - 5cms above the top of the box walls to allow for
settlement. Smaller material can be used to fill the voids on the top face, but excessive use of small
stones should be avoided. Fibre matting can be placed over the stones on the top of the gabion
to promoted vegetation growth;

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 70

ƒ The lids are then closed and stretched tightly over the stones, (carefully) using crowbars if necessary.
The corners should be temporarily secured to ensure that the mesh covers the whole area of the
box, the lid should then be securely woven to the tops of the walls, removing stones if necessary
to prevent the lid from being overstretched.

Further illustrations of the construction and use of gabions are shown in Plates E.7.17 to E.7.21.

Plate E.7.17: Filling gabion baskets Plate E.7.18: Gabion retaining wall
used as river bank protection

Plate E.7.19: Natural fibre matting inserted Plate E.7.20: Steel mesh gabions
in gabion protection works to encourage
rapid growth of vegetation cover

Plate E.7.21: Gabion river crossing

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 71

7.2 Brick and block masonry


Brick and block masonry materials are made in almost every Part of the world, and are frequently the
standard materials used for walling in building construction.

Fired clay bricks have been used for centuries for bridges (see Plates E.7.22 and E.7.23) and are one the
most flexible general building materials; their small and regular size make them suitable for incorporation
in any shape of structure. They are ideal for utilising local building labour and contractor skills. Good
quality fired brick is suitable for most types of road structure. The following requirements should be met
for its application:
ƒ Material selection and firing methods should ensure a consistent quality of brick;
ƒ Bricks should be laid with a suitable durable mortar (as stone masonry), that is however not as
strong as the bricks themselves;
ƒ Bricks should be laid bonded, to similar principles as stone masonry;
ƒ Bricks should be protected from the possible incidence of soaking in frost susceptible climates;
ƒ Abrasion situations such as drift surfaces should be avoided.

This makes fired clay bricks suitable for use in foundations, bridge piers, abutments, wing walls, arches,
culverts and retaining walls. Bricks can be a particularly important construction material in regions and
locations with shortages of hard stone resources.

Local brick making skills are established in many areas, and could be developed in new ones as a small
scale rural industry where there could be complementary demand in the road, building and infrastructure
sectors. In some locations only low grade bricks are produced to meet a relatively undemanding
requirement for general building bricks. In these cases some improvements in production such as kiln/
clamp design, firing temperature and period may be required to achieve bricks of suitable quality for the
more demanding structures applications. Bricks produced in ‘one burn clamps’ can have variable quality;
the bricks near the outside usually being less well burnt. Permanent kilns and industrial production usually
ensure more consistent quality products.

Fired clay bricks may be produced using agricultural wastes (such as rice husk) as the kiln fuel (Plate
E.7.24) as described in the gTKP report on practices established in Vietnam (Reference: Dzung).

Plate E.7.22: Fired clay brick is a very versatile material. This elliptical brick arch bridge over
the River Thames at Maidenhead, UK was constructed in 1838 and has two main spans of 38
metres. It is still in use today (designer: Isambard Kingdom Brunel) photograph by Jon Combe

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 72

Plate E.7.23: Fired clay brick structures exist in many countries and still
provide service for today’s traffic after centuries of use

Plate E.7.24: High quality burnt clay bricks produced in small ‘beehive’ kilns fired by
rice husk. Bricks can be fired using coal, wood or agricultural waste products

The range of masonry materials which are manufactured is wide. Burnt clay bricks (Plate E.7.25) are still
in many places the most widely available manufactured masonry materials. However, concrete blocks are
increasingly widely available and can often be cheaper for the same applications as burnt clay bricks.
Other masonry materials such as stabilised soil blocks and trass (a naturally occurring binder) - lime
blocks are found in different places, and may be suitable for less demanding applications such as culvert
headwalls and low retaining walls. Where natural stone is locally available, this may be better to use than
any manufactured material on environmental and transport costs grounds. The requirements, properties
and testing recommendations described for bricks and blocks need to be adapted for other masonry
materials.

Plate E.7.25: Simple burnt clay brick clamp

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 73

A wide range of soils is suitable for brick and block making. To make bricks, a suitable soil (called clay or
brick-earth) is mixed with water, formed into the desired shape in a mould, dried, and then set in a kiln
and fired at a sufficient temperature (usually 850-1000°C) to create permanent ceramic bonds between
the soil particles (Plate E.7.26).

Bricks are classified in various ways according to their intended use. A common classification recognises
three classes according to their durability: internal quality bricks or blocks (suitable only for protected
situations inside buildings); ordinary quality (suitable for external use in normal conditions of exposure
(walls protected by damp-proof courses and a coping); and special quality (suitable for unprotected
external uses such as parapets and earth retaining structures). Bricks and blocks may also be classified
according to strength characteristics or shape (see Figure E.7.7).

Concrete blocks are made from aggregates and cement, and mainly manufactured in large fixed or mobile
plants using heavy compaction or vibration, and sometimes steam curing. They can also be made on site
using individual moulds; a labour intensive
process which can result in quality variability
without adequate control processes. Solid
blocks have no holes, cellular blocks have
cavities which do not pass right through the
block, and hollow blocks have cavities
passing right through. Manufactured blocks
are made to satisfy standards requiring a
minimum crushing strength.

Plate E.7.26: Bricks baked in a kiln Figure E.7.7: Brick types

7.2.1 Properties of bricks and blocks

For all brick and block materials the principal requirements are:
ƒ Acceptable and handle-able size and small variation in dimensions;
ƒ Dimensional stability over time;
ƒ Strength;
ƒ Durability.

Chemical composition and limited water absorption are also important for clay bricks.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 74

Size
The standard size of clay bricks differs to some extent from location to location. In Ethiopia the standard
brick format is 250 x 120 x 60mm, although the average size of the actual brick is typically 10mm less
than this to allow for the mortar joint. For individual bricks some variation is acceptable and must be
allowed because of the differences of firing and moulding of individual bricks, but the average over a
large number of bricks should stay within about 3 to 4% of the standard size. For walls whose appearance
is important, the distortion of individual bricks should be limited, but normal distortions, even of hand-
made bricks, can usually be absorbed in the mortar joints.

The standard size for concrete blocks in Ethiopia is 400 x 200 x (100, 150 or 200)mm.

Dimensional stability
Burnt clay bricks change in dimension to a small extent over time as a result of moisture movements,
and temperature. There is an initial expansion of about 1mm per metre length, most of which occurs
within the first week after the bricks leave the kiln. Subsequent moisture movements are small, and
thermal expansion (about 0.15 to 0.25mm per metre for a 30°C temperature rise) is small compared with
other building materials. Expansion joints are normally allowed every 12m in facing brickwork in order to
accommodate these movements without causing cracking.

Blocks shrink after manufacture by about 0.5 to 1mm per m length of wall, which can be sufficient to
cause cracking if expansion joints are not used; expansion joints are normally required to be spaced at
8m centres in blockwork to allow for the initial drying shrinkage, and subsequent moisture and thermal
movements.

Strength
The compressive strength of individual bricks and blocks is usually much higher than is needed, but
the strength of panels is affected by their shape, how they are supported, and by the mortar used. The
compressive strength requirement depends on the loading on the wall. A minimum unit compressive
strength of 3.5N/mm2 may be adequate for walls which are not carrying large loads, and this is easily
achievable in masonry materials made by simple processes; but masonry units of strengths up to
50N/mm2 or even more can be manufactured for use in special conditions. The stronger masonry units
also tend to be less permeable, more resistant to frost and water erosion and thus more durable.

Water absorption
Water absorption is a concern for burnt clay bricks. It is a measure of the porosity of a brick, and should be
limited, especially if the bricks are to be used in exposed positions - parapets, piers and abutment walls.
A water absorption not greater than 15% by weight of the dry brick weight is required for acceptable
performance.

Chemical composition
For clay bricks, limitations on the content of certain salts are sometimes specified to reduce the problems
of efflorescence and sulphate attack. Limiting sulphate content to 0.5% can eliminate the problem of
sulphate attack (see below); alternatively sulphate resisting cement may be needed. Efflorescence is
unsightly but does not seriously affect the strength or durability of the masonry. Elimination of nodules
of lime (kankar) in the brick earth is essential; their expansion after the bricks have fired can damage the
brickwork.

Durability
The durability of a brick or block masonry wall depends as much on the climatic conditions, the extent
to which protection of the faces and edges is ensured by copings and damp-proof courses, and the
quality of the mortar as it does on the masonry units themselves. A key consideration is whether the
site may be exposed to frosts. No specific requirements for durability can be stated, but units satisfying
the requirements for strength and water absorption can usually be expected to perform satisfactorily if
properly protected, by design, from extreme exposure. Specific actions to limit susceptibility to frost and
chemical action for clay bricks are discussed below.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 75

Frost resistance
Water expands on freezing; where walls can become saturated with water which subsequently freezes,
the expansion can be very damaging to masonry work. Well-made bricks or blocks of low porosity are
generally able to withstand regular freeze-thaw cycles: but more porous and lower strength masonry units
may be damaged. Where frosts may occur, bricks or blocks which are not of the highest quality should
be protected from conditions in which they may become saturated. The use of damp-proof courses to
protect the base of the wall from rising damp, copings to protect the top, and weathering details to
prevent water from running down the face of the wall form the best protection from frost damage. All
mortar joints should also be well pointed to ensure that the finished face of the joints is dense.

Resistance to chemical action


Soluble salts in bricks, which may derive from the original clay used or from the kiln reactions, can cause
staining and efflorescence or deterioration of the mortar. Efflorescence is the crystallisation of soluble
salts at the surface of the brickwork, when bricks dry after a prolonged period of wetting. It is usually not
damaging and can be tolerated. If the bricks contain soluble sulphates, these may cause an expansive
reaction with Portland cement in the brickwork mortars, which will damage the integrity of the wall.
Sulphate attack may also occur as a result of sulphates in groundwater in contact with earth retaining
walls.

Abrasion and impact


Bridge structures, piers and abutments may be subject to abrasion (due to driving rain, wind-borne
sand or dust, or flood water). The possibility of vehicle impact from road or water should be considered.
Well-made masonry units will have adequate resistance to these actions, but they should be considered
in deciding the quality of bricks or blocks and mortar to use; and impact loads should be considered in
the design of the wall resistance. In some instances the design of the structure needs to be detailed to
minimise the risk of, or physically protect vulnerable components from, the impact from road or water
born traffic or debris.

Thermal and moisture movement


Thermal and moisture movements can cause expansion and contraction in brickwork, which can result
in cracking unless it is allowed for. Mortars should normally be designed to be weaker than the bricks or
blocks laid up in them, to enable high stress concentrations to be relieved - the recommended mortars
for various classes of bricks allow for this. Expansion joints should be provided through brickwork (and
any supported structure) every 12m; they should be 10mm wide, and filled with compressible material so
that they do not become inactive.

7.2.2 Field testing

Where possible, obtain information from the brick or block manufacturer regarding the Standard to which
the units conform, and details of the results of recent tests on strength, dimensional stability, water
absorption and chemical composition as appropriate. Failing such information, field testing will help
ensure that the bricks are of generally sound quality.

Quality of the raw materials used


The brick earth used for making burnt clay bricks should not contain iron pyrites, pebbles, or nodules of
lime or tree roots. The content of clay should be around 15% to 30%. A small quantity of lime is acceptable
as long as it is in a finely divided state. Soils from areas which are or have been saturated in salt or sea-
water should be avoided. Similarly aggregates for concrete blocks should satisfy the requirements for
good concrete, as set out later in this chapter.

General characteristics of clay bricks


A good clay brick should be sound, hard and well-burnt with uniform size, shape and colour, homogeneous
in texture, and free from flaws and cracks. A broken surface should show a uniform structure free from
holes or embedded lumps. Corners should be square, straight and well-defined. When struck against
another brick or with a small hammer, bricks should give a metallic ring, not a dull thud. When soaked
in water for 24 hours there should be no sign of softening or distortion. Before or after immersion, the
surface should not be able to be scratched with the finger nail.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 76

Strength: drop test


If no specific compressive strength requirement is required for load-bearing purposes, the strength of
a masonry unit may be crudely assessed by dropping it flat from a height of about 1.2 to 1.5m to the
ground, or striking one brick against another. In neither case should it break.

Strength: impact test


For a more controlled impact strength assessment, place a brick with its largest face downwards, resting
on timber battens 180mm apart. Drop a mason’s 2kg hammer from a height of 0.5m so that it strikes the
upper face midway between the battens. The brick should not break. If more than 2 in a sample of 20
bricks break in this test, the bricks should not be used. For bricks and blocks of different shapes a similar
test may be used, but the height of drop and span will vary.

Dimensions
No dimension should differ by more than 5mm from the standard size. The overall dimensions of a set
of 20 or more bricks or blocks, randomly selected, should not deviate by more than +/-4% (length and
width) or +/-3% (height) from the standard size.

Water absorption test for clay bricks


Take a sample of 5 bricks at random. Dry and weigh each brick (W1); then submerge in clean water at
ambient temperature for 24 hours. Wipe surface and weigh again (W2).

The water absorption (%) is the ratio:


100 x (W2-W1) / W1

For bricks to be used in normal conditions of exposure, water absorption should not exceed 15%. More
severe limitations will be required for bricks to be used in conditions exposed to permanent wetting and
drying (eg at the base of piers or abutments).

Durability
One way to test durability is through the construction of a test panel to be exposed to conditions similar
to the proposed work; but a period of some months’ exposure in severe weather will normally be needed
to assess performance adequately. Alternatively observing the performance of the same masonry units in
other building situations of comparable exposure (reference sites) can be a good indication of durability.
Some kind of exposure test should be used if the units are to be used in conditions where they will be
exposed to heavy frosts.

7.2.3 Mortars for brick and block masonry

The requirements for mortars are the same as those for stone masonry.

Brick and block masonry may be used for bridge piers (Plate E.7.27) abutments and wing walls, arch
culverts and wing walls, and small bridges.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 77

Plate E.7.27: Bricks in a bridge pier

7.3 Timber and organic materials


Timber, in the form of sawn sections or poles, is in many areas a highly cost-effective material to use for
load-bearing structures, even where there are concerns about the over-exploitation of tropical forests.
The quantities of timber required are relatively small and a good management regime will ensure or
arrange for planting of replacement trees for construction and maintenance needs. This can be stipulated
in contract documents. With proper selection of species; stress-grading to ensure efficient utilisation;
and attention to seasoning, preservation and subsequent maintenance; structures made from either
softwoods or hardwoods can have a design life comparable to that of steel or concrete structures. In
addition the appearance of timber structures fits the natural surroundings and its use can provide local
employment without the need for highly sophisticated technology in manufacture or preparation with
reduced transport costs. Sustainably produced timber can have attractive carbon footprint attributes.

The principal use of timber in low-cost road structures is for bridge decks, where its structural advantages
can be utilised most fully, and where it is more easily protected from moisture penetration. Use of timber
for running surfaces may make sense even when the supporting structure is of steel (Plate E.7.28), masonry
or concrete. Trussed or girder bridge decks can be made from cut sections of timber or from timber poles
(Plates E.7.29 and E.7.30). Timber has also been used for both bridge, piers and abutments (Plate E.7.31)
and for retaining structures, though in these uses a relatively shorter lifetime must be expected; and in
Tanzania, a successful programme to use timber for culvert linings has been in progress for some years.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 78

Plate E.7.28: Steel beams being moved into position using labour methods,
without the use of cranes; ready for completion with a timber deck

Plate E.7.29: Timber deck on recycled steel beams

Plate E.7.30: Completed timber deck

Plate E.7.31: Timber bridge deck, pier and abutments

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 79

7.3.1 Characteristics and utilisation of timber

Hardwoods and softwoods and their availability


Softwoods are derived from coniferous evergreen trees grown mainly in temperate forests. They
are relatively rapid-growing and the wood is of a generally relatively low density (typically less than
500 kg/m3) and moderate strength and easy to work. However, the wood is not normally very durable,
unless protected by preservatives. Globally, coniferous forests are very extensive, and are managed to
produce a sustained yield of timber. In areas of temperate climate, softwoods are therefore relatively cheap,
making timber structures highly cost-efficient.

Hardwoods are derived from broad-leaved (deciduous) trees which lose their leaves in winter; they are
found in both temperate and humid tropical climates. Compared with softwoods, they are relatively slow-
growing, and this results in wood which is denser (typically > 650 kg/m3), and of higher strength, though
sometimes difficult to work with normal hand-tools. Often hardwoods are highly durable even without the
use of preservatives. However, some hardwoods, such as balsa, are extremely light and have a low strength
(hardwood is a botanical rather than a mechanical classification of timber). Hardwoods of a number of
species from the tropical forests have been seriously over-exploited, and are or will soon become scarce.
Nevertheless in most tropical regions there is a sufficient supply of less well-known species which are
available locally at reasonable prices, and these may prove to be ideal for use in road structures in these
regions. Efforts are being made to introduce sustainable hardwood management practices.

The principal species suitable for road structures are shown in Tables E.7.4, E.7.5 and E.7.6 (with their
scientific and common name), which divide them into heavier and lighter hardwoods and softwoods.

Table E.7.4: Heavy hardwoods

Density >650kg/m3 when dried to 18% moisture content


Afrormosia (Pericopsis elata)
Ekki (Lophira elata)
Greenheart (Ocotea rodiaei)
Iroko (Chlorophora excelsa, regia)
Jarrah (Eucalyptus marginata)
Karri (Eucalyptus diversicolor)
Keruing (gurjun) (Dipterocarpus spp)
Opepe (Nauclea diderrichii)
Sapele (Entandrophragima cylindricum)
Teak (Tectona grandis)

Table E.7.5: Lighter hardwoods

Density <650kg/m3 when dried to 18% moisture content


African Mahogany (Khaya ivorensis, anthotheca)
Afzelia (Afzelia spp.)
Dahoma (Piptadeniastrum africanum)
Gum (Eucalyptus saligna)
Jacareuba (Calophyllum brasiliense)
Meranti (Shorea spp.)
Muminga (Pterocarpus anyolensis)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 80

Table E.7.6: Softwoods

Softwoods for bridge construction should generally have density


>420kg/m3 when dried to 18% moisture content
Cedar (Cedrus spp.)
Cypress (Cuppressus spp.)
Douglas fir (Psewdotsuga taxifora)
Kauri, East African (Agathis alba)
Parana Pine (Araucaria angustifolia)
Pine, Caribbean Pitch (Pinus Caribaea)
Pine, Scots or Redwood (Pinus sylvestries)

Forms of timber and timber products


Timber is most commonly utilised structurally
in the form of sawn sections. Timber is
generally sawn at sawmills, in or close to the
forests from which the trees are extracted,
and then supplied to timber wholesalers or
importers, who sort, grade and treat the timber
for supply to the users. There are however
many local variations, and hand-sawing is still
practised in some areas. Whether machine or
hand-sawn, logs are usually sawn by means
of a series of parallel cuts through the log,
which is referred to as flat-sawn. The resulting
sections have a tendency to some distortion
because of different shrinkage rates on the
upper and lower surfaces. The alternative
quarter-sawn logs will have less distortion,
but waste more of the log (see Figure E.7.8).
In sawing timber into rectangular sections
some of the log is inevitably wasted, and a
more economical way to use timber, which Figure E.7.8: Saw cuts
eliminates sawing and also preserves more of its
natural strength, is in the form of round poles as
illustrated in Plates E.7.32 and E.7.33.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 81

Plate E.7.32: Timber poles for piles


Plate E.7.33: Timber poles as deck
(not covered by this Manual) for a concrete
beams of a twin span bridge
box culvert in very soft ground

The poles can be used for piling or as part of a timber lattice structure. Larger logs can be used as
abutment, pier or deck members. Use of pole structures enables younger trees or thinnings from immature
forests to be used, and thus the timber is cheaper.

Bamboo, though botanically closer to grass than timber, can often be of very high strength and strong
enough to be used structurally. Bamboo bridges have been built for road traffic, but it is very difficult
to achieve good durability in bamboo structures and its use is not recommended in this manual without
further local research evidence.

Seasoning
Freshly cut timber contains a substantial proportion of water, up to 100% of its dry weight, and if used
in the green state it is subject to substantial shrinkage movement, as well as being prone to fungal
attack. Thus, for effective structural use timber must be dried so that its moisture content is close to the
equilibrium moisture content (between 10% and 20%, depending both on the type of timber and the
climatic conditions). This process, which has to be carried out with care to avoid distortion, is referred to
as seasoning. Seasoning also increases the strength and stiffness of the timber.

Timber preservation
Preservative treatment is needed to protect timber from fungal attack, insects and marine borers. There
are a number of chemical treatments available, and the success of the treatment depends on effective
choice of both the chemical substance used and the treatment process.

Chemical preservatives include:


ƒ Oil-based preservatives such as creosote;
ƒ Water-based preservatives such as copper/chrome/arsenite;
ƒ Organic solvent preservatives such as pentachlorophenol.

Stress grading
Because of the natural variability of timber, even of pieces from the same source, careful grading, piece
by piece, is essential to ensure safe and efficient use. Stress grading can be done either visually or
mechanically. Visual grading involves making a visual assessment of the extent of the principal factors
affecting strength - knots, fissures, grain slope, wane, distortion, and perhaps worm holes and fungal
decay, and classifying the timber according to predetermined measures of each which are acceptable in
the various grades. Some aspects of visual stress grading are described below. In machine grading, each
piece is subjected to a bending test under load in an automated process, and is graded according to its
deformation; a visual assessment is carried out at the same time.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 82

7.3.2 Properties of timber

Natural defects
Natural defects shown in Figure E.7.9 are features which develop in the living tree, which may affect its
structural usefulness. Some can be accommodated within limits.

The most important are:


ƒ Knots - parts of branches which have
become enclosed in the main tree; can
reduce strength in tension, can be difficult
to work;
ƒ Fissures - splitting separation of the fibres
due to a variety of causes including: stresses
in the standing tree (shakes), slits from
rapid drying, resin pockets (in resin-bearing
softwoods);
ƒ Wane - inclusion in the sawn timber of Part
of the original round surface of the log;
ƒ Insect holes;
ƒ Grain slope - the small angle between the
direction of the grain and the length of the
cut timber.

Several other types of natural defect are


unacceptable and should be eliminated from any
timber used structurally:
ƒ Brittleheart - this material is found in the
centre of some tropical trees, and should be
avoided because it is of low strength and
breaks with a brittle fracture;
ƒ Fungal decay - this is discussed below.

Shape
The processes of sawing and seasoning timber
create distortions which must be limited for
satisfactory use. The four principal types of
distortion encountered are bow, spring, twist and Figure E.7.9: Natural defects in timber
cup. Some suggested limits are given in the Table
below.

Moisture content
Moisture content needs to be limited to achieve the best structural properties and reduce shrinkage as
well as reduce susceptibility to fungal attack. Seasoning should reduce the moisture content to within 5%
of the equilibrium moisture content, which is in the range 10-12% for hot-dry regions, but may be 14-18%
for tropical rainforest regions.

Density
The density of timber depends on its type. Softwoods typically have densities in the range 350-480 kg/
m3, but for bridge construction those suitable have densities above 420 kg/m3 at 18% moisture content
are required. Tropical hardwoods typically have densities in the range 500 to 800 kg/m3 or even higher,
but there are many hardwoods with much lower densities. The foregoing tables divide the common
species of hardwoods into two classes: heavy hardwoods with densities above 650 kg/m3 when dried to
a moisture content of 18%; and lighter hardwoods with densities less than 650 kg/m3.

Strength and elasticity


Strength and stiffness are the most important properties from the point of view of structural utilisation,
and they are closely related; timbers with higher strengths generally also have higher modulus of elasticity.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 83

For structural design maximum allowable stresses in bending, tension, compression (both parallel to
and perpendicular to the grain) and in shear must not be exceeded. These are normally derived from
guidelines or codes of practice in which timbers are grouped into strength groups; values of the strength
and stiffness parameters are given for each group depending on the moisture content and extent of
defects in the timber. Design stresses for the three principal timber groups are shown in Table E.7.7.

Table E.7.7: Design stresses for the three principal timber groups

Heavy Lighter
Softwoods
Hardwoods Hardwoods
(N/mm2)
(N/mm2) (N/mm2)
Bending 15.1 8.6 5.4
Tension 9.0 5.0 3.2
Compression parallel to the grain 11.3 6.8 5.0
Compression perpendicular to the grain 2.2 1.8 1.5
Shear 2.2 1.1 0.9

Durability
The durability of timber relates primarily to its resistance to fungal attack and attack by insects or
marine borers. Durability is enhanced by good timber selection, effective seasoning and preservative
treatment, and maintenance after construction. It is also enhanced by good design, particularly measures
to ensure that timber is protected from water. The end grain and joints are particularly susceptible.

Fungal attack can cause both staining and decay. Some fungi attack cell contents only, rather than
the cell wall substance, and as a result, no structural degradation of the timber occurs. Decay is not
an inherent property of the material itself but depends on the availability of food (the wood itself),
moisture, air and favourable temperature conditions. Some species have more durable heartwood than
others and this is related to the toxic chemicals present in the cells and cell walls of the more durable
species.

The natural resistance of wood to decay can be increased by ensuring that its moisture content is below
18% (based on the oven-dry weight of the wood). In addition to using seasoned timber, the wood
should be protected from dampness by moisture barriers or flashing. If timber is in contact with the
ground, only the more durable heartwood or preservative-treated timber should be used.

In tropical climates, great damage is done to wood by subterranean termites. Termites must have
access to the soil or to some other constant source of moisture. They can severely damage timbers in
contact with the ground and may even extend attack to the roof timbers of high buildings.

Damage above ground may be prevented by ensuring that all means of access are eliminated. Metal
shields or stump caps, or poisoned soil barriers, are effective in preventing the passage of termites from
the foundations to other parts of the structure. Where shields are used, adequate clearance below deck
level should be provided to allow easy, and regular, inspection. In areas of severe infestation, the only
practical methods of control are, however, the use of termite-resistant or preservative-treated timbers.

Apart from termites, there are a number of other insects which attack timber. Moisture is an essential
element for some insects’ development and hence drying is an obvious protective treatment. However,
preservation is generally regarded as being a broad and more positive measure particularly where the
timber is to be used in structural applications.

Protection of timber submerged in salt water against attack by waterborn organisms is usually based
on the use of mechanical sheathing with resistant timbers, concrete or non-ferrous metal, or the use of

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 84

preservatives which are resistant to leaching, such as creosote. Some tropical woods possess a natural
resistance to such attack.

Shrinkage and thermal movement


Some shrinkage and expansion as a result in changes in the moisture content of the timber must be
allowed for in design. The important shrinkage movements are tangential and radial, that is across the
width of the timber; in these directions the movement can exceed 3% as a result of a change in relative
humidity from 90% to 60%. In the longitudinal direction the shrinkage movement is very small, less than
0.1%. The coefficient of thermal expansion is 30-60 x 10-6 per °C across the fibres, but less than one
tenth of this parallel to the fibres. Thermal expansion even of large structures is therefore not a problem.

Fire resistance
Timber is a combustible material and will ignite at temperatures of around 220 to 300°C; it also produces
toxic carbon monoxide and large quantities of smoke when ignited. When used in external conditions on
road structures the risks are from fire caused by fuel spillage in overturned vehicles and wildfires. However,
timber chars as it burns, at about 0.5-0.7mm per minute, which helps to insulate the interior; there is no
instant loss of strength in fire, nor a rapid expansion, and timber structures can safely carry their loads for
some time in a fire, enabling people to escape and the fire to be extinguished. Fire retardant and fire-
protection chemical treatments are available either as paints or for pressure impregnation, but they are
expensive, and the paints require maintenance. Fire protection is therefore not usually applied to external
structures for low volume roads.

7.3.3 Field testing

Some visual indicators for a good quality timber are:


ƒ The cellular tissue should be hard and compact;
ƒ The fibrous tissues should adhere firmly together and should not clog the teeth of the saw;
ƒ Depth of colour indicates strength and durability;
ƒ A freshly cut surface should be firm, shining and somewhat translucent, whereas a dull chalky
appearance is a sign of bad timber;
ƒ In resinous timbers those with least resin in the pores are the strongest and most durable; in non-
resinous timbers those with least sap are best;
ƒ A good timber is uniform in colour, with straight grains, free from dead knots, cracks and shakes,
and has regular annual growth rings.

Shape
The bow, spring, cup and twist of a piece of timber can be measured directly if the timber is placed
on a flat surface. An average of at least 10 measurements should be taken. Some limits to distortion
appropriate for tropical hardwoods and softwoods to be used structurally are shown in Figure E.7.10.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 85

Bow: X should not exceed 15mm per 2m length (in a piece of 75mm and greater in thickness)
Spring: Y should not exceed 7mm per 2m length (in a piece of 250mm or more in width)
Twist: Z should not exceed 10mm per 2m length
Cup: W should not exceed 1mm per 25mm of width

Figure E.7.10: Timber shape criteria

Visual stress grading


Visual stress grading involves making measurements or inspections of natural defects: slope of grain,
knots, fissures and resin pockets, wane and insect holes.

Minimum acceptable limits for all these characteristics are shown in Table E.7.8. Other visible defects
including bark pockets, compression failures, fungal decay, and brittleheart should not be permitted in
any structural timber.

Table E.7.8: Limits of visible defects for structural timber from tropical hardwoods

Property Acceptable limit for structural timber


Slope of the grain 1 in 11
Knots: size 25% of the thickness, up to 75mm
Knots: frequency One sizeable knot per metre of length
Moderate fissures (of greater than 1/3 the thickness but
Fissures and resin pockets less than the thickness): not to exceed in length 20% of the
length or 1.5 times the width.
Wane Not to exceed 25% the sum of the width and thickness
In a square of 100mm sides not more than 32 pinholes
Insect holes (<1mm), nor more than 4 shot holes
(<3mm) nor more than 2 holes of 6mm diameter

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 86

Determination of moisture content


The moisture content of wood can be measured by the oven test method. Small samples are cut from
the wood to be tested, the samples are weighed and then dried in the oven at 100 0C until their weight
becomes constant. They are then weighed again.

The moisture content, m, is calculated, as a percentage.

m % = 100 x weight of water = 100 x initial weight - final weight


dry weight final weight

Typical equilibrium moisture contents for different regions are shown in the foregoing text. The equilibrium
moisture content of the timber to be used should be determined by a laboratory; the oven test can then
be used as a check on the effectiveness of seasoning of timber delivered to site. Moisture content should
be kept within 5% of the equilibrium moisture content.

Strength and elasticity: load testing


The suitability of the structural properties of a timber are normally determined by the use of standard
tables of properties for the species to be used, coupled with stress grading to determine the classification
of the sections available. However, in certain circumstances, the structural properties of timber can be
checked by a direct load test. This is easiest to carry out when the timber is to be used in bending.

A pair of joists is set up between solid supports using the span length which will be used in the actual
structure. The joists are connected to each other by cross-bracing, and a deck is placed over them. The
deck is loaded uniformly, using heavy materials such as bricks or stone, until it reaches the design load.
The deformation at mid-span is then measured. Under the design load it should not exceed about 1/300
of the span. The load should then be increased to 50% above the design load, under which load the
timber should show no sign of failure.

7.3.4 Uses of timber

The principal use of timber in low-cost road structures is for bridge decks, where its structural advantages
can be utilised most fully, and where it is more easily protected from moisture penetration. Timber can
also be used for bridge abutments and retaining structures (though in these uses a relatively shorter
lifetime must be expected), and for culverts. Examples of the use of timber in structures are illustrated in
plates E.7.34, E.7.35 and E.7.36.

Plate E.7.34: Timber decked bridge

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 87

Plate E.7.35: Log abutments & deck

Plate E.7.36: Treated timber culvert

7.4 Plain and reinforced concrete


Plain and reinforced concrete are the widely used choices of material for a range of uses in road
construction. Concrete technology is now established almost universally, even if it is not always well
understood. Suitable raw materials to use as aggregates, forming the bulk of the material, are found almost
everywhere. Cement and reinforcing bars are widely manufactured to standards that are internationally
recognised. Concrete is sometimes the cheapest available option. However, high importation, production
or transport costs, and the high carbon footprint of both cement and steel can make locally produced
materials more attractive. When it is well-made, concrete is also a strong and durable material, leading
to a low maintenance requirement, important for rural structures. Concrete also has the particularly
important property of being able to resist the action of water.

Reinforced concrete is suitable for bridge decks, piers and abutments, as well as for box culverts and
culvert rings; plain concrete may be used for drifts and causeways, culvert rings up to 900mm diameter
and for the foundations of walls, piers and abutments made of masonry and timber.

Because concrete, unlike other structural materials, is generally made on site from its raw materials, an
important requirement for the use of concrete in structures is that both designers and builders, as well
as those responsible for long-term maintenance, understand its essential properties and characteristics.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 88

7.4.1 Materials for concrete

There are three essential constituents of concrete as illustrated in Figure E.7.11:


ƒ Cement is the active ingredient. It constitutes about 10%-15% of the concrete by weight. The
cement, in combination with water, forms a strong matrix which surrounds and binds the aggregate
together. As the concrete mix sets and hardens it gains strength and durability;
ƒ Water constitutes about 5% of the concrete by weight. Initially, it gives the concrete workability,
allowing it to flow and take up the shape in which it is moulded. Over time, the water combines
chemically with the cement in a process called hydration, which causes the concrete to set and
develop strength;

Figure E.7.11: The constituents of concrete

ƒ Aggregates are inert materials, usually of mineral origin, which constitute the bulk of the concrete
(about 75%-85%). They are usually chosen from local sources for low cost, but their size range,
shape, density, hardness and surface properties have important effects on the resulting concrete.

In making concrete, these three constituents are mixed together in appropriate proportions to make a
fluid mass, which is then placed in formwork, compacted to remove air, and finally allowed to set and
harden.

Plain concrete is relatively weak in tension, therefore steel reinforcement is used where tensile stresses
are expected. When reinforced concrete is being made, the reinforcement is formed into a cage or
grid, which is placed in the formwork before the concrete is placed. The following sections describe the
materials requirements.

Cement
The cement most commonly used for concrete is Ordinary Portland Cement (OPC). This is made in
factories in which a mixture of limestone (or other calcium-rich minerals) together with clay or shale is
fired at a high temperature, and the resulting cement clinker is ground to a fine powder. The operation
is highly controlled and the resulting cement is produced to a specification which defines the essential
properties including strength, setting rate and chemical composition.

Cement is normally delivered to site in 25 - 50kg paper bags. The cement must be kept totally dry until
it is to be used, otherwise it will begin to react with the water and be rendered useless. Cement should
therefore be stored off the ground in a shaded, dry and well-ventilated place (Figure E.7.12). If any lumps
of hardened cement are found in a bag, the cement in that bag should not be used for structural work.
Cement should typically be used within 6 months and therefore stored in a ‘first in – first out’ system.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 89

Figure E.7.12: Cement stored in dry conditions

Water
The mixing water used should be clean and free from salts. It can be taken from rivers, lakes or wells or
from a treated water supply. Salty water may be used for plain concrete, though it will affect the rate of
setting, but should not be used for reinforced concrete. River water containing sediments can be used if
the sediments are first allowed to settle out in a tank or drum until the water is clear.

Aggregates
The aggregate is divided into two parts: coarse aggregate and fine aggregate. The fine aggregate is
normally a naturally occurring sand, with particles up to about 2mm in size. The coarse aggregate is
normally stone with a range of sizes from about 5mm to 20mm (or sometimes larger); it may be a naturally
occurring gravel, or more commonly crushed or hand-broken quarry stone. In areas without hard stone
resources and with an establish fired clay brick industry, burnt bricks can be machine or hand crushed to
be used in concrete.

Aggregates must be entirely free from soil or organic materials such as grass and leaves, as well as fine
particles such as silt and clay, otherwise the resulting concrete will be of poor quality. Some aggregates,
particularly those from salty environments, may need to be washed to make them suitable for use. Tests
for aggregate quality are described in Section 7.4.4.

Both the coarse and fine aggregates need to contain a range of particle sizes, and are mixed together
in such a way that the fine aggregates fill the space between the coarse aggregate particles. A ratio by
volume of one Part fine aggregate to two parts coarse aggregate is generally used. Aggregates can be
crushed and screened by hand (Plate E.7.37) or by machine.

Aggregates should be stored in such a way that they do not become contaminated by soil, and that
rainwater can drain easily.

Plate E.7.37: Aggregate crushed and screened by hand

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 90

Reinforcement
Reinforcement is normally in the form of steel bars. Three characteristics are of primary importance:
enough strength that a small amount of reinforcement can be used to carry the tensile and shear forces;
enough ductility that the rods can be bent without breaking, and, if a member is overloaded, that the
structure will deform without failing; and sufficient bond between the reinforcing and the concrete that
forces can be transferred between them.

Two types of steel reinforcement are in common use: mild steel and hot rolled high-yield steel. Mild steel
bars are round, while high yield bars have a deformed surface to improve the bond with the concrete.
Typical reinforcement sizes range from 6mm to 30mm in diameter. Reinforcing steel is usually available
both in rod and mesh forms. Reinforcement bars are cut to the required length and bent to the required
shape; they are then tied together in the arrangements shown on the drawings using binding wire and
spacer blocks.

On site reinforcement should be kept straight until needed, and should be stored clear of the ground to
prevent contamination with soil. An example of a steel reinforcement cage being assembled on site for
the construction of a cut-off wall in a drift is shown in Plate E.7.38.

Plate E.7.38: Steel reinforcement cage being assembled

Concrete mixes
The proportions of the constituents may be varied to obtain the required properties. As a rule, the larger
the amount of water added to the mix, the more fluid and easy to cast in place it will be, but the lower will
be the final strength and durability. The ratio of water to cement should therefore be as low as possible
for the necessary workability of the concrete. Given this requirement, mixes with a larger proportion of
cement to aggregates will tend to be stronger and more durable.

Three principal types of concrete are required for use in low volume roads as shown in Table E.7.9.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 91

Table E.7.9: Concrete Types

Grades 20 and 25
This is concrete intended for use in reinforced structures and load-bearing
applications such as bridge decks and culvert rings. The grade indicates the
Structural concrete target crushing strength of cubes (N/mm2) at 28 days after casting. Maximum
aggregate size is normally 20 mm to allow the concrete to pass round the
reinforcement and give good compaction. Typical mix proportions for Grades
20 and 25 are given in the Table below.
This is appropriate for gravity structures where reinforcing steel is not used.
A large sized stone (up to 50mm) is permitted. For the construction of drifts
and causeways, larger pieces of stone (referred to as plums) may be set in
place before the concrete is poured, to act as fill. These should be of the same
Mass concrete
quality as the aggregate and have a maximum size not greater than three-
quarters of the depth of the concrete. The cement content for mass concrete
is higher than for lean concrete but lower than for structural concrete. Mix
proportions are 1:3:6.
This is a meagre mix with a low cement content. It is used for blinding the
foundation excavations for structures, where it acts as a clean working surface
prior to placing structural concrete. It is also used as a porous backing to
Lean concrete
structures and behind weep holes to allow water to migrate through without
washing soil particles through the structure. The mix proportions are 1:4:8 by
volume.

The nominal mixes shown in Table E.7.10 should achieve the strengths indicated, with good quality
graded aggregates, and water content just sufficient to give adequate workability. It is crucial that the
mix does not contain excess water as this will result in increased porosity in the final concrete, and
considerably reduced strength and durability.

Table E.7.10: Concrete grades, strengths and batching strengths

Cement/ Material required for 1m3 finished concrete


Expected 28 (guidance)
Class of fine agg./
day strength
concrete coarse agg. 50kg Cement Fine Coarse
N/mm2
(guidance) bags (kg) (m3) (m3)
Lean - 1:4:8 3.3 (166) 0.47 0.94
Mass 15 1:3:6 4.3 (215) 0.46 0.92
Grade 20 20 1:2:4 6.0 (300) 0.42 0.84
Grade 25 25 1:1.5:3 7.3 (365) 0.38 0.76

7.4.2 Production and placement

Mixing
Concrete may be mixed (or batched) by hand or by a mechanical mixer. When batching by volume is to
be used, the mix proportions should be measured using a gauge box with dimensions as shown in Figure
E.7.13. The gauge box has a volume of 0.036 m3, equivalent to one 50kg bag of cement.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 92

Batch with 1 bag cement


Class of Number of boxes Yield per
Concrete of aggregates batch
Fine Coarse (m3)

Lean 4 4 0.30

Mass 3 6 0.24

Grade 20 2 4 0.16

Grade 25 1.5 3 0.14

Figure E.7.13: Concrete mixes (guidance)

Aggregates and cement are thoroughly mixed together in the dry state, and then the water added
gradually while mixing until a uniform mass of the right workability is achieved. Concrete should be mixed
on a clean, hard, level and impermeable platform, or in a mixer.

Transporting
Concrete should be mixed as near as possible to the site of placement and may be transported using
trucks, wheelbarrows, or even using headpans for sites with difficult access. The wet mix should be
transported within 30 minutes to allow placing before setting commences.

Placing
The formwork or shuttering for the concrete must be clean, smooth faced, and secure from movement
or leakage when the concrete is poured. Formwork is normally constructed from timber and plywood,
especially where shapes are complex. Where the same shape is repeated (eg for culvert barrels or
headwalls) then steel formwork can be economical and efficient to use). The dimensions and widths of
the space to be filled must be carefully checked. Formwork construction should be planned to enable
later removal. Formwork must be strong and well secured so that it does not move or distort under the
pressures exerted by the wet concrete or the vibration operation. It must be complete without gaps for
the wet concrete to escape through. Any reinforcement must be well secured and positioned away from
the formwork with set mortar or plastic spacers to ensure that the correct cover is achieved. Purpose
made mould oil can be used to aid later removal of the formwork without damage to the concrete; used
engine oil may be used for this function.

The wet concrete should be placed in layers and rammed or vibrated immediately to form a dense well
graded mass with no air pockets. The layers should be built up and compacted into each other without
allowing joints of set concrete to form (except at predetermined construction joints). The concrete should
be placed in layers of thickness less than 300mm when hand ramming. This may be increased to 600mm
when a vibrating poker is used. Care must be taken not to disturb the formwork or any reinforcement
during placing and compaction. Over-vibration must be avoided as it can lead to segregation of the
concrete paste from the aggregates.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 93

The top of the placed concrete should be finished smooth with a mason’s trowel or float. However, any
day work joints (eg in a wall lift) should be left rough to ensure a good bond for the next layer of concrete.
Concrete should not be mixed or placed in ambient temperatures of less than 3˚C or above 38˚C.

Curing
Concrete hardens as a result of hydration of the cement with water. Fresh concrete contains more than
enough water to hydrate the cement completely but if the concrete is not protected against drying out,
the water content, especially near the surface, will be insufficient for complete hydration. This causes
cracking. Direct sunlight will speed up evaporation so temporary shading should be provided where
needed. Curing should start as soon as the concrete begins to harden (3-4 hours after placing). Suitable
methods include: sprinkling or flooding; covering with empty cement bags, hessian bags or other fabric,
sand, sawdust (50mm thick), grass or leaves, all of which should be kept wet. For faces cast against
formwork, the formwork may be loosened after one day and left in place, dampening from time to time.
All concrete should be cured for at least 7 days. During this time it should be protected from frost if
necessary.

Detailed local specifications for concreting procedures should be followed, as these can take account of
local raw materials, site practices and climate. Protection of the workforce from injury is a vital element of
good concreting practice.

Table E.7.11 provides guidelines for the placing, compacting and curing of concrete to assist in the
attainment of a quality material.

Table E.7.11: Recommendations for good quality concrete

Activity Recommendations of good practice


ƒ Forms and the shutters should be cleaned before placing the
concrete
ƒ Concrete should be placed in layers of 300mm depth
ƒ Concrete should not be placed in heaps, as this causes separation
of the stones from sand and cement
Placing concrete
ƒ Concrete should not be dropped from a height of more than 1.5m,
as this also causes separation of the stone from the sand and
cement
ƒ Reinforcement bars are to be placed inside the shuttering before
placing the concrete
ƒ Compacting is undertaken by tamping with a steel or wooden rod.
Compacting concrete It is important to remove all the air in the concrete as entrained air
reduces the strength of the concrete.
ƒ Curing means keeping the outside of the concrete moist (wet)
during the setting (hardening) of the concrete by:
‚ Wetting the concrete surface frequently
‚ Covering the surface with wet material (cloth, paper bags, sand
Curing concrete etc)
ƒ Hardening of concrete requires at least seven days. Curing prevents
cracks in the surface layer of the concrete.
ƒ As cement is normally one of the most expensive items in the
construction process, it should not be wasted.
Too much cement = costly Too little cement = low strength

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 94

7.4.3 Properties of concrete

Workability of fresh concrete


In its freshly mixed state concrete needs sufficient workability to enable it to be placed into the formwork
and compacted. The workability needed depends on the shape of the formwork to be filled, the amount
of reinforcement in it, and sometimes on the method of transportation. Workability is measured on site
by the slump test, which is described below. Table E.7.12 indicates the maximum workability suitable for
different situations.

Table E.7.12: Maximum slump values for particular uses

Concrete use Maximum slump


Lean concrete 100mm
Reinforced foundations 80mm
Other reinforced areas 50mm

Strength and stiffness


The strength of a concrete develops slowly as the cement hydration reaction continues. After 28 days, the
concrete will have attained most of its final strength, and this is the age at which the strength is specified
for use in design. Concrete mixes are designed to achieve a given 28 day strength in compression, as
measured by crushing tests on cubes or cylinders. Typical structural concretes have strengths in the range
25 to 40N/mm2. For high quality control concrete crushing test samples are made regularly on site, and
sent to a testing laboratory for testing at 28 days.

Tensile strength and stiffness also develop as the compressive strength develops. The tensile strength of
concrete is normally about one-tenth of its compressive strength. A quality control test which could be
used to assess the strength based on the tensile strength is suggested in field testing below.

Moisture movement
Wet cured concrete exposed to air will shrink over time. It will also expand and contract subsequently
as a result of changes in ambient humidity or exposure to rain or moisture. The extent of shrinkage
depends on the properties of the concrete and ambient conditions, but typically about 0.8 to 1.0mm
per m of drying shrinkage can be expected (in all dimensions) with subsequent variations of about one-
third of these values. This can cause unsightly cracking in concrete structures unless joints are provided at
intervals to allow it to occur. Additional (creep) moisture movements occur as a result of the load. Creep
continues over a long period of time (some months). Both creep and shrinkage can be restrained (though
not prevented) by the presence of reinforcement.

Durability
The durability of concrete depends on its resistance to the major causes of deterioration: corrosion of the
reinforcement, frost attack, sulphate attack, chemical attack, and deterioration of the aggregate-cement
bond. There are four principal agents of deterioration shown in Table E.7.13. Protection of the concrete
from these agencies of deterioration can be achieved by:
ƒ Good compaction - permeability of concrete is increased if compaction is poor or cracking occurs
as a result of poor curing;
ƒ Adequate cover to reinforcement - minimum cover is specified according to the environmental
conditions; greater for external surfaces, and surfaces which are to be tooled etc;
ƒ Use of low permeability concrete - by using well-compacted concrete with low water cement ratio,
which reduces the ability of water to move through the concrete;
ƒ Providing a minimum cement content - to create a sufficiently alkaline environment to inhibit
reinforcement corrosion, a minimum quantity of cement is needed. Nominal mixes provide an
adequate amount;
ƒ Minimise the risk of alkali silica reaction - by limiting the alkali content of the concrete or by using
non-reactive aggregates.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 95

Table E.7.13: Agents of concrete deterioration

Corrosion is caused by an electro-chemical reaction occurring in


the presence of water and air. It occurs when water gains access to
Corrosion (rusting) of
the reinforcement either through inadequate concrete cover to the
the reinforcement
reinforcement, or because of poorly mixed or poorly compacted concrete,
or as a result of cracking.
Frost attack is caused by expansion of water in the cement paste pores
resulting in reduction of the strength of the cement paste - concrete is
Frost attack
particularly vulnerable at early ages (up to 3 days) when its strength has
not developed.
Sulphates in soil, sea water and some aggregates will react with the
Sulphate attack
hydrated cement resulting in expansion and damage of the concrete.
Alkali-silica reaction So-called “concrete cancer” is a deterioration of the concrete as a result
of a reaction between alkaline fluids and reactive minerals in certain types
of aggregates.

Thermal movements
The coefficient of thermal expansion of concrete is about 10 to 14 x 10-6 mm/mm, ie about 3mm per m
for a 30°C temperature rise, about the same as for structural steels (12 x 10-6). Thus, for long concrete
structures such as multi-span bridges, expansion joints are needed to allow for seasonal temperature
changes.

7.4.4 Field testing

Presence of silt and clay in sand and coarse aggregates:


visual test
Rub a sample of the sand between damp hands, and note
the discolouration caused. Clean materials will leave the
hands only slightly stained. If the hands remain dirty after
the sand has been thrown away, it indicated the presence
of too much silt and clay.

Presence of silt and clay in sand and coarse aggregates:


bottle test
Half fill a clear bottle or tumbler with aggregates (Figure
E.7.14); add water until it almost reaches the top, shake
vigorously and then allow the aggregates to settle.
After about 30 minutes there should be no fine material
deposited on top of the aggregates and the water should
be clear. Salt may be added to the water (one teaspoon
per 0.5 litre) to speed the settlement. If the height of the
silt layer is more than 6%, the sand should be washed.

Figure E.7.14: Field fines test

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 96

Quality of reinforcement
Reinforcement to be used should be supplied to comply with the national standards. Before use it should
be checked to ensure that all bars are straight and free from loose scale, loose rust, oil, grease and dirt.

Quality of cement
Cement to be used should be supplied to the national standards. To check that it has not deteriorated
in storage, it should be tested by hand for hardened lumps. A small proportion of lumpy cement may be
removed by sieving.

Test for workability of fresh concrete: the slump test


The slump test is the standard way of making sure that concrete does not vary in consistency due to
variations in the water:cement ratio. To make the test the standard cone-shaped mould is required, and
a steel rod described in Figure E.7.15.

Mould shall be made of a metal not readily


attacked by cement paste and not thinner
than 1.5mm (eg galvanised steel). The
interior of the mould shall be smooth and
free from projections such as protruding
rivets and shall be free from dents.

The mould shall be in the form of a hollow


Part of a cone having the following internal
dimensions:
diameter of base: 200 ± 2mm
diameter of top: 100 ±2mm
height 300 ± 2mm.

Tamping rod: Made out of straight steel bar


of circular cross section, 16mm diameter,
600mm long with both ends hemispherical.

Figure E.7.15: Slump test mould

The procedure for the test is illustrated in Figure E.7.16. The cone has to be clean and dry inside and is
put on a smooth, hard surface. The cone is filled one-quarter full. Holding it firmly in place with the metal
feet, rod the concrete thoroughly 25 times. Then add more concrete to about half-way and rod it another
25 times, taking care to take the rod just through into the first layer. Next add the third layer filling the
cone three-quarters full, and rod again 25 times, going through into the layer below. Finally fill the cone
up, rod 25 times again, going well down into the third layer and smooth off the top. The top is smoothed
off level with the cone.

Wipe the metal plate it stands on clean and dry and wipe round the base of the cone. Then, carefully and
keeping it quite straight, lift the cone off and put it down beside the concrete. The concrete will collapse
to some extent - very dry concrete hardly at all, very wet concrete completely. Test it by measuring how
far it has collapsed.

To measure the slump, rest the rod across the top of the empty cone, so that it reaches over the concrete.
With a rule measure down from the underside of the rod to the top of the concrete always measuring
from the highest point on the concrete.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 97

Figure E.7.16: Slump test

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 98

Concrete strength: cube tests


Test cubes may be made on the site to check whether the concrete used on a job has the required
strength. The cubes are used to find the crushing strength of the concrete. It is essential that the cubes
are made with great care. They are generally sent away to a laboratory for testing. This test is only
recommended for structural concrete with a design strength of 20 N/mm2 or above. Cast cubes should
be cured as the constructed concrete by immersion in a tank of clean water for the initial period after
casting and tested 28 days after casting. In some cases additional cubes are cast and crushed after 7 days
to indicate whether the concrete is on track to achieve the required specification strength.

Strength: impact test


For a rough strength assessment for concrete of mass concrete grade, make a set of 10 briquettes from
plain concrete of dimensions 100 x 200 x 50mm. Place each brick in turn with its largest face downwards,
resting on timber battens 150mm apart. Drop a mason’s 2 kg hammer from a height of exactly 0.5 m so
that it strikes the upper face midway between the battens. The briquette should not break. If more than
1 in a sample of 10 bricks break in this test, the concrete is not of adequate strength and should not be
used.

7.4.5 Uses of concrete

Reinforced concrete is used for bridge decks, piers and abutments, as well as for box culverts and culvert
rings; and plain concrete may be used for drifts and causeways, and for the foundations of masonry and
timber walls, piers and abutments. Typical examples of where concrete can be used in small drainage
structures for low volume roads are shown in Plates E.7.39 to E.7.43.

Plate E.7.39: Concrete culvert rings

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 7 - 99

Plate E.7.40: Concrete drift under construction

Plate E.7.41: Reinforced concrete slab

Plate E.7.42: Flow spreader structure at the outlet of a mitre drain on a steep fragile
slope. Bio-engineering planting should be established downhill of such structures

Plate E.7.43: A simple reinforced concrete walled box culvert ready to


receive pre-cast top slab units (timber or reinforced concrete)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 100

8. STRUCTURE DESIGN

There is a large amount of energy stored in flowing water. A fast flowing river 0.5m deep can wash away
a car or pickup truck. Even at lower volumes and velocities, water can wash away road structures. A high
priority task in designing a road structure is therefore to minimise the disturbance to the water flow in the
channel, which then minimises the potential damage to the structure and scouring of the watercourse.

The vast majority of structural failures occur during flood periods and over 50% of these failures can be
attributed to scour. The initial section of this chapter deals with scour and how to design and construct a
structure to withstand scour effects.

There are often a number of elements which form a road structure. In some cases these are common to a
range of structures. After the section dealing with scour this chapter is broken down into sections which
each cover an individual structural element. Table E.8.1 shows the aspects which must be consulted for
the design of different structural elements for water crossing structures.

Table E.8.1: Guidance on design aspects

Vented Large-bore
Structural Item Drift Culvert Bridge
Drift Culvert
Foundations 9 9 9 9 9
Structural slabs 9 9 9
Cut-off walls 9 9 9 9 9
Pipes 9 9
Headwalls & wingwalls 9 9 9 9
Apron 9 9 9
Approach ramps 9 9 9
Downstream protection 9 9 9 9 9
Arches 9
Bridge design
ƒ general
9
ƒ deck
(arch
ƒ abutments
bridges)
ƒ piers
ƒ bearings & joints

8.1 Scour
Scour is the erosion of material from the river sides and bed due to water flow. Damage due to scour is
the most likely cause of structural failure (see Plate E.8.1). Minimising or eliminating the effects of scour
should therefore receive the most attention when designing any structure. Scour can occur during any
flow but the risk is generally greater during floods.

There are three major types of scour to be considered:


ƒ River morphology: these are long-term changes in the river due to bends and constrictions in the
channel affecting the shape and course of the channel.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 101

ƒ Construction scour: this is the scour experienced around road structures where the natural channel
flow is restricted by the opening in the structure. The speed of the water increases through the
restriction and results in more erosive power, removing material from the banks and bed.
ƒ Local scour: occurs around abutments and piers due to the increased velocity of the water and
vortices around these obstructions.

Plate E.8.1: Bridge damage due to scour of abutment

The latter two scour types are the most important to consider when designing a structure. The amount of
scour at a structure will be affected by the following factors:
ƒ Slope, alignment and bed material of the stream: the amount of scour is dependent on the
speed of the water flow and the erodability of the bed material. Higher water velocities result in
more scour.
ƒ Vegetation in the stream: any vegetation growing permanently in the stream can improve the
strength of the river bed, reducing scour. The vegetation can also reduce the speed of the water.
ƒ Depth, velocity and alignment of the flow through the bridge: the faster the flow, the more
scour will occur. If the flow is not parallel to the constriction more scour will occur on one side of
the constriction.
ƒ Alignment, size, shape and orientation of piers, abutments and other obstructions: water is
accelerated around these obstructions, creating vortices with high velocities at abrupt edges on
the obstruction, increasing the scour depth.
ƒ Trapped debris: debris can restrict the flow of water and cause an increase in water velocity. It is
important that structures are designed to minimise the chances of debris being trapped and to
ensure that inspections and maintenance are carried out after flood periods to remove any lodged
debris.
ƒ Amount of bed material in the water: if the water is already carrying a large amount of material
eroded from further upstream a greater amount of scour will occur at the structure.

Inspect the proposed structure site and the watercourse upstream and downstream for evidence of
existing scour, erosion or deposition in the watercourse and banks.

It is difficult to accurately predict the level of scour that may be experienced for a particular design. There
are many formulae for predicting the amount of scour around a structure but these formulae, in general,
require detailed knowledge of the river and bed characteristics. They are also based on empirical data
and will often give different design scour depths. Engineering judgement will be required. This Manual
proposes a number of ‘rules’ for designing to resist scour. It must be stressed that these rules are not
infallible and local knowledge should also be taken into account when designing a structure.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 102

8.1.1 Rule 1 - Provide minimum foundation or cut-off wall depths

Regardless of the required depth for foundations determined by the ground conditions and predicted
scour, the minimum foundation depths shown in Table E.8.2 should be provided. The depth is measured
from the lowest point in the bed of the watercourse at the crossing point. These depths can only be
reduced where firm rock is encountered at a shallower depth and the foundations are firmly keyed into
the rock.

Table E.8.2: Foundation depths

Structure Foundation Depth Cut-off wall depth


Drift Not applicable 1.5m
Relief culvert Not applicable 1.0m
1.5m
Watercourse culvert Not applicable
(headwalls and wingwalls)
Vented drift Not applicable 2m
Large bore culverts 3m 3m
Bridges 3m 3m

8.1.2 Rule 2 - Create a minimal constriction to the water flow

The amount of scour experienced at a structure is proportional to the restriction in the normal water flow.
If the flow is considered unconstrained then scour will not exist. If the site conditions permit, the opening
widths in Table E.8.3 should be provided to eliminate the effects of scour.

Table E.8.3: Opening widths

Peak flood flow rate 0.5 1 2 4 6 8 10 15 20 25 30 m3/s


Minimum width (W) 3.5 5 7 10 12 14 15 19 21 24 26 m

In some cases, particularly for bridges and larger flows, it will not be possible to provide the opening
widths shown in Table E.8.3 above. The design, particularly the level of foundations, should allow for a
lowering of the river bed level due to scour. The amount/depth of scour (as shown in Figure E.8.1) that
will occur depends on the following 3 factors:
ƒ Constricted flow width;
ƒ Maximum flow rate;
ƒ The type of material forming the sides and bottom of the watercourse.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 103

STREAM WIDTH
L

M.F.L.

1
D D

The depth of scour is therefore:


DS

Depth of scour = flood water depth FACE OF


ABUTMENT
at structure - original unconstrained
watercourse depth
W = THE ORIGINAL WIDTH OF THE STREAM

Ds = D1 – D L = THE DESIGNED WATERWAY. WHEN THE BRIDGE IS


ASSUMED TO CAUSE CONTRACTION, L IS LESS THAN W

D = THE ORIGINAL DEPTH OF STREAM


1
D = THE EXPECTED SCOUR DEPTH UNDER THE BRIDGE DUE
TO ITS CONSTRICTION

DS = ADDITIONAL DEPTH OF SCOUR DUE TO CONSTRICTION


OF BRIDGE

Figure: E.8.1: Scour allowances

The three following graphs (Figures E.8.2, E.8.3 and E.8.4) allow the prediction of the water depth in the
channel, which will allow the depth of scour to be calculated.

Figure E.8.2: Scour – Fine silt

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 104

Figure E.8.3: Scour - Fine sand

Figure E.8.4: Scour – Coarse sand

The depth of scour indicates the general level of erosion that will occur in the river bed. Additional local
scour will occur near bridge abutments and wingwalls and also at the edges of aprons. Table E.8.4 shows
the factor that the general scour should be multiplied by to calculate the depth of scour that may be
encountered near structural elements.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 105

Table E.8.4: Scour depth adjustment

All foundations should be constructed below the predicted depth of scour.


Predicted maximum depth of scour = depth of general scour x local scour multiplier

Local scour at structural elements Local scour multiplier


Long abutments parallel to water flow in straight channels 1.5
Abutments in curving channels and/or Part of structures with
2.0
multiple openings
Abutments and wingwalls where flow reaches structure at an
2.25
angle greater than 20 degrees
Ends of protective aprons or drift slabs 2.5

8.1.3 Rule 3 - Avoid the use of piers

If piers are absolutely necessary they should be aligned exactly in the direction of water flow.

Figure E.8.5 shows the likely depth of scour that may be encountered around piers that are aligned in the
direction of water flow. Scour around piers will be doubled for piers that are aligned 10-15o away from
the direction of water flow.

Figure E.8.5: Depth of scour around piers (from TRL ORN9)

Plate E.8.2 shows an example of the consequences of constricting a watercourse with a structure that is
too small leading to excessive scour allied to inadequate protection of the abutments

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 106

Plate E.8.2: Failure of structure due to a combination of constriction of the watercourse


(structure too small), scour and inadequate protection of abutments.

8.2 Foundations
The strength and durability of any structure will be determined by the quality of its foundation and the
bearing capacity of the soil (refer to Chapter E.2).

For small, simple structures such as drifts, culverts and vented fords it will be sufficient to construct the
structure on well drained, firm soil. Referring to the soil bearing capacity tables in Chapter E.5 these
conditions include any rock, clays and silts that are at least “firm” or sands and gravels that are at least
“loose”. These conditions can be determined on site by checking for footprints when walking over the
proposed location. If more than a faint footprint is left it will be necessary to improve the ground before
construction commences.

If the ground conditions are poor at the proposed level of the structure’s foundation it will be necessary
to continue excavation to firm material that can provide sufficient bearing capacity. The engineer then will
have three options for the construction of the structure:
ƒ Alter the design to lower the level of the foundations;
ƒ Replace the poor excavated material with new material that has a better bearing capacity (eg a well
graded sand and gravel) that is compacted into the excavation in 300mm layers;
ƒ Provide a piled foundation (not covered by this Manual).

For all structures it is necessary to start the construction on a well drained, level base. The excavations for
all structures, apart from those built on rock, should be dug an additional 300 mm below the proposed
foundation level. A 300mm layer of sand and fine gravel should then be placed and levelled in the
bottom of the excavation to provide a good base for the structure. Alternatively at least 10 cm of lean
concrete blinding should be laid to provide a firm clean working platform.

A rough method for calculating the load exerted by the foundations of a vented ford or large bore culvert
on the ground will be to calculate the load of the structural fill material and multiply by a safety factor
shown in Table E.8.5.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 107

Table E.8.5: Bearing safety factor

Material Load per metre of fill Safety factor


Concrete/gravel 25kN 1.5
Earth 20kN 1.5

For example:
The central section of a vented ford is 2m high (from its foundation level) and has masonry walls with an
earth fill inside. What is the foundation loading?

The load exerted on the soil below the structure will be: 2 x 20 x 1.5 = 60 kN/m2

Where a foundation is to be built on rock which may be sloping down to the watercourse (see figure E.8.6),
it will be necessary to form a level platform for the foundation. This may be achieved by either breaking
out the rock to give a level foundation or building up the foundation to level by placing concrete around
drilled and grouted mild steel bars. The preferred option which should be adopted, unless the rock is too
hard to break out, will be to break out a level platform. Sloping firm rock abutments are of course suitable
for arch bridge springings. In these circumstances the rock should be excavated approximately to a plane
roughly at right angles to the slope of invert of the arch at the springing. The face may be cut in steps to
increase bond between the structure and rock foundation.

CONCRETE

BEDROCK EXCAVATED TO
SLOPING BEDROCK FORM HORIZONTAL "SHELF"
OR "BENCH"

GROUTED STEEL
BAR ANCHORAGES SLOPING BEDROCK

A) BUILD ONTO ROCK WITH CONCRETE B) EXCAVATE INTO ROCK TO FORM


"SHELF" OR "BENCH"

Figure E.8.6: Construction on sloping bedrock

8.3 Structural Slabs

8.3.1 Drifts

The primary objective in the design of a drift is to provide a suitable surface for vehicles to drive across
while creating minimal disturbance to the water flow. Drift slabs should therefore follow, as closely as
possible, the bed of the watercourse. The drift slab surface should be no more than 200mm above the
existing bed level. However, it is desirable to construct the drift with a finished level at the same level as
the river bed. Slabs which are constructed more than 200mm above the existing bed level are likely to
cause severe erosion downstream of the drift, requiring frequent maintenance.

NOTE: There is one situation where it may be permissible to raise the finished level of the drift above
the river bed. If the site selected for the drift appears to suffer from silting the final level of the drift could
be raised 200-300mm above the natural river bed. This raising of the level will cause water to flow slightly
faster over the drift and reduce the potential for the drift to silt up.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 108

If the river is flowing in a channel with banks on each side it will be necessary to ensure that there is a
suitable approach slope from the road on each side to the drift in the bottom of the river bed. These
approach slopes should not be so steep that vehicles get stuck at the bottom of the drift. A maximum
gradient between 5 and 10% will be determined by the vehicles that are using the road. A gradient of
10% may be used if the only vehicles using the road are cars and light trucks. A gradient of 7.5% may
be used for medium size trucks and small minibuses and a gradient of 5% used if buses and large trucks
(>10tonnes) are expected to travel along the road. Allowance should be made for the fact that heavier
vehicles may use the road following improvement of the route.

Although vehicles may not be able to cross the drift during periods of high water it is essential that the
drift slab extends beyond the highest flood level to ensure that scour and erosion will not take place at
each end of the drift. It may, therefore, be necessary to construct the drift slab to the top of the river banks
at the end of the approach slope.

To reduce the cost of construction it may be possible to reduce the width of the drift slab so that it is
narrower than the normal road width. Vehicles would not be able to pass each other on the drift so the
designer must ensure that there is sufficient passing space on each side of the drift to allow vehicles to
wait and pass each other. To prevent vehicles driving off the drift and possibly getting stuck in the soft
or loose river bed, or vehicles attempting to pass each other on the drift, guide stones should be placed
along the edges of the approaches and across the drift (see Plate E.8.3).

Plate E.8.3: Guide stones at the edge of a drift

The width of the central or flat middle section of the drift should minimise disturbance to the water flow.
The construction of the road will cause a larger amount of water to flow across the drift due to water
flowing off the road along the side drains. Drifts should be constructed with the central flat sections of
the following length:

River crossings width of the watercourse


Relief and perennial stream drifts width of the dry bed: minimum dimension of 2m

Drift slab construction


There are four possible solutions for constructing the drift slab, in descending cost:
ƒ Concrete slab;
ƒ Cement bonded stone paving;
ƒ Dry pitched stone paving;
ƒ Gabions with gravel or broken stone.

The main factors affecting the choice of construction method are:


ƒ The nature of the river bed;

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 109

ƒ The expected volume and flow rates of the water;


ƒ The availability of different construction materials;
ƒ The cost of labour.

If large volumes of fast flowing water are expected it will be necessary to use a concrete slab or cement
bound stone paving as the water will erode gravel and dislodge hand pitched stones. In the cases of
slower flowing water or small streams hand pitched stone or gabions are likely to be acceptable and a
cheaper option.

8.3.2 Concrete slab

Although concrete slabs are the most expensive they are a long lasting, low maintenance solution. The
concrete slab should extend the full width of the drift (Plate E.8.4) between the cut-off walls with a
minimum thickness of 250mm. In areas where stone is locally available ‘plums’ may be put in the slab to
reduce the amount of cement required and hence reduce the overall cost.

Where plums are used they should not have a dimension greater than 75mm (100mm where the slab is
300mm or thicker) and should be placed as far as possible in the middle of the slab.

Plate E.8.4: Concrete slab drift

8.3.3 Cement mortar bonded stone paving

Stone paving will offer a cheaper alternative to a concrete slab in areas where masonry or locally
manufactured blocks of sufficient strength are available. The slab should be a minimum of 300mm thick
which may require more than one course of paving to be laid. The blocks should be laid in an arrangement
to ensure that the different courses interlock with each other.

8.3.4 Hand pitched stone

In areas where masonry stone is widely available this option is likely to be cheaper than constructing
a concrete slab (see Plate E.8.5). However, it is only suitable for low velocity flows and can take a
considerable length of time to construct for larger crossings. It is essential that the stones are well placed
to ensure that they are interlocked to prevent them being washed out by the water. The whole structure
can be washed away if the water can wash out one stone, as this weakens the remaining structure. Larger
stones are better than smaller ones as they are less likely to be washed away. The best stones to use are
angular and flat faced and should be placed on their edge, to give the greatest interlock between stones.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 110

Plate E.8.5: Hand pitched stone drift

8.3.5 Gabions and gravel

This option is likely to be the cheapest and quickest option for constructing a drift slab. Smaller stones
may be used in the gabion than for hand pitched stone and maintenance does not require specialist skills.
However, gabion baskets and gravel will be unable to withstand large flows of water. The drift basically
consists of a gabion basket on the down stream side which acts as a dam to prevent the gravel being
washed away (see Plate E.8.6). (note that the sand has been washed out on Plate E.8.6 but severe erosion
has not occurred).

Plate E.8.6: Gabion drift

Where gravel may be washed away; but there is a reasonable amount of gravel in the riverbed, it may
be possible to protect the riverbed and trap gravel and sand in the top of a gabion mattress to create
a vehicle running surface. Gabion mattresses are similar to gabion baskets except that they are a flatter
section; usually 250-300mm deep, and cover a wider plan area. Sand and gravel will tend to be trapped
on the top of the gabions which will prevent wear of the wire by traffic.

An additional measure to stabilise the face of the gabion and the retained material is to insert natural
fibre matting in the top and face of the gabion. This also encourage vegetation growth for improved
stabilisation (see Plate E.8.7).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 111

Plate E.8.7: Natural fibre matting inserted in the top and face of the gabion

8.3.6 Slab construction (vented fords and large bore culverts)

The number of options available for the type of slab will depend on its ultimate use. If the slab is to be
used on the top of a fill layer, as in the case of vented fords or causeways, it is likely that only a concrete
slab or cement bonded stone paving will be suitable. The slab should also have a 2-3% crossfall in the
direction of water flow to ensure that the deck drains quickly when overtopped and sand or silt is not
deposited on the running surface.

8.4 Cut-off walls


Cut-off walls, also called curtain walls, should be provided at the edge of a structure. They prevent water
eroding the material adjacent to the structure, which would eventually cause the structure to collapse.
The location of cut-off walls for the various structures is shown in Table E.8.6.

Table E.8.6: Cut-off wall locations

Structure Locations
Drift Upstream and downstream sides of drift slab
Culvert Edges of inlet and outlet apron
Vented ford Upstream and downstream sides of main structure and approach ramps
Upstream and downstream sides of approach ramps
Large bore culvert The foundations of the main structure should be built at a greater depth than
standard cut-off walls; below the possible scour depth
The foundations of the main structure should be built at a greater depth than
Bridge
standard cut-off walls; below the possible scour depth

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 112

The absence of cut-off walls at the inlet of HEADWALL

the structure could allow water to seep APRON


CULVERT
under the apron and structure causing
settlement and eventually collapse of
the structure. At the downstream end WATERCOURSE BED

of the structure the flowing water could WATER STARTS TO UNDERCUT


erode the material next to the apron, UNPROTECTED APRON

eventually eroding under the apron


and causing it to collapse. The benefits
of a cut-off wall are illustrated in Figure
E.8.7.

The depth of the cut-off walls will


depend on the ground conditions.
Where a rock layer is close to the UNDERCUTTING CONTINUES
RESULTING IN BROKEN APRON
ground surface the cut-off walls should
be built down to this level. If there is no
firm stratum near the surface the cut-
off walls should extend the minimum
dimensions listed in the previous
section on scour. The method of
construction of the cut-off wall should
be similar to the construction method
and material used for the remaining
parts of the structure, to facilitate the
CONSTRUCTION OF A CUTOFF WALL
construction and reduce cost. PREVENTS DAMAGE

8.5 Pipes
Figure E.8.7: Cut-off wall
Pipes will be required for culverts and
vented fords. This section initially covers the vertical positioning of culverts, followed by the sizing of
pipes and then other design issues including types of culvert and construction options.

8.5.1 Vertical positioning of pipes

The vertical positioning of culverts requires particular attention. The consideration of the natural vertical
alignment of the watercourse must take precedence over the vertical alignment of the road. Neglect of
this factor has led to many culverts being installed incorrectly, leading to excessive silting, erosion and
in some cases failure. It should be remembered that the water forces during peak flow will be actively
promoting the return to the natural watercourse alignment.

There are three basic culvert installation situations. The most appropriate culvert type will depend on the
outfall gradient. See also the section on setting out in Chapter E.9.
ƒ Type A: Flat outfall (less than 5%)
This culvert type should be used in flat areas and for watercourses with shallow gradients. In these
cases the road should be built up over the culvert with ramps 20-50m long or to comply with
national road vertical alignment standards. A culvert will silt up if it is positioned too low to avoid
the requirements of building up the road alignment.
ƒ Type B: Intermediate outfall (approx. 5 - 10 %)
This arrangement requires the culvert to be excavated slightly into the existing ground, although
the invert of the culvert at the inlet should be at the same level as the bed of the watercourse.
The outlet of the culvert will be below the existing ground level and will require an outfall ditch to
be dug with a gradient of approximately 4%. The road will still have to be built up with ramps or
alignment adjustment over the culvert to provide the minimum required cover.
ƒ Type C: Steep outfall (more than 10%)
The culvert can be installed without building up the road level. The culvert should be buried to
provide adequate cover over the pipe. A drop inlet will be required at the entrance to the pipe (see

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 113

below) and a short outfall ditch at the exit. On steeply sloping ground careful attention should be
given to preventing erosion downstream of the culvert. Further information on erosion protection
is given in a later section in this chapter.

8.5.2 Pipe sizing

Because of changing climatic conditions, debris and bed load in channels, changing land use patterns,
and uncertainties in hydrologic estimates, culvert size and capacity should be conservative, and should
be oversized rather than undersized. Ideally, a culvert will be of a size as wide as the natural channel to
avoid channel constriction. Channel protection, riprap, headwalls, and trash racks can all help mitigate
culvert problems, but none are as good as an adequately sized and well placed pipe. An oversized
culvert, designed to avoid pipe repairs or failure as well as prevent environmental damage, can be very
cost-effective in the long run. Also, the addition of concrete or masonry headwalls helps reduce the
likelihood of pipe plugging and failure.

A number of methods are available to assess the required culvert pipe size(s). These are described in the
following sub-sections.

The most appropriate method for sizing pipes is to carry out a design based on one of the three cases
shown. However, this design process requires data on the culvert catchment area and predicted rainfall
intensity. In the absence of other data Figure E.8.8 following suggests the size and number of pipes that
are required to give a suitable culvert capacity for the recommended storm return period. Figure E.8.8 is
based on gentle/rolling ground with medium soil permeability.

Figure E.8.8: Culvert pipe requirements

The design process for sizing pipes will depend on the particular flow characteristics of the water through
the pipes. There are three cases which must be considered as shown in Figure E.8.9. Proceed with the
following steps for the design of the pipe.

Step 1: Peak flood flow


The first stage in culvert pipe design is to estimate the maximum expected peak flood flow, which was
discussed in Chapter E.6, Section 6.1.

Step 2: Check for case 3


If case 3 exists it will not be necessary to carry out any further work, as the culvert size is determined
by the requirements of minimum diameter for cleaning. Table E.8.7 shows the maximum flow rates for
assuming case 3 flow exists for a 600mm diameter culvert with an invert on different gradients. For case
3 to exist the flow at the downstream end of the culvert must be uninhibited. This will require the outfall
from the culvert to have the same or greater slope than the invert of the culvert.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 114

Table E.8.7: Maximum flow rates

Maximum flow rates for 600mm diameter case 3 culverts


Invert slope Max flow rate
1% 20 l/s
2% 40 l/s
3% 50 l/s
4% 60 l/s
5% 70 l/s

ƒ Case 1 CENTRELINE OF ROAD


Case 1 has water backed up on the upstream side
of the culvert, but the water is able to flow freely
d1
away from the down stream side of the culvert. This
situation is likely to occur on sloping ground where
the outfall continues down the hillside.
h
d2
ƒ Case 2
Case 2 has water backed up on both the upstream CENTRELINE OF ROAD

and downstream sides of the culvert. The flow of


water through the culvert is less than in case 1 d1
(for the same size culvert) as the water backed up
downstream reduces the flow. This situation will
exist in flat areas where the water in the culvert
h d2
outfall flows slowly or ponds in the channel.

ƒ Case 3 CENTRELINE OF ROAD

Case 3, with no water backed up at either end of


the culvert, will only occur for low flow rates and
where the water can flow away from the culvert in d1

the downstream channel. If flow rates are low but


the outfall slope is shallow the culvert is likely to
h d2
operate under case 2.
Figure E.8.9: Pipe design cases
Step 3: Pipe dimensions
In order to design the pipe it will be necessary to guess a pipe size and invert level and gradient. These
dimensions will be used for the flow calculations and then compared with the predicted peak flood flow.
Through experience the designer will be able to make a good initial guess at the size and/or number of
culvert pipes required. For designing a culvert a first guess should be taken as one 600mm pipe. A fall
of 3-5% should be placed in the invert to ensure that water flows through the culvert without depositing
silt and other debris

Regardless of the design water flow, all pipes should have a minimum diameter of 600mm to
ensure that they can be manually cleaned when clogged

Step 4: Maximum upstream depth


During flood periods storm water will back up in the upstream channel of the culvert. The amount of
back up will depend on the culvert characteristics. The amount of back up permitted should be chosen to
ensure that the water does not flood cultivated land and property or overtop the road embankment and
culvert headwall. The depth of water due to backing up is measured for the stream bed and is shown as
d1 in the Figure 8.12.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 115

Step 5: Determine downstream characteristics


It will also be necessary to determine if the water is likely to pond and back up at the downstream end of
the pipe (see plates E.8.8 and E.8.9). Ponding will depend on the slope of the channel.

Plate E.8.8: Ponding at culvert outfall

Plate E.8.9: Ponding in outfall channel

Step 6: Driving head


The driving head is the potential energy which causes the water to flow through the pipe.

where
H is the driving head
Driving head = H
d1 is the upstream water depth
d2 is the downstream water depth
H = d1 + h - d2
h is the drop in culvert invert level
as shown in the previous design cases above

It is the difference between the water levels each side of the culvert.

Step 7: Friction factor


The length and roughness of the pipe will affect the flow rate. The friction factor determined from the
graph (in Figure E.8.10) is an indication of the resistance to flow due to the pipe’s characteristics and is
required to calculate the maximum flow in the pipe.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 116

Figure E.8.10: Friction factor

Step 8: Check maximum flow rate


Once the friction factor and head are known the maximum flow rate through the pipe can be obtained
from the graphs in Figure E.8.11.

Figure E.8.11: Flow Rate

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 117

Step 9: Check inlet restriction


For higher flow rates the rate of water flow through the culvert will be restricted by the entrance diameter
of the culvert. Check the maximum flow rate for the culverts obtained from Figure E.8.12 and compare it
with the flow rate obtained from step 8.

Figure E.8.12: Restriction

Step 10: Check acceptable flow rate


The maximum flow rate obtained in either step 8 or 9 should be compared with the maximum predicted
flow rate.

Where the maximum flow rate is larger than the predicted flow rate, the culvert design is acceptable.
The next design stages for the culvert should be carried out; selecting appropriate inlet and outlet
arrangements and confirming the type of pipe based on the assumptions made in the design steps.

If the maximum flow rate is less than the predicted flow rate the design is unacceptable. If the culvert
were to be constructed in this design the flood water would overtop the road causing it to be washed
out, or it would flood adjacent fields and properties. The design process must be carried out again from
step 3 making one of the following changes:
ƒ Adding another pipe of the same diameter;
ƒ Increasing the size of the pipe.

Nomogram Method
Pipe size, as a function of anticipated design flow (capacity) and headwater depth, can be determined
using the Nomograms shown in E.8.13, E.8.14 and E.8.15. These figures apply to commonly used culverts
of round corrugated metal pipe, round concrete pipe, and concrete boxes. Each of these figures applies
to pipes with inlet control where there is no constraint on the downstream elevation of the water exiting
the structure. In these circumstances the culvert acts as an orifice and the capacity can be determined
in a relatively simple manner on the basis of headwater height and inlet geometry (barrel shape, cross-
sectional area and the inlet edge). Barrel slope affects the inlet control performance to a small degree
but may be neglected. Ideally, the inlet water elevation (headwater depth) should not greatly exceed the
height or diameter of the structure in order to prevent saturation of the fill and minimize the likelihood of
the pipe plugging from floating debris.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 118

Figure E.8.13: Headwater depth and capacity for corrugated metal pipe
culverts with inlet control. (Adapted from FHWA, 1998)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 119

Figure E.8.14: Headwater depth and capacity for concrete pipe


culverts with inlet control. (Adapted from FHWA, 1998)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 120

Figure E.8.15: Headwater depth and capacity for concrete box


culverts with inlet control. (Adapted from FHWA, 1998)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 121

Simplified Formulae Method


The use of formulas (Table E8.8) is preferred by some designers to the use of the Nomograms shown
previous. The formulas are based on the following assumptions:
ƒ Inlet Control
ƒ Wingwall Angle = 450
ƒ Vertical Headwall

Table E.8.8: Simplified Formulae for Calculation of Discharge Capacity

Discharge Capacity Q (m3/s)


Type (with inlet control)
Hw/D=1.00 Hw/D=1.25 Hw/D=1.50
Concrete Pipe 1.3 x D2.5 1.9 x D2.5 2.2 x D2.5
Corrugated Metal Pipe 1.1 x D2.5 1.6 x D2.5 1.8 x D2.5
Arch Culvert (semi-circular) 2.3 x H2.5 3.4 x H2.5 4.0 x H2.5
Box Culvert 1.5 x B x H1.5 2.1 x B x H1.5 2.5 x B x H1.5
D :diameter of a pipe culvert (m) B :width of a box culvert (m)
Hw :headwater height (m) H :height of a box/arch culvert (m)

Tables for the hydraulic design of pipes sewers and channels Volumes I & II, 7th edition, published by HR
Wallingford (UK), may also be used where different conditions exist, or greater accuracy is needed. More
detailed information can also be found in FHWA Manual HDS-5, Hydraulic Design of Highway Culverts,
1998).

8.5.3 Pipe options

There are many different options available to the designer for constructing culvert pipes. The pipes can
be either precast or constructed in situ, circular or square openings, reinforced or unreinforced and built
from a variety of materials. In deciding which type of culvert to construct the designer has to assess the
advantages and disadvantages of each construction option. Careful consideration must be given to the
skills and resources available, the cost of each option, the prevailing site conditions for the region and the
advantages of choosing a few standard designs for the majority of the culverts to be constructed.

Precast pipes
Precast pipes are usually manufactured in a central yard and are then transported to site. This method
of construction has the advantage that the quality control for the construction of the pipe is likely to be
improved, but the two main disadvantages are the increased transportation costs (as illustrated in Figure
E.8.16) in bringing the pipes to site and the careful transportation and handling required to ensure the
pipes are not damaged. Concrete pipes should preferably be transported on end, on a bed of sand, to
minimise the risk of damage. Particular care is required in laying and jointing the pipes to ensure good
support to the lower third of the pipe circumference.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 122

Figure E.8.16: Transportation issues

In situ construction
Pipes constructed in situ can be made from a variety of materials. Careful supervision will be required on
site to ensure that the pipes are manufactured to sufficient quality, but the transportation costs may be
reduced when compared with precast pipes if their transport distances are substantial (Figure E.8.16).

ƒ Masonry culverts (arch and box)


Masonry culverts are generally constructed as box culverts for small sizes and arch culverts for
larger sizes (Figure E.8.17). There are three stages to constructing a wall and slab box culvert:
• Excavation & construction of the base;
• Construction of the walls;
• Laying the roof slab and backfilling the culvert.

The culverts can be constructed with different top slabs depending on the size of the culvert.
These slabs may be masonry, timber or precast concrete. The advantages and disadvantages of
masonry culverts are shown in Table E.8.9.

ROAD SURFACE

MINIMUM COVER 500 mm

MASONRY
400 - 500 mm

ARCH
SPRINGING
BACKFILL

MAXIMUM 2 METRES

FOUNDATION
CONCRETE
OR MASONRY

MASONRY ARCH CULVERT

600 - 800 mm

BOTTOM AND SIDE


WALLS OF MASONRY
WORK.
COVER OF WATER
RESISTANT
HARDWOOD BEAMS
OR REINFORCED
CONCRETE
BEAMS/SLAB.

Figure E.8.17: Masonry Culverts

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 123

Table E.8.9: Masonry culverts

Advantages Disadvantages
ƒ The use of locally available material reduces ƒ Arched culverts require dressed stone bricks,
the cost of construction blocks or mortared jointing
ƒ Simplicity of construction
ƒ Low level of maintenance required
ƒ Range of options available for the top slab on
box culverts

ƒ Concrete arch or box culverts


These can be constructed using the same principles as masonry culverts. Spans larger than 800mm
will require reinforcement design and detailing.

ƒ Timber Culverts
Option 1: Timber barrel
Timber barrel culverts shown in Plate E.8.10 are typically manufactured from shaped, treated
wooden planks with tongue and groove joints, held in position by steel bands or wire. Once the
culvert is in place and backfilled the steel bands are no longer required as the ground material
holds the pieces of the culvert in position. The bands can therefore rust away after the culvert has
been placed without the culvert collapsing. The advantages and disadvantages of these types of
culverts are shown in Table E.8.10.

Plate E.8.10: Timber barrel culvert

Table E.8.10: Timber barrel culverts

Advantages Disadvantages
ƒ Can provide cheap culvert if timber widely ƒ Professional wood treatment facilities
available required
ƒ Culverts can be assembled at site allowing
larger numbers to be transported on a lorry
ƒ Design life is over 25 years with treated wood
ƒ They are light and easy to handle
ƒ Culverts can withstand small ground
movements and settlement without losing
their structural integrity
ƒ Short working life if wood is badly treated

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 124

Option 2: Timber log culverts


A simple and quick method for constructing small relief culverts can be to use timber logs (see
Figure E.8.18 and Plate E.8.11). These culverts will usually be unlined, bare earth and will only
accommodate slow flows (up to 1 m/s). Figure E.8.15 shows the key dimensional requirements for
these types of culvert. This type of construction should only be viewed as a temporary culvert unless
the timber is properly treated. It can be a useful construction method for emergency maintenance
during the rainy season.

Advantages and disadvantages of timber log culverts are shown in Table E.8.11.

MARKER POSTS
HEADWALLS OF
POLES OR ROCK

500 mm
MIN

FULL ROADWAY WIDTH

CROSS SECTION
N.T.S.

POLE DIA. 200 mm


(WHEN STRIPPED OF BARK)

500 mm MIN.

SEATING: 500 mm SEATING: 500 mm

1000 mm MAX.

Figure E.8.18: Timber log culvert details

Plate E.8.11: Timber pole culvert during reconstruction

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 125

Table E.8.11: Timber log culverts

Advantages Disadvantages
ƒ Very quick and cheap to construct ƒ Very short life, especially if timber is
ƒ Minimal skills required for construction untreated
ƒ Unlined ditch very susceptible to scour
during heavy rains

ƒ Cast in-situ concrete culverts


These culverts use a timber or steel mould (Plate E.8.12) to form the pipe of the culvert. A rubble
concrete mixture is used to form the foundation of the pipe. The mould is then placed in position
and lean mix concrete poured around the culvert mould. Once the concrete has set the mould
is collapsed and removed. An example of a cast in-situ culvert is shown in Plate E.8.13 and the
advantages and disadvantages of this type of culvert are given in Table E.8.12.

Plate E.8.12: Reusable steel mould Plate E.8.13: Cast in-situ culvert
for cast in-situ culvert

Table E.8.12: Cast in-situ concrete culverts

Advantages Disadvantages
ƒ Low cost as mould can be reused many times ƒ Poor life expectancy if rubble concrete is not
ƒ Quick construction methods well placed or compacted
ƒ Low cement requirements due to use of rubble
concrete

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 126

ƒ Precast unreinforced concrete culverts


These culverts are usually manufactured in a casting yard and brought to site in units. They need
to be manufactured under good quality control conditions to ensure that they have sufficient
strength. This option is only worth considering for high production numbers where a large number
of culverts will be constructed in the same area. Examples of the steel casting mould and the
finished product are shown in Plates E.8.14 and E.8.15. The advantages and disadvantages are
shown in Table E.8.13

Plate E.8.14: Steel mould Plate E.8.15: Precast concrete culvert ring

Table E.8.13: Precast unreinforced concrete culverts

Advantages Disadvantages
ƒ The quality of the pipe can be ensured ƒ High cost for small batches
ƒ Do not require steel reinforcement ƒ Careful transportation required to ensure
ƒ Very good performance when bedding and they are not damaged or broken
backfilling has been carried out well ƒ Not suitable if site access route is in bad
ƒ Pipes up to 900mm dia. can be manhandled condition
by labour alone ƒ High transport costs due to their shape
ƒ Economic where a large number of identical ƒ Diameters greater than 900mm dia. can
pipes are required not be made due to strength and handling
problems
ƒ Pipe lengths are restricted to 1m to ensure
that they can be handled by labour alone

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 127

ƒ Steel culverts
Steel culverts will usually be constructed from pre-bent corrugated sheets (Plate E.8.16) which
are bolted together on site. They can be very expensive if a steel manufacturing capability is not
available locally in country. Imported steel culverts consume scarce foreign exchange resources.
Their advantages and disadvantages are shown in Table E.8.14.

Plate E.8.16: Steel culvert

Table E.8.14: Steel culverts

Advantages Disadvantages
ƒ The steel culverts can withstand small ground ƒ Requires the transport and possible import of
movements expensive steel sheets
ƒ Light sections easy to handle and install ƒ Secure storage of the sheets required to
ƒ The components for a number of culverts can prevent theft
be transported on one truck

8.5.4 Pipe inlets

The general design of headwalls and wingwalls is discussed elsewhere in this Chapter. However, there are
two design cases of pipe inlets that require special attention:
ƒ Pipes on steep slopes;
ƒ Pipes which are transferring large volumes of storm water from a side drain to the other side of
the road.

Pipes on steep slopes


The invert of a pipe should be placed on a 2-5% slope to ensure
that the flow is not too great to cause extensive scour but fast
enough to prevent debris and silt from being deposited in the
culvert. If the culvert is located on steeply sloping ground overall
height drop across the culvert may need to be much steeper
than 5%. If this case occurs the culvert should be designed
for the maximum desirable invert slope (5%) and a drop inlet
proposed. The drop inlet reduces the energy of the water
leaving the culvert, preventing extensive scour. Drop inlets can
also be used for relief culverts on long downhill lengths of side
drain (see Figure E.8.19)
Figure E.8.19:
Drop inlet on relief culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 128

Pipes transferring large water volumes


One of the most important design rules when constructing a road water structure is to disrupt the flow as
little as possible. Unfortunately this will usually not be possible for a culvert that is transferring water from
a side drain under the road. The water must make an abrupt right angle change in direction to enter the
culvert. For large flows there will therefore be a large amount of turbulence in the water and the potential
for scour. Plate E.8.17 and Figure E.8.20 indicates the following key features in the inlet design for large
flows:
ƒ Rounded wingwalls to ‘guide’ water into pipe;
ƒ Sloping wingwall on inside radius;
ƒ Lined channel sides and base which extend 5m up the channel;
ƒ Cut-off wall provided at the edge of the inlet;
ƒ Consider box culvert option as this will cause less restriction and turbulence.

Plate E.8.17: Drop inlet on stream culvert Figure E.8.20: L-shaped inlet

Plate E.8.18: Skew culvert crossing with splay wingwall to prevent water continuing in side drain

8.5.5 Pipe bedding and cover arrangements

Culverts pipes should be constructed on a firm foundation to ensure that they will not settle and crack.
The support for the pipe should be either 250mm of compacted crushed stone, granular material (with a
maximum stone size of 30mm) or 150mm concrete slab.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 129

If the culvert is constructed from precast units it will be necessary to place a bedding material, such
as sand, on the foundation to remove any irregularities and ensure an even support to the base of the
precast units. If the preferred design option is a masonry culvert the foundation for the walls can be
extended to form the base of the culvert (see Figure E.8.21).

Backfilling around the culvert is one of the most important stages in the construction. The quality of the
backfilling will determine the strength of a culvert to resist vehicle loads above it. The designer should
specify the material to be used to backfill around the culvert, which should be easy to compact and well
graded to promote drainage. Stones larger than 30mm should not be included in the backfill as they may
damage the culvert. The excavated material from the culvert construction may be used for backfilling if
it meets these criteria.

1.5 x OD

SELECTED FILL

120°

FIRM FOUNDATION SAND/GRAVEL LAYER


UNDER BASE OF CULVERT

NOTE:
C = COVER TO FINISHED ROAD LEVEL
= D = INTERNAL DIAMETER OF CULVERT (DESIRABLE)
= 3D
4 ABSOLUTE MINIMUM
OD = OUTSIDE DIAMETER OF CULVERT

Figure E.8.21: Pipe granular bedding and cover

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 130

Plate E.8.19: Poor quality jointing and bedding support for pipe culvert
As material is backfilled around the culvert it should be well compacted in layers of 150mm. Particular
care should be taken for the lower half of the pipe to ensure:
ƒ The material under the pipe is compacted with hand rammers;
ƒ Hand rammers do not damage the culvert;
ƒ The pipe is held at the correct level and does not ‘rise’;
ƒ Each side of the culvert is backfilled at the same rate to ensure that the culvert is not pushed out
of line.

The minimum desirable cover from the top of a culvert to the road surface should be the same as the
diameter of the culvert. If the conditions do not permit this depth of cover it may be reduced to 75% of
the pipe diameter.

The cover can be reduced to half the culvert’s diameter if the concrete bed, haunch and surround are cast
as shown in Figure E.8.22. The remaining cover should be good quality standard fill material and the road
should be surfaced with gravel or other material as appropriate.

SELECTED FILL

1.5 x OD

D/4 C

D/4

CLASS 15 CONCRETE
SURROUND

120°

FIRM FOUNDATION LEAN CONCRETE BED


AND HAUNCH

NOTE:
C = COVER TO FINISHED ROAD LEVEL
=D
2 MINIMUM
D = INTERNAL DIAMETER OF CULVERT
OD = OUTSIDE DIAMETER OF CULVERT

Figure E.8.22: Pipe arrangement with minimum cover

8.5.6 Multiple culverts and vented fords

The design principles for multiple culverts and vented fords are the same as single bore culverts. Where
more than one pipe is to be installed the minimum space between the centre line of adjacent pipes
should be at least 2 pipe diameters. Where space restrictions require the installation of pipes at closer
spacing the factors in Table E.8.15 should be used to reduce the flow rates through the pipes derived
previously in this chapter.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 131

Table E.8.15: Pipe spacing and flow reduction factors

Spacing between pipe centres Flow reduction factor


More than 2.0 pipe diameters 1.0
1.5 - 2.0 pipe diameters 0.9
Less than 1.5 pipe diameters Due to difficulties in ensuring adequate
compaction under and between pipes, bedding
of lean concrete should be used in these
circumstances.

The flow capacity of different culvert shapes and diameters should be checked according to the
characteristics of the site. The number and size of pipes should then be chosen to ensure that the sum of
all the individual pipe flows is greater than the design flow.

The design flow for a multi-bore culvert should be taken to be the maximum flood flow. As vented fords
are designed to be overtopped during peak flows the pipes should be designed to pass the normal
flow and small floods. Overtopping will only occur for the higher flow rates and the designer will have
to decide what level of flow the pipes will pass before overtopping occurs. The overtopping flow will
depend on the duration, size and regularity of high flows and the total number of pipes that can be fitted
into the structure.

Box Culverts - The design of box culvert options is not covered by this Manual. Refer to the ERA Bridge
Design Manual - 2011 and publications such as TRL Overseas Road Note 9.

8.6 Headwalls and Wingwalls

8.6.1 Culverts

Headwalls and small wingwalls are required at each end of a culvert and serve a number of different
purposes:
ƒ They direct the water in or out of the culvert;
ƒ They retain the soil around the culvert openings;
ƒ They prevent erosion near the culvert and seepage around the pipe which causes settlement.

The headwall can be positioned at different places in the road verge or embankment as shown in Figure
E.8.20.

ROAD CENTRELINE

CULVERT
HEADWALL
OPTIONS

a) b)
CULVERT

a) LONGER CULVERT, LOW HEADWALL COSTS


b) SHORTER CULVERT, HIGH HEADWALL COSTS

Figure E.8.23: Possible culvert headwall positions

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 132

The closer the headwall is placed to the road on an embankment the larger and more expensive it will
be. The most economical solution for headwall design will be to make it as small as possible. Although
a small headwall will require a longer culvert, the overall structure cost will normally be smaller. If, due to
special circumstances at a proposed culvert site, a large headwall with wingwalls is required it should be
designed as a bridge wingwall.

Where a road is not on an embankment the size of the headwall will be small regardless of position.
In this case the position of the headwalls will be determined by the road width and any requirements
of national standards. The headwalls should be positioned at least 1 metre beyond the edge of the
carriageway width to prevent a restriction in the road and reduce the possibility of vehicle collisions (see
Figure E.8.24).

HEADWALL
PARALLEL TO
ROAD

1 METRE MINIMUM

CARRIAGEWAY
WIDTH
1 METRE MINIMUM

Figure E.8.24: Position of culvert headwalls

Headwalls should project above the road surface by 300mm and be painted white so that they are visible
to drivers. There are a number of different layout options for culvert headwalls which are shown in Figure
E.8.25.

Headwall with drop inlet: This arrangement should be used when the road is on a steep side slope to
reduce the invert slope of the culvert (see Section E.8.5.4).

Headwall with L inlet: This arrangement should be used where the road is on a gradient and water is
to be transferred from the carriageway side drain on the high side of the road (see previous pipe inlets
section).

Headwall and adjacent works must be designed so that the culverts can be de-silted manually under
maintenance arrangements. This can be difficult with a drop inlet arrangement. Refer to chapter 10 on
maintenance.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 133

SIMPLE HEADWALL AND APRON HEADWALL WITH SPLAY WINGWALLS PARALLEL WINGWALLS

This arrangement can This arrangement should Where the road is on an embankment it
be used for simple low be used for larger flows is essential that wingwalls are provided
flow relief culverts. It is instead of the simple to prevent the risk of water seepage and
the cheapest option but arrangement. The subsidence of the embankment. This
prone to erosion for larger wingwalls increase the arrangement is likely to be the most
flows. protection from erosion. economical.

Figure E.8.25: Headwall and wingwall arrangements

8.6.2 Wingwalls - Larger structures

Wingwalls are used to retain the soil behind the abutments of bridges to help guide flows through the
structure in flood conditions and safely retain the backfill material without risk of erosion. There are two
basic reference layouts for wingwalls, either parallel to the road or parallel to the watercourse (see Table
8.16). However, wingwalls are usually constructed at an angle between these two arrangements. Wingwalls
should always be constructed to the toe (bottom) of the slope and not Part way down. Wingwalls that do
not extend to the bottom of the slope are likely to suffer from erosion around the ends.

Table E.8.16: Wingwalls

Wingwalls parallel to water course Wingwalls parallel to road


Foundations can be stepped but are harder to
Foundations on same level
construct
Wall more susceptible to erosion from
Wall mostly away from watercourse
watercourse
Wall size smaller than wall parallel to road Wall size longer than wall parallel to watercourse
Larger amount of fill to be moved, placed and Reduced amount of fill required to be moved,
compacted placed and compacted

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 134

The relative availability and cost of fill material and raw materials to construct the wingwalls will determine
the most appropriate arrangement. In general, to ensure the cheapest option, the design should ensure
the smallest wingwalls are chosen for the structure and its particular location. Where wingwalls are chosen
that run parallel to the road it is necessary to take suitable measures to prevent water in the carriageway
side drains causing erosion around the wall at their outfall. This usually requires a lined channel or cascade
at the base of the wingwall (Plate E.8.20). The two main factors affecting the overall design of a wingwall
are the construction material and the bearing capacity of the soil.

Plate E.8.20: Wingwall cascade

8.6.3 Stone, brick and blockwork walls

Stone, brick and blockwork walls should be built with a tapering back face to withstand the pressure
exerted by the fill material (see Figure E.8.26). The size of the wall will depend on its height, the bearing
capacity of the soil and if there is any surcharge (additional fill material above the wall). Any material used
in the wall should meet the requirements given in Chapter E.7.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 135

2
500 500
1

H
H
500 MIN. 500 MIN.

L L

WITHOUT SURCHARGE WITH SURCHARGE

Figure E.8.26: Stone, brick or blockwork wall with and without sloping backfill (surcharge)

Tables E.8.17 and E.8.18 provide a guide to the height of the wingwall with and without surcharge related
to the bearing capacity of the soil and the width of the base.

Table E.8.17: Height of wingwall without surcharge

Bearing capacity of the soil


H- Height of wingwall Medium
Low (75-125kPa) High (>250kPa)
(without surcharge) (125-250kPa)
L - Width of the base (mm)
1000 500 500 500
1500 900 800 800
2000 1700 1150 1150
2500 1450 1450
*Construction
3000 1750 1750
not possible
3500 without 2400 2000
ground
4000 3200 2300
improvement
4500 4200 2600
Notes:
* Ground improvement increases the bearing capacity of the soil through the addition of other materials to the ground eg
gravel or cement – this is outside the scope of this Manual.

Where wingwalls are constructed on medium or high bearing capacity soil, parallel to the road, and are only used to retain
road fill material to a height of up to 3 metres the wall may be constructed as follows:
1. Top of the wall to be 500mm wide
2. Vertical front face and 1:4 sloping back face (1 horizontal: 4 vertical)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 136

Table E.8.18: Height of wingwall with surcharge

Bearing capacity of the soil


H- Height of wingwall (with Medium
Low (75-125kPa) High (>250kPa)
surcharge) (125-250kPa)
L- Width of the base (mm)
1000 1000 950 950
1500 1500 1200 1200
2000 2000 1450 1450
2500 1750 1750
3000 *Construction not 2350 2000
possible without
3500 ground 3200 2250
4000 improvement 2550
Notes:
* Ground improvement increases the bearing capacity of the soil through the addition of other materials to the ground eg
gravel or cement – this is outside the scope of this Manual.

Where wingwalls are constructed on medium or high bearing capacity soil, parallel to the road, and are only used to retain
road fill material to a height of up to 3 metres the wall may be constructed as follows:
1. Top of the wall to be 500mm wide
2. Vertical front face and 1:4 sloping back face (1 horizontal: 4 vertical)

8.6.4 Gabion baskets

Gabion baskets may be used in areas where stones are available (Figure E.8.27).

In some areas there may be a problem of persons removing wire from the gabion baskets for other
construction purposes. If consultations through community groups cannot resolve this problem then
more robust steel mesh gabions may need to be considered.

WITH OR WITHOUT SURCHARGE 2


1
H: TOTAL WALL HEIGHT
L: DEPTH OF WALL AT HEIGHT n
FROM BASE
n: HEIGHT OF WALL MEASURED
FROM BASE

6
H

Figure E.8.27: Gabion baskets for walls

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 137

Table E.8.19 assumes that the gabion baskets have been filled according to the criteria outlined in
Chapter E.7 and have a height and width of 1 metre.

Table E.8.19: Height and width of gabion walls

Width of gabion wall at height ‘n’ above base


Height of
Bearing capacity
wall n (m)-
of soil
(m)
0 0.5 1 1.5 2 2.5 3 3.5 4
1.5 1 1 1 1
2 1 1 1 1 1
50 - 125
2.5 1.5 1.5 1.5 1 1 1
kPa
3 1.5 1.5 1.5 1.5 1 1 1
3.5 2 2 2 1.5 1.5 1 1 1
1.5 1 1 1 1
2 1 1 1 1 1

>125 2.5 1 1 1 1 1 1
kPa 3 1.5 1.5 1 1 1 1 1
3.5 1.5 1.5 1.5 1 1 1 1 1
4 2 2 1.5 1.5 1.5 1.5 1 1 1

8.6.5 Timber walls

Felled timber tree trunks as described in Section E.7.3 can be used to form a wingwall.

8.7 Aprons
An apron is required at the inlet and outlet of culverts and downstream of drifts and vented fords to prevent
erosion. As the water flows out of or off a structure it will tend to erode the watercourse downstream,
causing undercutting of the structure. Refer to the section on cut-off walls earlier in this Chapter. Aprons
should be constructed from a material which is less susceptible to erosion than the natural material in the
stream bed.

8.7.1 Drift aprons

Where the discharge velocity across the drift is less than 1.2m/s which may be experienced for relief drifts,
a coarse gravel layer (10mm) will provide sufficient protection down stream of the drift. For discharge
velocities greater than 1.2m/s more substantial protection will be required which utilises larger stones.
This is discussed in the section on downstream protection. The width of the apron should be at least half
the width of the drift and extend across the watercourse for the whole length of the drift.

8.7.2 Culvert aprons

Aprons should be provided at both the inlet and outlet of culverts (see Figure E.8.28). They should extend
the full width between the headwall and any wingwalls. If the culvert does not have wingwalls the apron
should be twice the width of the culvert pipe diameter. The apron should also extend a minimum of 1.5
times the culvert diameter beyond the end of the pipe. Cut-off walls should also be provided at the edge
of all apron slabs. The choice of apron construction is likely to depend on the type of material used for
construction of the culvert. It may be constructed from gabion baskets, cemented masonry or concrete.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 138

Dim A – Distance between wingwalls or 2 x culvert


dia. for culvert without wingwalls,
Dim B – 2x culvert dia.

Figure E.8.28: Culvert apron

8.7.3 Vented ford aprons

The apron for vented fords should extend the whole length of the structure including downstream of
the approach ramps to the maximum design level flood. The other design requirements for vented ford
aprons are the same as culvert aprons.

8.8 Approach ramps


The approaches to vented drifts, large bore culverts and bridges must allow vehicles to cross the structure
without loosing traction or getting stuck on the crossing. Ideally crossings should not have approaches
steeper than 10%. However, steeper approaches can be provided if governed by the local terrain.
Approaches steeper than 10% will require the running surface to have a thin concrete or cement bound
masonry slab to allow vehicles to maintain traction particularly during wet periods. The slab should be at
least 150mm thick and be constructed on a sand or compacted masonry/aggregate base.

The approach way is subjected to similar erosion characteristics as the main structure. It is therefore
necessary to surface the approach ways with the same material as the main structure, at least to the
height of the maximum flood level, to ensure damage does not occur. If the structure is designed to be
overtopped the approach ways must be constructed higher than the maximum flood level to ensure that
the water does not erode around the ends of the structure leaving it inaccessible.

It is also necessary to provide cut-off walls (Section E.8.4) along the sides of the approach ways to protect
against scour. The sides of the approach ways should be faced to ensure erosion does not occur. They
may be constructed from:
ƒ Masonry walls (most appropriate for higher walls);
ƒ Gabion baskets;
ƒ Concrete walls (for low walls up to 0.5 metre);
ƒ Timber logs (high maintenance required).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 139

The design of these walls will be similar to the design of wingwalls described in Section E.8.6. The fill
material in the approach way should be chosen from one of the three options shown in table E.8.20.

Table E.8.20: Fill material in the approach way

Well compacted sand


Rubble masonry Lean concrete mix with plums
and gravel
Sand and gravel may be readily If a well graded mix of sand A concrete mix of 1:4:8
available in the watercourse and gravel is not available it (cement, sand and aggregate)
around the crossing site. These may be more economic to use can be used with large plums
may be stockpiled during the rubble masonry rather than up to 200mm in size. This
initial stages of construction by breaking rocks to create a option will have the highest
labour. The material to be used well graded material. Broken cement requirement, and hence
as a fill should be well graded man-made bricks can be used cost. However, it may be the
and placed in 100mm layers in addition to, or instead of, most beneficial fill option if
which are well compacted natural stone provided they there are small quantities of
before subsequent layers are meet the requirements outlined sand, aggregate and large
placed. in chapter 7. Rubble masonry stone near the bridge site.
should be bound together with
a 1:8 cement-sand mortar.

The running surface of the approach way should be designed as a structural slab of either concrete or
cement bonded stone paving. The slab should also have a 2-3% crossfall in the direction of water flow to
ensure that the deck drains quickly after rainfall.

Approach ways will be susceptible to scour from water flowing from the carriageway side drains into the
water course due to the increased slope. A lined channel should therefore be provided at the edge of the
approach way to ensure that erosion does not occur. The approach ways should be constructed see Figure
E.8.29 separately from the main structure to allow for thermal expansion of the structure and slight ground
movements, particularly for the structural slab. If they were constructed integrally with the main structure
any slight settlement or thermal effects could cause cracks in the structure which would weaken it against
damage from water. The approach ways therefore require an end wall and cut-off wall next to the main
structure. The gap between the two structures should be very small (no greater than 10mm). The edges of
the approach ways should be marked

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 140

Figure E.8.29: Construction of approach ways

Figure E.8.30 shows an example of an approach way cross section with guide/kerb stones to show drivers
the location of the edge of the carriageway. Plate E.8.21 shows a masonry side drains at the edge of an
approach way to prevent erosion.

Figure E.8.30: Approach way cross section

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 141

Plate E.8.21: Masonry side drains at the edge of an approach way

8.9 Downstream Protection


A previous section on scour indicated that it is likely that
erosion of the watercourse will occur around the structure
due to a constriction of the water flow. The constriction
causes the water velocity to increase as it passes through/
over the structure and this high velocity can be maintained
well downstream of the structure. A previous section also
discussed the use of aprons downstream of a structure
to prevent erosion and undercutting of the structure
itself. However, in small constrained channels severe
erosion may still occur after the apron, particularly
where the watercourse is on a gradient. It is therefore
often necessary to provide additional protection to the
watercourse, to reduce the velocity of the water and
prevent erosion.

Plate E.8.22 shows a gully that has been formed due to


water eroding soft material downstream of a culvert as
the watercourse was unprotected. For slow flowing water
it is unlikely that any protection will be needed, but for
faster flowing water the maximum allowable velocity will
depend on the bed material and the amount of silt or
other material already being carried in the water. Other Plate E.8.22: Example of damage due
examples of erosion through lack of protection are shown to lack of downstream protection
in Plates E.8.23 to E.8.26

Erosion can occur in any channel regardless of the presence of any structure. It is therefore not possible to
state how far downstream of a structure channel protection should extend. However, the following issues
should be taken into account:
ƒ The general erodability of the bed, which will be based on the type of channel material and the
gradient;
ƒ The likelihood of damage to the structure if erosion occurs downstream;
ƒ The potential effects of erosion on downstream areas (eg. damage to buildings or farming land).

Maximum water flow velocities that can be tolerated without channel protection related to the type of
bed material are shown in Table E.8.21.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 142

Plate E.8.23: Minor erosion in watercourse upstream from culvert

Plate E.8.24: Serious erosion downstream from the same culvert due to
concentration of flow and lack of appropriate protection measures

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 143

Plate E.8.25: Severe erosion downstream of a relief culvert


due to inadequate protection measures

Plate E.8.26: Severe erosion downstream of a cross


culvert due to inadequate protection measures

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 144

Table E.8.21: Maximum water velocities

Maximum water velocities without


Bed material channel protection
Clear water Water carrying silt
Stiff clay 1 1.5
Volcanic ash 0.7 1
Silty soil / sandy clay 0.6 0.9
Fine sand / coarse silt 0.4 0.7
Sandy soil 0.5 0.7
Firm soil / coarse sand 0.7 1
Graded sand and gravel 1.2 1.5
Firm soil with silt and gravel 1 1.5
Gravel (5mm) 1.1 1.2
Gravel (10mm) 1.2 1.5
Course gravel (25mm) 1.5 1.9
Cobbles (50mm) 2 2.4
Cobbles (100mm) 3 3.5
Well established grass in good soil 1.8 2.4
Grass with exposed soil 1 1.8

There are many methods for providing protection to the watercourse. The choice of method will depend
on the availability or cost of different materials, the size of the watercourse and level of protection required.

8.9.1 Rip-rap

Rip-rap is the name given to stones placed in the river bed to resist erosion. In order to be effective the
stones used should be large or heavy enough that they will not be washed away during floods. Although
rip-rap may appear to consist of random rocks it should be well graded and placed as tightly as possible
to improve its resistance to erosion. The rocks used should also be strong and not likely to crumble.
Angular rocks, in general, have the best performance, due to the interlock that is formed between rocks.
Round rocks can be used if they are not to be placed on the sides of the watercourse which have a
gradient steeper than 1:4. Flat slab stones should also be avoided as they can be easily dislodged by the
water flow. Table E.8.22 shows the sizes of stone that should be used for rip-rap. It should be possible
for one or two labourers to place the majority of the stones with the few remaining larger stones being
placed by a small labour gang.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 145

Table E.8.22: Stone sizes for rip-rap bed protection

Minimum % of rock Thickness


Rock mass
Water velocity m/s Rock size dia. m meeting specified of rip-rap
kg
dimensions m
0.40 100 0%
Less than 2.5 0.30 35 50 % 0.5
0.15 3 90 %
0.55 250 0%
2.5 - 3 0.40 100 50 % 0.75
0.20 10 90 %
0.90 500 0%
3-4 0.70 250 50 % 1.0
0.40 35 90 %

8.9.2 Masonry slabs

In areas where outlets from culverts are on a steep slope it may not be possible to place rip-rap as it will be
washed down the slope. Masonry slabs, cascades or channels may be constructed on the steep section of
the outfall to control erosion (Figure E.8.31). As the water velocity will be high it will be necessary to use
mortar in the slab as hand pitched stones are likely to be washed out. It will not be necessary to make the
slab smooth as a rough slab will help to reduce the energy in the water. Large stones may be fixed in the
slab which project above the standard level to create more turbulence to slow the water speed. Masonry
cascades or step structures can incorporate a series of ‘ponds’ or sumps to help dissipate energy.

Figure E.8.31: Energy dissipating apron

In flatter areas, up to a 5% gradient, it should be possible for small watercourses to use hand pitched
masonry, providing it is well placed with any large flat stones bedded on their edges.

8.9.3 Gabions

Gabions can be used to protect the bottom or banks of a watercourse (see Figure E.8.31). As the stones
are confined by the wire cages much smaller stones than those used for rip-rap can be put in the cages.
The disadvantage of gabions is that they have the additional cost of the wire for the cages when compared

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 146

with rip-rap. However, the ability of single labourers to move and place the stones may outweigh the cost
of the wire. As gabions can be made in different sizes they can be used for a wide range of different
shaped watercourses. They can also withstand limited ground movements and therefore accommodate
any small changes in the river bed. If the bottom of the watercourse requires protection it will be possible
to make a gabion that is only 200 or 500mm thick to form a mattress over the watercourse bed. Figure
E.8.32 shows two methods for using gabions and mattresses for protecting the watercourse.

Figure E.8.32: Gabion protection on steep banks

Figure E.8.33 Illustrates how gabion can be tied together to form a protective mattress on slopes less
than 1:2. The size of the gabions will depend on the velocity of the water flow. For all flow velocities the
smallest gabion used is 0.5 x 0.5 x 1m.

Figure E.8.33: Gabion protection on shallow banks

Any mattresses in the bottom of the water course should be 200-300mm thick for water velocities up to
3m/s and 500mm thick for velocities over 3m/s. It is very important that they are securely wired together
to ensure that they do not slide down the bank and cause the water to erode the watercourse banks
behind them.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 147

The minimum size of the gabion baskets makes this option suitable only for larger watercourses. Plate
E.8.27 shows an example of gabions being used for erosion control downstream of a culvert in a seasonal
watercourse.

Plate E.8.27: Gabion erosion control structure downstream of a culvert

8.9.4 Vegetation

Vegetation is likely to be the best option for small watercourses as once established it slows down the
speed of the water flow and holds erodible soil together. It can also be a cost effective protection method
where suitable local plants are available. The use of vegetation to control erosion is sometimes called
bio-engineering. Bio-engineering covers a wide range of techniques that use vegetation, which include
the control of erosion and stabilisation of engineering structures. This Manual discusses the use of bio-
engineering to control erosion downstream of water crossings. It is not sufficient to randomly plant any
vegetation, as the conditions must be correct for the plants to grow and they must produce the desired
anti-erosion effect.

The most basic form of vegetation erosion control will be to allow the region’s natural grasses to grow
in the water channel. They may grow naturally without any assistance if they are already well established
in the channel. However, if some erosion has occurred in the channel it may not be possible for the
grass to establish itself without assistance. In these cases it will be necessary to cultivate the grass in a
nursery or near the site at the road side if it will not be damaged by vehicles or cattle. Once the grass is
established it can then be transplanted into the water channel. The replanting may be by individual plants
or by turfing techniques. Natural fibre matting may also help to establish plant growth. The timing of the
planting will be dependent on the rainy season. Plants need to get established in the watercourse while
there is moisture in the soil. It may be necessary to regularly water the plants until they are established in
their final situation. However, they are not able to grow during periods when the channel is full of water.
It is unlikely that the grass will grow in the base of the watercourse if water is flowing throughout the year.
In these cases it may be possible to plant the grass on the edges of the channel and an aquatic plant in
the base of the channel. The choice of plant will again be based on local knowledge, but it is likely that
plants found in other watercourses with similar conditions nearby will be the most appropriate. The local
agricultural or botanical institutions should be able to provide guidance on plant selection.

In areas where hand pitched stone is proposed to protect the channel downstream from a culvert it may
be reinforced with plants rather than cement or mortar, to bind the stones together.

Stones should be placed in the river bed in the same manner as for standard hand pitched stone slabs. Any
small gaps that remain between the stones should then be filled with soil and grass planted approximately
150mm apart. The exact distance will depend on the shapes and gaps between the stones. When the
grass is planted the workers should ensure that the roots are deep enough to enter the soil beneath the

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 148

stone pitching. In channels with a permanent water flow the grass should only be planted towards the
sides of the channel, as it will be unable to grow under water in the centre of the channel.

In Ethiopia, Wereda Agriculture Offices have adopted a wide scale use of bio-engineering to stabilise
slopes and prevent erosion. In these countries nurseries have been set up in each region to cultivate and
grow special grasses that are particularly good at resisting erosion. These nurseries are usually managed
by Wereda Agriculture Offices and in some areas by NGO’s and supply grasses and other plants to work
sites in the area. Vetiver grass is the most commonly used as it can grow in a wide variety of soil conditions
including those of very poor quality. It also develops a fibrous and deep root system which is ideal for
holding weak soil together and preventing erosion. Vetiver grass has successfully been used to prevent
erosion on steep roadside banks and at the edges of engineering structures. The cultivated grass shoots
are planted out in the area prone to erosion. The spacing of each shoot will depend on the perceived
erosion risk and will vary between 100mm for high erosion areas and 200mm for lower risk areas.

8.9.5 Steep channels

In areas where water is flowing down steep hillsides and crossing a road through a culvert, it is necessary
to provide protection to the slope above and below the road. This is particularly important when a road
is winding up a hill and a watercourse crosses the road a number of times, where it is not possible to
channel all the water down steep inclines at the hairpins. Water flowing downhill has a large amount of
energy which must be ‘lost’ if erosion is to be prevented. The most appropriate method in these cases is
to construct a step waterfall or cascade to dissipate the energy (see Figure E.8.34 and Plate E8.28).

The photograph and diagram show a step waterfall made from gabion baskets, but it would also be
possible to construct the structure from masonry if available. Regardless of the material chosen the
structure should be built into the hillside by excavating the necessary material. Care must be taken to
ensure that the sides of the channel extend outwards far enough to ensure that the water is contained in
the channel.

Plate E.8.28: Gabion basket step waterfall


Figure E.8.34: Gabion
basket step waterfall

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 149

8.9.6 Drain protection

Along each side of a road there should be a drain to assist in removing water from the carriageway
and transferring it into the nearest watercourse. In flat terrain these drains can be earth or gravel lined
however, where gradients are greater than 2% they will require protection to prevent fast flowing water
eroding the ditch. The most effective method of preventing erosion is to use scour checks, which are
mini dams constructed in the drain. These scour checks form barriers to the water flow, causing silt to
be collected behind each scour check and hence forming a series of steps in the drain which help to
dissipate the water energy.

Further guidance is provided in Chapter D.5 of the Manual.

8.10 Arches
It is often difficult to define the difference between large
bore culverts and arch bridges (Plate E.8.29). Regardless
of the name given to the structure, it will normally only be
required where a road crosses a well defined watercourse
and/or large flows are expected. This Manual defines a
large bore culvert as a structure with arches up to 2.5 metre
diameter. There are two design issues to be resolved if this
type of structure is to be constructed.

ƒ Some form of permanent wall will be required on the


upstream and downstream sides of the structure and
on the base of the archway to retain the enclosed fill.
ƒ A large amount of fill material will be required to
complete the construction.

An example of a semi-circular arch bridge is shown in Plate Plate E.8.29: Masonry Arch
E.8.30

Plate E.8.30: Semi-circular Arch bridge

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 150

8.10.1 Arch shape

An arch resists the dead weight and traffic loads by compressive forces in the arch ring. This results in
very large forces at each end of the arch which must be resisted by the foundations. If the arch is not
semicircular these forces will have a horizontal component which is harder for the foundations to resist
than vertical forces alone. It is therefore recommended that only semicircular arches are used unless
specialist engineering support is available for the design. The magnitude of the forces at the end of the
semi circular arch shown in Figure E.8.35 will be equal to half the total weight of the arch and fill material,
plus the weight of any traffic. The design of semi-circular arches should allow for an element of horizontal
loading particularly during construction and placing of fill material. As the arch load will be concentrated
in the foundations at each end of the arch these structures should only be built on ground which has a
good bearing capacity.

SEMI-CIRCULAR ARCH PARABOLIC-CIRCULAR ARCH

THRUST FROM ARCH IS VERTICAL THRUST FORM ARCH HAS A


SIMPLE BEARING SOLUTION HORIZONTAL COMPONENT
COMPLEX FOUNDATION

Figure E.8.35: Arch forces

8.10.2 Formwork reuse

Depending on the type of materials used to build the arch, formwork may be required during construction.
Temporary formwork can be very expensive when compared with the cost of the construction materials.
Where possible it should therefore be designed to be reused on future bridges in order to reduce the
overall cost and unnecessary resource use.

8.10.3 Bridge/culvert layout

Once the designer has chosen to construct an arch bridge/culvert he will have to decide on the size of the
arch or arches for the structure. The choice will depend on the particular characteristics of each potential
site but Table E.8.23 highlights the different options. If the designer wishes to use piers then reference
should be made to a later section in this chapter which discusses the design of piers.

Table E.8.23: Small Versus Large arches?

Small arches Large arches


Easier to construct using labour based Formwork may require cranes to manoeuvre
techniques components into place
Piers will be required to be constructed in the It may be possible to span the whole watercourse
water course in wider rivers as multiple spans with one arch and avoid the need for piers in the
may be required watercourse (reducing scour problems)
The load exerted by a large arch will require
The bearing pressures exerted by the piers will
ground conditions that can withstand very high
be lower than for large arches
bearing pressures

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 151

8.10.4 Construction sequence

The first stage of building an arch structure is to construct the foundations and any piers that may be
required. The arch formwork can then be put in place and the arch constructed. The side wall construction
should only commence once the ring is fully completed. The placing of fill material above the arch can
proceed as the side walls are built. The placing of fill in layers about 1m below the constructed fill height
would serve as a platform for the artisans who are laying the stonework for the side walls. Guide stones
should be included on each side of the deck to mark the edge of the carriageway. These could be integral
with the side walls or be formed with the deck surface. The options for the design of the deck surface will
be the same as for the approach ways discussed previously.

8.10.5 Arch materials

There are a number of different material options


available for the construction of walls and temporary
or permanent shutters for an arched bridge. Some
of these options can be used in both the walls and
arch, while others are only suitable for forming the
arch (see Figure E.8.36).

Stone, bricks and blockwork can be used to form the


walls of the structure. The choice of material should
be made based on the cost and availability of each
material. Any material that is used should conform
to the specifications given in Chapter E.7. If Part of
the wall is in the water flow the material should be
Figure E.8.36: Use of tyre in formwork
hard enough to resist erosion. The walls should be
constructed with a tapered back face, similar to the
characteristics of wingwalls discussed in a previous section.

Stone, bricks or blocks can also be used to construct the arch of the structure. Some form of temporary
framework will be required during construction. This temporary formwork is likely to cost as much as the
stonework used in the bridge itself. This
option is therefore only likely to be viable if
the formwork will be reused for additional
spans or on other structures. The most
appropriate formwork will usually be a
wooden frame covered in wooden planks or
sheets, although large truck tyres may be
used to hold timber sheets in place for smaller
arches. Reusable steel formwork may also be
used, especially if a large number of culverts
of the same diameter are to be constructed.
Once constructed the arch gets its strength
from its uniform shape with all components in
compression on the arch face. It is therefore
important that the formwork used is good Figure E.8.37: Two course arch
quality and rigid, to ensure that the arch does
not deform during construction.

All stonework used in an arch should be placed as shown in Figure E.8.37. The arch should consist of a
minimum of 2 courses of masonry which should be interlocking where possible. In addition the minimum
thickness of a semi circular arch ring is shown in Table E.8.24 below.

It is not possible to get the level of interleave shown in Figure E.8.34 if using bricks. The strength of brick
arches can only be ensured if a good bond is achieved between the brick and mortar. As the arch will

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 152

be very strong and rigid once it has been completed there should be a simple method for releasing the
formwork without damage in order that it can be used again.

Table E.8.24: Minimum arch ring thickness

Arch span (m) 1 2 3 4 5 6


Ring thickness (m) 0.2 0.3 0.35 0.40 0.45 0.5

An alternative to stone or brickwork for


the construction of the arch is to use
corrugated metal sheets (Figure E.8.38).
The advantage of these sheets is that
they act as permanent formwork to be left
in place, becoming Part of the finished
structure, and preventing the need to use
expensive temporary formwork. Although
corrugated metal sheets are likely to have
a higher purchase and transport cost
than stonework this additional cost may
be offset by the elimination of temporary
formwork and the possibility to use lower
grade fill, lean concrete or stonework and
skills in the construction of the arch over
the corrugated sheets.
Figure E.8.38: Corrugated metal sheet arch
Corrugated metal sheets will need to be pre-bent to the correct radius for the arch by the supplier. They
can then be bolted together at the bridge site to form the arch. To ensure that the arch does not distort
when the fill is placed and compacted, the foundations or piers should restrain the corrugated metal,
preventing it from flattening out. This requires a ledge to be constructed to hold the sheets in place.

8.10.6 Fill options

There are three fill options that can be used in arch bridges which were discussed in the fills for approach
ways in the section shown in Table E.8.19:
ƒ Well compacted gravel;
ƒ Weak concrete mix with plums;
ƒ Rubble masonry.

8.11 Bridge design


This section covers the design of bridge decks appropriate for use on low-volume (traffic) roads in rural,
often remote, and urban areas. It includes guidelines for the design and construction of support abutments
and piers (see Plate E.8.31). A bridge is basically an extension of a road, albeit a more sophisticated and
expensive part. At a cost of up to 100 times or more than that of an equivalent length of road, however, it
is important that careful attention be paid to its design and construction. Bridges are critical elements of
the road system. A bridge collapse not only disrupts the serviceability of the whole of the road network
but it can also endanger life to a much greater extent than other components of the road. The possible
consequences of structural failure must be taken into account and given due emphasis in the design
process.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 153

Plate E.8.31: Two span concrete deck bridge with masonry pier and abutments

In this section, as in the rest of this Manual, emphasis is placed on relatively low-technology, labour based
solutions, as these tend to be the most economic and socially beneficial in rural areas of Ethiopia. The
previous text in this Manual is generally applicable to both single and two lane traffic small structures. The
following pages generally cover bridges spanning less than 10m and carrying a single lane of low volume
traffic. For single lane bridges, an appropriate deck width between kerbs or width limiting obstacles
is 4m which is sufficient for most commercial farm and public transport vehicles. This can be reduced
where certain vehicles are physically prevented from using the bridge and use is confined to motorcycles,
bicycles, pedestrians and animals. Extrapolation of the contents of this Manual to larger bridge spans or
for heavier traffic is not advisable. In these situations, a full engineered solution is required and reference
should be made to the ERA Bridge Design Manual - 2011, Overseas Road Note 9 (TRL 2000) or other
appropriate documents.

8.11.1 Choice of bridge site

An appropriate choice of location is important if an effective bridge solution is to be obtained, in terms of


cost of construction, maintenance and service life. The ideal site would have low flood levels, high solid
banks (preferably rock), a non-skewed crossing and straight approach roads. Normally, however, some
compromise is required.

8.11.2 Loading

Careful consideration must be given to the type, volume and weight of vehicles which will use the road.
It is often stated that “if a heavy truck can physically use the road, then at some stage it will”. Generally,
bridges must also be designed to carry the heaviest load expected. This is particularly important for
decks, less so for abutments and piers. Modern bridge loading specifications are generally applicable
to structures which experience high volumes of traffic (>10,000 vehicles per day). The economics are
such that bridges built to these specifications cannot be justified for the majority of low cost roads
used to service rural areas. Note that many low-volume rural roads in Ethiopia rarely experience vehicles
greater than 6 tonnes: this limit covers cars, light buses, pick-up trucks, cattle wagons, etc. In particular
circumstances this may not be sufficient, for example, near stone or gravel sources or factories which
produce heavy goods. Where heavier traffic (>6 tonnes gross vehicle weight) is likely to be a regular

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 154

occurrence proper engineering design by suitably qualified engineers is required. This is beyond the
scope of this Manual and reference should be made to the ERA Bridge Design Manual - 2011 and
documents such as Overseas Road Note 9 (TRL 2000).

8.11.3 Scour

The site of bridges must be carefully chosen to take local conditions into account to ensure durability
and functionality, including alignment. Chapter E.5 gives details of the general principles involved in
site selection and appraisal. For bridges, this is crucial if future problems and maintenance costs are to
be minimised. The type of site investigation required to take the watercourse into account is outlined in
Chapter E.6. The detrimental effects of scour on bridges and support systems must be recognised; in fact
this is the most likely cause of structural failure in bridges around the world.

In most cases problems can be minimised, and often avoided completely, by appropriate choice of form
and location for the crossing.

8.11.4 Drainage

Every form of bridge requires some water management to ensure that water does not pond on the deck,
which could cause a traffic safety hazard, rotting of timber, corrosion of reinforcement or deterioration
of masonry. For solid decks a transverse camber of 1 in 40 and a 1 in 100 longitudinal fall is sufficient to
prevent ponding. Where kerbs are present some means of disposing of water from the deck is required.
For timber decks, a 20mm gap between planks is sufficient to allow adequate drainage. For solid decks,
scuppers should be considered and should be carefully located and detailed to discharge excess water
through the deck without causing erosion, staining or maintenance problems. The careful detailing of
road side drainage outfalls at the bridge site is essential to avoid erosion problems.

8.11.5 Maintenance

In bridge design, there is a trade-off between initial construction cost and on-going maintenance costs,
and bridges which are cheapest to build can end up being the most expensive when whole life costs
are considered. Maintenance of a bridge must be considered at the design and construction phase. The
designer should make allowances for access for inspection and should recommend a maintenance plan
which includes extent and frequency of inspection, and any routine works required. These maintenance
costs and their practical arrangements should always be considered when selecting the preferred design
solution.

In general, it is a good idea to design bridges to minimise future maintenance actions and costs. This
is because maintenance is often neglected, particularly in rural areas where traffic levels are low and
financial/physical resources and logistics may be severely constrained or challenging. It should be
remembered that routine maintenance will ALWAYS be required. This involves regular brief inspections,
including preventative maintenance such as clearing of drains and removal of debris or garbage, on
an annual basis. This gives a clear indication of the performance of the bridge and the progress of any
deterioration. Provided adequate guidance and a means of recording the results of the inspection are
provided, these inspections do not require qualified engineers. However, a more detailed inspection at
intervals of about seven years by a qualified engineer is recommended. The detailed cost of the bridge
structure options should include the expected costs of the maintenance regime inspections over the
design life of the structure in present day costs, and also an estimate of the likely routine maintenance
activities. These should be estimated from maintenance records for existing similar structures.

Abutments and piers are often constructed within the watercourse. These should be designed and
constructed to keep to an absolute minimum their effect on water flow. This minimises the possibility of
scour and helps to avoid expensive maintenance work. In general and where possible deck soffits should
be constructed a minimum of 300mm above the highest expected waterline. For timber decks this should
be increased to 1000mm. For further guidance see TableE.8.24. Structure design and flood return period
considerations are addressed in Section E.3.3.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 155

In some cases, the construction of a low level bridge or vented ford might be appropriate where normal
water depth exceeds fordable depth, and where dry access is not required all of the time. These are
bridges which allow flooding approximately once a year for up to three days at a time. Most modern
vehicles can drive through 150mm of water and this may be an acceptable economic solution for very
low volume (traffic) roads. Scour protection should be provided to cope with the 50-year flood level
where practical. This should be sufficient to prevent scour or even complete washout of both the deck
and support system. Construction materials must be carefully chosen to prevent deterioration with time.
In particular, the flood water must be prevented from flowing around the structure or flowing down the
road. This can best be ensured by proper location of the bridge; retrospective work to keep flood water
within the original channel can be very expensive, if not impossible.

8.11.6 Choice of structure

The selection of structure type is discussed in Chapter E.4.

8.11.7 Choice of materials and form of construction

The general properties of construction materials and how to identify and evaluate them are outlined in
Chapter E.7. For bridges as for other road structures, the choice depends primarily on local conditions
and on the availability of materials and labour and the costs of the feasible options. However, greater care
is required in the selection of appropriate materials for bridge structures as the materials will be called
upon to take greater loads, and local weaknesses or defects may lead to total collapse of the bridge.
It is probable that in order to minimise the total cost of the structure, maximum use should be made
of local materials and labour. Any choice of materials and form of construction may have maintenance
implications and these should be included in the overall assessment of the options.

Reinforced concrete is generally considered to be the most economic material for construction of
bridge spans up to 30m (see Plate E.32). This is because of the long life expectancy, good durability
characteristics and low maintenance costs. However, while well-constructed concrete is very durable
and requires very little maintenance, construction requires a high level of technical skill as well as the
availability of good quality materials. The guidelines in Chapter E.7 must be followed if good quality
structural concrete is required. Bad site practice and poor workmanship can lead to a very poor structure
which can cause loss of stability and early collapse. Typical faults include use of dirty water; sand and
aggregate; inadequate mixing; placing and compacting of concrete; inaccurate fixing and positioning of
reinforcement or formwork; and storing of cement in humid conditions. Mix design (ie the proportions
of cement, sand, coarse aggregate and materials to be used) is very sensitive to mistakes. Labourers
often do not realise the consequences of poor practice and close supervision should always be carried
out when structural grade concrete is required. If there are local shortages of formwork, steel fixing and
structural concreting skills (which often have to be imported into a rural area), it may be more appropriate
to adopt designs that utilise locally available building skills such as carpentry and masonry.

Plate E.8.32: Reinforced concrete deck on masonry abutments and pier

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 156

Each region tends to have its own local construction artisans (eg blacksmiths, carpenters and
stonemasons,and materials (stone, brick, wood and gravel). These will affect the economics and local
resources should be used where possible, although other factors may also influence the final choice,
for example a local policy may influence preferences. The construction of stone or brick masonry arch
bridges is labour intensive but these are the most durable and, arguably, the most aesthetically pleasing
bridge forms. Simple arches are also technically the simplest form of bridge structure to construct with
relatively limited supervision requirements. If suitable materials and stonemasons are available, this may
be the most effective long-term solution.

Timber as a primary structural material has its advantages. Its low weight, low cost, general availability,
and ease of construction make it attractive in many remote situations where it is grown locally. Timber can
be assembled using non-skilled labour and in adverse weather conditions. It requires some protection
against deterioration and insects, particularly in hot humid climates. Timber requires deeper sections
than steel or concrete mainly because of its lower stiffness. Experience in North America, where there are
many timber bridges, suggests an average life of 50 years, although with good maintenance, the life can
be considerably greater.

Timber as a structural material has some major disadvantages which should be considered. All timber can
rot and be eaten by insects. Some degree of protection such as creosote is required and this should be
re-applied periodically through the life of the structure as required to ensure maximum life. Immersion
in creosote or other preservative for several days prior to assembly provides long-lasting preservation.
As timber is light it can easily be washed or blown away. All timber decks should be tied down at
supports and these fixings should be inspected at regular intervals. Timber is easily set on fire, either by
accident or maliciously. Garbage, driftwood, weeds, etc. should not be allowed to accumulate under the
structure. When timber, either in the form of sawn sections or logs, is used for structural purposes it is
very important to have a clear understanding of the strength and durability obtained from the particular
material available. Seasoned timber free of defects and properly preserved should always be used. See
Chapter E.7 for more details including tests to evaluate prospective timber sources.

Durable local stone in compression is the most economical material of construction when whole life
maintenance costs are included. General properties of different stone are given in Chapter E.7. Alternatively
bricks can be used but for bridge structures it is important that they are consistent in strength and quality.
Chapter E.7 gives some background on the expected properties of locally produced bricks.

8.11.8 Foundations

Foundations for piers and abutments are discussed earlier in this chapter. Bridges are usually constructed
on sub-soil with an allowable bearing capacity greater than 300kN/mm2. This is easily achieved in gravel,
compact sand and strong clay. A simple check to indicate this minimum capacity is:
ƒ A man’s weight bearing on a 30mm diameter bar only penetrates 100mm;
ƒ A 2m rod driven into the ground with a 3kg hammer experiences increasing resistance.

On softer soils, a bridge may not be appropriate and another site or form of structure should be
considered. Bridges can be constructed on very soft soils using piles. Timber piles can be driven using
fairly rudimentary equipment and manual or animal power. Where piles are used, design and supervision
should always be carried out by a suitably qualified engineer. Where bearing capacity is limited, it should
be noted that gabion abutments are lighter than concrete and spread the load well.

8.11.9 Arch bridges

Arch bridges usually provide the best solution in consideration of the level of maintenance required.
Spans greater than 10m require a properly engineered solution and reference should be made to the
ERA Bridge Design Manual - 2011, Overseas Road Note 9 (TRL 2000) or other appropriate documents for
design and construction. This Manual is appropriate only for spans less than 10m. Section E.8.10 deals
with large bore culverts and provides general information on the construction of masonry arch structures.
The following paragraphs refer to arch bridges appropriate for low volume roads suitable for pedestrians
and vehicles less than 10 tonnes.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 157

Arch bridges can be built in different forms and shapes. The key elements of an arch bridge are shown
in Figure E.8.39. The wedge shaped blocks, stones or bricks which form the barrel or ring of the arch are
called voussoirs. These are usually placed symmetrically around a centre stone or key-stone. In fact, the
key-stone has no special function and is an aesthetic rather than a structural requirement. The block in
the abutment on which the arch barrel sits is called a skewback and the surface between the skewback
and the end of the arch barrel is called the springing. The highest point of the arch is called the crown
and the lower sections are the haunches. The upper and lower boundary lines of the arch ring are called
the extrados and intrados respectively. The outer walls which retain the fill are the spandrel walls and they
become the wingwalls at either side of the arch.

SPANDREL ARCH AXIS


MASONRY CROWN

EXTRADOS

DEPTH (d)

INTRADOS RISE (R)

SKEWBACK
SPRINGING
ABUTMENT ABUTMENT

SPAN (S)

Figure E.8.39: Arch bridge details

Arches can be constructed using any good quality stone or brick. Wedge shaped stone can be used
without mortar but it is more common to use regular shaped rectangular stone or brick placed with a
good quality mortar forming the slightly wedge shaped joints between each unit. The use of mortar can
reduce the stresses in the stone by as much as 30% and should always be used if possible. If bricks are
used, a high standard is required; they must be fired to a good engineering quality and be consistent in
shape and strength.
Arch bridges are heavy structures and care should be taken to ensure that the foundation has sufficient
bearing capacity. Foundations are usually relatively shallow spread footings or onto solid rock where
this exists at the springing. It is essential that there is sufficient resistance in the abutments to resist the
substantial horizontal spreading forces inherent in an arch design. Excavation must be taken down to firm
material. In soft soils, timber, concrete or steel piles may be required beyond the scope of this Manual. A
cofferdam can be used to provide a temporary dry working area.

Piers in multi-span arch structures are usually thick structural components with widths about 25% of the
arch span. These are massive enough so that individual arches of multi-arch bridges are self-supporting.
Piers can be made using a double outer layer of bricks or blocks and the cavity filled with clay or rubble.
However, it is good practice to make the piers of solid masonry where possible, particularly for smaller
bridges.

For the arch barrel, extensive support is required during construction and it is likely that supporting
falsework will be placed in the river bed. There are obvious related seasonal storm or flood risk
considerations. Formwork will normally be made from timber of sufficient strength, fixed to give the
correct shape to the arch (see Plate E.8.33). As the Section on arches suggests, other material can be
used, eg. corrugated iron sheets. The formwork and supporting falsework must be firmly positioned and
able to take the weight of the masonry and workmen. It must be devised in such a way that it can easily
be removed once the arch has been constructed.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 158

Plate E.8.33: Masonry arch under construction with wooden formwork

Distortion of the arch during construction must be avoided as this can have serious implications on the
strength and stability of the completed bridge. The formwork should not distort or move noticeably due
to workmen moving over it. It is economical to reuse the formwork and this should be kept in mind when
devising the installation and method of removal after construction. To avoid having supports in the river
bed, formwork arching between the abutments can be used but this would not usually be required for
small span arches of normal height.

As access to the river bed may be required for a long period of time, arches may not be suitable where
floods occur frequently.

Arch bridges are suitable where high clearances are required. As the section above suggests, the simplest
arch shape is a semi-circle which avoids horizontal thrust forces at the springings. It also provides maximum
headroom and simplifies the geometric layout. Other shapes such as ellipses are used to reduce the
height of large span bridges; these are considered to have a potential weakness at the quarter points.
Any arch form where the ring is not vertical at the support will induce horizontal forces in the abutments
or piers which must be resisted.

The thickness of the ring or barrel of the arch is the main factor affecting the strength of a well constructed
bridge. Small arches may be built using a single layer of bricks laid radially providing a ring thickness of
215mm for a standard brick size. For larger arches the ring thicknesses shown in Table E.8.23 should be
followed. Because of the arch shape, the thickness of the mortar will vary through the depth of the ring.
Most arches are made using two or more concentric rings with mortar providing the only bond. A header
or stretcher bond may also be used, ie a brick laid radially to provide a key between the rings. For larger
spans, the number of rings can be increased towards the springings. It is recommended that skewed
arches are avoided.

Once the arch ring has been completed the fill material is put in place. A large amount of fill is required.
Any local material of consistent quality can be used, for example the material excavated during the
construction of the foundations. Strength is not a requirement, its only function being to distribute the
load uniformly to the arch barrel. However, well compacted fill can add considerably to the strength
of an arch bridge. Refer to the section on approach ways for appropriate materials and compaction
requirements. A well drained granular fill is the best material, being flexible enough to allow the bridge
to tolerate some degree of movement. It is recommended that the arch formwork is only removed once
all the fill material is in place. Plate E.8.33 shows an example of a masonry arch bridge under construction.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 159

Plate E.8.34: Completed masonry arch bridge with splayed wing walls on hard rock foundations

For brick arches, it is also recommended that the formwork be removed after the mortar has fully
hardened, about seven days, to avoid distortion of the arch while the mortar is still soft. For stone arches,
this period can be reduced.

Spandrel and wingwalls retain the fill material and stiffen the arch ring at its edges. They should be
thickened at the base to provide better stability. For larger spans it may be helpful to have wingwalls
sloped outwards in plan for extra stability.

8.11.10 Deck

The deck, or superstructure, is that Part of a bridge which carries the roadway. Its function is to transmit
the load safely to the abutments and piers, without damage to the bridge structure or undue distortion
of the deck. For bridges with spans less than 10m, the only loads that need to be considered are the
dead load of the deck itself, including parapets and any other bridge “furniture”, and the live load due
to traffic or pedestrians.

It is always a good idea to carry out a design check if possible. A simple analysis can be carried out,
assuming the deck is a simply supported beam. The loading to be used should consist of the heaviest
vehicle likely to use the bridge and a uniformly distributed load of 5kN/m2 of deck area to represent
pedestrian loading (including cycles and animals). The maximum expected stresses can be obtained
and compared with the strength of the material used. Maximum deflections can also be calculated once
the deck details have been established. In general, it is a good idea to limit the maximum expected
deflection to 1/100th of the span to avoid damage at the deck joints.
The deck can take many structural forms depending on local conditions and availability of materials and
labour. Arch bridges have been described in Section E.8.10; other types of bridges include reinforced
concrete slab bridges, beam bridges (reinforced concrete, timber, steel), and truss bridges (timber or
steel). The following gives general information on how different materials can be used to provide low
cost bridge decks.

Material - Concrete
Precast concrete beams are likely to be the most economical construction material. However, for small
spans (<6m), simple cast in situ reinforced concrete slabs are likely to be the most economical solution.
For larger spans, beams will generally be required. A span to depth ratio of about 12 will generally be
sufficient, although decks should not be constructed less than 300mm thick. As previously mentioned,
reinforced concrete is a material requiring certain technical expertise and requires care in construction
if an effective structural material is to be produced. Best practice as described above should always be
followed and supervision of unskilled workers is necessary if structural grade concrete is to be produced.
Reference should be made to the ERA Bridge Design Manual - 2011 and Overseas Road Note 9 (TRL
2000) for further information.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 160

Material – Timber
The weight of timber is about 25% that of concrete and timber is therefore quite an effective construction
material. All timber should be obtained from suitable hardwood which is generally available in tropical
forest areas, and should be treated using creosote etc. to prolong life. There are three basic elements to
a timber girder deck:
ƒ Road bearers: These support the surface of the deck and are often called beams, girders or
stringers, although trusses can also be used. The road bearers form the main structural elements
of the deck and are described in more detail below.
ƒ Floor planking: These are the boards which are nailed to the stringers to form the surface of the
deck. These boards spread the wheel load to the girders. As the girders are generally spaced at
less than 1m the individual pieces of floor planking do not need to be too long. A depth of 75-
100m is normally sufficient.
ƒ Wheel tracks or running boards: These are boards which are fixed to the deck in the direction
of traffic flow on which the vehicle wheels run. They provide protection to the floor planking from
wear and tear from heavy vehicles. The geometry of the tracks must be such as to accommodate
the wheel base of all vehicles likely to use the bridge. For most cases, tracks 1200mm wide with
a gap of 800mm between inside edges should be sufficient. In some cases, a cover of asphalt
or sand can be applied to prevent damage from heavy vehicles. Worn out or damaged running
boards, flow planks and girders should be replaced to avoid progressive damage and injury to
bridge users. A beneficial additional detail is to fix a ‘threshold’ plank laterally across the road at
each end of the running boards. This detail will help to reduce the vehicle impact loadings on the
ends of the running boards (this location is particularly susceptible to loosening of the running
board fixings).

The design and suitability of the final product is very dependent on the type and grade of timber available.
General advice is difficult because of the wide range of timber available. An example of a timber deck (in
need of maintenance and repair) is shown in Plate E.8.35.

Most codes refer to sawn timber of consistent quality. In the following, it is assumed that a supply of well-
seasoned hardwood timber is available, which is free of rot or insect infestation. It also assumes that, in
the worst case, the bridge will be loaded with light vehicles (< 6 tonnes in weight). Where heavier vehicles
are expected, more attention should be paid to structural details and reference should be made to the
ERA Bridge Design Manual - 2011, Overseas Road Note 9 (TRL 2000) or similar documents to define the
size and spacing of main structural elements.

The road bearers can consist of either a number of girders spanning between supports or a pair of trusses
along the edges of the bridge with transverse stringers carrying the deck. Simple girder bridges are
easier to construct and require less skilled labour but are only suitable for short spans. For longer span
bridges trusses provide a more efficient use of timber but these require specialist skills for design and
construction. In particular the joints and connections require careful attention. Design of timber truss
bridges should only be carried out by a suitably qualified engineer.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 161

Plate E.8.35: Timber deck with floor planking and edge beams in need of repair

Timber girders can be constructed from either sawn timber sections or from the original logs depending
on the source of timber available. The factors affecting the strength of girder decks are:
ƒ Type of timber (quality, strength);
ƒ Depth of member;
ƒ Width of member;
ƒ Spacing.

It is possible to design the timber deck for a particular type of timber but this will require detailed knowledge
of its properties. Where sawn timber is available commercially, this information may be obtainable from
the supplier. Chapter E.7 presents the general properties of different timber broadly classified into soft,
medium, hard and very hard wood and gives samples of the tree species. This highlights the fact that
strength is closely related to timber density.

Generally sawn timber is easier to use and fix in place because of the regular shape and flat surfaces. It
is also easier to examine for defects such as knots or insect damage which can seriously reduce strength.
Where minor flaws exist, the timber can be used provided the flaw is placed as close to the top of the
girder as possible to reduce its effect on strength. Where sawn timber is not available, logs can be used.
These require more care in selection for quality and size, positioning and fixing in place.

Table E.8.25 provides the size and spacing of sawn timber girders required for various spans. These are
appropriate for pedestrians and light vehicles only (up to 6 tonnes). For heavier vehicles, the tables in
Overseas Road Note 9 should be used. Note that wide spacing makes fixing of deck planks more difficult.

Table E.8.25: Sawn timber girder bridge deck for 6 ton vehicles

Span Timber size* Girder spacing


(width x depth - mm) (m)
5 150 x 300 0.5
8 200 x 400 0.8
10 200 x 400 0.5
12 250 x 500 1.0
Note:
* All timber to have a density greater than 450kg/m3

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 162

Logs are best used round but with the top shaven to carry the deck (Plate E.8.36). The bark should be
stripped and each log checked for soundness and defects. Properly seasoned logs should be used.
Particular care should be taken to ensure that the timber has not been attacked by insects. As with all
timber, logs should be treated with creosote or other preservative agent preferably by immersion for
several days. Painting is not sufficient protection. The ends of the logs are particularly vulnerable as they
are often in contact with soil. Moisture and garbage often collect at supports and can cause rotting. The
logs should be closely matched for size and positioned with the top surfaces in the same plane and to
accommodate any variations in log diameter with the large diameter at alternate ends on adjacent logs
(refer to Chapter E.9).

Plate E.8.36: Log stringers from underside of bridge deck

Running boards can be placed directly on top of the logs although deck planking is recommended if
pedestrians and animals are to use the bridge regularly. In general, three or four logs of about 300mm
diameter are sufficient to span up to 10m to carry a single lane of light traffic. Again, for heavier traffic,
the tables in TRL Overseas Road Note 9 should be used.

One common problem with timber decks is excessive spacing of the longitudinal stringers. Excessive
deflection of the stringers under vehicle loading can cause surface damage to the timber at the supports
which can lead to rotting and early deterioration of the deck. The deflection can also cause the deck
planks to work loose leading to damage, rot or even complete loss. A general recommendation for
heavily trafficked bridges is that the stringers be placed as close as is reasonable for the available timber
sizes to avoid excessive differential movement across the deck. This can be relaxed for low-volume roads.
Stringers should be placed so that the tops are at the same level; this ensures that deck planks bear
evenly across the deck. If one stringer is higher than the rest, the underside should be trimmed where it
bears on the support or the seating for that stringer should be lowered. This avoids having to trim the
whole top length of the timber. Floor planks 50x100mm make a very effective deck. These can be laid on
edge and nailed to the preceding one to make a very stiff solid slab 100mm thick.

Where joints are made using nails or screws, the minimum spacing distances shown in Table E.8.26
should be used (in terms of the nail diameter) to minimise the chance of damage to the timber and
premature failure of the joint.

Table E.8.26: Nailing requirements

Location of nail Number of nail diameters


Edge distance parallel to grain 20 diameters
Edge distance perpendicular to grain 5 diameters
Distance between lines of nails 10 diameters
Distance between adjacent nails in a line 20 diameters

Material – Steel

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 163

Steel beams with a concrete or timber deck make a very effective bridge. Steel beams are imported,
expensive and may be difficult to transport. However, they may be available from demolished steel truss
bridges or buildings. A concrete deck can be cast on top of the beams (composite construction). This
must be made integral with the steel beams either by encasing the beams in concrete or using shear
keys fixed to the top of the beam at 100mm spacing and penetrating 50mm into the concrete deck.
The deck can also be constructed using soil, rubble or lean concrete provided a method of supporting
and retaining the fill is devised. This could consist of transverse arches supported by the bottom flange
over which fill material is compacted. The arches can consist of brick or stone masonry, metal plates or
concrete.

Steel beam decks tend to rattle and vibrate excessively due to inadequate fixing at the supports. Beams
can be fixed to timber abutments using screws or nails driven through holes in the bottom flange. If a
timber deck is used the planks should be fixed securely to the beams.

If available and of suitable length, old railway lines can be used to form a bridge deck. Because of
difficulty of fixing to abutments and attaching deck planks, the rails can be encased in concrete so that
the rails act as reinforcement. This also protects the railss from corrosion.

8.11.11 Abutments BALLAST WALL

Abutments provide the support system for the deck BEARING SHELF
WING WALL

and retain the soil under the approach road (see


ABUTMENT
Figure E.8.40 for details). They can be built using
various forms and materials. The main function is to
EMBANKMENT
transfer the loads from the deck to the supporting WEEP HOLES
foundations. They are also located at the transition
between the approach embankment and the bridge
deck. Effective abutments should provide good
performance and stability to the bridge structure ABUTMENT WITH WINGWALLS
as a whole. The form of the abutment will depend
on foundation material and on the deck type. The
bearing capacity of typical soils and rock are given in
Chapter E.6; this will dictate the size of the abutment
and the bearing area required. RETAINING WALLS

The material used for abutment construction depends


ABUTMENT
primarily on the availability of local material. It is
recommended that concrete or masonry be used to
make abutments where possible. Mass concrete can WEEP HOLES

be used provided the concrete is of sufficient quality


and the abutment is of sufficient size.

Timber abutments may be considered acceptable ABUTMENT WITH SEPARATE RETAINING WALLS

for low volume road structures but their vulnerability


to deterioration and short service life should be
recognised. Gabions can also be used (see Figure Figure E.8.40: Abutment details
E.8.41) providing fill material of suitable size and
resistance to water damage is available. They have
the advantage of providing natural drainage to the
approach road. However, they are susceptible to
damage and settlement due to scour and should be checked regularly to ensure that the wire has not
corroded. Gabion abutments are not suitable for situations of paved road surfaces due to the settlement
risks.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 164

Figure E.8.41: Gabion Abutment Log Deck Bridge

Abutments should be built away from the watercourse if possible to avoid scour problems, even if it
means an increase in length of bridge. High abutments are expensive and it may be more cost effective
to increase the span if smaller abutments can be constructed further back from the watercourse.. Further
information about the options for filling behind abutments is provided in the section on approach ways.

Abutments experience lateral loads resulting from the action of the backfill material. The most critical
loading situation is often when the abutment has been constructed to full height but before the deck is
constructed to provide propping support. To achieve this it may be convenient to delay completion of
the backfilling operation until after the deck has been placed.

8.11.12 Piers

Piers can be the weakest parts of bridges and are most susceptible to damage by scour. The number
of intermediate piers should be minimised and they should be omitted completely if possible. If it is
necessary to include piers they should be oriented exactly in the direction of the water flow to minimise
the obstruction and water turbulence. Typical pier shapes are shown in Figure E.8.42.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 165

Figure E.8.42: Pier shapes (Plan view)

The section on scour presents a simple estimation of the potential scour depth that may be expected
around a pier. The footing should be placed well below this depth unless a firm rock foundation is
encountered. The shape of the pier will affect the amount of scour and designers should always aim to
construct piers with cross-sections which will minimise their effect on the water flow.

Design procedures are similar to those for abutments and the


guidelines given above should be followed however, performance
and stability requires more attention.

Piers are required to support the deck of a bridge or the base of an


arch. They may therefore be called upon to carry large vertical loads
to the foundations through footings. Footings may be considerably
larger than the piers if the ground conditions are poor. The form and
shape of the pier will depend on the bearing capacity of the foundation
material. The bearing capacity of typical soils is given in Chapter E.5,
Section 5.1.5.

Stonework or brick masonry (Plate E.8.37) is the most suitable for pier
construction due to its ease of construction, durability and resistance
to scour. It can also be used to create permanent formwork for the Plate E.8.37: Brick pier
pier and allow the use of other fill material in the middle (refer to the
section on fill material in approach ways). Reinforced concrete piers will tend to be more expensive than
masonry due to the increased temporary works required and the probable need to transport to the location
the steel, and the shuttering and steel fixing skills. Timber would be a third choice although it requires
frequent inspection and maintenance. Timber must be braced due to its lower strength capabilities; this
will ensure lateral forces due to the water flow can be resisted. Gabions are not recommended for use as
piers due to scour and settlement risks.

8.11.13 Bearings and joints

On major bridge spans (>20m) bearings and joints are required to allow movement of the structure due
to temperature or imposed loading without causing structural damage. For bridges of less than 10m
spans, these movements are small enough to be catered for by simple bearings, such as a sheet of felt
or rubber placed between the beams and the abutment, or can be resisted by stresses in the structural
elements. Nevertheless, bridge movements cannot be ignored and should be considered as Part of the
design and construction of the bridge.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 166

Movements arise from vehicle loading, pedestrians, temperature, wind and earthquakes. Wind and
earthquake loading are major considerations for long span bridges and are not normally considered for
bridges with spans less than 50m. Where high winds and earthquakes are expected, however, detailing
should be such that lateral and lifting forces are resisted by suitably tying down the deck and structural
elements. Vibrations from pedestrians, and particularly from vandalism, can cause problems on “lively”
structures, and decks should be prevented from jumping off their supports. Simple upstands at the
supports on either side of the deck will be sufficient to prevent lateral movements in most cases. Steel or
timber dowels can also be used where appropriate.

It is difficult to construct a road continuously over a bridge and the construction joints cause many
problems even in well-designed structures and paved roads. The ingress of moisture and differential
movements between the bridge structure and the backfill material invariably causes progressive damage
which adversely affects vehicles as well as the bridge. On low volume roads where vehicle speeds are low,
the effect of this is not serious and routine maintenance is sufficient to maintain a smooth ride. In some
cases, however, it may be a serious problem and a proper drainage system may be required to prevent
major damage.

8.11.14 Parapets

Generally, bridges are constructed with parapets to prevent people from falling over the edge or to
provide containment for vehicles in the case of accidents. For low volume roads, however, these are often
not necessary. Some form of kerb to prevent vehicles from slipping over the edge (see Plate E.8.38) or to
provide some degree of protection to pedestrians should always be considered.

Where significant flows of pedestrians or animals use the bridge regularly, handrails are required (see
Plate E.8.39) particularly where a hazard such as a dangerous drop (greater than 2m) exists. Handrails
should be 1m high and are most conveniently made from timber. Where children are expected to use
the bridge regularly, a mesh type of barrier may also be necessary to prevent them climbing or falling
through the parapet.

Plate E.8.38: Raised kerbs on a vented causeway allow water to pass over the structure
with minimum disturbance, but provide protection from vehicles driving off the structure

Plate E.8.39: Timber bridge with handrails

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 8 - 167

8.12 Other design issues

8.12.1 Debris control

During a flood vegetation and other debris will be carried in the water. The designer must make sure that
this debris will not either damage the structure itself or cause a blockage in the water flow which then
damages the structure. In the case of bridges it is particularly important that the water does not overtop
the deck, as it not designed to withstand the water flow. Table E.8.27 below provides minimum clearances
that should be provided between the maximum water level and the bottom of the bridge deck.

Table E.8.27: Minimum deck clearances

Discharge (m3/s) Minimum clearance (mm)


< 0.3 150
0.3 - 3.0 450
3.0 - 30 600
> 30 1000

8.12.2 Road signage

Bridges, drifts and any other structures causing a restriction in the road
width should be well marked by signs to warn approaching drivers (see
Plate E.8.39). Depending on the visibility along the road the sign should be
placed between 50 and 100m back from the obstruction and about 1.5m
from the edge of the road. Fixings should be robust and tamper proof. If
theft of metal signs/components is a problem at the structure location, then
signs should be painted on a masonry backing. On surfaced roads, surface
markings may be an option.

8.12.3 Carbon footprint

It is likely that there will be increasing concern regarding the sustainable


use of resources and the carbon footprint of road works and particularly
structures, both in the initial construction and life cycle of the infrastructure.
The designer should accommodate any current national and regional
requirements in the planning and design of the structures.

8.12.4 Climate change

It is unlikely that there will be significant Climate Change implications for


the design of small structures in the typical design lives being applied under Plate E.8.39:
the guidance provided in this Manual. Road narrows sign

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 168

9. CONSTRUCTION

The purpose of this chapter is to provide guidance on the actual construction of structures, from the
preparatory work, through the various site activities to the completion of site works. It includes aspects
of programming, construction, supervision and monitoring of works, whether the structure is built by a
contractor, a road authority work force or a work group set up specifically for the task.

Not all issues dealt with in this chapter will arise during the construction of a structure, especially a
small one. Checklists are provided and where appropriate the text refers to other documents for further
reference and information.

9.1 Preparatory work


Culverts
The limited resources and costs involved, and usually standardised nature of culverts, will often mean that
the amount of preparatory work may be limited. However some aspects of the preparatory work in the
following sections for larger structures may be relevant.

Bridges, drifts and large culverts


The size, resources and funding required for larger structures will usually necessitate considerable
preparatory work before the actual site works can begin.

It is assumed that structural survey and design will be carried out in accordance with the guidelines
elsewhere in this Manual and with any locally established standards. It is also assumed that cost estimates,
detailed drawings and bills of quantities will be prepared for the works.

If the work is contracted, appropriate contract documentation should be prepared in accordance with
local standards and procedures. When a contractor will be appointed, local contractor classification,
tendering, selection and award procedures should also be complied with. Arrangements should be in
place for resolution of any disputes that may arise through the contract.

(Note: Inappropriate contract documentation has often been used in the past for relatively simple
structures. Fortunately, simplified model documentation and guidelines are now available from
organisations such as FIDIC and ICE, which are more appropriate for small structures contracts. Contract
documentation should be appropriate, equitable and acceptable in the local legal environment. The ERA
Low Volume Roads contract documentation and specifications are most appropriate for LVR structures in
Ethiopia.)

Arrangements for management, supervision, testing, approval and audit of the works should be
established. All of these issues should be clearly documented and known to the parties involved in the
construction process. If there is any doubt about the responsibilities, adequacy or arrangements for any
of these issues, then professional advice should be sought to rectify the situation.

The construction of any structure for a public road involves risks and responsibilities which must be
appreciated and should be assigned to the most appropriate parties.

Inadequate attention to some aspects of the work can result in a structure not fit for its purpose, waste of
resources, or even serious damage or eventual loss of the structure.

Structure costing
It will usually be necessary to prepare a detailed costing of the structure, either for internal budgeting and
funding purposes, or for contracting out the work. This will normally be achieved through preparation of
a Bill of Quantities which can be priced by the client/promoter and by a contractor.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 169

Sample bills of quantities are provided in Volume 2. The following checklist indicates the components
which should be included in any complete costing of a structure.
Table E.9.1 below (checklist for preparing a construction programme) may be used as the basis for
developing a Bill of Quantities. Bills of Quantities in a national standardised format, with activity related
items, will assist clients and contractors in pricing works and assessing value for money.

Table E.9.1: Checklist of cost components for detailed costing of a structure

Direct costs Overheads


ƒ Materials ƒ Supervisory and technical staff
ƒ Unskilled labour ƒ Survey and setting out
ƒ Skilled labour ƒ Main office, workshop costs
ƒ Equipment purchase ƒ Supervision vehicles
ƒ Equipment operating costs ƒ Transport to and from site
ƒ Equipment hire ƒ Site camp and stores
ƒ Tools ƒ Security measures and facilities
ƒ Temporary works ƒ Communications (telephone, mail)
ƒ Services hired in ƒ Insurances, bonds
ƒ Banking and other charges
ƒ Training
ƒ Protective clothing and safety
ƒ Traffic control/signs
ƒ Testing
ƒ Welfare, pensions, social costs
Contingency/risks
(eg unforeseen additional work, late payment, delays)
Profit
The contractor should normally be expecting to make in the order of 10% profit on his work, after
covering ALL other costs. However this percentage will be affected by local market conditions,
competition and perceived risks

9.1.1 Planning of site works

Good planning of the site works is essential, particularly as many structures sites are remote from
organisational bases, sources of some materials and skilled manpower, and communications can be
difficult. Poor planning can lead to serious delays and increased costs. Table E.9.2 provides a checklist
for planning of site works and Table E.9.3 provides a checklist for preparing a construction programme.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 170

Table E.9.2: Checklist for planning site works

1. List all construction and support activities and prepare a construction programme (using a bar
chart) based on the Bill of Quantities, expected productivities and logical sequence of activities.
2. Prepare resource plan and cash flow requirements.
3. Plan in recognition of the seasonal watercourse conditions and expected flood conditions. Plan
adequate arrangements for damming, diverting or control of water.
4. Ensure compliance with all laws and regulations regarding recruitment, labour, (permanent/
casual) employment, gender and disadvantaged groups opportunities, payment, security for
payment to labourers, conditions of work.
5. Plan compliance with environmental requirements, particularly with regard to materials
exploitation, replacement of felled timber, watercourse pollution and waste disposal.
6. Inspect site. Check site survey. Review designs and documentation for compatibility and with the
actual site conditions. Clarify any inconsistencies
7. Plan and arrange land (acquisition/lease/use) and setting up site, camp and stores. Cement to be
stored in a secure, dry and well ventilated place.
8. Ensure adequate site access arrangements, particularly if the structure is being built in advance of
the road works.
9. Plan water supply, other services requirements and sanitation arrangements
10. Plan site security (particularly against theft of handtools & materials; cement is particularly
susceptible)
11. Ensure availability and accessibility of funds and contingency finance.
12. Ensure payment arrangements for (sub)contractors, and suppliers are in place.
13. Plan staffing, identify skills locally available or required to be imported to the site area,
accommodation, logistics, transport to site, recruitment and training of workforce.
14. Arrange for supplies of materials to site.
15. Plan safe and adequate temporary arrangements for traffic and pedestrians where replacing an
existing structure or facility.
16. Plan actual/contingency arrangements for de-watering and shoring of foundations.

Table E.9.3: Checklist for preparing a construction programme

The construction programme will involve some or all of the following activities:
1. Clear trees, bush and scrub dispose of safely.
2. Primary setting out and establishment of reference points.
3. Remove topsoil, stockpile for re-use or disposal.
4. Dig catchwater drains to protect site and any side drains.
5. Remove/bury nearby/break surface boulders (see below).
6. Detailed setting out and establishment of levels and profile boards.
7. Excavate foundations and any cut-off trenches.
8. Temporary shoring, watercourse diversions, piling, cofferdams, de-watering/drainage.
9. Drill and blast any solid rock.
10. Replace “soft spots” in ground, clean and prepare foundation area.
11. Construct foundations.
12. Construct temporary works for superstructure.
13. Erect abutments, piers, deck, wingwalls.
14. Fix deck timbers and running boards where applicable.
15. Erect kerbs, parapets barriers and safety structures.
16. Install drainage layers and features against structure.
17. Backfill against and adjacent to the structure, compacting each layer according to the
specifications. Particular attention to be paid to all compaction within 5 metres of the structure.
18. Construct road pavement/surfacing and markings, road shoulders.
19. Construct road drainage features.
20. Construct gabions and erosion control measures.
21. Lay topsoil/turves and planting.
22. Install traffic warning signs if necessary.
23. Clear site, remove surplus materials and leave tidy.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 171

The productivity standards shown in Table E.9.4 may be useful in estimating the resources and time
required for each activity.

Table E.9.4: Recommended productivity standards

Site clearance (bush clearing, tree felling, etc) 100 – 350m2 / worker day
Removal of tree stumps 1 / worker day
Soil excavation (and stockpiling alongside) 2 - 5m3 / worker day
Rock (fractured) excavation 0.8m3 / worker day
(solid rock will require drilling and blasting/
splitting)
Loading 8.5m3 / worker day
Haulage by wheelbarrow
0 - 20m 8.5m3 / worker day
20 - 40m 7.0m3 / worker day
40 - 60m 6.5m3 / worker day
60 - 80m 5.5m3 / worker day
80 - 100m 5.0m3 / worker day
100 - 150m 4.5m3 / worker day
Install only 600 or 900mm diameter culvert lines
0.8 - 1.2 lin.m / worker day
(including excavation and backfill)
Mix and place concrete 1.0m3 / worker day
Erect masonry work 1.0m3 / worker day

Productivity depends on a number of factors, including worker nutrition, fitness, experience and
motivation, site organisation, tool quality and condition, and climate. Individual small structures sites
do not allow much scope for improvement of performance with experience due to the short time spans
involved for individual activities. New workers under training will also be less productive. Poor quality and
condition of handtools can affect productivity by up to 25%.

The following list shows the range of skills which may be required on a structures site.
ƒ Surveying and setting out;
ƒ Drilling and blasting;
ƒ Piling/cofferdam;
ƒ Carpentry;
ƒ Masonry;
ƒ Temporary works;
ƒ Steel bending and fixing;
ƒ Concreting;
ƒ Equipment maintenance.

The more specialist skills may need to be imported into the project area. Some skills may be taught
through on-the-job training. This will involve costs and loss of productivity. Workers not from the area of
the structure site may require temporary accommodation and incur costs relating to travel and allowances.
Table E.9.5 provides a checklist of hand tools and site equipment that could be required to execute the
works.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 172

Table E.9.5: Checklist of handtools and site equipment

Handtools Equipment

Ranging rods Hand drills Culvert moulds


Spirit level /Abney level / water
Plugs and feathers
tube level
Stringlines, pegs Plate compacter
Profile boards & travellers Masons trowels Pedestrian vibrating roller
Plumb bob Masons hammers Water bowser
Tape measures Spirit levels Water pump
Straight edges Concrete mixer
Felling axes Lifting tackle Batching boxes
Tree felling saws Buckets Vibrating poker
Bush knives Mortar pans Piling equipment
Brush hooks Mixing boards Hydraulic excavator
Ropes Water containers/drums Compressor and air tools
Screeding boards Craneage
Pick axes Pointing tool Aggregate crushing eqp.
Mattocks Aggregate screens
Hoes Hand rammers Supply and site transport
Crowbars Rakes/spreaders Formwork/moulds
Shovels
Sledge hammers Slump test equipment Traffic signs and barriers
Concrete cube moulds and
Wheelbarrows
curing tank
Head pans/baskets Soil density testing equipment Safety helmets and equipment
Earth ‘stretchers’
Sandbags for water control
Carpenters tool kits

9.2 Site works


The works will be organised according to the activities in the construction programme. Guidance on the
individual activities is provided in the previous chapters of this Manual. However some specific aspects
warrant further explanation.

Temporary works should be designed to withstand the watercourse (eg flood or debris) conditions
expected and temporary loading situations.

Works must be carried out according to the specifications and drawings. Quality control arrangements
should ensure compliance in accordance with specifications. The control of cement requires particular

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 173

attention; this is a valuable commodity and it is difficult to detect reduced inputs until after construction.
Batching should be carefully controlled and the making and curing of cubes should be closely supervised.
It will normally be necessary to arrange for cubes to be crushed at a reliable laboratory remote from the
site. The laboratory must be consulted to ensure that the testing can be carried out according to the
standards specified. It is advisable to visit the laboratory to assess the standard of service provided.

The slump test (see Chapter E.7) should be used to control concrete workability and check the water:
cement ratio. Treatment of permanent timbers should be closely supervised or guaranteed.

9.2.1 Simple setting out techniques

The following simple setting out techniques are illustrated in the respective figures:
ƒ Levelling with a water hose (Figure E.9.1);
ƒ Setting out a right angle (Figure E.9.2);
ƒ Use of batter boards (Figure E.9.4).

Figure E.9.1: Levelling with a water hose Figure E.9.2: Setting out a right angle

Figure E.9.3: Use of batter boards

9.2.2 Setting out culverts and drifts


The setting out should be according to the design. The principal setting out requirements are the
establishment of the centreline of the barrels (for culverts), the extent of the structure (ends and corners)
and the inlet/upstream and outlet/downstream invert/slab levels (see Figure E.9.4).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 174

Figure E.9.4: Culvert arrangement A (flat outfall)

Wooden pegs should be used to establish key positions and levels. For minor culverts and drifts where no
levels are provided, the invert of the culvert or drift slab should follow the level of the existing watercourse
as closely as possible. The guidance in Figures E.9.5 and E.9.6 and the procedure for setting out a culvert
shown in Figures E.9.7 and E.9.8(a) and (b) will minimise the possibility of silting or erosion of a culvert due
to installation at an incorrect level.

Figure E.9.5: Culvert arrangement B (intermediate outfall)

Figure E.9.6: Culvert arrangement C (steep outfall)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 175

If the culvert site is flat, check the watercourse gradient for 20 metres downstream from the location of the
culvert outlet. Use boning rods and Abney Level, or line and level, for this purpose. If the gradient is less
than 5% (1 metre fall in 20 metres), then construct the culvert in Arrangement A with the culvert inverts
as close to existing ground/water course level as possible. Also construct Arrangement A, if the height of
embankment fill (measured from ground level to edge of road running surface) at culvert site is at least 1.1
metres. Otherwise proceed with the following steps to install Arangement B and C.

SETTING OUT OF 600 mm Ø CULVERT - ARRANGEMENT B OR C

MAIN DIMENSIONS ARE FOR ROADWAY WIDTH OF 5.5 m. (DIMENSIONS IN BRACKETS ARE FOR
CROSS SECTION WITH ROADWAY WITH OF ‘w’ METRES)

PROCEDURE STEP BY STEP EXAMPLE/EXPLANATION


STEP 1 OUTLET INLET
2.75 m 2.75 m
(0.5 w) (0.5 w)
Fix the centreline of the culvert. Establish two
pegs (peg A and peg B) at the location of PEG B PEG A
both roadway edges and at proposed finished
roadway level. Make sure that pegs are on the
same level (use line and level or Abney level). PEG AT PROPOSED
ROAD LEVEL

STEP 2
ROADWAY
5.50 m
Measure distance between peg A and B
(5.50 m, or ‘w’ for other cross sections). (w)

STEP 3
OUTSIDE DIAMETER OF
Calculate the minimum depth (d) to be CULVERT or 600 mm 0.72 m
excavated from proposed road level to OVERFILL (MINIMUM COVER) + 0.45 m
underside of culvert pipe at the inlet to TOTAL DEPTH (d) 1.17 m
ensure adequate cover (at peg a).
STEP 4 GRADIENT:
DIFFERENCE IN LEVEL:
Calculate the difference in culvert level between 4% x 5.50 m 4% = 0.22 m
peg a and b with the chosen culvert gradient 100% (FOR ROAD WIDTH w,
(4% is normally selected as the ideal gradient). = 0.04 w)

STEP 5 ROAD WIDTH (m) 5.50 (w)


INLET DEPTH 1.17 m 1.17 m
Calculate the depth to be excavated from DIFFERENCE IN LEVEL + 0.22 m 0,04 w
proposed road level to the underside DEPTH AT OUTLET 1.39 m y = (1.17
of culvert at the outlet (at peg b). + 0.04 w)

Figure E.9.7: Procedure for setting out a culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 176

PROCEDURE STEP BY STEP EXAMPLE/EXPLANATION

STEP 6
SECOND PEG
220 mm
Raise level peg A by the same measurement (0.04 w)
that you have calculated under step 4
establishing a second peg (difference in level).
FIST PEG

STEP 7

Find the end of the outlet-drain by using boning rods and a stick or rod of length 1.39 m (‘y’ for other
cross section road widths) (see sketch below): Walk S and rod away from B until the tops of A, B
and S are in line
A
B SECOND
S
PEG
C/L
1.39 M
1.39 M ROD (‘Y’)
(‘Y’) 4%

OUTLET DRAIN
MAX. 20 m. CULVERT

If the length of the outlet drain SB is less than 20 metres then establish the drain outlet peg at ground
level at point S. Construct the culvert in Arrangement C. (ie the road alignment will not need to be
raised). Establish the excavation level for the underside of the culvert pipe by measuring vertically down
1.39 m (‘y’) from peg B and the top of the second peg at point A. The excavation pegs should be 5.50
m apart (‘w’ for other cross section road width).
STEP 8

If the drain outlet cannot be found within 20 metres of the culvert outlet then place a peg at ground
level at point S, 20 metres away from the culvert outlet point B. Adjust the boning rod at points S until
the tops of the 3 boning rods A, B and S are in line. Measure the distance from the bottom of the
boning rod S to the ground level: z metres. The road level at A and B will have to be raised by
1.39 - z metres (y - z for other cross sections). To fix the culvert inlet excavation level, measure (1.39 - z)
metres (y - z for other cross sections) down from the top of the second peg at point A. To fix the culvert
outlet excavation level, measure (1.39 - z) metres (y - z for other cross sections) down from the peg at
point B (these pegs should be 5.50 m apart, or w for other cross sections). This is Arrangement B. The
road will need to be raised as indicated above and the vertical alignment raised or a suitable ramp
constructed either side of the culvert.

NOTE:
1. Where pipes will be bedded on imported material, excavation levels will have to be lowered by the thickness of bedding
material.
2. For 900 mm Ø culverts the dimension in step 3 is increased to 1.05 + 0.70 = 1.75 m. In step 5 the depth at outlet will
be 1.97 m (1.75 + 0.04 w).
Figure E.9.7 (continued)

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 177

(a)

(b)

Figure E.9.8(a) and (b): Setting out culvert profiles

Plate E.9.1 shows an example of the setting out of a culvert apron and Plate E.9.2 shows a completed
culvert and inlet apron.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 178

Plate E.9.1: Construction of culvert inlet apron with guide pegs and strings

Plate E.9.2: Completed culvert inlet apron

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 179

9.2.3 Setting out bridges and large structures

Benchmarks and reference points must be established. They should be well marked and protected. The
structure centreline should be set out. Front faces of abutments and centrelines of piers should be set
out. All setting out pegs should be located well back from the main working areas and be protected by
(preferably brightly painted) timber markers. Two pegs established in a line both sides of a structure will
mean that lines can be re-established if one of the pegs is accidentally disturbed. Right angles can be set
out using the ‘3:4:5 triangle’ rule (see Figure E.9.9 for simple setting out techniques).

Figure E.9.9: Basic setting out techniques for structures

9.2.4 Excavations

Foundations (if not on bedrock) are usually the most risky stage of the construction process. Whatever
site surveys were carried out beforehand, the actual foundation conditions cannot be determined until full
excavation takes place. Unforeseen problems may occur. If problems are not adequately tackled at the
foundation stage, they can be particularly difficult or expensive to rectify later.

Particular care must be taken to ensure the safety of workers and suitable precautions as illustrated in
Figure E.9.10. Excavations can collapse particularly in weak ground or where groundwater seeps into the
excavation. Temporary shoring should be used if there is any risk of this.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 180

Figure E.9.10: Safety issues

Foundations should be kept as dry as feasible and covered as soon as possible. If weak or soft spots are
uncovered then they should be replaced with imported suitable material such as clean gravel or crushed
stone. The bearing capacity of the actual foundation soil can be checked using simple apparatus such as
the DCP.

A sump and drainage channels may be necessary to keep the foundation dry. Buckets or a water pump
may be used to remove the water from the sump see Figure E.9.11.

If the ground conditions are worse than expected then the design engineer should be consulted (his
availability at this critical stage should be ensured during the construction planning process).

Figure E.9.11: Pumping arrangements

If steel reinforcement is incorporated in the foundations it is advisable to use a blinding of lean mix
concrete directly on to the trimmed soil to form a clean working platform for arranging and fixing the
reinforcement. For other works compacted gravel or crushed stone can be used for a clean working
surface.

If the foundation is on bedrock a good key must be ensured. This can be achieved by chipping the
complete rock surface and exposing a rough clean rock face. The drilling of holes and insertion of dowels
may be necessary to ensure a good key. Sloping bedrock should be excavated in steps (benched) to
provide a stable foundation.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 181

Culverts on busy existing roads may be built in two halves with adequate traffic control and safety measures.
Otherwise a diversion must be arranged with adequate warning signs and traffic control measures. Table
E.9.6 provides a supervisory check box of precautions that should be taken during excavation works.

Table E.9.6: Supervision check box - Excavation

ƒ Inspection and approval by engineer/senior technician


ƒ Keep as dry as possible
ƒ Trim to correct levels (and falls for a culvert or drift)
ƒ Ensure firm foundation - remove soft spots and replace with good material
ƒ Culvert excavations should be no wider than necessary to install culvert

9.2.5 Temporary works

Scaffolding, shuttering and temporary works should be constructed by skilled artisans to designs prepared
taking account of the local materials and loadings expected.

9.2.6 Shuttering/formwork and steel reinforcement for concrete work

Formwork and steel reinforcement (where specified) for concrete should be constructed in accordance
with current standards and guidelines. Each stage of the work should be thoroughly checked before
concreting is permitted. Checks should include cleanliness, soundness and quality of formwork and to
ensure that it will not move or leak under the loading of fresh wet concrete and workers. Details of the
formwork for concrete slabs are shown in Figure E.9.12.

Timber chamfer fillets (eg 20mm x 20mm timber sawn at 45o) should be fixed on all external 90o angles
to ensure smooth finished edges to the concrete.

Figure E.9.12: Formwork detail

The steps for formwork for concrete walls are shown in Figure E.9.13.

9.2.7 Steel reinforcement

Table E.9.8 provides a check list of precautionary measures for steel reinforcement with details of spacers
for the reinforcement given in Figure E.9.13.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 182

Figure E.9.13: Wall Formwork detail

Formwork should be coated with mould oil to allow easy striking of formwork after the concrete has set.
Linseed oil, old engine oil or other cheaply available oils may be suitable for this. Table E.9.7 provides a
checklist of precautions that should be taken in the erection of formwork for concreting.

Table E.9.7: Supervision Check Box - Shuttering/formwork

ƒ Faces of sawn timber/plywood/sheet steel securely fixed and supported on a timber/bamboo/


steel framework
ƒ Erected to the correct levels, alignment and tolerances, strong enough to support the weight of
wet concrete and operations without distorting/settling
ƒ No gaps or holes in faces for wet concrete to escape
ƒ Oiled to assist with striking/removal
ƒ Weep holes/scuppers/fixings/joints etc. in correct locations
ƒ Clean with all debris removed prior to concreting

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 183

Table E.9.8: Supervision check box - Reinforcement steel fixing

ƒ Ensure steel bar grades, sizes, numbers, spacing and shapes according to design drawings
ƒ Ensure minimum cover to steel from soffits, walls and top surfaces (50 - 100mm)
ƒ Ensure minimum overlaps between bars
ƒ Ensure steel and shuttering is clean with no loose rust or contamination (slight corrosion assists
bonding of concrete to reinforcement)
ƒ Reinforcement should be cut to the specified lengths, bent cold and provided with the minimum
laps specified. Overlapping and lapped bars should be bound tightly with binding wire at ALL
points.

Figure E.9.14: Reinforcement spacers

The spacers supporting reinforcement should be securely fixed so that they will not move during
concreting. Spacers can be made from mortar cubes with binding wire cast in. The amount of cover to
the reinforcement (usually 50 mm in moderate conditions to 100 mm in extreme exposure conditions) is
important for reinforced concrete durability.

9.2.8 Concrete work

The requirements of the specifications


and guidelines elsewhere in this Figure 9.11
Manual should be followed. Water Batch Box
for concrete should be clean and free
from contaminants such as salt and silt.
Water with solids in suspension should
be allowed to stand in barrels so that
the sediments settle out. Water that is
drinkable is usually fit for concrete works.
Concreting is usually not permitted
at ambient (shade) air temperatures
below 30oC. Hot weather concreting
must also be avoided as the uncured
concrete will dry out too quickly. Usually
concreting is not permitted in ambient
(shade) temperatures above 400oC. In
ambient temperatures well below this
figure structures and aggregates in
direct sunlight can rise to unacceptable
temperatures. Shading and timing of
concreting during cooler hours can be
important countermeasures.
Figure E.9.15: Batch box

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 184

Concrete strength and durability depends particularly on the correct mix proportions as specified
elsewhere in this Manual. Batching boxes (Figure E.9.15) or weighing methods should be used to ensure
correct quantities. Volume batching should be carried out using batching or gauge boxes of volume
equivalent to one 50kg bag of cement. A box of internal dimensions 400 x 300 x 300mm will have the
correct unit volume of 0.036m3.

Hand (see Figure E.9.16 and Plate E.9.3) or machine mixing of concrete is acceptable. Either method must
ensure complete mixing of the components. For Mechanical mixing the order of adding the materials to
the drum should be: coarse aggregate, cement, fine aggregate, water. Water should be added at the
specified water:cement ratio. Less water will not allow the materials to be properly mixed and placed.
Too much water will lead to segregation and a weak concrete. If the aggregates are already wet, then
adjustment of the quantity of added water will be necessary. The slump test should be used (see Figure
E.7.16 in Section E.7.4.4) to check the water:cement ratio and workability of the wet concrete.

Figure E.9.16: Concrete hand mixing

Plate E.9.3: Hand mixing mortar on a clean surface

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 185

Wet concrete must not be transported long distances as segregation will occur. It must not be dropped
more than 1.5 metres as this will also encourage segregation. Concrete should be vibrated with a
mechanised immersion poker placed throughout the fresh wet concrete to agitate the particles into a
compact matrix and expel the excess air. Over vibration must be avoided as this will result in segregation.
The vibrator should be kept away from the formwork. For minor works and standard concrete grades it is
acceptable to use lengths of reinforcing rods to agitate the concrete if a vibrating poker is not available.
Good supervision is required to ensure a methodical process.

Finished concrete surfaces should be tamped and screeded with excess concrete removed and disposed
of. Exposed surfaces should be finished smooth with a steel or wooden trowel. The finished surface of the
concrete must be protected from rain within the first two hours after placing. Curing with water should
commence three hours after casting. Concrete must be adequately cured for quality and durability and
adequate arrangements must be made for this for at least the first 7 days after casting. Sacking or sand
or other suitable material should be used to cover the concrete and retain the curing moisture/water. The
surface of the concrete should be kept damp with repeated wetting during the curing period.

Formwork must not be removed until the concrete is strong enough to support its loading. Removal must
be carried out carefully as point loads can damage “green” concrete.

A checklist for supervision or concreting works is shown in Table E.9.9.

Table E.9.9: Supervision check box - Concreting

ƒ Permissible air temperatures for placing


ƒ Clean water for concrete mix
ƒ Correct mixing proportions (check especially that correct cement quantities used and that batches
are weighed or gauge/batching boxes are used)
ƒ Correct water:cement ratio (check workability with slump test)
ƒ Concrete placed and compacted within 30 minutes of mixing
ƒ Cure continuously (keep all surfaces damp) for at least 7 days

9.2.9 Precast concrete


Where possible and where transport arrangements allow, precast units can be built at a central location
(eg culvert pipes or reinforced concrete beams). This should allow the benefits of efficient production
arrangements and greater quality control.

Precast culvert rings should be carefully lowered into position using ropes or straps to control the
operation. Particular care must be taken to prepare the bed to ensure uniform support of the pipe. Final
adjustment of position should be carried out with the aid of crowbars. Joints should be mortared and
protected (for example with banana leaves) prior to backfilling. After backfilling the joints should be
inspected internally and repaired if necessary. Precast box culvert deck slab units or even bridge deck
beams may be cast on site and then placed in final position. Crane or heavy lifting and moving equipment
will be required for precast deck beams.

9.2.10 Timber stave culverts (Plate E.9.4)

These pre-treated timber culverts may be


transported in component form to the site. The
efficiency of the system means than a number
of culverts can be transported in one truck load.
The timber culvert is assembled with its binding
hoops at ground level alongside the culvert
site. It is then gently lowered into its permanent
location. Particular care must be taken to ensure
that the culvert does not move during backfilling
and compaction.

Plate E.9.4: Timber stave culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 186

As with all precast unit systems, particular care is required to ensure a completely uniform bedding,
surround and backfill operation to avoid later settlement or failure problems.

9.2.11 Masonry work

In masonry work, the corners and ends should be constructed first, with particular attention to verticality
and alignment. String lines can then be stretched between the initial work to guide and ensure smooth
faces and coursing of the subsequent work. Adequate bonding should be ensured with vertical joints
staggered. A timber template may be fabricated to assist with constructing irregular shapes (see Figure
E.9.17 and Plate E.9.5). Mortar must be used within 30 minutes of mixing. Masonry work should be cured
as concrete.

Figure 9.14
!

Templates

Figure E.9.17: Masonry Templates

Plate E.9.5: Example of Masonry works

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 187

A checklist for masonry work is shown in Table E.9.10.

Table E.9.10: Supervision check box - Masonry

ƒ Stone must be clean, sound and firm


ƒ Vertical joints staggered
ƒ Edges and corners true, and appropriate stones selected for these locations
ƒ Faces true without irregularities, dishing or bulges
ƒ All joints to be pointed to ensure effective load transfer and minimise water penetration

9.2.12 Timber superstructure

Although timber may be used for bridge abutments and piers, its eventual replacement can involve
considerable work when resources may not be readily available. Timber beams, decks and running boards
are comparatively low cost and easy to replace, and are discussed in the following text.

Sawn timber or logs can be used for bridge


deck beams or bearers. Timber plank decking
or running boards can be fixed to the timber
bearers or to steel beams. All timbers should
be carefully treated prior to installation to
achieve acceptable service life. A small
diameter pole or timber can be used to launch
the first main beams across the bridge span
as shown in the diagram, using log rollers and
ropes if necessary. Once two main beams are in
place then these can be used as a platform to
manhandle the remainder of the main beams
into place (see Figure E.9.18)

Log bearers should be selected which are


as straight as possible and do not taper Figure E.9.18: Timber beam launching
substantially over their length. Log bearers will
not all be of exactly equal diameter. After all bearers are in place they should be levelled with string
lines and spirit level so that all top faces are reasonably level. They should then be packed with stones/
bricks and mortared into position at the abutments and piers (Figure E.9.19). The ends of the logs should
be concreted or mortared to discourage insect access. If logs vary in diameter along their length, then
adjacent logs should be laid with their thick sections at opposite ends of the deck to provide uniform
deck stiffness and strength.

Figure E.9.19: Log packing

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 188

There will usually be high and low points on the log beams. These are either trimmed off with an axe or
built up with packing to achieve a level area to which the deck planking can be fixed. Packing timbers
should be at least 300mm long, treated and securely nailed to the bearers before the deck planking is
laid.

Decking timbers are fixed by nailing or bolting at right angles to the main beams (even on a skew bridge).
These timbers will spread the vehicle loads to the main beams. Each deck timber should be fixed to every
beam as close to its centre as possible (Figure E.9.20).

Figure E.9.20: Fixing deck timbers

If the timber is particularly hard the nail holes may need to be pre-drilled to a diameter slightly smaller
than the nails to avoid the timber splitting yet achieve a secure fixing.

Excess timber should be trimmed from the ends of the deck planking.

Running boards are fixed to the decking by nailing or bolting. Heads should be recessed to avoid damage
to vehicle tyres. Running boards should be laid to form a running surface for each wheel of at least 1.1m
wide for safety. Joints in running boards should be staggered and cut square. All board ends must bear
on a decking timber and be securely fixed to it and trimmed (see Figure E.9.21). A threshold board should
be fixed across the end of the running boards at each end of the bridge. This will absorb the initial impact
of a vehicle and protect the running boards.

Figure E.9.21: Trimming running boards

Kerb timbers should be fixed to the edges of the deck and parapets provided for pedestrian safety (see
Figure E.9.22).

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 189

Figure E.9.22: Kerb timber fixing

9.2.13 Earthworks/Backfilling

Prior to backfilling, weep holes in masonry and concrete walls should be backed with a lean concrete
plug. This will be porous to allow ground water drainage, but will prevent the backfill material from
washing out.

Backfilling (Figure E.9.23) and compaction should be carefully and methodically carried out. Adjacent to
structures it is usually not possible to use heavy compaction equipment; in confined spaces small plant
or hand tools must be used. The backfilled area adjacent to the structure is particularly susceptible to
settlement in contrast to the relatively rigid structure. Good procedures and supervision are required to
minimise the risk of later settlement so that particular care must be taken within 5 metres of the structure.
The control of layer thicknesses is important.

Figure E.9.23: Backfilling

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 190

Control of compaction should either be by “end-product” specification, or “method” specification.


Nuclear density testing is quick and convenient and avoids the need for laboratory testing and time
delay using the sand replacement density method. Method specifications rely entirely on the presence
and integrity of an inspector throughout the filling and compaction process.

Table E.9.11: Supervision check box - Backfilling

ƒ Backfill with selected suitable material in even layers of no more than 150mm thickness and
thoroughly compacted. Material to be moistened if necessary to aid compaction.
ƒ No large stones or rocks to be placed directly against structure.

9.2.14 Safety measures

Markers, chevrons and/or reflectors should be installed at the ends of the bridges to clearly show the
extent of the structure and the vehicle path. Kerbs and ends of the structure should be painted white.
Warning signs should be installed on the bridge approaches in accordance with the local traffic sign
recommendations.

9.3 Site administration


The activities shown in Table E.9.12 will be required to be carried out in support of the site works. It is
important to keep accurate records of the actual works carried out to compare to the planned progress,
resource use and expenditure. This will help to ensure value for money. Any problems encountered
should be recorded along with explanations of how they were overcome. This will assist in explaining any
delays or cost over runs and help to improve future planning of structures.

Table E.9.12: Checklist of site administration tasks

‚ Set and review individual/gang task rates


‚ Daily muster roll and work achievement record for site labour force
‚ Daily diary of works achieved; problems encountered and methods of solving them should be
recorded
‚ Update work programme
‚ Daily checks on site stores, tools, materials, re-order as necessary
‚ Daily checks and service of site equipment
‚ Testing of materials, inspection and quality control
‚ Prepare payrolls
‚ Arrangements for payment of labour force
‚ Keep a careful record of all costs
‚ Reporting of progress to client/senior management
‚ Safety and first aid arrangements

Weekly and daily programmes should be prepared based on the Bills of Quantities, the overall works
programme and expected local productivities. Adjustments will be required to be made continuously
based on actual experience. Weekly reports should be prepared for management monitoring purposes.
Key indicators should be used to monitor the progress of the work, such as cement or worker days used
against the quantities planned.

For structures, ‘as-built’ drawings should be prepared; these should particularly record differences from
the original design, and important details such as actual foundation levels and concrete strengths. If there
are no changes, this is also valuable information.

A cost analysis of the completed structure should be carried out to enable cost estimating of future
structures to be more accurate.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 9 - 191

A final inspection of the completed structure should be carried out prior to handing over to the authority
that will be responsible for its maintenance. The following chapter discusses maintenance arrangements.

It is advisable for an independent performance audit to be carried out on a completed structure to review
the works. This should verify the structure’s ‘fitness-for-purpose’ and value for money.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 192

10. MAINTENANCE

Cross drainage structures usually account for a high proportion of the total cost of a road. They are the
potential weak points in a road network due to the possible damaging effects of floods and high water
flows being concentrated at the points where the water crosses the road. The failure of these structures
would result in high replacement costs and long delays due to the closure of the road. It is particularly
important therefore, that sufficient attention is given to structures to ensure that they are maintained in
good condition.

A culvert, bridge or other structure is an integral Part of the road, and its condition will affect the level
of service that the road provides. A structure should be designed so that no major repair works should
be required during its ‘design life’ (eg replacement of abutments, piers or deck structural members).
Eventually major works may be required such as a complete new timber bridge deck or safety barrier
replacement. However, the structure should be designed to provide many years of service through its
design life with only minor maintenance.

Importantly, if the maintenance is not carried out, there can be serious consequences for road users. It
can result in increased safety hazards, reduced quality of service or even loss of the structure and severing
of the transport link.

It is usually not possible with the resources available in developing countries to devise a ‘maintenance-
free structure’ for a watercourse crossing. However, application of the design and construction guidelines
contained in this Manual should reduce maintenance requirements to an acceptable and manageable
level.

Conversely, poor design or construction will result in an abnormally high requirement for maintenance, or
even eventual loss of the structure.

There are a number of aspects which should be appreciated in devising appropriate management and
maintenance arrangements for structures. This applies to consideration of an individual structure, or a
large number constructed at various locations on a road network.

ƒ Structures will often need no maintenance for periods of many months or sometimes even years.
ƒ Deterioration or damage to a structure can progress slowly (eg corrosion, attack by insects), or
suddenly (eg in a flood or vehicle accident).
ƒ The need for repairs may not be obvious to road users or through casual observation from the
road. However, the deterioration can progress, if not checked, to result in the need for major works
at great cost and requiring substantial unplanned resource mobilisation.
ƒ The resources for maintenance and repair of a typical structure are required intermittently, not
continuously.
ƒ It is usually most efficient to provide maintenance resources only when the structure requires
maintenance or repair works.

It is important to ensure the maintenance of a structure so that it remains in its intended condition,
providing the service and benefits to road users and the community that it was designed for. It is an asset
that needs to be managed.

10.1 Managing the structure


The maintenance works required to be carried out on a structure will range from basic seasonal clearing
of silt and debris to ensure it continues to function properly, through to replacement of components of
the structure when they are worn out or damaged. It can be expected that ALL structures will normally
require at least some basic maintenance each year.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 193

It is necessary to set up a management system to ensure that the structure stays in a condition that
it is able to carry out its function in a safe manner. In essence this ‘system’ should identify when work
needs to be carried out. From this assessment the maintenance funding and works can be arranged and
supervised to ensure that the maintenance is completed satisfactorily.

A system of inspections is required to identify any damage or deterioration of the structure, or problems
adjacent to the structure which may threaten its stability.

The key components of a structures management system are:


ƒ An inventory of all structures (ie What is the asset? What are its key features? These are management
records which generally do not change with time, except for new structures or after major structural
changes to an existing one);
ƒ An inspection system (to determine the condition and repair needs);
ƒ Arrangements for specifying, arranging, supervising, recording/reporting and paying for the
works. Arrangements should also be in place for checking the ‘value for money’ of maintenance
operations and expenditures.

TRL Overseas Road Note 7 provides comprehensive guidelines on the inspection and documentation of
inventory and condition information on structures. A paper based system is quite adequate. Computer
systems can help if the number of structures being managed is substantial and the operating environment
can support the maintenance of the computer system itself, including arrangements for the ongoing costs
and skilled resources required. In a limited resource environment it can be difficult to justify and secure
the recurring costs of administration, computer support personnel and inevitable software and hardware
upgrades required for a computer system. Ie there can be an undue and unnecessary dependence on
external resources.

Certain maintenance activities such as de-silting and removal of debris should be carried out under
a routine programme of works. For example, before the rainy season all silt should be removed from
culverts, their inlets and outlet channels. After the rains, and particularly after individual floods, silt
and debris should be cleared from structures to avoid later damage due to blockages or diversion/
concentration of water.

These routine clearing operations are an ideal opportunity to carry out an inspection of a structure. With
the scarcity and expense of engineering personnel, it is possible to train persons with limited education
(eg the gang leader) to carry out inspections and to alert engineering staff to situations that require their
action.

Inspections of ALL structures should be carried out after a flood situation as this is the most likely time for
damage to have occurred. Particular attention should be paid to identifying any movement, especially
at joints, cracking/spalling and assessing whether erosion has occurred around abutments and piers, or
at the ends of aprons. Where water is permanently standing against the structure, probing with ranging
rods, poles or plumb lines should be carried out to identify unseen scouring. A boat or raft may be
required for this inspection.

All structures, from culverts to bridges, should receive a documented routine inspection at least once
each year. As indicated above these can be carried out by relatively unskilled personnel if the appropriate
training is provided. Inspection records should be carefully filed for future reference. Even a report of ‘no
defects’ is important management information.

The management of a structure costs money and, even before a structure is built, the ongoing provision
of the funds and resources for the management (including inspections) as well as the maintenance of the
structure should be assured.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 194

10.2 Maintaining the structure


Structure maintenance activities can be grouped into regular routine maintenance (Table E.10.1) and
periodic major operations.

Table E.10.1: Structure - Routine

ƒ 1.Cleaning/clearing
‚ sweeping
‚ de-silting
‚ unblocking
‚ removal of vegetation and flood/wind borne debris).
‚ (This includes inlets and outlet channels as well as culvert openings themselves)
ƒ Repair of loose/missing connectors and fixings
ƒ Replacement of damaged/missing planks or kerbs
ƒ Painting
ƒ Wood preservation
ƒ Pointing/repair of masonry
ƒ Repair of parapets, marker posts, safety barriers and features/signs

Table E.10.2: Structure - Periodic

ƒ Random stone filling


ƒ Retaining wall repairs
ƒ Riverbed scour repairs
ƒ Gabion repairs
ƒ Structural repairs to the following defects:-
‚ structural timber decay, splitting or insect attack
‚ bulging masonry
‚ cracked concrete or masonry
‚ honeycombed concrete
‚ spalling concrete
‚ serious rust or chemical stains
‚ exposed or corroding reinforcement or pre-stressing steel
‚ damp patches on the concrete
‚ seriously corroded structural steelwork
‚ damaged/distorted structural steelwork
‚ loose structural rivets, bolts or other fixings
‚ cracks in structural steelwork
‚ settlement of deck, piers, abutments or wingwalls
‚ expansion joint or bearing defects
‚ erosion requiring piling works

Major repairs will generally require technical expertise for the design and supervision of remedial work.

Maintenance works should be planned, organised and supervised using the guidelines set out in the
previous construction chapter. Maintenance records should be kept for each structure, which include:
ƒ Estimates of work proposed;
ƒ Details of work carried out;
ƒ Date of completion of the repair;
ƒ Supervisor’s quality control reports;
ƒ Actual costs of repair.

Storage of information should be on a structure by structure basis so that the complete history can be
easily viewed.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 195

Further guidance on maintaining structures is provided in the PIARC International Road Maintenance
Handbook, Volume 4. The handbook includes advice on the defects, resources and maintenance methods
involved.

10.3 Common maintenance requirements


Damage due to scour and erosion is the most likely cause of major or unrepairable damage to a structure.
Once scour or erosion around a structure starts the damage can increase very rapidly. It is therefore
essential that maintenance is carried out quickly to prevent further structural damage.

10.3.1 Drifts

The drift must maintain a firm roadway across the width of the river which is not covered by debris or
eroded by the flood water. The face of the river embankments should also be protected against scour and
erosion. It may be possible to encourage the growth of vegetation along the banks to improve the bank
stability and prevent erosion. The common maintenance issues to address are:
ƒ Cracking of the slab;
ƒ Undercutting on the downstream side;
ƒ Erosion at ends of slab where it is not extended above high flood levels;
ƒ Lack of downstream protection;
ƒ Guidestones knocked off.

10.3.2 Culverts

The most common maintenance problem associated with culverts is blockage due to silt and other debris.
Hence regular cleaning of the culverts is essential as illustrated in Plates E.10.1 and E.10.2. A blocked
culvert can result in damage to the road in 3 ways:
1. Water can seep into the subgrade of the road and reduce its strength. The road will tend to subside
and the road surface will break up.
2. The water can undermine the head and wingwall of the culvert causing it to collapse. The road
embankment will then be unsupported and rapidly subside.
3. In an extreme case the water level may continue to increase until the water floods over the road.
The road may then become impassable and major damage occur as the water erodes the road and
culvert. Ultimately the road will be washed away and a large gully will be scoured across the road.

Water discharging from culverts with excessive velocity will erode the stream bed and possibly undermine
the whole structure. It is therefore essential to provide some form of protection to the beds below the
outlet of a culvert. The protection is usually in the form of a masonry apron. It may also be necessary to
prevent erosion of the watercourse itself further downstream of the culvert. Bio-engineering planting may
be an appropriate and low-cost solution.

Plate E.10.1 and Plate E.10.2: Culvert cleaning tool

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 196

The main cause of blockage of culverts is by water carried debris and vegetation (Plate E.10.3). Larger
debris tends to collect at the entrance to the culvert causing blockage at the headwall, while silt is
deposited in the culvert barrel. If unchecked this silt can build up until it fills the culvert barrel. Long grass
at the outlet of a culvert can cause silting at the outlet and eventual blockage of the culvert. It is therefore
necessary, particularly before the rainy season, to clean culvert barrels, inlets and outlets to allow water
to flow freely through the culvert.

Plate E.10.3: Box culvert partially blocked by vegetation

Any material removed from a culvert should be disposed of downstream of the culvert to prevent it
washing back into the structure. Other common defects that require maintenance to be carried out on
culverts include:
ƒ Downstream erosion;
ƒ Headwall knocked down or damaged;
ƒ Outfall channel eroded or silted;
ƒ Undercutting of the culvert outfall apron;
ƒ Ponding of water at the inlet and/or outlet causing subsidence of the road embankment.

10.3.3 Vented drifts and large bore culverts

The common maintenance requirements with vented fords and large bore culverts are similar to culverts
and drifts (Plate E.10.4). In addition to the issues discussed above the following defects may need to be
corrected during maintenance:
ƒ floating debris, such as tree branches, can block the culvert barrels;
ƒ cracking and breaking of roadway slab;
ƒ cracking and breaking of structure faces.

Plate E.10.4: Cleaning the drop inlet and barrel of a large bore culvert

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 197

10.3.4 Bridge Maintenance

The substantially larger investment in bridges, compared to smaller structures, justifies greater attention
to inspection and maintenance of these vital road components. The consequences of failure due to lack
of maintenance can be routes being severed for considerable period with serious adverse social and
economic outcomes. It is therefore important to establish a regime of regular bridge inspections and
active maintenance response works.

Inspections should be formulated on a regular basis to identify and initiate maintenance repairs for the
following possible defects:
ƒ Minor Defects (non-structural)
• Accumulation of dirt or soil on bridge deck;
• Blocked scuppers;
• Stones, soil or dirt in joints or around bearings;
• Vegetation or soil in weep holes or in backfill drainage outlets;
• Flood debris at or under bridge;
• Wind-blown debris on or under bridge.
ƒ Minor Defects (structural)
• Loose or missing nailed/bolted connectors or fixings;
• Damaged running boards or deck planks;
• Rusty steel, faded paint;
• Untreated wood;
• Defective masonry joints.
ƒ Minor Defects (Safety)
• Damaged safety barrier or parapet;
• Damaged warning signs.

The following defects may require major repair works and specialist expertise to ensure appropriate
remedial works. The inspection system should initiate mobilisation of the necessary expertise when these
defects are identified:
ƒ Major Defects
• Scour adjacent to structure;
• Structural timber decay, splitting or insect attack;
• Bulging masonry;
• Cracked concrete or masonry;
• Honeycombed concrete;
• Spalling concrete;
• Serious rust or chemical stains;
• Exposed or corroding reinforcement;
• Damp patches on the concrete;
• Seriously corroded structural steelwork;
• Damage/distorted structural steelwork;
• Loose structural rivets, bolts or other fixings;
• Cracks in structural steelwork;
• Settlement of deck, piers or abutments;
• Erosion requiring piling works;
• Vehicle Impact damage (especially to steel panel bridges).

Further guidance on bridge and structures maintenance is available from the PIARC International Road
Maintenance Handbook Volume IV.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 10 - 198

Plate E.10.5: Serious vehicle impact damage to steel panel bridge

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 11 - 199

11. REFERENCES

The references below may be used to supplement the information contained in this Manual. In order
to assist readers in the selection of relevant additional information the following information has been
provided with each reference;
ƒ list of the topics covered
ƒ a brief review of the issues discussed in each reference
ƒ contact details of selected publishers

Many of the documents and further information on specific topics may be accessed on the global
Transport Knowledge Partnership website: www.gtkp.com

Berger L, Greenstein J, Arrieta J, 1987, Guidelines for the Design of Low Cost Water Crossings, TRR
1106, Transportation Research Board, Washington
(Finance, Bridges, Design, Materials) pp10
ƒ This paper reviews the different designs and materials used for the construction of bridges on low
volume roads in Central America. The article indicates that the standards used on these roads need
only be suitable for vehicles up to 10 tonnes, resulting in major cost savings when compared with
full specifications. The use of gravelled fords, split deck concrete bridges and timber bridges are
discussed. The paper outlines a design for timber bridge decks.

Brandon T, 1989, River Engineering – Part 2, Structures and Coastal Defence Works, Institute of Water
and Environmental Management, London
(Design, Hydraulics, Maintenance) pp332
ƒ This book is the second of two volumes and covers the design, construction and maintenance of
water structures. It is primarily concerned with the design of river control structures such as locks,
weirs and sluices, but also includes limited discussion on river protection, culverts and bridges.
There are two useful chapters on maintenance issues and construction planning and management.

Clark J, Hellin J, 1996, Bio-Engineering for Effective Road Maintenance in the Caribbean, Natural
Resources Institute, Chatham, UK
(Environmental, Materials, Erosion Protection, Slope Stability) pp122
ƒ This book discusses the use of vegetation for the control of erosion and stabilising slopes, indicating
the functions different types of vegetation can perform. Six simple techniques are described, along
with the vegetation species that may be used, which are useful in the road sector for drainage control.
The book also contains a large section which gives background details of eleven species which
are suitable for bio-engineering. These species are normally found in the Caribbean, however the
description of the species and specification of different planting material should allow practitioners
in other areas to make use of the information.

Dzung Bach The, Petts Robert, 2009, Report on Rice Husk Fired Clay Brick Road Paving, Vietnam, global
Transport Knowledge Partnership
ƒ This report documents the experiences in Vietnam of the established practice of burning high
quality clay bricks for building and road works use. The flexibility of the process suggests that other
agricultural wastes could be used to produce high quality clay bricks for structures and other road
works uses.

Ethiopian Roads Authority Bridge Design Manual - 2011


ƒ This manual is the principal design reference for highway and road structures in Ethiopia.

Farraday R, Charlton F, 1983, Hydraulic Factors in Bridge Design, Hydraulics Research, Wallingford
(Bridges, Design, Hydraulics, Erosion Protection) pp102
ƒ This book explains in fairly simple terms the different hydraulic issues which need to be addressed
when designing bridges over rivers. It describes the data which needs to be collected and a step by
step design process which must be undertaken to ensure that bridges will be able to withstand the

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 11 - 200

loads exerted by the water and changing flow patterns due to scour of the river bed. Each chapter
is extensively referenced.

Flavell D. (ed), 1994, Waterway Design, Austroads, Sydney


(Bridges, Culverts, Design, Hydraulics) pp138
ƒ This book provides guidance on the selection of design floods required for the various aspects
of the design of waterway structures and the hydraulic design of bridges, culverts and floodways.
It also provides information for the design of works required to protect these structures from the
effects of scour.

Gupta D P, 1997, Manual on Route Location, Design, Construction and Maintenance of Rural Roads,
Special Publication 20, Indian Roads Congress, New Delhi
(Design) pp108
ƒ This book primarily covers the design and construction of roads, however it has two chapters
covering drainage and cross drainage structures. Other sections of the book highlight design issues
which are affected by highway structures. This book would be a useful reference if structures were
to be designed on a new road.

Heyman J, 1980, The Estimation of the Strength of Masonry Arches, Proc. Institution of Civil Engineers
Part 2 Dec 1980
(Design) pp921-937
ƒ This paper discusses the development of the simplified method for estimating the strength
of masonry arches by the military load classification. It discusses the mathematical proof of the
assumptions made and explains that the strength of an arch is closely related to its span and crown
thickness. The paper suggests that nomographs could be used to predict strength with correction
factors used to account for span/rise ratio, mortar condition and quality of material used.

Hindson J, 1983, Earth Roads: Their Construction and Maintenance, IT Publications, London
(Culverts, Design, Drifts, Erosion Protection, Site Construction) pp124
ƒ This book covers the design and construction of earth roads for traffic up to about 50 vehicles per
day. It concentrates on the control of water through drainage control measures. The first half of the
book deals with the theory of road design which includes splashes, drifts and culverts. The second
half of the book deals with the techniques of construction offering different solutions for different
topological conditions. The second half of the book also includes a section on maintenance.

ILO, 1991, Stone Masonry, (Training Element and Technical Guide for SPWP Workers Booklet 2), UNDP/
ILO, Geneva
Materials, Site Construction) pp84
ƒ This booklet covers the design, construction and maintenance of small masonry structures which
include culverts and small headwalls. It may be used as a technical manual for site personnel or as
the basis for a training course for site supervisors in the use of masonry for construction.

ILO, 1986, Gabions, (Training Element and Technical Guide for SPWP Workers Booklet 3), UNDP/ILO,
Geneva
(Materials, Site Construction) pp84
ƒ This booklet is similar to its predecessor covering stone masonry. It does not specifically deal with
highway structures but highlights the uses of gabions and how they should be constructed.

Jayanetti, L. 1990, Timber Pole Construction, IT Publications, UK


(Materials) pp64
ƒ Although not focused on road structures this book provides information to designers of timber pole
bridges.

Jones T, Parry J, 1993, Design of Irish Bridges, Fords and Causeways in Developing Countries, Highways
and Transportation (Jan 1993), Institution of Highways and Transportation, London
(Drifts) pp28-33

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 11 - 201

ƒ This article explains the differences between the different types of low level water crossing. It
discusses site selection and the materials which may be used for different crossings. The article
also describes with a series of photographs and diagrams the key design points of the different
structures.

Kadam S.P, 1993, Vented Paved Dips for Rural Roads, Indian Highways
(Design, Hydraulics, Drifts) pp14
ƒ This paper discusses the use of vented fords on rural roads to allow for monsoon rains. It concentrates
on the issues of scour downstream of the structure but also discusses the hydraulics of a vented ford
when it is being overtopped. The optimum dimensions for a structure are provided which include a
standard construction drawing.

Khanna, P.N. , 1996, Indian Civil Engineer’s Practical Handbook, 15th Edition, Engineer’s Publications,
New Delhi.
(Design, Site Construction)
ƒ This book provides a wide range of practical information for engineers involved in the design,
planning and construction of civil engineering projects. The handbook contains a chapter specifically
covering the engineering aspects of roads and road structures, in addition to chapters covering
surveying, setting out, material properties and basic design principles.

Lal G, 1995, Guidelines for the Design of Small Bridges and Culverts, Special Publication 13, Indian
Roads Congress, New Delhi
(Bridges, Culverts, Design, Hydraulics, Drifts, Erosion Protection) pp176
ƒ This book covers the complete design of small bridges and culverts from the collection of the initial
design data to the preparation of construction drawings. It concentrates on the mathematics of the
estimation of maximum water flows and scour around structural supports. However, other empirical
results and solutions are also described throughout the book to simplify the design process where
applicable.

Morris, J, 1995, Earth Roads, Avebury


(Bridges, Culverts, Drifts, Materials, Maintenance, Erosion Protection, Site Construction) pp304
ƒ This book is a practical guide for managers and engineers of agricultural estates to provide guidelines
and advice on how roads can meet the needs of their commercial operation. It concentrates on earth
and other unsealed roads in developing countries, but has extensive sections covering bridges and
culverts. The majority of the solutions discussed make use of timber which is likely to be available as
a by-product from the agricultural operations.

PIARC, 1994, International Road Maintenance Handbook, Transport Research Laboratory (for the World
Road Association (PIARC)), UK
Vol. 1 Maintenance of Roadside Areas and Drainage,
Vol. 2 Maintenance of Unpaved Roads,
Vol. 3 Maintenance of Paved Roads,
Vol. 4 Maintenance of Structures and Traffic Control Devices (Maintenance)
ƒ These four handbooks are aimed at the supervisors of road maintenance contracts. They explain
the causes and the measures required to prevent road deterioration. Each maintenance task is
addressed in turn with simple text and illustrations to show the labour and tools required to carry
out the task.

Shadmon, A, 1989, Stone; An Introduction, IT Publications, London


(Materials), pp184
ƒ This book provides a good introduction to the extraction and use of both field and cut stone. It
describes the different types of stone and outlines tests that can be carried out to determine the
tensile and compressive strength of stone samples.

Spence, R, and Cook D, 1983, Building Materials in Developing Countries, John Wiley and Sons.
(Materials) 356pp

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 11 - 202

ƒ This book provides practical information about various building materials commonly
available in developing countries. The majority of the information is focused on housing
construction. However, a significant proportion of the information available will be useful
to designers of road structures.

Stern, P. et al, 1983, Field Engineering, IT Publications, London


(Bridges, Culverts, Design, Maintenance, Site Construction) pp272
ƒ This book is aimed at individuals working on engineering projects in rural areas. It has a
detailed section dealing with site survey and setting out techniques. In addition to sections
on roads, simple river crossings and bridges, it also covers water supply, sanitation and
small dams.

Stulz, R and Mukerji, K, 1993, Appropriate Building Materials, SKAT, St Gallen, Switzerland.
(Materials) 456pp
ƒ This book provides technical data and practical information about various building
materials for low cost construction. The majority of the information is focused on non road
structures. However, a significant proportion of the information available will be useful to
designers of road structures.

Thagesen B (ed), 1996, Highway and Traffic Engineering in Developing Countries, E & FN
Spon, London
(Culverts, Design, Maintenance, Site Construction) pp485
ƒ This textbook covers the planning, design, construction, maintenance and management
of roads in tropical developing countries. It contains a section on drainage design which
covers hydrology and hydraulic design and another section which discusses maintenance
strategies and management.

TRL, 1988, Bridge Inspectors Handbook Vol. 1 & 2, (ORN7), Transport Research Laboratory,
Crowthorne, UK
(Bridges, Culverts, Maintenance) pp40 & pp250
ƒ The object of these two volumes is to allow a district engineer to establish and operate
an effective bridge and culvert record system. The guide explains the principles of record
keeping and contains a series of proforma record sheets. The pocket size handbook (Vol.
2) deals with the actual inspection, highlighting, through the use of photographs and
drawings, the items which should be checked and recorded.

TRL, 1992, A Design Manual for Small Bridges, (ORN9), Transport Research Laboratory,
Crowthorne, UK
(Bridges, Culverts, Design, Hydraulics, Drifts, Materials, Maintenance, Erosion Protection, Site
Construction) pp223
ƒ This manual prepared by TRL offers a comprehensive set of guidelines to highways
engineers for the design of small bridges and culverts. It covers the whole process from
the planning stage to the final preparation of detailed specifications and drawings. It is
intended for practising engineers who may not be highway specialists. The designs which
are discussed in the manual are appropriate for relatively large roads or traffic flows and
predominantly utilise reinforced concrete.

TRL, 1997, Principles of Low Cost Road Engineering in Mountainous Areas (ORN16), Transport
Research Laboroatory, Crowthorne, UK
(Culverts, Environmental, Erosion Protection, Slope Stability) pp150
ƒ This manual describes and explains techniques for designing, constructing and maintaining
roads in mountainous areas. It contains sections on drainage and retaining walls as well as
methods to control erosion and maintain slope stability.

Tufnell R, 1995, Dry Stone Causeways, Appropriate Technology Vol. 22 No. 1, IT Publications,
London
(Drifts, Materials) pp3

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


E - Chapter 11 - 203

ƒ This article explains in simple practical terms, with sketches, how to build a masonry causeway or
vented ford.

Watkins L and Fiddes D, Highway and Urban Hydrology in the Tropics, Pentech Press, London

Selected publishers’ addresses:

ƒ Austroads:
Suite 2, Level 9, 287 Elisabeth St, Sydney, NSW 2000, Australia
Fax: +02 264 1657, www.austroads.com.au/

ƒ Global Transport Knowledge Partnership (gTKP), Chemin de Blandonnet 2, 1214 Vernier, Genève,
Switzerland, www.gtkp.com

ƒ Hydraulics Research:
Hydraulics Research Station Ltd, Wallingford, Oxfordshire, UK,
www.hrwallingford.co.uk

ƒ International Labour Organisation (ILO):


4 route des Morillons, Geneva, CH 1211, Switzerland,
www.ilo.org/public/english/

ƒ Indian Roads Congress (IRC):


Sector 6, (Near RBI Quarters), R K Puram, New Delhi- 110022, India,
www.irc.org.in
Institute of Agricultural Engineers:
The Bullock Building, University Way, Cranfield, Bedford, MK43 0GH,
UK, www.iagre.org

ƒ Institution of Civil Engineers (ICE):


1 Great George St, Westminster, London, SW1A 3AA, UK, www.ice.org.uk
ƒ Intermediate Technology Publications / Practical Action Publishing:
The Schumacher Centre for Technology & Development, Bourton on Dunsmore, Rugby,
Warwickshire CV23 9QZ, United Kingdom, www.practicalactionpublishing.org/publishing

ƒ Natural Resources Institute (NRI):


University of Greenwich at Medway, Central Avenue, Chatham Maritime, Kent, ME4 4TB
United Kingdom, www.nri.org

ƒ Transportation Research Board (TRB):


The National Academies, 500 Fifth Street, NW, Washington, DC 2000, USA,
www.trb.org

ƒ Transport Research Laboratory (TRL):


Crowthorne House, Nine Mile Ride, Wokingham, Berkshire, RG40 3GA,
United Kingdom, www.trl.co.uk

12.

PART E: EXPLANATORY NOTES AND DESIGN STANDARDS FOR SMALL STRUCTURES


FEDERAL
EDERAL DEMOCRATIC REPUBLIC OF

ETHIOPIAN ROADS AUTHORITY

IT Y
E TH

OR
IO

IA
NR UT
H
P

OADS A

DESIGN MANUAL FOR LOW VOLUME ROADS


TRAIL BRIDGES MANUAL - PART F
FINAL DRAFT, JULY 2011
Part F
TRAIL BRIDGES MANUAL

Introduction

Part F
Trail Bridges Manual Suspended Trail Bridge

Suspension Trail Bridge


F - ii

F TABLE OF CONTENTS

F. TABLE OF CONTENTS.......................................................................................................... ii
F. LIST OF TABLES....................................................................................................................v
F. list of figures................................................................................................................vii
1. INTRODUCTION................................................................................................................. F.1
1.1 Survey and bridge site selection.......................................................................................F.1
1.2 Preparation for survey.......................................................................................................F.2
1.3 General data collection.....................................................................................................F.2
1.3.1 Location of Bridge Site.......................................................................................F.3
1.3.2 Nature of crossing and fordability .....................................................................F.3
1.3.3 Traffic volume ....................................................................................................F.3
1.3.4 Width of walkway................................................................................................F.4
1.3.5 Local participation..............................................................................................F.4
1.3.6 Transportation distance......................................................................................F.4
1.3.7 Availability of local materials..............................................................................F.5
1.3.8 Availability of local bridge builders....................................................................F.5
1.3.9 Temporary crossing.............................................................................................F.5
1.4 Bridge site selection..........................................................................................................F.5
1.4.1 General condition...............................................................................................F.6
1.4.2 River condition....................................................................................................F.8
1.4.3 Slope and bank condition...................................................................................F.9
1.4.4 Evaluation of the bridge site............................................................................F.14
1.4.5 Classification of soil and rock ..........................................................................F.15
1.4.6 Identification of soil and rock ..........................................................................F.17
1.5 Topographic survey.........................................................................................................F.18
1.5.1 Survey area.......................................................................................................F.18
1.5.2 Setting of bridge centerline..............................................................................F.19
1.5.3 Survey methods................................................................................................F.20
1.5.4 Survey by abney level.......................................................................................F.20
1.5.5 Survey by Theodolite........................................................................................F.25
1.5.6 Topographic maps............................................................................................F.32
1.5.7 Photographs.....................................................................................................F.36
1.5.8 Survey report....................................................................................................F.36
1.6 Calculation of quantity and cost estimate.......................................................................F.36
1.7 Bridge maintenance........................................................................................................F.37
1.7.1 Introduction......................................................................................................F.37
1.7.2 Routine maintenance........................................................................................F.37
1.7.3 Major maintenance..........................................................................................F.38
2. SUSPENDED TRAIL BRIDGE............................................................................................ F.39
2.1 Presentation of the suspended type...............................................................................F.39

PART F: TRAIL BRIDGE MANUAL


F - iii

2.2 Technical features and limitations of suspended type....................................................F.39


2.3 Basic design concept......................................................................................................F.42
2.3.1 Loadings...........................................................................................................F.42
2.3.2 Construction materials......................................................................................F.43
2.3.3 Structural analysis and design..........................................................................F.46
2.4 Design of standard Suspended bridge...........................................................................F.51
2.4.1 The major bridge components.........................................................................F.51
2.4.2 Design procedure.............................................................................................F.51
2.4.3 Designing the position of the bridge foundations...........................................F.52
2.4.4 Cable Design....................................................................................................F.57
2.4.5 Design of bridge foundation structures............................................................F.59
2.4.6 Other structures................................................................................................F.65
2.4.7 General arrangement drawing..........................................................................F.71
2.4.8 Design example................................................................................................F.72
2.5 Bridge standard drawings...............................................................................................F.73
2.5.1 Introduction and overview of drawings............................................................F.73
2.5.2 Concept of the standard drawings...................................................................F.75
2.5.3 Relationship between construction and steel drawings...................................F.79
2.6 Calculation of quantity and cost estimate.......................................................................F.81
2.6.1 Implementation procedure ..............................................................................F.81
2.7 Construction....................................................................................................................F.82
2.7.1 Bridge layout....................................................................................................F.82
2.7.2 Foundation excavation.....................................................................................F.84
2.7.3 Local material collection...................................................................................F.86
2.7.4 Transportation and storage of the materials.....................................................F.87
2.7.5 Masonry and Stone dressing work....................................................................F.89
2.7.6 Cement works...................................................................................................F.92
2.7.7 Cable hoisting and sag setting.........................................................................F.98
2.7.8 Finalizing the cable anchorage.......................................................................F.100
2.7.9 Walkway fitting...............................................................................................F.100
2.7.10 Water management backfilling and general finishing works..........................F.101
2.8 Steel Drawings...............................................................................................................F.111
2.9 Construction Drawings..................................................................................................F.112
3. SUSPENSION TRAIL BRIDGE ........................................................................................ F.114
3.1 Background...................................................................................................................F.114
3.2 Presentation of the suspension bridge type..................................................................F.114
3.3 Technical features and limitations.................................................................................F.115
3.4 Bridge Design................................................................................................................F.116
3.5 Basic design concept....................................................................................................F.118
3.5.1 Loadings.........................................................................................................F.118
3.5.2 Construction materials....................................................................................F.119

PART F: TRAIL BRIDGE MANUAL


F - iv

3.5.3 Structural analysis and design........................................................................F.123


3.6 Design of standard suspension bridge..........................................................................F.140
3.6.1 The major components of the suspension type bridge..................................F.140
3.6.2 Design procedure...........................................................................................F.140
3.6.3 Designing the position of the walkway & tower foundations........................F.142
3.6.4 Cable design & standard tower selection.......................................................F.147
3.6.5 Design of the main cable anchor....................................................................F.153
3.6.6 Walkway and tower foundation design .........................................................F.158
3.6.7 Suspender list.................................................................................................F.161
3.6.8 List of drawings...............................................................................................F.161
3.6.9 Other structures..............................................................................................F.161
3.6.10 General arrangement drawing........................................................................F.170
3.7 Design example.............................................................................................................F.171
3.8 Bridge standard drawing...............................................................................................F.172
3.8.1 Introduction & overview of drawings..............................................................F.172
3.8.2 Construction drawings....................................................................................F.172
3.8.3 Concept of the standard drawings.................................................................F.176
3.8.4 Relationship between construction and steel drawings.................................F.180
3.9 Construction..................................................................................................................F.181
3.9.1 Bridge layout..................................................................................................F.181
3.9.2 Foundation excavation...................................................................................F.185
3.9.3 Cement works.................................................................................................F.186
3.9.4 Bridge erection...............................................................................................F.191

PART F: TRAIL BRIDGE MANUAL


F-v

F LIST OF TABLES

Table F.1.1: Soil classification........................................................................................................... F.15


Table F.1.2: Rock Classification......................................................................................................... F.17
Table F.2.1: Selection of RCC deadman & gravity soil anchor block in flat ground......................... F.63
Table F.2.2: Selection of RCC Deadman & Gravity Soil Anchor Block in Hill Slope......................... F.64
Table F.2.3: Selection of RCC Single Drum Anchor in Hard Rock.................................................... F.64
Table F.2.4: Selection of RCC Double Drum Anchor in Hard Rock.................................................. F.64
Table F.2.5: Selection of RCC Single Drum Anchor in Fractured Hard Rock/Soft Rock................... F.64
Table F.2.6: Selection of RCC Double Drum Anchor in Fractured Hard Rock/Soft Rock.................. F.65
Table F.2.7: Selection of RCC Deadman Anchor in Fractured Hard Rock/Soft Rock........................ F.65
Table F.2.8: Suspended Bridge Construction Drawings................................................................... F.73
Table F.2.9: Suspended Bridge Steel Drawings............................................................................... F.74
Table F.2.10: Legend for the drawing numbers and suffixes:............................................................. F.75
Table F.2.11: Quantities for various Types of Cement Works............................................................. F.93
Table F.2.12: Calculating Elevation of Low Point for Cable Hoisting................................................ F.99
Table F.2.13: Selection of Gravity Soil Anchor Block in Flat Ground................................................ F.107
Table F.2.14: Selection of Gravity Soil Anchor Block in Hill Slope .................................................. F.108
Table F.2.15: Selection of RCC Single Drum Anchor in Hard Rock.................................................. F.108
Table F.2.16: Selection of RCC Double Drum Anchor in Hard Rock................................................ F.108
Table F.2.17: Selection of RCC Single Drum Anchor in Fractured Hard Rock/Soft Rock................. F.108
Table F.2.18: Selection of RCC Double Drum Anchor in Fractured Hard Rock/Soft Rock................ F.109
Table F.2.19: Selection of RCC Deadman Anchor in Fractured Hard Rock/Soft Rock...................... F.109
Table F.3.1: Selection of Cables and Standard Tower.................................................................... F.149
Table F.3.2: Selection of the main cable deaman anchor in soil & flat ground.............................. F.157
Table F.3.3: Selection of the main cable deadman anchor in soil & flat ground............................ F.157
Table F.3.4: Selection of the main cable anchor block in soil & hill slope...................................... F.157
Table F.3.5: Selection of the main cable anchor block in soil & hill slope...................................... F.158
Table F.3.6: Selection of the Main Cable Anchor Block in Hard, Soft or Fractured Rock............... F.158
Table F.3.7: Selection of the Main Cable Drum Anchor in Hard Rock............................................ F.158
Table F.3.8: Selection of the Main Cable Drum Anchor in Fractured Hard Rock or Soft Rock....... F.158
Table F.3.9: Selection of walkway & tower foundations in soil....................................................... F.160
Table F.3.10: Selection of walkway & tower foundations in hard rock.............................................. F.161
Table F.3.11: Selection of walkway & tower foundations in fractured hard rock or soft................... F.161
Table F.3.12: Selection of windguy cable anchor block in soil......................................................... F.167
Table F.3.13: Selection of windguy cable anchor block in all types of rock..................................... F.167
Table F.3.14: Selection of windguy cable drum anchor in hard rock................................................ F.167
Table F.3.15: Selection of windguy cable drum anchor in fractured or soft rock............................. F.167
Table F3.16: Construction Drawings for Main Cable Anchors......................................................... F.173
Table F.3.17: Construction drawings for walkway and tower foundations ...................................... F.174

PART F: TRAIL BRIDGE MANUAL


F - vi

Table F.3.18: Construction Drawings for Windguy Cable Anchors ................................................. F.174
Table F.3.19: Summary of All Construction Drawings...................................................................... F.175
Table F.3.20: Summary of All Steel Drawings................................................................................... F.176
Table F.3.21: Quantities for Various Types of Cement Works........................................................... F.187
Table F.3.22: Selection of Cables & Standard Towers...................................................................... F.204
Table F.3.23: Selection of Main Cable Deadman Anchor in Soil & Flat Ground.............................. F.209
Table F.3.24: Selection of Main Cable Deadman Anchor in Soil & Flat Ground.............................. F.209
Table F.3.25: Selection of Main Cable Deadman Anchor in Soil & Hill Slope.................................. F.210
Table F.3.26: Selection of Main Cable Anchor Block in Soil & Hill Slope......................................... F.210
Table F.3.27: Selection of Main Cable Anchor Block in Hard, Soft or Fractured Rock..................... F.210
Table F.3.28: Selection of Main Cable Drum Anchor in Hard Rock.................................................. F.210
Table F.3.29: Selection of Main Cable Drum Anchor in Fractured Hard/Soft Rock.......................... F.210
Table F.3.30: Selection of Walkway & Tower Foundation in Soil...................................................... F.212
Table F.3.31: Selection of Walkway & Tower Foundation in Hard Rock........................................... F.212
Table F.3.32: Selection of Walkway & Tower Foundation in Fractured Hard Rock or Soft Rock....... F.213
Table F.3.33: Selection of Windguy Cable Anchor Block in Soil...................................................... F.215
Table F.3.34: Selection of Windguy Cable Anchor Block in All Types of Rocks............................... F.215
Table F.3.35: Selection of Windguy Cable Drucm Anchor in Hard Rock.......................................... F.215
Table F.3.36: Selection of Windguy Cable Drum Anchor in Fractured of Soft Rock........................ F.215

PART F: TRAIL BRIDGE MANUAL


F - vii

F list of figures

Figure F.1.1: Process of Survey for D-type and N-type Bridges.......................................................... F.1
Figure F.1.2: Ratio of coarse grains.................................................................................................... F.16
Figure F.1.3: Bridge axis profile......................................................................................................... F.34
Figure F.1.4: Contour plan................................................................................................................. F.35
Figure F.2.1: Suspended type bridge, Keleta, Oromia region........................................................... F.39
Figure F.2.2: The typical profile of Suspended type bridge.............................................................. F.40
Figure F.2.3: The typical plan of Suspended type bridge.................................................................. F.41
Figure F.2.4: Cross-Section & lay of Wire Rope................................................................................. F.44
Figure F.2.5: Bulldog Grip.................................................................................................................. F.44
Figure F.2.6: Cable Geometry............................................................................................................ F.46
Figure F.2.7: Walkway Structure......................................................................................................... F.47
Figure F.2.8: Foundation in Soil......................................................................................................... F.48
Figure F.2.9: Major Bridge Components & Parameters..................................................................... F.51
Figure F.2.10: RCC Deadman & gravity soil anchor block on flat topography.................................... F.59
Figure F.2.11: RCC Deadman & gravity soil anchor block on slope topography................................ F.59
Figure F.2.12: RCC Single Drum Rock Anchor Block in Hard Rock...................................................... F.60
Figure F.2.13: RCC Double Drum Rock Anchor Block in Hard Rock.................................................... F.60
Figure F.2.14: RCC Single Rock Anchor Block in Fractured Hard Rock & Soft Rock............................ F.60
Figure F.2.15: RCC Double Drum Rock Anchor Block in Fractured Hard Rock & Soft Rock................ F.61
Figure F.2.16: RCC Deadman Anchor Block in Fractured Hard Rock and Soft Rock........................... F.61
Figure F.3.1: Suspension type bridge, Endamayno, Tigray Region................................................. F.114
Figure F.3.2: A Typical Profile of Suspension Type Bridge............................................................... F.116
Figure F.3.3: A Typical Plan of Suspension Type Bridge.................................................................. F.117
Figure F.3.4: The Geometry of the Suspension Bridge ................................................................... F.123
Figure F.3.5: Main Cable Deadman Anchor Block with Turnbuckle on Flat Topography................ F.153
Figure F.3.6: Main cable anchor block with direct cable connection on flat topography ............... F.154
Figure F.3.7: Main cable gravity anchor block with turnbuckle on hill slope topography............... F.154
Figure F.3.8: Main cable gravity anchor block with direct cable connection on hill slope .............. F.154
Figure F.3.9: Main cable drum anchor block in hard rock................................................................ F.155
Figure F.3.10: Main cable drum anchor block in fractured hard rock or soft rock............................. F.155
Figure F.3.11: Layout of windguy arrangement................................................................................. F.162
Figure F.3.12: Gravity anchor block in soil......................................................................................... F.163
Figure F.3.13: Gravity anchor in rock................................................................................................. F.164
Figure F.3.14: Drum anchor block in hard rock.................................................................................. F.164
Figure F.3.15: Drum Anchor Block in Fractured Hard Rock & Soft Rock............................................ F.164
Figure F.3.16: Assembly & Layout of Tower No. 4............................................................................. F.193
Figure F.3.17: Typical Design of Suspender....................................................................................... F.216

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 1

1. INTRODUCTION

This Manual provides technical guidelines for the construction of unstiffened pedestrian suspended
and suspension type cable bridges (catwalk). It applies for suspension bridges, which require higher
towers or pylons on both sides of the bridge. This Manual follows the actual step-by-step processes of
constructing a suspended (D-type) and suspension (N-type) bridges. It starts with preparatory work for
the site assessment and survey, design and cost estimate and ends with practical guidelines for bridge
construction and maintenance. The Manual is structured into three Chapters. Chapter 1 deals with the
technical and social issues which are common to D-type and N-type bridges. Chapter 2 and 3 deal with
details of specific technical issues of D-type and N-type bridges, respectively. All relevant specific forms
are included at the last section of each chapter. The drawings and Bill of Quantity calculation sheet and
specific steel and construction drawings are found in a separate folder. The drawings are done using a
software called ‘’Claris Draw’’ and one has to install this software in order to open the drawings.

This Manual along with the Forms and the Drawings is intended to give quick and reliable technical methods
of surveying, designing and constructing simple pedestrian bridges for engineers and technologists.

1.1 Survey and bridge site selection


Careful Surveys and Bridge Site Assessments are the basis for proper planning and designing and form
the main source for successful bridge construction. The main objective of the Survey and Bridge Site
Assessment is to identify the proper bridge site by considering socio-economic as well as technical points
of view. Survey and Bridge Site Assessment is done in the two steps:

Step 1 is Social Feasibility Survey and Step 2 is Technical Survey. Both surveys are of equal importance.
The social feasibility survey establishes community ownership and responsibility, and the technical survey
ensures that bridge construction is sound and safe.

Bridge
Request

Socio-feasibility survey

No Is the
End bridge
feasible

Yes

Technical Survey

Study of general condition


Bridge site selection
Select another site

Study of River condition

Study of Slope & Bank

Evaluation

Unfavourable Questionable and


Result no alternative site

Conduct detailed
Favourable

geological survey
Refer SBD Survey Manual

Favourable
Topographic Survey
Result

Survey Report
Unfavourable

Design Reject
the site

Figure F.1.1: Process of Survey for D-type and N-type Bridges

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 2

Step 1: Social Feasibility Survey


A Social Feasibility Survey is necessary to justify the construction of a requested bridge. For ranking and
prioritizing the vast number of requests, the socio-economic indicators which have utmost importance
are:
¡¡ Level of local participation;
¡¡ Size of area of influence;
¡¡ Size of traffic flow;
¡¡ Socio-economic benefits produced by the proposed bridge.

The first step for conducting a social feasibility survey is to introduce the participants, the survey team and
other groups who will be involved in the process of bridge construction. This is best done in the form of
a mass meeting right at the spot, or nearby the place, where the bridge is going to be built. Agenda of
the mass meeting should consist of:
¡¡ Verification of the proposed bridge site with official documentation together with the community;
¡¡ Explanation of the bridge building process and the role of the community:
• Phase I: collection of local materials (sand, gravel, stones and boulders), dressing of stones
and excavation work.
• Phase II: carrying (portering) of construction materials from the nearest road head to the site
• Phase III: masonry and concrete work, cable pulling and fitting
¡¡ Explanation of the self-help nature of the project
¡¡ Evaluation and explanation of the bridge location regarding technical limitations and requirements,
costs and situation of local traffic
¡¡ Assessment of capacity of the partners, funds & technical support from outside

One of the major indicators reflecting the real need of the bridge is the degree of participation and the
commitment demonstrated by the local community or beneficiaries in the construction of the requested
bridge. These indicators are assessed and measured from different points of view depending on the need
and purpose of the bridge. However a Social Feasibility Survey is not included in this Manual.

Step 2: Technical Survey


The technical survey includes: Bridge site selection and Topographic Survey of the selected bridge site

1.2 Preparation for survey


Preparatory works to be completed before going to the field for the survey:
¡¡ Collect maps with tentative location of the bridge and any available background information.
¡¡ Collect the survey equipment.

Materials required along with the survey equipment are:

For Survey by Abney Level For Survey by Theodolite


ƒƒ Abney Level, Survey Form & Checklist ƒƒ Theodolite, Tripod & Staff
ƒƒ Measuring Tape (50 or 100m and 3m) ƒƒ Measuring Tape (50m and 3m)
ƒƒ Red Enamel Paint and Paint Brush ƒƒ Red Enamel Paint and Paint Brush
ƒƒ Marker Pen, Scale and A3 Graph Paper ƒƒ Marker Pen, Scale and A3 Graph Paper
ƒƒ Camera and Film Roll ƒƒ Camera and Film Roll
ƒƒ Hammer ƒƒ Hammer
ƒƒ Ranging Rod (prepared at site) ƒƒ Survey Form and Checklist
ƒƒ Calculator, Note Book & Pencil ƒƒ Calculator, Note Book & Pencil
ƒƒ Nylon Rope (min. 50m) Masons Thread ƒƒ Thread and Plumb Bob

1.3 General data collection


General data is required for needs assessment and construction planning of the proposed bridge. The
general data and information shall include:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 3

ƒƒ Location of bridge site ƒƒ Transportation distance ƒƒ Nature of crossing and fordability


ƒƒ Availability of local materials ƒƒ Traffic volume ƒƒ Availability of local bridge
ƒƒ Temporary crossing ƒƒ Local participation builders

1.3.1 Location of bridge site

Describe the location of the bridge site:

Left Bank Right Bank


Region
Woreda
Village
Draw a bridge site location map covering the proposed bridge’s area of influence. The map should
contain information on:
¡¡ River system with names and river flow direction;
¡¡ Location of proposed bridge and traditional crossing point;
¡¡ Location of the nearest bridge (approximate walking distance from the proposed bridge site);
¡¡ Existing trail system and, if required, specify length of new trail for access to the proposed bridge;
¡¡ Location of the villages, health posts, schools and other important places with approximate
distances to the bridge site.

1.3.2 Nature of crossing and fordability

Examination of the present crossing situation is necessary to determine the need and the priority of the
requested bridge. Assess period of time the river cannot be crossed in one year.
a. Whole year
b. Some months per year only
c. Some days during high flood only

Situation (a) should be given first priority for construction and least priority given to situation (c).

Study the type of crossing facility available at present and also the location of the nearest bridge. Assess
whether the available crossing facility or the existing nearest bridge is sufficient for the crossing or that a
new bridge is necessary.

1.3.3 Traffic volume

Traffic volume at the crossing is one of the key indicators in the need assessment of the bridge. Information
should be collected by two methods. Count traffic volume at the traditional crossing point for at least one
day. And then interview the local people to form a broader impression of the traffic volume throughout
the year.

Average Number of Traffic per Day


Porters
Goods Traffic
Pack Animals
Persons
Non-goods Traffic
Animals
Determine the purpose of the traffic by interviewing the persons crossing and the local people. This will
indicate the importance of the crossing.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 4

Access to Yes No
Schools
Hospital / Health posts
Markets
Road head
Farming
Others (specify)
The most important crossing is one which provides access to schools, hospital and health posts.

1.3.4 Width of walkway

The standard width of walkway in this handbook is 70 cm or 106 cm. In most cases the 70 cm walkway
is sufficient. In stances of heavy traffic, mule and pack animal passage carrying bulky goods, or if the
crossing is on a main trail, a 106 cm walkway is necessary.

Discuss this issue with the local people, informing them that more work, especially collection of stones,
is required for the 106 cm walkway.

Recommended Width of Walkway: 70 cm 106 cm

1.3.5 Local participation

The commitment and participation of the local people in the construction of the proposed bridge will
truthfully indicate the need of the bridge. The stronger the commitment and participation, the higher is
the need of the bridge.

Compute the tentative preliminary number of mandays as:


¡¡ Mandays for skilled Labor: = 1.3 x span [m] + 400
¡¡ Mandays for unskilled Labor: = 5 x Span [m] + 1300

Assess the availability of local participation for bridge building from within the concerned local community.

By Whom Type of Participation


Local Community
Woreda
Zone
Kebele
Individual
Others (specify)

1.3.6 Transportation distance

Information on the transportation distance from nearest road head to the site is required for planning the
construction of the bridge.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 5

Name of nearest road Distance from site up to road head


Type of Transport
head. km Porter days Mule, days
Served by Truck
Served by Tractor
Other

1.3.7 Availability of local materials

Assess the availability of local materials needed for the bridge construction. Identify the nearest collection
place for these materials.

Description Moving distance, km Remarks


Stones
Natural Gravel
Sand
Wood
Other

1.3.8 Availability of local bridge builders

In the villages nearby there may be local construction personnel or bridge builders who have already built
some structures or bridges. Their skill can be used in construction of the proposed bridge. If such people
are available, record their names.

Names Skill (Mason, Bridge Fitter) Village / Address

1.3.9 Temporary crossing

Is a temporary crossing necessary


during the construction of the bridge? Yes No
If yes, what kind of temporary
crossing do you propose? Ferry Other

Temporary Bridge, Span mt

1.4 Bridge site selection


The main purpose of the technical field survey is to select the appropriate bridge site. The site should
optimally serve the local people. The selected site must be economically justified and have along life
span:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 6

¡¡ Fulfill the general condition;


¡¡ Have stable bank and slope conditions;
¡¡ Have favourable river conditions;
¡¡ Have shortest possible span.

1.4.1 General condition

The bridge site should fulfill a number of general conditions:


¡¡ Traditional crossing point;
¡¡ Minimum free board;
¡¡ Maximum bridge span;
¡¡ Space for the bridge foundations.

Use this checklist to evaluate the general condition:

Features Condition
Traditional Crossing Point
The bridge site should be selected at or near to the traditional
crossing point.

Favourable:
Selected site is at or nearby
the traditional crossing point

Unfavourable:
Selected site is far from the
traditional crossing point

ƒƒ For minor river detour from the traditional crossing point is not
acceptable.
ƒƒ For major rivers, detour up to 500m up stream and 500m down
stream from the traditional crossing point may be acceptable.
Bridge Span
The bridge span in this standard is limited to 120 m span.

Favourable:
span, l is equal or shorter than
120 m

Unfavourable:
span, l is longer than
120 m

ƒƒ Measure tentative span


ƒƒ Compare with the limit

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 7

Features Condition
Level Difference between two Banks
The level difference h between the two foundation blocks should
not be more than l/25.

Favourable:
h is equal or less than l/25

Unfavourable:
h is bigger than l/25

ƒƒ Locate the tentative position of the bridge foundations at both


banks
ƒƒ Measure the level difference h between the foundations of two
banks.
ƒƒ Compare with the condition.
Space for Foundation
Foundation should be placed at least 3m behind the soil slope and
1.5m behind the rock slope from the front edge of the riverbank.

Favourable:
Condition can be fulfilled

Unfavourable:
Condition can not be fulfilled

ƒƒ Measure tentative span


ƒƒ Compare with the limit
Slope Profile
The bridge foundation should be placed behind the line of the
angle of internal friction. (Angle of internal friction is the angle of
slope of soil or rock at which it is still stable and does not slide).

Favourable:
Condition can be fulfilled

Unfavourable:
Condition can not be fulfilled

ƒƒ Draw a slope line of 35o (angle of internal friction) in case of a


Soil slope and 60o in case of a Rock slope.
ƒƒ Foundation should be placed behind this line.
ƒƒ Check if these conditions can be fulfilled

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 8

Features Condition
Free Board
The Freeboard between the lowest point of the bridge and the
highest flood level should not be less than 5m. For this, sufficient
clearance between the lower foundation saddle and HFL should be
maintained.

Favourable Unfavourable

Clearance between lower


foundation saddle and HFL is:

not less than: less than:


span
up to 50m 7.5m 7.5 m
51 - 70m 8.0m 8.0 m
71 - 90m 9.0m 9.0 m
91 - 110m 10.0m 10.0 m
110 - 120m 11.0m 11.0 m
ƒƒ Identify HFL by local observation and villagers’ information.
ƒƒ Calculate available clearance and compare with the
requirement.
ƒƒ Exception: At flat or wide river banks free board may be
reduced.
At gorges free board may have to be increased.

1.4.2 RIVER CONDITION

The selected bridge site must have favourable river conditions. Accordingly, a bridge should be located:
¡¡ On a straight reach of the river;
¡¡ Beyond the disturbing influence of larger tributaries;
¡¡ On well defined banks.

Use this checklist to evaluate the river condition:

Features Condition
River Flow
In order to protect the bridge from sudden over-flooding and
strong erosion, the bridge site should not be located near the
confluence area of two rivers. Favourable:
Bridge site far from river
confluences

Unfavourable:
Bridge site near river
confluences

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 9

Features Condition
River Bed Favourable:
The river bed at the selected bridge site should be stable without River bed is not erosive, not
the possibility of erosion or filling up with bed load (boulders, filling up
gravel, silt, or sand)
Unfavourable:
River bed is erosive or filling up

1.4.3 Slope and bank condition

A bridge should be located at a site with safe and stable slope and bank conditions. The surveyor must
identify any potential instability features or failure modes of the soil or rock slope and along the bank.

If the slope and bank is soil, potential instability features and failure modes are:
¡¡ Bank erosion;
¡¡ Toppling instability of the bank;
¡¡ Erosion of the slope;
¡¡ Land slide.

If the slope and bank is rock, potential instability features and failure modes are:
¡¡ Plain failures in a rock slide along the slope.
¡¡ Wedge failure leading to the fall of rock mass.
¡¡ Toppling leading to the fall of rock blocks.
¡¡ Rotational slide is similar to the landslide in a soil slope. Such failure is likely when the material of
the rock is very weak (soft rock) and the rock mass is heavily jointed and broken into small pieces.

To avoid the above instability features, use this checklist to evaluate the slope and bank of the selected
site:

Features Condition
If the River Bank or Slope is SOIL
Bank Profile
The bank profile should be smooth.
Favourable:
Bank profile is smooth to
partially cut out

Unfavourable:
Bank profile is strongly cut out

Smooth Partially cut out Strongly cut out

River Bank Contour


The bridge site should be located at the straight reach of the river
to avoid the river from undercutting or bank erosion. Favourable:
River contour is straight or
convex

Unfavourable:
River contour is concave

Straight Convex Concave

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 10

Features Condition
Bank Erosion Favourable:
The River bank should not show any sign of erosion. No sign of fresh erosion

Unfavourable:
Presence of fresh erosion
Slope Profile
Favourable:
The slope profile should be smooth.
Slope profile is smooth to
partially cut out

Unfavourable:
Slope profile is strongly cut out
Smooth Partially cut out Strongly cut out
Transverse Slope Favourable:
The transverse slope should be smooth. Transverse slope is smooth to
partially cut out
Transverse slope strongly cut out
Unfavourable:
Transverse slope is strongly cut
Smooth Partially cut out Strongly cut out out

Slope Inclination (Soil Slope)


The slope inclination should be less than 35o.

Favourable:
Slope inclination is equal or
smaller than 35o

Unfavourable:
Slope inclination is bigger than
35o

Estimate the slope inclination and compare it with the condition. If


the site has an unfavourable slope inclination, it can still be
selected provided the general condition of slope profile is fulfilled.
River Undercutting Favourable:
The bridge site should be free from river undercutting which may There is no river undercutting
lead to landslide.
Unfavourable:
River undercutting is active
or there is potential for river
undercutting

Inclined Trees
The selected site should not have inclined trees, which indicate an Favourable:
active landslide. Inclined trees are not present

Unfavourable:
Inclined trees are present

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 11

Features Condition
Seepage or Swampy Area Favourable:
The bank slope should not have any seepage or swampy area, No seepage or swampy area is
which may lead to slope instability. absent

Unfavourable:
Seepage or swampy area is
present
Gully Erosion
Favourable:
No signs of gully erosion should exist within the vicinity of the
No sign of gully erosion or only
selected site.
light gully erosion
ƒƒ Observe if any rivulets are within the vicinity of the selected
Unfavourable:
site.
Heavy gully erosion exist
ƒƒ If rivulet exists, examine the dimension of the gully cutting.
Slipped (Slump) Soil Mass
The bridge should not be located on already slipped soil masses.
Slope with slipped soil mass

Favourable:
There are no back scars or
signs of soil mass movement

Unfavourable:
There are back scars or signs of
soil mass movement

ƒƒ Examine and identify any indication of soil mass movement.


This can be done by observing traces of back scars on the
slope.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 12

Features Condition
If the River Bank is ROCK Favourable:
Plain Failure Plain failure will not take place,
Plain failures lead to the slide of rock layers along the slope. The if:
rock bank/slope of the selected site should not have any feature of ƒƒ Bedding/fracture plain is
plain failure. sub-parallel to opposite to
the slope;
Bedding plain is parallel to the slope and plain failure is active. ƒƒ Bedding/fracture plain is
Site is extremely unfavourable! parallel to the slope, but
inclination is less than 35o.

Unfavourable:
Plain failure will take place, if:
ƒƒ Bedding/fracture plain is
parallel to the slope and
inclination is greater than
35o
ƒƒ Presence of old slided rocks
Plain Failure Model

Identify bedding/fracture plain (layers of rock)


ƒƒ Check its direction and inclination
ƒƒ Compare with the condition

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 13

Features Condition
Wedge Failure
Any form of wedge failure leads to sliding of rock masses. The
rock bank/slope should not have wedge failures or potential
wedge failures.
Traces of wedge failure Favourable:
Wedge failure will not take
place, if
ƒƒ There are no fracture plains
facing each other
ƒƒ There are two or more
intersecting fracture plains
but the inclination of its line
of intersection is less than
35o
ƒƒ There are two or more
intersecting fracture plains
but the inclination of its line
of intersection is opposite
to the slope

Unfavourable:
Wedge failure will take place, if
ƒƒ There are two or more
intersecting fracture plains
and the inclination of its
intersection line is more
than 35o to the slope
ƒƒ Presence of old slided
wedge
ƒƒ Identify if there are fracture plains facing each other
(intersecting)
ƒƒ Check the inclination of the intersection line
ƒƒ Compare with the condition

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 14

Features Condition
Toppling Failure
The rock bank/slope should not have any features of toppling
failure.
Potential toppling failure Favourable:
Toppling will not take place,
if
ƒƒ The Rock slope is less than
60o;
ƒƒ There is no formation of
rock blocks;
ƒƒ There is a formation of rock
block but b/h (width of the
block/height of the block) is
bigger than 1.

Unfavourable:
Toppling will take place, if
ƒƒ There is a formation of
vertically elongated rock
blocks in a steep slope
bigger than 60o and the
blocks are tilted towards the
slope;
ƒƒ Identify if there is formation of vertically elongated rock blocks ƒƒ Old toppled rock blocks are
(cubes) due to vertical and horizontal fracture planes/joints; present.
ƒƒ Estimate inclination of the slope;
ƒƒ Estimate inclination and orientation of the rock block and
compare with the conditions.
Translational Failure (Slide) Favourable:
The rock slope should not have any potential of rotational failure. Sliding will not take place, if
the
Translational failure (sliding) of soft rock slope ƒƒ Slope is hard rock;
ƒƒ Slope is soft rock but not
weathered;
ƒƒ Slope is soft rock and
weathered but not steeper
than 40°.

Unfavourable:
Sliding will take place, if
ƒƒ The slope is highly
weathered soft rock and
failure (slide) model steep than 40°;
ƒƒ Identify type of rock and its weathering grade; ƒƒ Back scars or old slide are
ƒƒ Estimate inclination of the slope; present.
ƒƒ Compare with the conditions.

1.4.4 Evaluation of the bridge site

After completing investigation of the site, categories the bridge site as:
Good All or most of the features are favourable and if the surveyor is confident about the
stability of the slopes. Proceed with further survey work.
Bad Most of the features are unfavourable. Reject site.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 15

Questionable Most of the features are favourable and some are unfavourable. The site is questionable.
In this case, further detailed investigation by an experienced geo-technical engineer is
necessary..

As far as possible, the bridge site should be selected at a location where protection works will not be
required. If protection works are unavoidable determine the required special structures like retaining wall,
drainage channels, etc. A tentative design with dimensions and location of these structures should be
illustrated in a sketch showing a plan view and a typical section. But it is best to avoid bridge sites, which
require river protection works.

1.4.5 Classification of soil and rock

Identification of Soil and Rock types is required for appropriate foundation design.

Table F.1.1: Soil classification

Soil Parameters
Angle Applicable
Soil Type How to Identify Bearing of Unit Foundation
Capacity Internal Weight, g Design
[kN/m2], Friction, [kN/m3]
φ°
Estimate the
percentage (%)
of coarse grains
Coarse larger than 6 mm.
Grained Soils Gravelly If, more than half 400-600 32-38
More than half 19
Soils of the coarse (400) (35)
of the materials fraction is smaller
are of individual than 0.06 mm
grains visible grain size, the soil
to the naked is Sandy Soil

Deadman Anchor
eye (grain size
bigger than If, more than half
0.06 mm) of the coarse
Sandy 200-300 31-37
fraction is larger 18
Soils (200) (33)
than 6 mm, the
soil is Gravelly Soil
Prepare moist soil
ball from the soil
sample and cut it
Fine Grained Silty with a knife. 50-200 30-32
Soils 17
Soils If, the cut (150) (30)
More than half surface is dull or
of the materials scratched, the soil
are individual is Silty Soil
grains not
visible to the Prepare moist soil
naked eye ball from the soil
(grain size sample and cut it
smaller than with a knife. 100-200 9-25
Clay 16
0.06mm) If, the cut surface (100) (22)
is smooth and
shiny, the soil is
Clay.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 16

For estimating the percentage (%) of coarse grains use Figure F.1.2.

Figure F.1.2: Ratio of coarse grains

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 17

Table F.1.2 Rock Classification

Rock Parameters

Rock Type
Degree of Angle of Applicable
How to How to Bearing
Examples Fractures or Internal Foundation
Identify Identify Capacity
Weathering Friction, Design
[kN/m2],
φ°

Rock has
no sign of
weathering
Rock is
or only faint 1500- Drum
sound and 35-50
signs of 2000 Anchor in
fresh to fairly (40)
Gives weathering (1500) Hard Rock
Quartzite weathered
up to
Hard Rock

metallic
Limestone 1-5 cm
sound
Granite, thickness
after
Dolomite
hammer In the
etc.
blow Highly rock exists
Drum
fractured widely open
35-50 Anchor in
rock and cracks, 1500
(40) Fractured
fresh to fairly fractures
Rock
weathered and
bedding
Drum
No sign of 25-40 Anchor in
Fresh 1300
weathering (30) Fractured
Rock
Phylite Gives
Slate dull Most of
Soft Rock

Siltstone sound the original


Claystone after rock has
Schist hammer been
Fairly to highly 600-750 25-40 Deadman
etc. blow seriously
weathered (650) (30) Anchor
altered
Rock can
be broken
by hand

1.4.6 Identification of soil and rock

Excavate a test pit with a depth of up to the estimated foundation level (but not less than 2.0m) or up to
the bedrock at the proposed foundation locations. If the bank/slope is soil, investigate each layer of soil
in the pit and classify the soil according to the Soil Classification chart, filling in the soil investigation table.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 18

Example for Soil


Location: Main Anchorage Foundation at Right Bank

Depth Soil Parameter


from Bearing Angle of
Sketch Soil Type Remarks
Surface, Capacity Internal
[m] [kN/m2], Friction, φ°

0.0 Top soil

0.3 Sandy Soil

Foundation
1.1 Gravelly Soil 400 35 Design
Parameter

If the bank/slope is rock, investigate the rock type according to the Rock Classification chart, filling in the
rock investigation table as per the example.

Example for Rock


Location: Main Anchorage Foundation at Left Bank

Soil Parameter
Degree of Applicable
Rock Type Fracture/ Bearing Angle of Foundation Remarks
Weathering Capacity Internal Design
[kN/m2], Friction, φ°
Highly Drum
fractured anchorage
Hard Rock 1500 40
and fairly foundation in
weathered fractured rock

1.5 Topographic survey


After final selection of the bridge site, the surveyor proceeds with the topographic survey.
The purpose is to:
¡¡ Provide a topographic map of the bridge site with details relevant to the bridge design;
¡¡ Establish axis pegs and bench marks for use during construction of the bridge.

1.5.1 Survey area

Area to be covered by the topographic survey:


¡¡ For bridges without windguy arrangement,
• A profile along the bridge axis covering up to 25m behind the main anchorage blocks.
¡¡ For bridges with windguy arrangement,
• A profile along the bridge axis covering up to 25m behind the main anchorage blocks and
a topographic plan covering the area of 10m upstream and 10m downstream from the
tentative location of the windguy foundations.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 19

1.5.2 Setting of bridge centerline

Fix the bridge centerline with two permanent axis points A on the left bank and B on the right bank.
Permanent axis points A and B should be fixed on rock out crop along the bridge centerline, if available.
If rock out crop is not available, these points should be marked on a boulder sufficiently embedded into
the ground:

Additional survey points along the centerline should be fixed to survey the bridge axis profile. These
survey points should be fixed at breaking points of slope and terraces, which will accurately indicate the
topography of the bridge axis. The profile should cover 25m behind the main anchorage block up to the
edge of the river flow.

Draw a sketch of the profile/cross section of the bridge axis (centerline) with axis points A and B, with all
the survey points and topographic features, including tentative position of the bridge foundations, low
water level and high flood level.

Profile/Cross Section (Example)

Draw a plan view with the bridge axis (centerline), axis points A and B, with all the benchmarks and fixed
objects like trees, houses etc. Give distances and directions from the reference points so that the axis
points and benchmarks can be located during the construction. A plan view is necessary only when a
windguy arrangement needs to be considered in bridge design.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 20

Plan (Example)

1.5.3 Survey methods

There are two options for conducting the topographic survey. Depending upon the span and type of
bridge, a profile along the bridge axis or a more detailed survey including contour lines will be necessary.
In general Windguy Arrangement is not required for bridges with span up to 120m.

¡¡ A detailed profile along the selected bridge axis is sufficient for bridges without windguy
arrangement. A topographic profile can be made by the Abney level, however for fixing precise
levels a Level Instrument is necessary.
¡¡ For bridges requiring a windguy arrangement a more detailed topographic survey is necessary,
from which a detailed contour plan can be plotted. This type of survey should be done by a
Theodolite.

1.5.4 Survey by abney level

¡¡ The main function of the Abney Level is to measure the vertical angle ϕ. By measuring the slope
distance d between the survey points with a measuring tape, the horizontal distance D and the
vertical difference of elevation ΔH can be calculated.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 21

Measurement of Vertical Angle:


Illustrations on the principal of measuring the vertical angle by the Abney Level:

The Procedure of measurement:


¡¡ The Surveyor stands at point A with the Abney level.
¡¡ The Assistant stands at the target point C with a stick (or ranging rod). The target mark at the stick,
which the Surveyor sights must be at the same height above the ground as the Abney Level. For
this, the height of the ranging rod should be equal to height up to Surveyor’s eye level.
¡¡ The Surveyor holds the Abney Level to the eye and sights towards the target at point C, centering
the cross hair against the target.
¡¡ The index arm is then adjusted until the bubble is centered against the target and cross hair.
¡¡ When the bubble is centered horizontally and the cross hair is aligned with the target, read the
vertical angle on the arc.

Measurement of Slope Distance d:


The slope distance d between the survey points is measured with a measuring tape. Distances larger
than 30m should be divided into sub-distances. The total distance can then be calculated by adding the
sub-distances. Slope distances should be measured twice and the mean value should be taken as the
accurate slope distance.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 22

Calculation of Horizontal and Vertical Distance and Elevations of Survey Points:

Horizontal distance D between a and c D = d x cos ϕ


Vertical distance ΔH between a and c ΔH = d x sin ϕ
Elevation of c H = Elevation of a ± ΔH
Add +ΔH, if sighting is upward and subtract -ΔH, if sighting is downward.

d = slope distance from a to c


ϕ = vertical angle from a to c

To take the profile along the bridge axis, the Surveyor should first set the exact centerline. There are two
methods of setting the centerline.

By Nylon Rope and Plumb Bob:

This method is accurate only for spans up to 50m. Survey points along the bridge centerline are fixed
with the help of a nylon rope and plumb bob.

The Nylon rope is stretched along the axis point A of the left bank and B of the right bank. Care should be
taken that the tape or nylon rope is hanging freely and does not touch any obstacles. The survey points
are then fixed along this rope with the plumb:

By Bamboo or Wooden Sticks or Ranging Rods:


This method is applied for span above 50m. In this method the survey points along the bridge centerline
are fixed with the help of Bamboo or Wooden Sticks or Ranging Rods. Fix Stick at each axis point A and
B in vertical position. Now the surveyor can aim at other points along the bridge centerline line of A and
B. By fixing in line additional survey points behind and in front of A and B, more points can be gained
along the bridge centerline ranging:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 23

Bridge Axis Profile:


Proceed with the survey of the bridge axis profile after having fixed the centerline.

1. The survey starts from the fixed permanent axis points A or B and proceeds to other survey points
M, N, O, P, 1, 3, 4, 5 or S, T, U, V, 4, 5, 6 (refer example sketch of bridge axis profile)
2. Measure the vertical angles and slope distances between these survey points of the centerline. It
is important that the sighted target is on the same height above the ground as the Abney Level
while taking the readings
3. Measure all points M, N, O, P, 1, 2, 3 starting from the permanent axis point A, as described in
the second step above
4. Similarly, measure all points S, T, U, V, 3, 4, 5, 6 starting from the permanent axis point B
5. Measure vertical angles from A to B and B to A to check the accuracy in vertical angle readings

Before calculating the horizontal and vertical distances, it is necessary to determine the accuracy of the
measurement of the vertical angles. This can be done by comparing the measured vertical angle from
A to B with the vertical angle from B to A. Both angle readings should be equal. Differences in these
readings indicate an error in the angle measurements and needs correction.

The error factor ‘E’ is calculated with the formula:


[ϕAB]-[ϕBA]
E= ; Corrected angle ϕ′ = ϕAB ± E or ϕ′ = ϕBA ± E
2

Example: Error Correction of Measured Vertical Angles

measured vertical angle from A to B ϕAB = 0°50’ (Downhill sighted)


measured vertical angle from B to A ϕBA = 1°30’ (Uphill sighted)
error factor E = (1°30’ - 0°50’)/2 = 0°20’
corrected angle ϕ’AB = 0°50’ + 20’ = 1°10’
corrected angle ϕ’BA = 1°30’ - 20’ = 1°10’

All measured vertical angles should be corrected as:


Corrected angle ϕ′ = [ϕ] + E for downhill sighted (-) angles
Corrected angle ϕ′ = [ϕ] - E for uphill sighted (+)angles
Compute the horizontal distances and elevations of the corresponding survey points with the corrected
vertical angles as per the example.

Example: Calculation of Horizontal Distance D and Elevation H of Survey Points:

Elevation of A = 100. 00 m
measured vertical angle from A to M, ϕAM = +7°0’ (upward sighting)
corrected vertical angle from A to M ϕ’AM = +7°0’- 0°20’ = +6°40’
measured slope distance A to M d = 13.35 m
horizontal distance between A and M D = d x cos ϕ’ = 13.35 x cos 6°40’ = 13.26m
vertical distance between A and M ΔH = d x sin ϕ’ = 13.35 x sin 6°40’ = +1.55m
Elevation of M H = Elevation of ‘A’ + ΔH = 100.0 + 1.55
= 101.55m

Enter the measurements and calculations into the Abney Level survey sheet.

Measuring the River Width:


In certain cases, it might not be possible to directly measure the river width from one bank to another
bank by tape. In such a situation, the river width should be measured by indirect method as described in
the example.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 24

How much is the width of the river l = Distance from L to R?

The procedure of the measurement is as follows:


¡¡ Set a base line R - B perpendicular to the line L - R along the river. This can be done easily by 3-4-5
method (refer Setting of a Right Angle).
¡¡ On this base line R - B, mark the mid point C so that RC is equal to CB.
¡¡ Set again a base line B – A perpendicular to the line R – B similarly to step1.
¡¡ Mark exactly point A by ranging through point L – C, so that all these three points lie in the same
line of sight.
¡¡ Measure length B – A. This length will be equivalent to the river width L - R = l.

Setting of a right angle:


One simple method to set a right angle from a point of a base line is the 3-4-5 method. One measuring
tape and 3 wooden pegs are needed.

The procedure of the measurement is as follows


¡¡ The first person should hold ‘0’ and ‘12’m mark of the measuring tape at point R.
¡¡ The second person holds the tape at the 3m mark, and a third person at the 8m mark of the tape.
Stretch all these sides of the tape. A right angle triangle will be formed with sides of 3, 4, and 5m.
¡¡ Line R – B’ is now perpendicular to the line R – L’.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 25

1.5.5 Survey by Theodolite

When the span of the bridge is more than 120m or when a windguy arrangement needs to be included
in the bridge design, the survey is conducted with a theodolite.

For proper use of a theodolite, refer to the respective instruction manual that comes with the theodolite.

Profile along bridge axis:


Fix the bridge centerline. Measure the distance between the axis points A and B by horizontal triangulation
method. Triangulation is done by measuring all three angles of a triangle and length of one side.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 26

For accuracy double triangulation is necessary. The procedure is:

Ist Triangulation
1. Set out a peg at C in such a way that the distance B - C can be easily measured. The length ‘d’ should
be at least 20% of the distance A - B
2. Measure distance B - C = d accurately with a measuring tape. Measure this distance several times
and calculate the mean distance.
3. Set up theodolite at B and measure the horizontal angle ABC = β from face left and face right
4. Set up theodolite at C and measure the horizontal angle ACB = γ from face left and face right
5. Set up theodolite at A and measure the horizontal angle BAC = α from face left and face right
6. Sum up these angles (δ=β+γ+α), which should be theoretically equal to 180° or 200g. If, the sum is
not equal to 180° or 200g, the difference Δ should be equally distributed to all the three angles so
that the sum becomes 180° or 200g
7. Calculate distance A – B = D with the trigonometric formula,
d x sinγ
D = sinβ

IInd Triangulation
1. Repeat the same procedure as above and calculate distance A – B = D’
2. Calculate final distance
D + D’
D= 2

Use the standard form “Triangulation” for recording the readings and calculation.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 27

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 28

Elevation of Axis Points and Benchmarks:


It is necessary to establish the elevations of the Axis Points A and B and Benchmarks. This is done by
vertical triangulation indicated in this manual.

The procedure is as follows:


1. Select a first Bench Mark BM I on a rock or big boulder near to the axis point A and fix it as 100.00m
2. Select bench mark BM II near to the axis point B
3. Measure the horizontal distance D from A to BM I accurately with a tape
4. Measure the horizontal distance D from B to BM II accurately with a tape
5. Take the distance D between axis points A and B from the triangulation
6. Set up the theodolite at the axis point A and measure the vertical angle to axis point B and the
vertical angle to BM I. Take the middle hair reading Z and measure the instrument height I.
7. Set up the theodolite at the axis point B and measure the vertical angle to axis point A and the
vertical angle to BM II. Take the middle hair reading Z and measure the instrument height I.
8. Set up the theodolite at BM I, measure the vertical angle to axis point A. Take the middle hear
reading Z, and measure the instrument height I.
9. Set up the theodolite at BM II, measure the vertical angle to axis point B. Take the middle hair
reading Z, and measure instrument height I
10. Calculate the s for all readings:
Vertical Distance V = D x tan ϕ
or
D
V=
tan β

Elevation difference ΔH = V - Z + I for upward vertical angle reading


ΔH = V + Z - I for downward vertical angle reading
11. Calculate Elevations of A, B, BM I and BM II, starting from BM I to A to B to BM II

Elevation of A = El. of BMI ± ΔH


Elevation of B = El. of A ± ΔH
Elevation of BM II = El. of B ± ΔH

Insert the readings and the calculations in the Survey form of “Summary of Triangulation and
Elevations of Pegs and Benchmarks”.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 29

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 30

Topographic Detail Survey:


The topographic detail survey is necessary to represent the topography of the bridge site by means of a
map (plan) with the contour lines.

The topographic detail survey uses the tachometric method. Tachometric survey is done by Theodolite
with stadia hairs (having constant value of 100) and leveling staff.

Checking the Stadia Hair:


Check the stadia hair of the theodolite before doing the detail survey by tachometry. For this, measure
a distance of about 40m using stadia readings and compare them with actual tape measurements. If the
difference between the stadia measurement and the tape measurement is more than 0.2%, calculation
of horizontal and vertical distance needs to be corrected. The distances should be corrected for error Δ
as per Formula.

D = (100l ± Δ) x cos2 ϕ V = (50l ± Δ) x sin2 ϕ

Δ is calculated before the survey as per procedure:


1. Put the theodolite on horizontal ground and level it
2. Level the telescope of the theodolite so that the vertical angle is 0
3. Put pegs at approximate distances of approx. 10, 20, 30, 40 and 50m
4. Measure accurately the distance between the vertical axis of the theodolite and the pegs by tape
5. Take the stadia hair readings by theodolite at each peg
6. Calculate the horizontal distance to each peg by tachometric calculation
7. Determine the difference (error) between the tape measurement and the tachometric measurement
for each peg
8. Plot the graph for Δ correction

Example: Δ - Corrections

Top hair Bottom hair Difference Distance Tape Measurement Correction


l1 l2 l = l1- l2 D’ = l x 100 Distance, D Δ = D – D’
(cm) (cm) (cm) (m) (m) (cm)
118.70 108.30 10.40 10.40 10.48 +8
135.90 114.20 21.70 21.70 21.87 + 17
140.75 109.80 30.95 30.95 31.33 + 38
160.25 118.05 42.20 42.20 42.66 + 46
120.20 66.45 53.75 53.75 54.36 + 61

The graph is used for the calculation of the tacheometric error for the horizontal distances

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 31

Tachometric Survey:
All topographic details are taken by the tacheometric survey. Tachometric details are mainly taken from
the axis point A and B (theodolite stations). If the area of survey cannot be covered by these two points,
details should be taken from other additional points. The survey points (staff points) should be taken
at break points of slopes, terraces, fields and other features representing the actual topography of the
ground. Survey points should also include other details such as houses, trees, foot trails, rocks, river
banks, high flood level, water level at survey time etc.

The procedure of survey is:


1. Set up the theodolite on the axis point A. Measure the instrument height I
2. Fix the 0 reading of horizontal circle along the bridge axis towards B as illustrated in the sketch

3. Take for every survey point (staff point) the readings of the horizontal circle, the vertical circle, the top
hair, the middle hair and the bottom hair, after proper sighting to the respective survey points.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 32

Vertical Angle, b = 94g 06c Top Hair, l 1 = 2.455


Horizontal Angle a = 214g 97c Bottom Hair, l 2 = 1.844
Middle Hair, Z = 2.15
4. Record the readings into the “Tachometry” survey sheet.
5. Set up the theodolite on the axis point B. Measure the instrument height I.
6. Fix the zero reading of the horizontal circle along the bridge axis towards A.
7. Take the details, which were not covered from axis point A, the procedure from step 3-4.
8. Calculate the horizontal and vertical distances and elevations of the survey points with the help of the
tachometric formulas given in the “Tachometry” survey sheet.

1.5.6 TOPOGRAPHIC MAPS

As per the field survey data, it is necessary to prepare the topographic maps in the scale 1:100 or 1:200
¡¡ Profile along the bridge axis;
¡¡ Contour plan of the bridge site in scale (only when windguy arrangement is necessary).

Profile Along the Bridge Axis:


Plot the profile along the bridge axis as per steps (refer the example in this manual):
1. Choose the scale of the drawing. Vertical and horizontal scale should be the same.
2. Choose the datum level so that the points with lowest and highest elevations are within the drawing
area.
3. Choose the position of the axis point A so that most farest survey point of Right Bank and left bank
from the axis point A are within the drawing area.
4. Plot the axis point B as per its elevation and horizontal distance from axis point A.
5. Draw the survey points of the bridge axis according to the horizontal distance and elevations as per
the data from the “Bridge Axis Profile by Abney Level” survey sheet or the “Tachometry” survey
sheet from axis point A. Refer also to the sketch of the bridge profile prepared during the field survey.
6. Similarly, draw the remaining survey points of the bridge axis from axis point B.
7. Join all the survey points by straight lines. This will represent the bridge axis profile.
8. Draw horizontal lines with the elevation of the high flood level and the water level at the time of
survey.

Contour Plan of the Bridge Site:


The contour plan represents the overall topography of the bridge site by means of contour lines. A
contour line is a continuous line passing through points of equal elevation.

A contour plan is necessary only when a wind guy arrangement is to be considered in the bridge design.
In most of the cases of short span trail bridges, wind guy arrangement will is not necessary, and a contour
plan is not required for bridge design.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 33

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 34

Figure F.1.3: Bridge axis profile

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 35

Figure F.1.4: Contour plan

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 36

1.5.7 PHOTOGRAPHS

Photographs of the bridge site to support its technical feasibility / topography and facilitate the bridge
design.

Take the photographs:


¡¡ An overall view of the bridge site from upstream indicating approximate location of bridge
foundations and the axis line;
¡¡ An overall view of the bridge site from down stream indicating approximate location of bridge
foundations the axis line;
¡¡ View of the right bank from left bank with an approximate location of bridge foundations
¡¡ View of the left bank from right bank with an approximate location of bridge foundations
¡¡ An overall top view (if possible)
¡¡ A close up view of the axis points and the bench marks
¡¡ A view of the soil test pits at the location of the bridge foundation blocks
¡¡ Other relevant photos

Take above photographs from the positions. If one picture does not cover the necessary area, take
several pictures from the same spot with sufficient overlapping.

Present all the photographs systematically with respective captions.

1.5.8 Survey report

The technical survey report consists of:


¡¡ Filled in Survey Forms and Checklist;
¡¡ Topographic map
• Profile along the bridge axis in scale;
• Contour plan of the bridge site in scale (only if windguy arrangement is necessary)

1.6 Calculation of quantity and cost estimate


Calculation of quantities and cost estimate is required for the purpose of planning and implementation.
¡¡ Quantity Calculation
• Calculate the quantities of the cables from Cable Design Form. Fill in the Quantity Calculation
Sheet of Wire Rope (Cables). This sheet will show the cable lengths of each diameter and the
total weight of the cables.
• Calculate the quantities of the Steel Parts and Steel Deck from the corresponding steel
drawings. Fill in the Quantity Calculation Sheet of Steel Parts and Steel Deck.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 37

• Calculate the quantities of the Earth Works from the General Arrangement Drawing. Fill in
the Quantity Calculation Sheet of Construction.
• Calculate the quantities of other construction works from the corresponding Construction
Drawings. Fill in the Quantity Calculation Sheet of Construction.
• Prepare a list of construction materials according to the calculated quantities of construction
works.
• Calculate the transportation weights for cables and other construction materials not available
locally.
• Calculate the quantities of Works and Labor.
¡¡ Rate analysis
Prepare the rate analysis for fabrication of steel parts, steel decks and road transportation (items of
external support to the community) as per unit quantity, unit cost and standard norms.
¡¡ Abstract of cost
Compute the abstract of cost of the bridge as per the quantities of works (from Quantity Calculation
Sheets) and the rates (from Rate Analysis) for each item of works, and summarize the cost as per
the category of works.
¡¡ Summary of estimated cost
• Bridge cost: Calculate the Estimated Bridge Cost by summarizing the Abstract of Cost.
Also calculate the bridge cost per m span.
• Contribution: Estimate the expected contribution from different partners.
• Breakdown of the contribution:
Break down the contribution as Local Contribution and Outside Contribution.
¡¡ Summary of actual cost
In the majority of cases, the actual bridge cost will not be the same as estimated. Therefore, calculate the
actual bridge cost after completion of the bridge.

1.7 Bridge maintenance

1.7.1 Introduction

Maintenance of trail bridges is very crucial for keeping the bridge and foot trails functional throughout
the year. It is extremely essential to guarantee their permanent and safe use, maintain them in usable
condition, and to preserve the investment made in these bridges. In order to determine the required
maintenance, regular inspection of the bridge should be made after completion of the construction work.

The bridge maintenance work consists of the two categories:


¡¡ Routine maintenance
¡¡ Major maintenance

A brief description of these two maintenance categories is given in the sub-chapters.

1.7.2 Routine maintenance

Routine maintenance is a preventive type of maintenance and should be done regularly. It is important
to protect the bridges from getting big and irreparable damages and assures long-term use by keeping
them in serviceable condition. After completion of the bridge construction, routine maintenance should
be carried out on regular basis. In general, the works under routine maintenance are simple in nature.

The routine maintenance work includes the important tasks:


¡¡ Cleaning around the most important bridge elements
Cleaning and removing all sorts of debris, dirt, plants and bushes in and around the drainage
channels, the cable anchorage terminals, the tower base, the area around foundations, the area
below the bridge entrance and the bridge access trails.
¡¡ Fixing and re-tightening of bridge parts
Fixing and re-tightening of walkway wire mesh, nuts and bolts, bulldog grips, etc., which are loose.
¡¡ Repairing the walkway deck
Re-tightening of loose nuts and bolts of steel decks and J-Hooks.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 1 - 38

¡¡ Minor repairing of gabion boxes for bank and slope protection purposes
Inspection and checking of the slope and riverbank protection structures and execution of minor
repair work.
¡¡ Reporting of the bridge condition
Inspection and checking the general condition of all the bridge parts and structures and reporting
to the concerned Woreda Administration and/or Rural Roads Authority and seek their necessary
support in case of big landslides, bank erosion, etc., which may damage the bridge foundations
and structures or even cause the collapse of the bridge.

Routine maintenance work can be carried out either by forming a Bridge Maintenance Committee (BMC)
or by appointing a bridge warden. In both cases one trained person must be assigned for regularly
inspecting the bridge. S/he should preferably live close to the bridge and should be equipped with some
basic tools.

Primarily the concerned Users Committee are responsible for ensuring that routine maintenance is done.
The Rural Roads Authority, who bears the overall responsibility, shall monitor the maintenance and shall
support the Users Committees for cases beyond their capacity.

1.7.3 Major maintenance

Major maintenance (MM) work includes all works, which need proper - planning, survey, design and cost
estimates. A certain level of knowledge and skill is required to execute the major maintenance of the
bridges.

The major maintenance work includes the tasks:


¡¡ Replacing rotten wooden planks with galvanized steel decks;
¡¡ Replacing rotten wooden crossbeams with galvanized steel beams;
¡¡ Repairing of wind guy arrangements/system;
¡¡ Repair, adjustment or replacement of suspenders including adjustment of camber of for suspension
bridges;
¡¡ Re-painting of all non-galvanized steel parts;
¡¡ Re-tensioning of all loose cables and adjusting bridge alignment;
¡¡ Coaltar treatment of all non-galvanized threads;
¡¡ River bank and slope protection works.

Major maintenance responsibilities are gradually operationalized at the district level by imparting
technical know-how, methods and practices for carrying out maintenance work and providing material
support. The Rural Roads Authority are becoming better prepared to implement the major maintenance
and are responsible for the execution independently or with the support of the Helvetas Ethiopia Trail
Bridge Section (TBS), or other concerned bridge building agencies.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 39

2. SUSPENDED TRAIL BRIDGE

2.1 Presentation of the suspended type


The Suspended Cable Bridge design presented in this Manual has been adapted from the successful
experience of Nepal. The major structural elements are steel wire ropes, which are anchored by gravity
blocks or rock anchors at either side of the river. The superstructure is completely unstiffened and thus
allows “some” reasonable degree of lateral, vertical and torsional vibrations. For economic reasons the
design allows a choice between two options for the width of the walkway. The 70cm walkway is mainly
applicable for pedestrian traffic, whereas the 106cm walkway should be applied for crossings where pack
animal traffic is also expected.

2.2 Technical features and limitations of suspended type


The Short-Span Trail Bridge Standard as presented in this manual conforms to mainly Indian Standard,
but also to Swiss and German Standards, codes and norms. All its components fulfill the necessary safety
factors by applying the loadings prescribed in the standard bridge design.

All the construction materials conform to international specifications. Exposed steel parts are painted or
hot dip galvanized, and should not be altered unless proven to fulfill standard norms.

For practical, economical and safety reasons the span range for the Short-Span Trail Bridge Standard
presented in this Manual is limited to 120m.

Longer spans are possible but would require special engineering input. As with every standard design, not
all site conditions are covered with this standard. It is especially not suitable in unfavourable geological
site conditions. At such sites, as mentioned above for longer spans, engineers’ input is mandatory.

This type of bridge has downward sagging walkway. Sagging walkway cables are suspended below
their anchorage. Cables are anchored in to the main anchorage foundation at both banks. The main
components of this bridge are: Walkway cables and Handrail cables, Walkway system and Main anchorage
foundations. This type of bridge is selected where the bridge foundations can be placed at sufficiently
high position giving required free board from the highest flood level. Suspended type is more economic,
simple to design and construct than suspension type bridge.

Figure F.2.1: Suspended type bridge, Keleta, Oromia region

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 40

¡¡ Layout and sections

Figure F.2.2: The typical profile of Suspended type bridge

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 41

Figure F.2.3: The typical plan of Suspended type bridge

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 42

¡¡ Walkway section
There are two types of walkway width: 70cm & 106cm

2.3 Basic design concept

2.3.1 Loadings

For designing a bridge structure loadings as per Long Span Trail Bridge standard is followed.
¡¡ Live load
Live load for smaller span up to 50m is equivalent to 400 kg/m2 and for longer span as per formula,
50 50
P = 300+100. kg/m2 or 3+ kN/m2
l l

¡¡ Dead load
Dead load includes the weight of all permanent components of the bridge structure.

Width of Walkway 34 cm 70 cm 106 cm


Dead Load (without weight of 25 kg/m or 42 kg/m or 57 kg/m or
Handrail- and Walkway Cables) 0.25 kN/m 0.41 kN/m 0.56 kN/m

¡¡ Wind load
The design wind load is taken as a uniformly distributed load based on a wind speed of 160km/h acting
horizontally to the walkway. This corresponds to a wind pressure of 1.3 kN/m2 acting on the lateral bridge
area of 0.3 m2 per meter span. By maintaining a wind coefficient of 1.30 (acc. to Swiss Standard) the
actual lateral design wind load is 0.50 kN (1.3 x 1.3 x 0.3) per meter span. The foundation structures are
sufficient to resist this design wind load.

Wind load affects also the dynamic behavior of the bridge. However, practical experience has proven
that bridge-spans of up to 120 m, no significant dynamic effects due to wind load has taken place. For
minimizing dynamic wind-effects on the bridge sufficient dead load, a steel deck and most favourable
span to sag ratios have been introduced. Therefore, no lateral stabilizing measures (windguy system)
is considered in this standard suspended bridge design. However, for special cases (spans more than
120m or extreme windy areas exceeding wind speed of 160km/h) there is a provision for fixing a windguy
system.
¡¡ Snow load
The probability of occurrence of a full load on a bridge loaded with snow is low. Moreover, the design
live load itself is comparatively high. Therefore, it is considered that the snow load is already covered by
the design live load.
¡¡ Temperature effect
Loading (cable forces) according to temperature effect is negligible in comparison with other loading
conditions. Therefore, temperature effect has not been considered for the standard bridge design.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 43

¡¡ Seismic load
There is low probability of a full live load occurring at the time of an earthquake. Moreover, the design
live load itself is comparatively high. Therefore, it is assumed that the seismic load is already covered by
the design live load. A separate loading combination with seismic load is not considered.

2.3.2 Construction materials

The construction materials used for the standard trail bridges are wire ropes (cables), steel parts,
steel fixtures and fasteners (thimbles, bulldog gripes, nuts and bolts), concrete and stone masonry.
Specifications of these materials are based on Indian Standard (IS). However, for frequently referred
material, specifications are given for quick reference.
¡¡ Steel Wire Ropes (Cables)
For more convenient stock keeping, handling and transportation only three cable diameters are applied.
This also reduces the number of steelparts, logistics and eventually bridge costs.

Steel wire ropes should comply with all the requirement of:
IS 1835 – 1977 Steel wires for ropes
IS 6594 – 1977 Technical supply conditions for wire ropes and strands
IS 9282 – 1979 Specification for wire ropes and strands for suspension bridges
IS 9182 (part II) –1979 Specification for Lubricants for Wire Strands and Ropes

Nominal diameters: 13 mm 26, 32 mm

Rope Construction 7 x 7 (6/1) 7 x 19 (12/6/1)


Elongation non pre-stretched pre-stretched
Lay RHO, right hand ordinary lay
Core WRC, Wire Strand Core
Tensile strength of wire 160 kg/mm2 or 1.57 kN/mm2
Preforming Preformed
Coating Galvanized ‘A’ heavy
Impregnation Non-drying type and non-bituminous Lubricant

Nominal Diameter of Minimum Permissible


Weight Metallic Area
Cable Breaking Load in load in
[mm2]
mm kg/m N/m Tones kN Tones kN
13 0.64 6.3 73 10.5 103 3.5 34
26 2.51 24.6 292 39.3 386 13.1 129
32 3.80 37.3 442 59.6 585 19.9 195
Modulus of Elasticity, E = 110 kN/mm2 = 11 t/ mm2

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 44

Figure F.2.4: Cross-Section & lay of Wire Rope


¡¡ Bulldog Grips
Bulldog Grips are used for cable terminal to secure the cable ends.

Bulldog grips should conform to IS 2361-1970, specifications for Bulldog Grips. The bridges must be
dropforged and suitably scored to grip a round strand rope of right hand lay having six strands. Bridges,
U-bolt and nuts should be hot dip galvanized with minimum zinc coating of 40μm. The size of the bulldog
grip is equal to the size of the cable to be anchored or connected.

Figure F.2.5: Bulldog Grip

¡¡ Structural Steel
Steel grade should be of standard quality Fe 410 and structural steel should comply with the requirement
of:
IS 226 – 1975 Structural Steel
IS 800 – 1984 General Construction in Steel

The steel should have the mechanical properties:


Tensile Strength : 410 N/mm2
Yield Stress : 250 N/mm2
Modulus of Elasticity : 200 kN/mm2
Elongation : 23%
¡¡ Fasteners
Bolts, nuts and washers should be of grade C, property class 4.6 and should comply with the requirement
of:
IS 1363 - 1984 (Part 1) Hexagonal Head Bolts and Nuts
IS 1367 – 1983 Threaded Fasteners

All the fasteners should be hot dip galvanized with minimum zinc coating of 40μm.
¡¡ Reinforcement Steel
Reinforcement Steel should be of steel grade Fe415, high yield deformed bars and should comply with
the requirement of:
IS 1786 – 1985 High Strength Deformed Steel Bars for Concrete Reinforcement
IS 456 – 1978 Plain and Reinforced Concrete

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 45

The Reinforcement Steel should have the mechanical properties:


Yield Stress = 415 N/mm2
Modulus of Elasticity = 210 kN/mm2
¡¡ Rust Prevention
Rust prevention of steel parts should be done by Hot Dip Galvanization according to:
IS 2629 – 1966 Recommended Practice for Hot Dip Galvanization of Iron and Steel
IS 4759 – 1984 Specification for Hot-Dip Zinc Coating on Structural Steel

Minimum Mass of Coating Minimum Thickness of


Products Coating
g/m2 N/m2 μm
Structural Steel 610 6.0 80
Threaded work, nuts and bolts 300 3.0 40

¡¡ Wire Mesh Netting


Wire for wire mesh netting should comply with the requirement of:
IS 280 - 1978 Mild Steel Wire for General Engineering Purposes
IS 4826 - 1979 Hot-dipped Galvanizing Coatings on Round Steel Wire

Diameter of wire should be 12 SWG (2.64mm) and zinc coating should not be less than 270 g/m2. The
average tensile strength of the wire should not be less than 380N/mm2.
¡¡ Concrete
Concrete should comply with all the requirements of:
IS 456 – 1978 Plain and Reinforced Concrete
IS 269 – 1989 Ordinary Portland Cement
IS 383 - 1970 Coarse and Fine Aggregate

Concrete Grades used in the standard design are:


Concrete 1:3:6 (M10) for miscellaneous use
Concrete 1:2:4 (M15) for structure
¡¡ Masonry
Masonry should comply with all the requirements of:
IS 1597 – 1967 Code of Practice for Construction of Stone Masonry
IS 2250 – 1981 Preparation and Use of Masonry Mortars

Stone masonry used in the standard design are:


Chisel Dressed Stone Masonry in 1:4 cement : sand mortar
Hammer Dressed Stone Masonry in 1:6 cement : sand mortar
Dry Stone Masonry.
¡¡ Unit Weight of Construction Materials
The unit weight of the construction material used in the standard design is given in the table.

Unit Weight,
Materials
kg/m3 kN/m3
Concrete 2200 22.0
Stone Masonry 2100 21.0
Steel 7850 78.5
Soil 1800 18.0
Water 1000 10.0

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 46

2.3.3 Structural analysis and design

Statical analysis is based on calculation of forces and stresses in the structures due to the external loadings.
Calculated forces and stresses are compared with the permissible loads and stresses of these structures
with some safety of factors.

The standard suspended bridge has been designed as per the statical analysis and basic design concept.
¡¡ Cable Design
A cable hanging between two supports and carrying a uniformly distributed load “q” forms a parabola.
Thus Main Cables and Walkway Cables are assumed to be of parabolic geometry. In suspended type
bridge, the main cables (walkway cables) and handrail cables carry the load equally proportional to their
sectional area.

Figure F.2.6: Cable Geometry

¡¡ Maximum permissible Height Difference of Walkway Cable Saddles,


Span(l)
h=
25
¡¡ Sag of the Cable at Midspan in Dead Load
for Spans up to 80m:
Span(l)
b=
20

or for Spans over 80m:


Span(l)
b=
22

The fundamental equations for the calculation of the cable forces are:
¡¡ Total Horizontal Tension,
g·l2
H=
8·b
¡¡ Total maximum Tension at the higher Foundation Saddle,
H
Tmax =
cos βmax

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 47

¡¡ Cable Inclination at higher Foundation Saddle,


4b+h
βmax = a tan( )
l

Example:
4 l/22+ l/25 0.1818·l+0.04l
βmax = a tan a tan a tan 0.2218 = βmax = 12.51º
l l

The Safety Factor for Cable Breaking Load is taken as fs ≥ 3.

As per above design concept, a simplified cable combination chart has been developed and standardized
for suspended type bridge for different walkway width and span up to 120m.

These cable combinations depend only on the span and the chosen walkway width and do not depend on
the site condition. The cable combination has been designed to satisfy the worst topographic condition,
i.e. considering maximum allowable level difference between the cable saddles of right and left banks.
Therefore, in more favourable topographic conditions, the cable may seem to be over designed, but, as
per analysis this has no significant effect (less than 1%) on the overall cable strength.

¡¡ Steel Parts
Steel parts of the suspended type bridge are mainly subject to axial tension, bending stress, shear stress
and bond stress. All the steel parts are designed for these forces and stresses with a safety factor of
fs = 1.6.

All the steel parts have been standardized according to the cable combinations and do not depend on
the site condition.
¡¡ Walkway Structure
Walkway structures (Steel deck, Cross beams, hangers) are designed for local loadings as steel parts with
safety factor of, fs = 1.6. Concentrated load P = 150 kg (weight of a porter with load) positioned along
the walkway as well as across the walkway has been considered.

Figure F.2.7: Walkway Structure

¡¡ Main Anchorage Foundations


There are mainly three basic types of main anchorage foundations:
• Foundation on Soil
• Foundation on Hard Rock
• Foundation on Fractured Hard Rock or Soft Rock

All foundations are designed as per the static analysis and principle of soil / rock mechanics. Soil and rock
parameters are determined by the site investigation. mode of failure and respective safety factors are
considered in the design:
• Sliding FSL ≥ 1.5
• Toppling FT ≥ 1.5

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 48

• Bearing Capacity FBC ≥ 2.0


• Maximum Ground Pressure σmax < Bearing capacity of soil or rock
The standard design is without windguy system. Therefore, foundations are designed to take the wind
load satisfying the worst case from the two load cases.
• Load case A = dead load + full wind load
• Load case B = full load + 1/3 wind load

¡¡ Foundation in Soil
RCC deadman and gravity soil anchor block has been designed as a main anchorage foundation on soil.
Consider the design concept of the foundation.

Figure F.2.8: Foundation in Soil

The foundations have been designed to satisfy the safety against sliding with the theoretical basis.
Safety factor against sliding:
Re sisting Force R’v·tanϕsl
Fsl = = >1.5
Driving Force R’H

Where, R’V = RV · cosα + RH· sinα RV = W1 + W2 +…+ Wn + Eav - Tv


R’H = RV · sinα + RH · cosα RH = TH +Eah

The gravity load has also been maintained to resist the uplift force at the dead man beam by providing
minimum width (B1min) and height (H1min).

Safety against uplifting:

Tmax · sinα = (b+B1min) · 0.5H1min · Lbeam

B1min = b + H1min · tan30°

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 49

The position of the foundation behind the slope line of the internal angle of friction guarantees the safety
against the Ground Shear failure.

Further, a curtain wall with cement stone masonry of sufficient thickness has been provided to resist the
lateral pressure and bulging effect. The lateral pressure has been calculated as lateral pressure against
rigid wall as per elastic theory with the formula.

¡¡ Foundation in Hard Rock


RCC Concrete Drum with rock anchoring is used for foundation on sound hard rock. The sizes of drum
and anchorage rods are designed to withstand the shear due to the horizontal tension of the cables. It is
also designed to resist the uplift force at Drum.

Required number of anchor rods, N is calculated as:

N is adopted from whichever is higher from above.

¡¡ Foundation in Fractured Hard Rock or Soft Rock


Drum anchorage without rock anchoring is used for foundation on fractured hard rock or fresh soft rock.

The size of drum is designed to withstand the shear failure due to horizontal tension of the cables.
Determine the Drum size and number of reinforcement bars as:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 50

¡¡ Stone Masonry Tower


Stone Masonry Tower is designed to be safe against overturning and maximum pressure on foundation.
The ground bearing capacity (shear failure) is controlled by placing the tower foundation behind the
critical slope line.

Forces on top of the Tower:

Forces at Tower Base and Soil Pressure


Safety factor against overturning:

Safety against Ground Bearing Pressure:

Eccentricity should be within the permissible limit so that there is no negative pressure on the foundation.
It is achieved by meeting the condition.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 51

¡¡ Reinforced Concrete Works


All the reinforced concrete works are designed as per the principle of RCC design in working stress. For
detail refer to the relevant literature of Reinforced Concrete design.

2.4 Design of standard Suspended bridge

2.4.1 The major bridge components

Figure F.2.9: Major Bridge Components & Parameters

2.4.2 Design procedure

For designing a suspended standard bridge:


¡¡ Draw the bridge profile from the survey data;
¡¡ Fix the position of the bridge foundations and the span;
¡¡ Select walkway width;
¡¡ Select walkway cables and handrail cables from Design Form;
¡¡ Design walkway tower;
¡¡ Design main anchorage foundation;s
¡¡ Transfer data to the bridge profile and prepare the General Arrangement Drawing;
¡¡ Compile and fill in the standard design drawings;
¡¡ Calculate the quantities of works and prepare the Cost Estimate.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 52

2.4.3 Designing the position of the bridge foundations

Fix the position of the bridge foundations and the actual span of the bridge in the bridge profile. This
bridge profile will be the basis for the layout of the bridge at the construction site. Fulfill criteria while
fixing the position of the bridge foundations.
Criteria for fixing the Bridge Foundations
¡¡ The Bridge Foundations should be placed at least 3 meter back form the soil slope and 1.5 meter
back from the rock slope
¡¡ The Bridge Foundations should be placed behind the line of angle of internal friction of the soil or
rock. This angle is 35° for soil and 60° for rock.
¡¡ Level difference between the walkway cable saddles of two banks, h should not be more than
span/25
¡¡ Walkway tower height should be as small as possible. However, walkway cable saddle should be at
least a height of 1.3 meter from ground but should not be at a height more than 3.0 meter.
¡¡ Free board F, between lowest point of the bridge in dead load case and the high flood level should
be not less than 5.0 m

Criteria for fixing the location of the bridge foundation

Procedure for fixing the Bridge Foundations

According to the above criteria,


¡¡ Draw the bridge axis profile on an A3 size paper in a scale 1:200 (for up to 50m span) or 1: 400
(for span above 50m) with all details like axis points A and B, HFL, WL and tentative location of the
walkway towers at both banks based on the survey data.
¡¡ Fix location of the walkway towers at both banks.

Case-1: When Tentative Position of the Towers has been fixed during the Survey.

Step 1: Fix the Front of the Tower and check with Slope Line.

Mark front of the tower as fixed during the survey. Check position of the front of the tower as per minimum
required distance from bank edge and slope line. If minimum required distance from bank edge is not
sufficient, shift its position backward. Towers should be located behind the slope line. If tower is out of
the slope line, shift its position backward.

Step 2: Fix the Position of the Walkway Saddle and Bridge Span, l.

Mark position of the walkway cable saddles. Walkway cable saddles should be at height of 1.3m at soil
slope and 0.8m at rock slope from the ground at tower front. Thus, fix its elevations, El, Eh & l.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 53

Step 3: Check the Level Difference, h.

Level difference ‘h’ between walkway cable saddles of two banks should be less than l/25. If ‘h’ is found
more than the limit:
Rise the elevation of the walkway cable saddle of lower bank by increasing saddle height (in the series of
1.3, 2.3, 3.3 meter) but not more than 3.3 meter from the ground level in case of Flat slope
Or
Shift the position of the tower of lower bank backward to gain the required walkway cable elevation in
case of Hill Slope.
Or
Lower the elevation of the walkway cable saddle of higher bank. Avoid deep earth cutting.

Step 4: Calculate the vertical Distance, fmin and check the Free Board, fb.

Calculate vertical distance fmin between the lowest point of the bridge and walkway cable saddle of lower
bank,
(4 · l-20h)2
Fmin =
320 · l

Draw line of lowest point of the bridge.

Check available free board between lowest level of the bridge and high flood level.

Freeboard should be not less than 5.0m. If free board is not sufficient:

Raise the elevations of walkway cable saddles at both banks. This can be done either by raising the tower
height in case of the flat ground or by placing the tower further back in case of the hill slope.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 54

Step 5: Finalize the Bridge Profile.


Finalize bridge profile with final span and elevation of the walkway cable saddles.

Case-2: When the Position of the Towers has not been fixed during the Survey.

Step 1: Fix the Free Board Line and Front of the Towers.

Mark minimum free board level. Minimum free board from high flood level is 5.0 meter.

Fix the position of the front of the tower maintaining minimum required distance from bank edge and
slope line.

Step 2: Fix the approximate Bridge Span, l and minimum Level of Walkway Cable Saddles.
Calculate approximate bridge span as distance between the tower fronts.
Mark minimum level of walkway cable saddles as per required sag of the cable,
l
b=
20

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 55

Step 3: Fix the Position of the Walkway and Cable Saddles.

Walkway cable saddles should not be below the minimum level.

Walkway cable saddles should be 1.3 meter above the ground level at soil slope and 0.9 meter at rock
slope.

Thus, fix the elevations of the walkway cable saddles, El and Eh.

Step 4: Check the Level Difference, h.

Level difference ‘h’ between walkway cable saddles of two banks should be less than l/25. If ‘h’ is found
more than the limit:

Rise the elevation of the walkway cable saddle of lower bank by increasing saddle height (in the series of
1.3, 2.3, 3.3 meter) but not more than 3.3 meter form the ground level in case of flat ground.
or
Shift the position of the tower of lower bank backward to gain the required walkway cable elevation in
case of hill slope.
or
Lower the elevation of the walkway cable saddle of higher bank. Avoid deep earth cutting.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 56

Step 5: Calculate the vertical Distance fmin and check the Free Board fb.
Calculate vertical distance fmin between the lowest point of the bridge and walkway cable saddle of
lower bank.

Draw actual line of lowest point of the bridge


(4 · l-20h)2
Fmin =
320 · l
Check available free board between lowest level of the bridge and high flood level.

Freeboard should be not less than 5.0 meter. If free board is not sufficient:

Raise the elevations of walkway cable saddles at both banks. This can be done either by raising the tower
height in case of the flat ground or by placing the tower further back in case of the hill slope.

Step 6: Finalize the Bridge profile.

Finalize bridge profile with final span and elevation of the walkway cable saddles.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 57

2.4.4 Cable Design

Designing the cable for a bridge involves selecting required numbers and diameter of the handrail and
walkway cables for given span and selected walkway width.

To design the cable:


¡¡ Select the appropriate walkway width (70cm or 106cm) according to the nature of the traffic and
type of trail.
¡¡ Fix the span of the bridge and height difference of cable saddles of the right bank and left bank
from the bridge profile.
¡¡ Select cables from the Table: Selection of Cable according to the span and selected walkway width.

Design the cable structure

1. Survey data & calculation of freeboard

1. Span of the bridge l = ................... m


2. Saddle Elevation of the Walkway Cable on the higher Side Eh = ................... m
3. Saddle Elevation of the Walkway Cable on the lower Side El = ................... m
4. Difference in Elevation h = Eh - El = h = ................... m
(max. permissible height: hmax = l/25)
5. Dead Load Sag: for Span up to 80m
l
bd= = bd = ................... m
20
for Span over 80m
l
bd= = bd = ................... m
22
6. Fmin in Dead Load case (at the lowest point of the cable)
(4 · bd -h)2
fmin= = bd = ................... m
16 · bd
7. Highest Flood Level Hfl = ................... m
8. Free Board (min. 5.00m) fb = El - Hfl = Fb = ................... m

(if freeboard is less than 5.00m, try either to raise the saddle elevations or to adjust the span,
but keep the ratio between span and sag always fixed at l/bd =20 or l/bd =22 )

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 58

2. Selection of Cables
Select a cable combination according to the span and walkway width of the bridge from the table.
Always select the higher cable combination when the span is in between two values.

Maximum Span for


Cable Combinations Weight of
Walkway Width:
all Cables
Handrail Walkway gh
70cm 106cm
Cables Cables
span [m] span [m] nos Ømm nos Ømm [kg/m]
50 40 2 26 2 26 10.04
90 60 2 26 2 32 12.62
100 75 2 26 4 26 15.06
120 105 2 26 4 32 20.22
---- 120 2 32 4 32 22.80

3. Calculation of Cable Length

dia Backstay Length * Cutting Length**


Type of Cable Nos
[mm] [m] [m/pc]
Fixation Cable 13 2 …………. …………
Handrail Cable …….. 2 ………… …………
Walkway Cable …….. …… ..………. …………
Notes:
* Backstay Length = Cable length between saddle center and center of dead man or drum as per foundation drawing
(both banks) + 6.0m.
** Cutting Length = 1.1 x Span + Backstay Lengths

4. Calculation of (fmin) Hoisting Sag


This calculation has to be made after tower and foundation work is completed

1. Actual span measured in the field l = ................... m


2. Saddle Elevation of the Walkway Cable on the higher side Eh = ................... m
3. Saddle Elevation of the Walkway Cable on the lower side El = ................... m
4. Difference in Elevation h = Eh - El = h = ................... m
(max. permissible height: hmax = l/25)
5. Dead Load Sag: for Span up to 80m
l
bd= = bd = ................... m
20
for Span over 80m
l
bd= = bd = ................... m
22
6. Hoisting sag bh = 0.95 x bd = bh = ................... m
7. Fmin in Dead Load case (at the lowest point of the cable)
(4 · bd -h)2
fmin= = fmin = ................... m
16 · bd

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 59

2.4.5 Design of bridge foundation structures

Design of a bridge foundation structure is mainly to select the standard anchor block types for right bank
and for left bank and fill in the required data in the selected drawings.

Standard anchor block (bridge foundation structure) types have been developed for all possible cases up
to span 120m. The design concept and statical analysis has been used.

There are basically seven types of anchor blocks depending upon the soil or rock type, whereof the
typical designs are illustrated:

Figure F.2.10: RCC Deadman & gravity soil anchor block on flat topography

Figure F.2.11: RCC Deadman & gravity soil anchor block on slope topography

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 60

Figure F.2.12: RCC Single Drum Rock Anchor Block in Hard Rock

Figure F.2.13: RCC Double Drum Rock Anchor Block in Hard Rock

Figure F.2.14: RCC Single Rock Anchor Block in Fractured Hard Rock & Soft Rock

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 61

Figure F.2.15: RCC Double Drum Rock Anchor Block in Fractured Hard Rock & Soft Rock

Figure F.2.16: RCC Deadman Anchor Block in Fractured Hard Rock and Soft Rock

To select an Anchor block type:


¡¡ Define the walkway width.
¡¡ Define the span of the bridge from the bridge profile
¡¡ Define the topography of the ground where the anchorage block will be placed as flat or slope.
¡¡ The topography is defined as flat if the ground slope is less than 10°, and slope if the ground slope
is more than 10°.
¡¡ Define the soil or rock type from the survey form and checklist
¡¡ Define the tower height from bridge profile . Tower height = Height of walkway cable saddle from
the ground + 1.1m in case of soil bank and flat topography. Tower height = 2.4m (fixed) in case of
soil bank and slope topography. Tower height = 2.0 m (fixed) in case of rock bank.
¡¡ Select the anchor type and the corresponding drawing from the selection tables according to the
above design data.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 62

Design the Anchor Types

I. Design Data

• Walkway Width, WW (70 or 106cm): ............................cm

• Bridge Span: ............................m


Right Bank Condition
Geology:
Soil
If Soil, how is the Ground Surface?

What is the Soil Type? Flat Hill slope

Tower Height from Ground up to (up to 15° slope) (more than 15° slope)
H.C.Saddle (data from bridge profile): or
Gravelly Sandy Silty

2.4m 3.4m 2.4m

If rock, what is the rock type?


Hard rock Hard rock Soft rock

(only a few fractures) (highly fractured)


Tower Height
2.0m in case of rock

Left Bank Condition


Geology:
Soil
If soil, how is the ground surface?

What is the Soil Type? Flat Hill slope

(up to 10° slope) (more than 10° slope)


or
Gravelly Sandy Silty

2.4m 3.4m 4.4m

If Rock, what is the Rock Type?


Hard rock Hard rock Soft rock

(only a few fractures) (highly fractured)


Tower Height
2.0m in case of rock

II. Selection of Anchorage Types


Select appropriate anchorage type at Right Bank and Left Bank according to the above design data.
Procedure for Selection:
¡¡ According to the Soil/Rock type and Slope of the ground, refer to respective tables for selection
of Anchorage Block.
for Soil and Flat Ground : Table F.2.1
for Soil and Hill Slope : Table F.2.2
for Hard Rock : Table F.2.3 or Table F.2.4

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 63

for Fractured Hard Rock or Soft Rock:


Span up to 90m (WW = 70cm) and up to 60m (WW = 106cm) : Table F.2.5 or Table F.2.6
Span Range 91-120m (WW = 70cm), 61-120m (WW = 106cm) : Table F.2.7
¡¡ In the table match the design data:
Selected Walkway Width  Bridge Span  Tower Height  Soil type  Select the corresponding
Anchor Type and Drawing No. for right bank and left bank respectively.

III. Anchor Type Selection Tables


¡¡ In soil and flat ground:

Table F.2.1: Selection of RCC deadman & gravity soil anchor block in flat ground

Span Range, m Tower


Foundation Block Drawing
Height
Walkway: 70cm Walkway: 106cm Soil Type Type No.
[m]
2.4 1F 21Dcon
Up to 45m Up to 30m 3.4 All 2F 22Dcon
4.4 3F 23Dcon
2.4 4F 24Dcon
46 - 90 31 - 60 3.4 All 5F 25Dcon
4.4 6F 26Dcon
2.4 7F 27Dcon
91 - 120 61 - 75 3.4 All 8F 28Dcon
4.4 9F 29Dcon
2.4 10F 30Dcon
– 76 - 90 3.4 All 8F 28Dcon
4.4 11F 31Dcon
2.4 12F 32Dcon
– 91 - 105 3.4 All 8F 28Dcon
4.4 13F 33Dcon
Gravely 12F 32Dcon
2.4
Sandy, Silty 14F 34Dcon
– 106 – 120
3.4 15F 35Dcon
All
13F 33Dcon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 64

¡¡ In Soil and Hill Slope:

Table F.2.2: Selection of RCC Deadman & Gravity Soil Anchor Block in Hill Slope

Span Range, m Tower


Foundation Block Drawing
Height
Walkway: 70cm Walkway: 106cm Soil Type Type No.
[m]
Up to 60m Up to 40m 2.4 All 1S 41Dcon
61 – 90 41- 60 2.4 All 2S 42Dcon
91 - 120 61-75 2.4 All 3S 43Dcon
Gravely 4S 44Dcon
– 76 - 90 2.4 Sandy 5S 45Dcon
Silty 6S 46Dcon
Gravely, Sandy 7S 47Dcon
– 91 - 105 2.4
Silty 8S 48Dcon
Gravely, Sandy 8S 48Dcon
– 106 - 120 2.4
Silty 9S 49Dcon

¡¡ In Hard Rock for all Span Range:

Table F.2.3: Selection of RCC Single Drum Anchor in Hard Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.0 1HRS 61Dcon


91 – 120 61 - 120 2.0 2HRS 62Dcon
When slope is too steep and there is not enough space for single drum anchorage system, select the
double drum system from Table F.2.4.

Table F.2.4: Selection of RCC Double Drum Anchor in Hard Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.0 1HRD 63Dcon


91 – 120 61 - 120 2.0 2HRD 64Dcon

¡¡ In Fractured Hard Rock/Soft Rock for Span Range up to 90m (WW = 70 cm) and 60m (WW =
106cm):

Table F.2.5: Selection of RCC Single Drum Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.0 1FRS 65Dcon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 65

When slope is too steep and there is not enough space for a single drum anchorage system (Table F.2.5),
select the double drum system from Table F.2.6.

Table F.2.6: Selection of RCC Double Drum Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.0 1FRD 66Dcon

¡¡ In Fractured Hard Rock/Soft Rock for Span Rang of 91- 120m (WW = 70 cm)and 61-120m (WW =
106cm):

Table F.2.7: Selection of RCC Deadman Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

91-120 61-120 2.0 2FRD 67Dcon

Selected Anchorage Foundation Type and corresponding Drawings from the Tables above:
Right Bank: Anchor Type...............................Drawing No...........................
Left Bank: Anchor Type...............................Drawing No...........................

IV. List of drawings


Select the required Steel Drawings and Construction Drawings according to the walkway width, selected
cables and selected Anchorage Block types.

Prepare a General Arrangement drawing for individual bridge design.

2.4.6 Other structures

Besides the bridge structure, some other adjacent structures may be required for overall bridge stability
and for safety measures. These are the structures.

¡¡ Windguy Arrangement
¡¡ Retaining Structures
¡¡ Slope Protection works
¡¡ River Bank Protection

¡¡ Windguy Arrangement
Generally Windguy Arrangements are not required for bridges with span of up to 120m. Therefore, no
lateral stabilizing measures (windguy system) is considered in this standard suspended bridge design.
However, for special cases (spans more than 120m or extreme windy areas exceeding wind speed of
160km/h) there is a provision for fixing a windguy system. Refer the Steel Drawing for the windtie cable
clamps from Drawing No. 11A and respective Construction Drawings for different types of windguy cable
anchorages from Drawing Nos 51Acon, 52Acon, 53Acon, 54Acon, 57Acon and 58Acon.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 66

The Layout of the Windguy Arrangement

The Geometry of the Windguy Arrangement

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 67

The design of the Windguy Arrangement is to:


¡¡ Select the windguy cable and windties cable
¡¡ Select the windguy cable anchor blocks
¡¡ Calculate geometry of the windguy arrangement

Selection of Windguy Cable and Windties Cable:


Select the windguy & windtie cables from the table according to the span of the bridge.

Selection of Windguy Cable and Windtie Cable

Span, m Windguy Cable, mm Windtie Cable, mm


Up to 150 26mm 13
150 - 200 32mm 13

Selection of Windguy Cable Anchor Type:


There are each two types of gravity Soil Anchor Blocks, gravity Rock Anchor Blocks and Drum Anchors
depending on the soil or rock type and the diameter of the Windguy Cable.

Select the Windguy Cable Anchorage Type from Tables:

Selection of Windguy Cable Gravity Anchor Block on Soil

Windguy Cable Ø [mm] Block type Drawing No


26 Soil Block 51Acon
32 Soil Block 53Acon

Selection of Windguy Cable Gravity Block on Rock *

Windguy Cable Ø [mm] Block type Drawing No


26 Rock Block 52Acon
32 Rock Block 54Acon

Selection of Windguy Cable Drum Anchor on Hard Rock*

Windguy Cable Ø [mm] Block type Drawing No


26 or 32 Hard Rock Drum 57Acon

Selection of Windguy Cable Drum Anchor on Fractured Hard Rock or Soft Rock*

Windguy Cable Ø [mm] Block type Drawing No


26 or 32 Soft Rock Drum 58Acon
* If both banks are rock, select the Drum Anchor for one bank and the Gravity Rock Anchor for the other bank.

If one bank is rock and the other bank is soil, select always the Drum Anchor for the rock bank.

Retaining structures

Retaining structures are necessary to retain the earth (soil, fractured rock and weathered soft rock) behind
the anchorage blocks of the bridge. There are many options of retaining structures. But for the trail bridge

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 68

construction most feasible are retaining walls. Retaining walls can be of gabion boxes, rubble masonry
and dry stone masonry. For the Short-Span trail bridge construction dry stone retaining wall or breast wall
are preferable, since they require only local materials.

The choice between retaining wall and breast wall depends on different factors, such as available space
behind the blocks, required height of the protection, soil conditions etc.

Retaining walls are used when the earth to be retained is loose soil with large protection height.

Retaining Walls

Banded Dry
Type Dry Stone
Stone / Masonry

Section

Top width, Wt 0.6 - 1.0m 0.6 - 1.0m


Base width, Wb 0.5 - 0.7H 0.6 - 0.65H
Front batter varies varies
Back batter varies vertical
Inward dip of
1:3 1:3
foundation, n
Foundation depth ≥0.50m ≥0.50 - 1m
Range of height, H 1 - 6m 6 - 8m
Hill slope angle, a < 35º 20º

Breast walls are used when earth to be retained is fractured or weathered rock or compact soil with
temporarily unstable nature.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 69

Banded Dry
Type Dry Stone
Stone / Masonry

Section

Top width, Wt 0.50m 0.50m


Base width, Wb 0.29H 0.3H 0.33H > 0.5H
Back batter 3:1 4:1 5:1 3:1
Inward dip of foundation, n 1:3 1:4 1:5 1:3
Foundation depth ≥0.50m ≥0.50m ≥ 0.50m ≥0.50m
Range of height, H <6m <4m <3m 3 - 8m
Hill slope angle, a 35 - 60º 35 - 60º

Slope protection measure


Slope protection measure depends on the factors influencing slope instability. It is recommended to
select the bridge site, where there are no slope instability features. However, often it is necessary to drain
out the surface runoff and seepage water from the slope as a slope protection measure.

Water should be collected as closely as possible from its origin and safely channeled to a nearby
watercourse. The surface drainage can be catch drain on the slope or drainage around the anchorage
foundation or combination of both. The choice depends on the position of the anchorage foundation and
the profile of the natural terrain.

The drain should be open type.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 70

To avoid self-scouring, the drain outlet should be protected.

In seepage area, sub-surface drainage is required around the anchorage foundation. Consider the typical
layout design of the sub-surface drain.

A typical design of the sub-surface drain is:

Bio-Engineering
Surface drainage alone may not be sufficient for unstable slopes. The most effective method for stabilizing
such slopes is bio-engineering in combination with light civil structures such as catch drains, check dams,
cascades etc. This is a cheap and easy method. The main concept of this method is to grow trees, plants
such as shrubs or grasses. Deep rooted and fast growing trees and plants are most suitable for this
purpose. Proper selection of plant types is most important and should be based upon local experience.
Some of the vegetation measures are:
¡¡ Planted grass lines: contour/horizontal or down slope/vertical or random planting;
¡¡ Grass seeding;
¡¡ Turfing;

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 71

¡¡ Shrub and tree planting


¡¡ Shrub and tree seeding
¡¡ Fascines (bundles of live branches are laid in shallow trenches)

River bank protection


River protection works are of temporary nature and costly. This requires frequent maintenance to keep
the structure functional. Therefore, avoid bridge site, which requires river protection works as far as
possible. This is a complex subject and cannot be covered by this Manual.

2.4.7 General arrangement drawing

General Arrangement (GA) drawing shows the overall plan and profile of the bridge. The GA should
reflect the major components of the bridge and its geometry including the elevations of the foundations
at the right and left bank. The GA is required for overall view of the designed bridge and also for layout
of the bridge for construction.

Draw the GA on the same bridge profile and mention the data on the plan and the profile of the bridge:
¡¡ Span and dead load sag;
¡¡ Distance from the axis point A and B (which, were fixed during survey) and the center of the tower,
¡¡ Cable elevations at saddles, elevation of the lowest bridge point and all bridge foundation levels,
¡¡ Over all dimensions of the bridge structures and its elevations.

The completed GA should be sufficient for the layout of the bridge.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 72

2.4.8 Design example

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 73

2.5 Bridge standard drawings

2.5.1 Introduction and overview of drawings

The Bridge Standard Drawings represent the centerpiece of the Short Trail Bridge Standard. They are
composed as a unit component system and are categorized in two categories:
¡¡ Suspended Bridge Construction Drawings
¡¡ Suspended Bridge Steel Drawings

Both drawing categories are linked with each other and depending on the bridge design the required
drawings are selected.

Table F.2.8: Suspended Bridge Construction Drawings

Drawing Titles Drawing Nos.


FITTING Walkway Fitting for 70 cm Walkway Width 19Dcon70
DETAILS Walkway Fitting for 106 cm Walkway Width 19Dcon106
TOWER CSM Tower & RCC Core for 70 cm Walkway Width 20Dcon70
DETAILS CSM Tower & RCC Core for 106 cm Walkway Width 20Dcon106
RCC Deadman & Gravity Soil Anchor Block in Flat Ground 21Dcon - 26Dcon
GROUND

for 2 Walkway Cables (6 Drawings)


FLAT
SOIL ANCHORS

RCC Deadman & Gravity Soil Anchor Block in Flat Ground 27Dcon - 35Dcon
for 4 Walkway Cables (9 Drawings)
RCC Deadman & Gravity Soil Anchor Block in Hill Slope 41Dcon &
for 2 Walkway Cables 42Dcon
SLOPE
HILL

RCC Deadman & Gravity Soil Anchor Block in Hill Slope 43Dcon - 49Dcon
for 4 Walkway Cables (7 Drawings)
RCC Single Drum Anchor in Hard Rock for 2 Walkway Cables 61Dcon
HARD ROCK

RCC Single Drum Anchor in Hard Rock for 4 Walkway Cables 62Dcon
ROCK ANCHORS

RCC Double Drum Anchor in Hard Rock for 2 Walkway Cables 63Dcon

RCC Double & Single Drum Anchor in Hard Rock 64Dcon


for 4 Walkway Cables
RCC Single Drum Anchor in fractured Rock for 2 Walkway Cables 65Dcon
FRACTURED
ROCK

RCC Double Drum Anchor in fractured Rock for 2 Walkway 66Dcon


Cables
67Dcon
RCC Deadman Anchor in fractured Rock for 4 Walkway Cables
Gravity Soil Block for Cable ø 26mm 51Acon
Anchor Drawings

Gravity Rock Block for Cable ø 26mm 52Acon


for Windguy

optional
Cables

Gravity Soil Block for Cable ø 32mm 53Acon


Gravity Rock Block for Cable ø 32mm 54Acon
RCC single Drum Anchor in Hard Rock, Cable ø 26 or 32mm 57Acon
RCC single Drum Anchor in Fractured Rock, Cable ø 26 or 32mm 58Acon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 74

Table F.2.9: Suspended Bridge Steel Drawings

Drawing Titles Drawing Nos.


Crossbeam for 2 Walkway Cables for walkway width = 34 cm* 01D*

Crossbeam for 2 Walkway Cables for walkway width = 70 cm 02D


WALKWAY
Crossbeam for 4 Walkway Cables for walkway width = 70 cm 02D4
CROSS BEAMS
Crossbeam for 2 Walkway Cables for walkway width = 106 cm 03D

Crossbeam for 4 Walkway Cables for walkway width = 106 cm 03D4


Steeldeck Standard Panel, length = 198 cm / width = 34 cm 08A

STEEL DECK Steeldeck Standard Half Panel, length = 98 cm / width = 34 cm 09A

Steeldeck Special Panel, length = 223 cm / width = 34 cm 10A


Saddles & Reinforcement for
RCC Deadman & Gravity Soil Anchor for 2 Walkway Cables 20D2
SOIL

Saddles & Reinforcement for 20D4


REINFORCEMENT

RCC Deadman & Gravity Soil Anchor for 4 Walkway Cables


SADDLES &

Saddles & Reinforcement for 20D4S


RCC Deadman Anchor in fractured Rock for 4 Walkway Cables
ROCK

Saddles & Reinforcement for 60D2


Drum Rock Anchor for 2 Walkway Cables

Saddles & Reinforcement for 60D4


Drum Rock Anchor for 2 Walkway Cables
Windties Cable Clamps for Windguys Cable ø 26 or 32 mm 11A
optional
Windguys Cable Anchorage for one Cable End ø 26 or 32 mm 50A
* Also a narrow walkway of 34cm width (panel wide only) has been developed, but is not used very often and is, therefore not
considered in this Handbook; but can be used, if dimensions of anchor blocks are designed accordingly.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 75

Table F.2.10: Legend for the drawing numbers and suffixes:

Drawing No. Suffix Bridge or Drawing Type


11 A For All bridge types

DRAWINGS
STEEL 02 D For suspended bridge types
02 D4 For 4 walkway cables
02 D4W Suitable for Windguy cables
20 D4S Special
42 Dcon construction drawings
20 Dcon 70 For 70 cm walkway width
CONSTRUCTION

Anchor Drawings (Block Types)


DRAWINGS

3F
Block Type 3 in Flat Ground
5S Block Type 5 in Hill Slope
1HRS Block Type 1 in Hard Rock for Single Drum
1HRD Block Type 1 in Hard Rock for Double Drum
1FRS Block Type 1 in Fractured Rock for Single Drum
Block Type 1 in Fractured Rock for Deadman
1FRD

2.5.2 Concept of the standard drawings

Steel drawings

Each Drawing is providing the necessary information and specifications for manufacturing the desired
steel parts. Depending on the width of the walkway, the size of walkway cable and the span the empty
spaces in the materials list have to be filled in and the total weight has to be calculated.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 76

Section Quantity Weight


Part No. Working Drawing
[mm] [nos] kg/pc total Kg

Angle
(spacer)
40/40/5
l = …..

Width of walkway L X Weight / pc

70 cm 910 740 2.73 kg

106 cm 1270 1100 3.81 kg

5 holes Ø 14

fill in weight and length per piece (kg/pc according


to width of walkway and multiply by the quantity for
computing total Kg.

Ri-Bar
8 Ø 16
Width of walkway L Weight / pc
...R
l = ….. 70 cm 2100 3.32 kg
106 cm 2500 3.95 kg

Bulldog
9 Grip ...D
Ø …. for fixing first suspender at handrail cable
Ø 26 or 32 MS forged. according to ISI
standard, galvanized.

Plain
10 Rod * Erection Hooks ...R
Ø 20 needed at one
bank only

fill in the Ø and corresponding weight fill in the weight of erection


of the handrail cable hooks if needed.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 77

for fixing first & joining Fixation


13 Ø 13 12 1.4D

MS forged ISI standard


Cable Ø 13 mm

Bulldog Grips
Ø
14 ......
2 for Handrail Cable Ø 26 or 32 mm ...D

Ø
15 ......
2 for Walkway Cable Ø 26 or 32 mm ...D

A = ……………………….................. kg D = ……………………….......kg g = 10.64 kg


Total transportation Weight Total Structural Steel to be
B+C+D+R+0.16 kg Steel = (u+g) galvanized

C = 0.32 kg D = .........................kg R = .........................kg


Nuts, Bolts, Washers Bulldog Grips Reinforcement Steel

fill in the of the required Bulldog


compute total weights from weight
Grips plus the weight from the table
column of material list
below

Fill in :
Bridge Name
Bulldog Grips Weight
Cable No, Bank &
for two
Ø mm (kg/pc) Total kg Span
cables

26 10 1.10 11.00 Width of Walkway

32 12 1.30 15.60
No of required foundation

From the total weights of each drawing the grand total for each steel category has to be added up as
follows:

A: Means the entire weight of steel including galvanization to be transported to the site.
B: This is the total structural steel raw or untreated. This includes steel profiles, plates and flats but
not reinforcement bars and other steel items.
C: This is the weight of Nuts, Bolts & Washers (galvanized weight)
D: This is the weight of Bulldog Grips or Thimbles, if required (galvanized weight)
R: This is the raw weight of Reinforcement Steel or Plain Rods they are never galvanized.
The total transportation weight A = B + C + D + R
g: The little g indicates the weight of structural steel to be galvanized. This weight is part or can be
the sum of the total structural steel (B), but is not an additional weight.

Above distinction is made for quotation purposes, because the price per kg (or piece) is varying greatly
among each other. Reinforcement steel is much cheaper and Nuts & Bolts are much more expensive than
structural steel.

The weight of steel to be galvanized is necessary, to obtain the price for galvanization separately, without
the cost of the steel as such.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 78

Usually steel drawings are not necessary at the construction site but for assembly and identification of the
steel parts a copy of each steel drawing should be available. Also for maintenance at a later stage copies
of the steel drawings are useful.

Construction Drawings
The construction drawings are the actual site drawings of which one complete set is absolutely necessary
at the site. Depending on the required width of the walkway the corresponding “Walkway Fitting” &
“CSM Tower” drawings have to be selected (either 70 cm or 106 cm).

¡¡ Walkway Fitting Drawing;


(for 70 or 106 cm walkway)
¡¡ Details of CSM Tower & RCC Core.
(for 70 or 106 cm walkway)

The CSM Tower & RCC Core, drawing no. 20Dcon70 or 20Dcon106, are identical for all bridges.

For the actual anchorage arrangements there are two main categories of drawings:
¡¡ Soil Anchor Drawings;
¡¡ Rock Anchor Drawings.

Both drawing types are complete designs and are fulfilling respective parameters selected in the design
form. Also in both drawings the necessary quantities of construction materials are already calculated.
These have to be filled in respective tables given in the cost estimate Form.

The Soil Anchor Drawings are sub-divided in to:


¡¡ Soil Anchor Block in Flat Ground;
¡¡ Soil Anchor Block in Hill Slope.

In flat ground with a gradient of max 10o the block types 1F – 15F are applicable, whereas in slopes over
10o the block types 1S – 9S are to be applied.

It is absolutely necessary to fill in the Elevation and Cable diameters as indicated in the drawings. The
levels are to be determined in the topographic profile of he survey whereas the cable diameter can be
taken from the design form.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 79

2.5.3 Relationship between construction and steel drawings

Each Construction Drawing has related Steel Drawings. Respective related drawing numbers are
mentioned on the drawing itself and also respective steel parts numbers are indicated on the construction
drawing for easy reference.

The Construction Drawing “Walkway Fitting”, Nos 19Dcon70 or 19Dcon106 is showing the superstructure
of the bridge. The related Steel drawings are the corresponding Steel Crossbeam and Steel deck Drawings.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 80

The Steel Drawings for Saddle and Reinforcement, Nos 20 & 60 are related to the corresponding
Construction Drawings Nos 20Dcon70 (70 cm walkway), or 20Dcon106 (106 cm walkway).

Furthermore, depending of the soil conditions (soil or rock), the Steel Drawing Nos 20 & 60 are related
either to Anchor Drawings “Soil” or “Rock”.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 81

2.6 Calculation of quantity and cost estimate


Calculation of quantities and cost estimate is required for the purpose of planning and implementation

2.6.1 Implementation procedure

Use the Cost Estimate Form to calculate the quantities and to prepare the cost estimate.
¡¡ Quantity Calculation
• Calculate quantities of cables from Cable Design. Fill in the Quantity Calculation Sheet of
Wire Rope (Cables). This sheet will show the cable lengths of each diameter and the total
weight of the cables.
• Calculate quantities of Steel Parts and Steel Deck from the corresponding steel drawings. Fill
in the Quantity Calculation Sheet of Steel Parts and Steel Deck.
• Calculate quantities of Earth Works from General Arrangement Drawing. Fill in the Quantity
Calculation Sheet of Construction.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 82

• Calculate quantities of other construction works from the corresponding Construction


Drawings. Fill in the Quantity Calculation Sheet of Construction.
• Prepare list of construction materials according to the calculated quantities of construction
works.
• Calculate transportation weights for cables and other construction materials locally not
available.
• Calculate quantities of Works and Labor.
¡¡ Rate Analysis
Prepare the rate analysis for fabrication of steel parts, steel decks and road transportation (items of
external support to the community) as per unit quantity, unit cost and standard norms.
¡¡ Abstract of Cost
Compute the abstract of cost of the bridge as per the quantities of works (from Quantity Calculation
Sheets) and the rates (from Rate Analysis) for each item of works and summarize the cost as per
the category of works.
¡¡ Summary of Estimated Cost
• Bridge Cost: Calculate the Estimated Bridge Cost by summarizing the Abstract of Cost. Also
calculate bridge cost per m span.
• Contribution: Estimate the expected contribution from different partners.
• Breakdown of the Contribution: Break down the contribution as Local Contribution and
Outside Contribution.
¡¡ Summary of Actual Cost
In the majority of cases, the actual bridge cost will not be the same as estimated. Therefore,
calculate the actual bridge cost after completion of the bridge.

2.7 Construction

2.7.1 Bridge layout

The Bridge Layout is to fix the bridge position and foundations at the site as per design.

Procedure for General Bridge Layout (refer to General Arrangement ‘GA’ Drawing):
¡¡ Find the existing pegs and benchmarks.
¡¡ Measure the horizontal distance between axis pegs A (L) and B(R) and compare with the
measurement given in the General Arrangement.
¡¡ Check the elevations of the axis pegs A (L) and B(R) and compare with the elevations given in the
GA.

¡¡ If the horizontal distance between axis pegs A (L) and B(R) and its elevations is not similar to the
measurements given in the GA, readjust the design according to the actual measurements
¡¡ If the horizontal distance between axis pegs A (L) and B(R) and its elevations is identical to the
measurements given in the GA, fix the position of all foundation blocks.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 83

Procedure for Detailed Foundation Layout:


¡¡ Align the centerline of the bridge by joining the permanent points with mason threads or by
ranging between axis pegs ‘A’ and ‘B’.

¡¡ Mark the front of the tower foundation on the bridge centerline (peg 1) with reference to the axis
peg. The distance between front of the tower foundation and axis peg is given in the GA.

¡¡ Check the location of the front of the tower to ensure it has sufficient distance (minimum 3 m for
soil slope and 1.5m for rock slope) from the bank edge.
¡¡ Measure the length of the foundation from peg 1 and fix peg 2. Set up two additional centerline
pegs at safe distance for the excavation works (peg 3 and 4).

¡¡ Draw offset line (right angle) through peg 1 by 3-4-5 method. Starting from peg 1 set out pegs 5,
6, 7, and 8 for the reference line of the front edge.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 84

¡¡ Draw offset line through peg 2 for reference line of the back edge. Starting from peg 2 fix the pegs
9, 10, 11 and 12. Similarly, fix the reference line of the tower foundation with pegs 13, 14, 15, 16.

¡¡ Determine the reference line at the downstream edge with the help of pegs 5 & 9, for the upstream
edge, use pegs 6 & 10 (peg 19 & 20).

¡¡ Fix the elevation line (datum level) and indicate the depth of the excavation work for tower and
dead man or drum anchorage as per elevations in the GA and Anchorage Block drawings.

2.7.2 Foundation excavation

In Soil:

Foundations should be excavated with slopes to provide stability of the cut slope. The cut slope in soil
should not exceed to 3:1 (V:H). The foundation should be excavated stage wise.

¡¡ 1st Stage – See all excavation.

¡¡ 2nd Stage – Foundation excavation for tower.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 85

¡¡ 3rd Stage – Construction of tower as per design.

¡¡ 4th Stage – Final Excavation for dead man beam

In all of the above excavation stages, excavation depth should be accurately maintained. For this, establish
an elevation line (datum level) and measure the foundation depths with fixed stick.

All the excavated soils should be safely disposed without damaging the existing vegetation at down
hillside, thus not effecting the environment.

In Rock:
Rock excavation is necessary to prepare the platform for the drum anchorage. Rock should be excavated
manually without blasting.

Excavation in rock is done by first drilling holes to weaken the rock and then using the crowbars to break
up and dig out the rock parts. The cutting can be carried out by forming steps.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 86

1st Stage 2nd Stage

2.7.3 Local material collection

The required local materials for the bridge construction are sand, gravel (river gravel or broken aggregate),
and stones/boulders.

Stone/Boulders: The best stone collection is from the rock quarry. The rock should be unweathered, hard
and dense with metallic sound.

In unavoidable case, boulders from the river deposits can also be collected. However, this can be used
only for filling purpose (broken stone filling). In any case, stones from rock quarry are necessary for
masonry works.

Sand: Sand can be collected from river deposits or from a quarry. The quality of the sand should be
assessed before sand collection. Check visually the content of the impurities such as mica, clay, loam,
mud organic materials etc. If such impurities are unavoidable, it is recommended to wash the sand before
use. Sand containing significant quantity of mica should be rejected. The grain size of the collected sand
should not be too fine.

Fill a bottle with sand and water and shake vigorously and leave to settle. If the sand is clean the
sedimentation will be less than 5mm after two hours. And the water above will only be lightly cloudy.

Gravel: Gravel can be collected from river deposits or by breaking


boulders into the necessary size. The required sizes and
their proportion should be
5 to 20mm - 40%
20 to 40mm - 60%

Gravel should be of hard rock origin. Gravel of unsuitable rock such as mica, marl and sand stone should
be rejected. Likewise flat and flaky particles should also be rejected. The collected gravel should be free
from organic contaminants like clay, loam, mud or stone dust etc.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 87

2.7.4 Transportation and storage of the materials

Material other than local materials has to be transported from road head to the site by porter or other
means. These materials are mainly Cement, Steel Parts and Wire Ropes.

Cement Transportation and Storing: Utmost care should be given for transportation and storing of the
cement. The prime importance is the proper packing of the cement before transportation to make it
watertight and airtight. For this, cement bags as received from the market or factory should be double
packed by additional packing with Nylon Bags and plastic layer inside. Re-opening the bags (especially
when transporting by mules) is not permitted before use at the site.

The conditions must be met for the storing of the cement:


¡¡ Cement must always be stored under a roof with adequate protection from rain. A raised plank
floor is necessary to prevent cement from damp.
¡¡ Storage must be arranged in such a way, that the oldest cement can be used first.

Steel Parts Transportation and Storing: There is a great chance of damage of steel parts during loading/
unloading and transportation. The most common damage is:
¡¡ Deformation of cross beams and steel decks due to mishandling during loading and unloading,
¡¡ Deformation of suspenders and reinforcement bars due to mishandling during loading and
unloading

The steel parts should be loaded or unloaded carefully to avoid above damages. Do not allow steel parts
to fall from a height. Suspenders should be bound together with the crossbeam.

Similarly, the conditions must be met for the storing of steel parts to avoid any damage.

¡¡ Galvanized and non-galvanized steel parts must always be stored under a roof with adequate
protection from rain and should not be in contact with the ground.
¡¡ Galvanized steel parts should not be transported and stored together with salt or acid.
¡¡ Steel parts should be stacked and stored element/component-wise separately, avoiding the mix
up of different elements. Thus, any element or component can be easily located during the bridge
erection.
¡¡ All fixtures (nut/bolts, washers, thimbles and bulldog grips) should be packed/marked and stored
separately according to its sizes.
¡¡ Steel parts, particularly suspenders and reinforcement bars, should not be permitted to bend
during portering and storage.

Wire Rope Transportation and Storing: It is vital to handle and transport the cable carefully to avoid
any defects like kinks, splices and broken strands. Consider some examples of defects on cables due to
mishandling and improper transportation.

Also pulling or dragging the cable along the road for transportation is not permitted.

To avoid such defects, follow handling and transportation methods.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 88

¡¡ Method of Unreeling Light Cables with the Help of a Reel Support

Wrong Correct Correct

¡¡ Method of Unreeling Cables by Unrolling Each Loop Taken from the Reel.

Before cable cutting, the cable ends should be tightened by a binding-wire (seizing) to avoid loosening
of the cable wires.

¡¡ Method of Transportation by Porters

Such kinds of wrong cable transportation should not be practiced.

Wrong method of transportation

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 89

2.7.5 Masonry and Stone dressing work

Requirements for Building Stones: Building stones must be of high strength, density and durability. A
good building stone should be hard, tough, compact grained and uniform in texture and color.

Crystalline stones are superior to non-crystalline stones. Metamorphic rocks are more durable than
sedimentary rocks. Sedimentary rocks have been formed by water sediments of clay, sand or gravel,
which got cemented together by lime, silica etc. Originally metamorphic rocks are either of volcanic or
sedimentary origin but have subsequently been formed and shaped by movements of the earths’ crust
imposing high pressure and heat.

A good building stone absorbs no or very little water and must be free from decay, cracks and sand-holes.

Quarrying: Rocks for stone masonry works should be broken from a quarry by crowbars and wedges.
Natural fractures and bedding planes of stratification are the weak features of rock. These natural joints
are taken as advantage to break and separate one block from the other.

It is advised that only when natural joints do not exist those artificial fissures be made by drilling a line of
holes in rows along the desired breaking line. By inserting conical wedges and driving them in succession
with a hammer the rock will crack along the face of the holes.

However it is generally worthwhile to search for quarries with existing natural joints like dominant bedding
planes, since the broken stones are much easier for dressing.

Boulders fallen from rocky slopes can also be used as building stones.

Building stones are first dressed to obtain two parallel planes, and then outward faces must be dressed
well with the help of the square bottom.
outward face

Stones from the riverbed are generally very hard and durable and can be used as filling stone, but are not
suitable for stone dressing.

Stone Dressing: Broken stones from the quarry are to be dressed by hammer or chisel to required sizes.
Depending on the function of the stone in stone masonry construction the types of stones have to be
prepared:
¡¡ Corner Stone: The corner stone is placed at each corner of the stone masonry structure.
Recommended sizes are:
Length : 30 - 85 cm
Width : min. 30 cm
Height : min. 10 cm

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 90

¡¡ Face Stone: Only one face of this stone faces outside; recommended sizes are:
Length : 30 - 75 cm
Width : min. 30 cm
Height : min. 10 cm
¡¡ Bond Stone: Like the face stone only one face is outside, but the bond stone extends to the
interiors of the structure. Bond stones, also called through stones, go right
through walls of up to 85 cm thick or more. Recommended sizes are:
Length : 45 - 85 cm
Width : min. 30 cm (face side)
Height : min. 10 cm
¡¡ Coping Stone: Copingstones are put on top or at steps of stone masonry structures. They
should be larger and heavier than the stones:
Length : as large as possible
Width : as large as possible
Height : min. 10 cm
¡¡ Filling Stone: Filling stones do not need to be dressed and are placed in the inner part of the
stone masonry structure mainly to gain gravity load.

Stone Masonry Laying: There are many different kinds and types of stone masonries. For constructing
anchor blocks and towers, only coursed (in layers of equal height) stone masonry is applied.

There are two types of stone masonry used for bridge construction:
¡¡ Coursed Random Rubble Stone Masonry
The stones are hammer-dressed, except the inside face. Gaps between beds and joints shall not
exceed 12 mm. All Face Stones tail into the wall twice their height.

Bond Stones running right through the wall are inserted at least at every 150 cm intervals.
¡¡ Coursed Block Stone Masonry
The stones are chisel-dressed at all faces, except the inside face. Joints are dressed at right angles
to the face. Gaps between beds and joints should not exceed 6 mm.

All Face Stones tail into the wall twice their height.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 91

Bond Stones running right through the wall are inserted in each course at least at every 150cm intervals.

Course Stone Masonry must be made in layers of equal height. Individual layer heights may vary but
should never be less than 10 cm. Alternate joints shall be made between the layer above and below.

In a reasonably well made stone masonry the inner friction between the beds amounts to approx. 35°.

The verification of corners as well as faces has to be checked carefully with the plumb-bob.

Correct use of plumb-bob

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 92

The Strength of stone masonry structures depends mainly on the qualities.

…bigger… The Strength of Stone Masonry is… …smaller…

… with
rectangular
Form or … with irregular
stones. Shape stones.

… the more
… the less stones
are used. Number stones are
used.

… the rougher the Roughness … the smoother


joints are. of joints the joints are.

… the bigger
… the smaller the
beds are. Bed the
beds are.

… the more … the slimmer


compact the
Height & the stones
stones are. Width are.

Bond Across … the worse


…the better the the
bond across is. (in plan bond across
view) is.

… the higher … the lower the


strength of the
Strength strength of the
mortar is. of Mortar mortar is.

2.7.6 Cement works

(A). Composition and Mixtures


Cement concrete is a mixture of 4 components:
¡¡ Cement: Ordinary Portland Cement commonly used for general construction works;
¡¡ Sand;
¡¡ Gravel;
¡¡ Water.

Cement is very sensitive to humidity and moisture; therefore it should never be stored for a long time. In
the rainy season cement bags have to be packed in additional sealed plastic bags plus additional nylon
bags for protecting the cement against water and the plastic bags against damage.

Sand should be clean, sharp, angular, hard and durable. Sand must be well washed and cleaned from
mud or any organic material before use. Well-graded sand should be used for cement works. All or most
of the sand should pass through a 3 mm sieve or mesh wire. However sand should not be too fine, only
max. 15% of the sand can be smaller than 150 microns, which is like dust.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 93

Gravel should be clean, hard, angular and non-porous. Usually riverside gravel makes the best aggregate
for preparing concrete. The corn size of gravel should be smaller than 40 mm (1½ inches) but bigger than
5mm.

Water from rivers or lakes is usually suitable for making cement mixtures. Do not use water from ponds
or swamps; this water may contain a lot of organic materials.

The main characteristics of any cement work are given by the mix proportions of their components:
¡¡ Cement Mortar = Mix between Cement & Sand
¡¡ Cement Concrete = Mix between Cement, Sand & Gravel

Of course, Water is added in both cases, but the mix proportions of cement, sand and gravel give the
main characteristics of any cemented work.

Mixing above components thoroughly is of utmost importance. Hand mixing should be done on a clean
watertight platform. Cement and Sand should first be mixed dry, and then gravel added. Now the whole
mixture should be turned over 3 times dry. Then mixing should take place for at least 5 minutes by slowly
sprinkling water until the concrete is of a uniform color. The table depicts the most commonly used mix
proportions and required quantities:

Table F.2.11: Quantities for various Types of Cement Works

Mix. proportions Dry required quantities for one cubic meter wet:
Type of
Cement Stones or
Cement bags Sand Gravel
Work Cement : Sand : Gravel kg Boulders
@ 50 kg [m3] [m3]
[m3]
1 : 1 – 20.4 1020 0.71 – –
1 : 2 – 13.6 680 0.95 – –
Cement
1 : 3 – 10.2 510 1.05 – –
Mortars
1 : 4 – 7.6 380 1.05 – –
1 : 6 – 5.0 250 1.05 – –
Cement
Plaster
1 : 4 – 0.18 9 0.024 – –
(20 mm
1 : 6 – 0.12 6 0.024 – –
includes
12% waste)
uncoursed
1 : 4 2.66 133 0.37 – 1.2
stone
Cement 1 : 6 1.75 87.5 0.37 – 1.2
masonery
Stone
Masonries uncoursed
1 : 4 2.28 114 0.32 – 1.25
stone
1 : 6 1.50 75 0.32 – 1.25
masonery

Cement 1 : 4 : 8 3.4 170 0.47 0.94 –


Concretes 1 : 3 : 6 (M10) 4.4 220 0.46 0.92 –
(plain or 1 : 2 : 4 (M15) 6.4 320 0.45 0.90 –
reinforced) 1 : 1½ : 3 (M20) 8 400 0.42 0.84 –

"Plum" 1 : 3 : 6
2.64 132 0.28 0.54 0.50
Concrete with 50% boulders

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 94

The amount of Water should be about 50% or half the volume of cement. One 50 kg bag of cement has
a volume of approx. 35 liters, which is equal to approx two kerosene tins.

Concrete and Mortars should be placed in its final position within one hour! After placing it should
be well compacted by rods in order to remove any air pockets. For a concrete of high quality good
compaction is essential. This may mean extra work during placing, but on no account should more water
be added for reducing compacting work. Concreting should never be done if it is raining.

Curing means keeping completed cement works wet until its setting process is completed. If concrete
works are not continuously kept wet during its setting process, cement mortars, cement stone masonry
work and especially concrete does not develop its full strength. Curing should be done for at least 28
days.

For increasing the strength of concrete, ripped Tor-Steel bars are added which makes Reinforced Cement
Concrete or RCC.

(B). Concrete Work for Cement Stone Masonry (CSM) Towers


The CSM Towers or Limb Walls are concreted together to form one solid unit (See Drawing Nos. 20Dcon70
& 20Dcon106). The core and the connection of both the towers are made in R.C.C 1:2:4, whereas the
limb walls are made in CSM 1:4.

Section through CSM Tower (Bridge Entrance)

¡¡ Placing the Saddles for the Walkway Cables


The saddles for the walkway cables are to be placed in between the towers. The position of the saddles
has to be checked thoroughly and the levels can be controlled with the help of a transparent plastic pipe
filled with water (Level Pipe).

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 95

Concreting Work for CSM Tower

Checking Level of Walkway Cable Saddles

¡¡ Construction of Towers and Placing Saddles for Handrail Cables


The Towers or Limb walls support the handrail cables. The limb walls are made out of cement stone
masonry 1:4 with a R.C.C. core.

The handrail cable saddles are to be placed on top of the “hump” of the limb wall. Make sure that the
position and shape of the “hump” is correct so that the handrail cable touches the saddle plate only.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 96

(C). Constructing the Deadman Beam


The Deadman Beam is a soil anchor cast in reinforced cement concrete R.C.C that lies buried under
the gravity structure. The handrail and walkway cables are placed around the reinforcement bars before
concreting the beam. At one bank the cables are inserted into a polyethylene (PE) pipe, so that the
cables can still be moved while sag setting. Tensioning cable must be tightened with the bridge of
bulldog grip. Use leftover plastic or cloth from cement bags to cover open parts of the pipe so that no
concrete can flow into the pipe.

Nos & spacing of Bulldog Grips


Cable dia mm Nos G
13 5 10 cm
26 6 15 cm
32 3 20 cm

Place and Fix Reinforcement Bars, Stirrups and Erection Hooks.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 97

The Deadman Beam is casted in concrete 1:2:4

The fixation cable can be anchored at the temporary erection hook or fixed at one of the walkway cables.

(D). Constructing Drum Anchorages in Rock


There are two types of Drum Anchorages in Rock:
¡¡ R.C.C Drum Anchor in Hard Rock;
¡¡ R.C.C Drum Anchor in Soft or highly fractured Rock.

Drum Anchorages in Hard Rock are made by drilling holes of 32 mm diameter ( = diameter of crowbar)
into the rock. Clean boreholes from dust and debris by flushing them with water.

Fill the holes with cement mortar 1:1 before the anchor rods are inserted. The formwork for the drum
is made by a chitra (bamboo mat) or plain G.I. sheet inside lined with a plastic sheet. Use binding wire
around the chitra to prevent the bamboo mat bulging during concreting.

Drum Anchorages in soft or fractured rock are not done by drilling holes but by excavating a round pit
instead. The Anchor Reinforcement has to be placed into the pit and is fixed with the help of stirrups. The
excavated pit is then filled and well compacted with concrete 1 : 2 : 4 up to ground level. At this stage
the anchor rods should protrude (stand out) by approx. 40 cm. The formwork for the drum is made by a
chitra (bamboo mat) inside lined with a plastic sheet bound together with a binding wire.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 98

2.7.7 Cable hoisting and sag setting

Cables are hoisted and the prescribed sag set after the Deadman Beams or the Drum Anchors have been
concreted. Please note that it takes 4 weeks until cast concrete develops its full strength. Therefore, the
final cable pulling is done after a minimum of 4 weeks.

(A). Calculation of Hoisting Sag


Before starting any hoisting work, the actual span l from saddle to saddle of the bridge and the actual
difference of elevation h between the walkway cable saddles have to be measured first and filled into
Design Form: Calculation of Hoisting Sag .

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 99
Short Span Trail Bridge Standard

Table F.2.12: Calculating Elevation of Low Point for Cable Hoisting


Table 13.2 Table for Calculating Elevation of Low Point for Cable Hoisting
1. Actual Span measured in the Field l = …………… m

2. Saddle Elevation of the Walkway Cable on the higher side Eh = …………… m

3. Saddle Elevation of the Walkway Cable on the lower side El = …………… m

4. Difference in Elevation h = Eh - El = h = …………… m


5. Dead Load Sag


For Span up to 80 meters: bd = = bd = …………… m
20


For Span over 80 meters: bd = = bd = …………… m
22

6. Hoisting Sag bh = 0.95 x bd = bh = …………… m

(4 ⋅ b h − h) 2
7. f min in hoisting case fmin = = fmin = …………… m
16 ⋅ b h

8. f max in hoisting case fmax = f min + h = fmax = …………… m

9. Elevation of Cable low point in hoisting case = El - f min = _________ m

¡•¡ Mark
Markthe
thecalculated
calculatedelevation
elevationofofthe
thecable
cablehoisting
hoistingsag
sag(low
(lowpoint)
point)on
onaaprepared
preparedstick,
stick,tree
treeororatatthe
tower foundation.
the tower foundation.
¡•¡ Nowset
Now setupupthe
theAbney
AbneyLevel
LevelororLeveling
LevelingInstrument
Instrumentatatthe
theElevation
Elevationofofthe
thecable
cablehoisting
hoistingsag
sag
soso
that
that the line of sight can easily see the mark and the low point of the cable. Setting up the
the line of sight can easily see the mark and the low point of the cable. Setting up the Leveling Leveling
Instrument at the calculated Elevation has to be done by trial and error and may take several
Instrument at the calculated Elevation has to be done by trial and error and may take several attempts.
attempts.

(B).
(B). Cable Hosting
Cable Hosting
Cables are first pulled by hand and for final sag setting with the help of the cable pulling machine or tirfor,
Cables
whichare first at
is fixed pulled by handhook.
the erection and for final sag setting with the help of the cable pulling machine or tirfor,
which is fixed at the erection hook.
¡¡ Pull the cable until it reaches a level of about 20 cm higher than the calculated Elevation. Each
cable should be left in this “over-pulled” position for at least 12 hours. “Over-pulling” is done to
• Pull the cable until it reaches a level of about 20 cm higher than the calculated Elevation. Each cable
prevent any later relaxation of the cable, which may lead to a tilted walkway.
¡¡
should be left in this "over-pulled" position for at least 12 hours. "Over-pulling" is done to prevent
For actual and precise sag setting first firmly clamp the special cable belonging to the tirfor machine
any laterbackstay
to the relaxation of theofcable,
portion whichtomay
the cable lead to aThen
be pulled. tiltedfixwalkway.
the tirfor machine at the erection hook
• For
andactual
insertand
theprecise
special sag setting
cable firstthe
through firmly clamp themachine.
cable-pulling special cableNowbelonging
apply forceto until
the tirfor machine
the special
tocable
the backstay
is firmly portion of the cable
under tension. Nowtoloosen, do not
be pulled. Thenremove,
fix the the
tirfor machine
bulldog at the
grips. erection
The hook and
cable should
now be
insert the held by the
special tirfor
cable machine
through the only. Slowly release
cable-pulling machine.some force
Now by carefully
apply moving
force until the lever
the special of is
cable
the cable-pulling machine until the desired pre-calculated Elevation has been
firmly under tension. Now loosen, do not remove, the bulldog grips. The cable should now be held reached. When this
byis the
the tirfor
case immediately
machine only. retighten
Slowlythe bulldog
release grips,
some then
force bycompletely release the
carefully moving the lever
tensionof applied
the cable-
by the tirfor machine.
pulling machine until the desired pre-calculated Elevation has been reached. When this is the case
immediately retighten the bulldog grips, then completely release the tension applied by the tirfor
The cable should now hang in proper hoisting position. If the low point has gone below the hoisting
machine.
Elevation the whole process has to be repeated.

The cable should now hang in proper hoisting position. If the low point has gone below the hoisting
Elevation the whole process has to be repeated.

110 PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 100

That means the cable has again to be over-pulled and then slowly released.

Check also that parallel cables have equal hoisting sag.

2.7.8 Finalizing the cable anchorage

After the cables have been pulled and the hoisting sag is firmly set the Cable Anchorage has to be fully
completed before any fitting works for the walkway can start.

(A). Rust Protection for the Cable


To achieve optimal rust protection, paint the cables in the gravity structure with coal tar and then cover
with 20 x 20 cm cement concrete 1:3:6. Before painting, the bulldog grips need to be checked and
retightened if required.

(B). Completing the Gravity Structure


The actual gravity structure on top of the Dead Man Beam or Drum Anchors is constructed according
to the respective construction drawing. The side and back walls as well as the top are made of coursed
cement stone masonry 1:6, whereas the inside is filled with broken stones. The cement stone masonry
work for the walls has to be made with hammer dressed stones of equal layer height. The broken stones
for filling the inside should not be thrown but laid and interlocked as for as possible.

Only after the gravity structure is completed can fitting work for the walkway structure start.

2.7.9 Walkway fitting

The fitting work for the walkway must only start after the gravity structure of the cable anchorage has
been completed. Walkway fitting is simple and self-explanatory. Refer also to the construction Drawings
No. 19Dcon70 or 19Dcon106 respectively.

Points must be observed:


¡¡ Start fitting from one bank only
¡¡ First fit crossbeam, steel panels or wooden planks as close as possible to the bridge entrance.
¡¡ Always start fitting walkway deck with a “Half Panel” then continue with “Standard Panels” only.
¡¡ Fix J-bolts at crossbeams first loosely and hang pre-bent suspender over handrail cable.
¡¡ Avoid accidents by bolting panels loosely immediately after placing.
¡¡ Maintain equal vertical distance between handrail and walkway cable by using a support guide
(“Tokche”) made of wood or bamboo.
¡¡ Always finish walkway fitting with “Special Panels” and cut off extra length by hacksaw.
¡¡ Check and retighten all Nuts and Bolts after completion of walkway fitting.
¡¡ If wood is used for the bridge deck, the planks should be 2 meter long and min. 4 cm thick and
should be fitted in staggered way. Use washers below Bolt Heads. Distance between Crossbeams
is 1 meter.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 101

Fencing is woven on the spot with gabion wire (12SWG) between the handrail cable and the fixation
cable. First fix the fixation cable by pulling it through the bottom eye of the suspender along either side
of the walkway, and then join it with the short piece at the other end of the bridge.

…for Fence weaving

2.7.10 Water management backfilling and general finishing works

(A). Water Management


The life expectancy of the bridge largely depends on proper water management.

Any water seepage encountered during excavation should be intercepted as close as possible to its
origin and channeled safely to a nearby watercourse. Especially vulnerable is the place behind the
Deadman Beam! If in doubt, or in case of unusual humidity of water seepage, provide a drain behind
the Deadman Beam with side outlet. Sometimes water seepage occurs during rainy season only. Inquire
with the local people.

Divert surface water and provide drainage channels as necessary. Do not hamper existing irrigation
channels, rather improve and adjust them with some cement works. Discuss solutions with local people
and decide on the spot.

As a general rule divert water as far away from bridge foundations as possible.

For managing surface water well, also fill the gaps around completed anchor blocks well above the
existing surface. Back filling prevents surface water to flush out excavations.

Do Back Fill !!!


(B). Finishing Work
Provide finishing structures like retaining walls, staircases, small trail improvements, adjustments to nearby
houses etc., if it adds functional value to the bridge.

Never do cement pointing or other non-functional works.

Also check the vegetation and plant life in the vicinity of the bridge. Plant some new trees if possible,
especially if some had to be cut for bridge construction.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 102

Suspended Trail Bridge Design Form

1. Cable Design

2. Anchor Block Design

3. Bridge Standard Drawings

Bridge Name:.......................................................................................

River Name:.........................................................................................

Woreda Name:.....................................................................................

Designed by:........................................................................................

Date:....................................................................................................

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 103
Short Span Trail Bridge Standard

Bridge No:………… Bridge Name :…………… Woreda:……………………..Span: ………. m,

Walkway Width (70 or 106): ………………cm

1. Cable Design for Suspended Bridge Type

A. Survey Data & Calculation of Freeboard

Span 

 /2  /2
Eh
h

E
sa
gb
Walkway cable
f
min

HFL

Fb
1. Span of the Bridge  = ................... m
2. Saddle Elevation of the walkway cable on the higher side Eh = ................... m
3. Saddle Elevation of the walkway cable on the lower side E = ................... m
4. Difference in Elevation h = E h - E = h = ................... m
(max. permissible height: hmax = /25)

5. Dead Load Sag (fixed ratio) 


bd = = bd = ................... m
(in the middle of the bridge) 22

6. f min in Dead Load Case (4 ⋅ b d - h) 2


fmin = = fmin = ................... m
(at the lowest point of the cable) 16 ⋅ b d
7. Highest Flood Level HFl = ................... m

8. Free Board (min. 5.00m) Fb = E - HFl - f min = Fb = ................... m

( if freeboard is less than 5.00m, try either to raise the saddle elevations or to adjust
the span, but keep the ratio between span and sag always fixed,  / bd =22 )

114

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 104

B. Selection of Cables
Select a cable combination according to the span and walkway width of the bridge. Always select the
higher cable combination, when the span is in between two values.

Maximum Span for Weight of all


Cable Combinations
Walkway Width: Cables
70cm 106cm Handrail Cable Walkway Cable gh
span [m] span [m] nos Ø mm nos Ø mm [kg/m]
50 40 2 26 2 26 10.04
90 60 2 26 2 26 2.62
100 75 2 26 2 26 15.06
120 105 2 26 2 26 20.22
---- 120 2 32 2 32 22.80

Above cable combinations are calculated for the specifications:


Cables: construction 7 x 19, wire strand core, 160 kg/mm2 tensile strength of wire
Safety Factor: minimum 3 or higher than 3
Live load p = (300+5000/span) kg/m2 according to Standard of the Suspension Bridge Division
or p = 400 kg/m2 if the span is 50.0 m or less
Sag to Span Ratio: = 22 in dead load case; this is a fixed ratio
Max. Permissible Height Difference of Saddles: h =span/25

Example:
width of walkway = 70cm; span = 88m
Selected cable combination:
Handrail Cables 2 Ø 26mm
Walkway Cables 2 Ø 32mm
Weight of Cables = 12.62kg/m

Selected Cable combination and parameters from the table above:

HRC Handrail Cables: nos 2 Ø .......mm


WWC Walkway Cables: nos….. Ø .......mm
Weight of all Cables per meter gh …....kg/m

C. Calculation of Cable Length

Backstay Length * Cutting Length**


Type of Cable dia (mm) Nos
[m] [m/pc]
Fixation Cable 13 2 ………. ……….
Handrail Cable ….. 2 ………… …………
Walkway Cable …….. - .…..……. .…..…….
* Backstay Length = Cable length between saddle center and center of dead man or drum as per foundation drawing
(both banks) + 6.0m.
** Cutting Length = 1.1 x Span + Backstay Lengths

PART F: TRAIL BRIDGE MANUAL


Walkway Cable …….. - .…..……. …………
F - Chapter 2 - 105
*Backstay Length = Cable length between saddle center and center of dead man or drum as per
foundation drawing (both banks) + 6.0m.
**Cutting
D. Length = 1.1
Calculation x Span&+fmax)
of (fmin Backstay Lengths
Hoisting Sag
D. Calculation of ( fmin & fmax )Hoisting Sag
This calculation has to be made after tower and foundation work is completed
This calculation has to be made after tower and foundation work is completed

1. Actual Span measured in the Field  = …………… m

2. Saddle Elevation of the walkway cable on the higher side Eh = …………… m

3. Saddle Elevation of the walkway cable on the lower side E = …………… m

4. Difference in Elevation h = Eh - E = h = …………… m



5. Dead load sag bd = = bd = …………… m
22

6. Hoisting Sag bh = 0.95 x bd = bh = …………… m

(4 ⋅ b h − h) 2
7. f min in hoisting case fmin = = fmin = …………… m
16 ⋅ b h

8. f max in hoisting case fmax = f min + h = fmax = …………… m

9. Elevation of Cable low point in hoisting case = E - f min = m

Span  E
h
h

E

f
max
f

walkway cable
min

Mark Low point


Elevation of
of cable cable hoisting sag

116

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 106

Anchorage Block Design (Foundation) for Suspended Bridge Type

A. Design Data
Fill in the design data

ƒƒ Walkway Width, WW (70 or 106cm):............................cm


ƒƒ Bridge Span:............................m
ƒƒ Bank Condidtion:

Right Bank
Geology: Soil or Rock

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 10° slope) (more than 10° slope)

What is the Soil Type? Gravelly Sandy Silty

If rock, what is the rock type? Hard Rock Hard Rock Soft Rock
(data from Survey Form and (only few fractures) (highly fractures)
Check list)
Tower Height from Ground up to
H.C. Saddle (data from bridge 2.4m 3.4m 4.4m
profile):
Left Bank
Geology: Soil or Rock

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 10° slope) (more than 10° slope)

What is the Soil Type? Gravelly Sandy Silty

If rock, what is the rock type? Hard Rock Hard Rock Soft Rock
(data from Survey Form and (only few fractures) (highly fractures)
Check list)
Tower Height from Ground up to
H.C. Saddle (data from bridge 2.4m 3.4m 4.4m
profile):

B. Selection of anchorage blocks


Select appropriate anchorage blocks at Right and Left Bank according to the above design data.
Procedure for selection:

¡¡ According to the Soil/Rock type and Slope of ground, refer to respective table for selection of
Anchorage Block:
for Soil and Flat Ground : Table F.2.13
for Soil and Hill Slope : Table F.2.14
for Hard Rock : Table F.2.15 or Table F.2.16
for Fractured Hard Rock or Soft Rock,
Span up to 90m (WW = 70cm) and upto 60m (WW = 106cm) : Table F.2.17 or Table F.2.18
Span Range 91-120m (WW = 70cm), 61-120m (WW = 106cm) : Table F.2.19

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 107

¡¡ In the table match the design data: selected Walkway Width  Bridge Span  Tower Height 
Soil type  select the corresponding Block Type and Drawing No. for right bank and for left bank
respectively.

Selection Tables

In Soil and Flat Ground:

Table F.2.13: Selection of Gravity Soil Anchor Block in Flat Ground

Span Range, m Tower


Foundation
Height Block Type DrawingNo.
Walkway:70cm Walkway:106cm Soil Type
[m]
2.4 1F 21Dcon
Up to 45m Up to 30m 3.4 All 2F 22Dcon
4.4 3F 23Dcon
2.4 4F 24Dcon
46 - 90 31 - 60 3.4 All 5F 25Dcon
4.4 6F 26Dcon
2.4 7F 27Dcon
91 - 120 61 - 75 3.4 All 8F 28Dcon
4.4 9F 29Dcon
2.4 10F 30Dcon
– 76 - 90 3.4 All 8F 28Dcon
4.4 11F 31Dcon
2.4 12F 32Dcon
– 91 - 105 3.4 All 8F 28Dcon
4.4 13F 33Dcon
Gravely 12F 32Dcon
2.4
Sandy, Silty 14F 34Dcon
– 106 - 120
3.4 15F 35Dcon
All
4.4 13F 33Dcon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 108

In Soil and Hill Slope:

Table F.2.14: Selection of Gravity Soil Anchor Block in Hill Slope

Span Range, m Tower


Foundation
Height Block Type DrawingNo.
Walkway:70cm Walkway:106cm Soil Type
[m]
Up to 60m Up to 40m 2.4 All 1S 41Dcon
61 - 90 41 - 60 2.4 All 2S 42Dcon
91 - 120 61 - 75 2.4 All 3S 43Dcon
Gravely 4S 44Dcon
– 76 - 90 2.4 Sandy 5S 45Dcon
Silty 6S 46Dcon
Gravely, Sandy 7S 47Dcon
– 91 - 105 2.4
Silty 8S 48Dcon
Gravely, Sandy 8S 48Dcon
– 106 - 120 2.4
Silty 9S 49Dcon

In Hard Rock for all Span Ranges:

Table F.2.15: Selection of RCC Single Drum Anchor in Hard Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.4 1HRS 61Dcon


91 – 120 61 - 120 2.4 2HRS 62Dcon

When slope is too steep and there is not enough space for single drum anchorage system (Table F.2.15),
select the double drum system from Table F.2.16.

Table F.2.16: Selection of RCC Double Drum Anchor in Hard Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.4 1HRS 63Dcon


91 – 120 61 - 120 2.4 2HRS 64Dcon

In Fractured Hard Rock/Soft Rock for Span Range up to 90m (WW = 70 cm) and 60m (WW = 106cm):

Table F.2.17: Selection of RCC Single Drum Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.4 1FRS 65Dcon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 109

When slope is too steep and there is not enough space for single drum anchorage system (Table F.2.17),
select the double drum system from Table F.2.18.

Table F.2.18: Selection of RCC Double Drum Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

up to 90 up to 60 2.4 1FRD 66Dcon

In Fractured Hard Rock/Soft Rock for Span Rang of 91- 120m (WW = 70 cm) and 61-120m (WW = 106cm):

Table F.2.19: Selection of RCC Deadman Anchor in Fractured Hard Rock/Soft Rock

Span Range, m Tower Height


Block Type Drawing No
Walkway: 70cm Walkway: 106cm [m]

91-120 61-120 2.4 2FRD 67Dcon


Selected Anchorage Foundation type and corresponding drawings from the table above:

Right Bank: Block Type………………..., Drawing No………………………


Left Bank: Block Type………………..., Drawing No……………………….
If rock, what is the rock type?

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 110

ƒƒ Walkway Width, WW (70 or 106cm): 70 cm


ƒƒ Bridge Span: 88m
ƒƒ Bank Condidtion:

Right Bank

Geology: Soil or Rock

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 10° slope) (more than 10° slope)

What is the Soil Type? Gravelly Sandy Silty

If rock, what is the rock type? Hard Rock Hard Rock Soft Rock
(data from Survey Form and (only few (highly
Check list) fractures) fractures)

Tower Height from Ground


up to H.C. Saddle (data from 2.4m 3.4m 4.4m
bridge profile):

Left Bank

Geology: Soil or Rock

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 10° slope) (more than 10° slope)

What is the Soil Type? Gravelly Sandy Silty

If rock, what is the rock type? Hard Rock Hard Rock Soft Rock
(data from Survey Form and (only few (highly fractures)
Check list) fractures)

Tower Height from Ground


up to H.C. Saddle (data from 2.4m 3.4m 4.4m
bridge profile):

==> Selected Anchorage Blocks


Right Bank: Block Type 5F, Drawing No. 25Dcon
Left Bank: Block Type 1HRS, Drawing No. 61Dcon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 111

Bridge Standard Drawings


Select the required Steel Drawings and Construction Drawings from the drawing list.

2.8 Steel Drawings

Drawing Title Drawing No Required Drawing


02D or 02D4 or
Walkway Cross Beams ………….
03D or 03D4
RB:
Saddle and Reinforcement for RCC Deadman ………….
20D2 or 20D4
and Gravity Soil Anchor LB:
………….
RB:
Saddle and Reinforcement for RCC Deadman ………….
20D4S
Anchor in Soft or Fractured Hard Rock LB:
………….
RB:
Saddle and Reinforcement for Drum Rock ………….
60D2 or 60D4
Anchor LB:
………….
Steel Deck 11A, 12A and 13A 11A, 12A and 13A

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 112

2.9 Construction Drawings

Required
Drawing Title Drawing No
Drawing
Walkway Fitting 19Dcon70 or 19Dcon106 ………….
Details of Cement Stone Masonry Tower & RCC
20Dcon70 or 20Dcon106 ………….
Core
RB:
RCC Deadman and Gravity Soil Anchor Block ………….
21Dcon……….35 Dcon
for Flat Ground LB:
………….
RB:
RCC Deadman and Gravity Soil Anchor Block ………….
41Dcon……….49Dcon
for Hill Slope LB:
………….
RB:
………….
RCC Single Drum Rock Anchor in Hard Rock 61Dcon or 62Dcon
LB:
………….
RB:
………….
RCC Double Drum Rock Anchor in Hard Rock 63Dcon or 64Dcon
LB:
………….
RB:
RCC Single Drum Rock Anchor in Soft or ………….
65Dcon
Fractured Hard Rock LB:
………….
RB:
RCC Double Drum Rock Anchor in Soft or ………….
66Dcon
Fractured Hard Rock LB:
………….
RB:
………….
RCC Deadman in Soft or Fractured Hard Rock 67Dcon
LB:
………….

Designed by: ……………………………………... Date: ………………..

Cable hoisted by: ………………………………… Date: ………………..

Example:

==> Selected Drawings

Steel Drawings

PART F: TRAIL BRIDGE MANUAL


F - Chapter 2 - 113

Drawing Title Drawing No Required Drawing


Walkway Cross Beams 02D or 02D4 or 03D, or 03D4 02D
Saddle and Reinforcement
for RCC Deadman and 20D2 or 20D4 RB: 20D2
Gravity Soil Anchor
LB: x
Saddle and Reinforcement
for RCC Deadman Anchor
20D4S RB: x
in Soft or Fractured Hard
Rock

Saddle and Reinforcement LB: 60D2


60D2 or 60D4
for Drum Rock Anchor RB: 60D2
LB: x
Steel Deck 11A, 12A, 13A 11A, 12A, 13A

Construction Drawings

Drawing Title Drawing No Required Drawing


Walkway Fitting 19Dcon70 or 19Dcon106 19Dcon70
Details of Cement Stone Masonry
20Dcon70 or 20Dcon106 20Dcon70
Tower & RCC Core

RCC Deadman and Gravity Soil RB: 25Dcon


21Dcon……….35 Dcon
Anchor Block for Flat Ground LB: x

RCC Deadman and Gravity Soil RB: x


41Dcon……….49Dcon
Anchor Block for Hill Slope LB: x

RCC Single Drum Rock Anchor in RB: x


61Dcon or 62Dcon
Hard Rock LB: 61Dcon

RCC Double Drum Rock Anchor in RB: x


63Dcon or 64Dcon
Hard Rock LB: x

RCC Single Drum Rock Anchor in RB: x


65Dcon
Soft or Fractured Hard Rock LB: x

RCC Double Drum Rock Anchor in RB: x


66Dcon
Soft or Fractured Hard Rock LB: x

RCC Deadman in in Soft or RB: x


67Dcon
Fractured Hard Rock LB: x

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 114

3. SUSPENSION TRAIL BRIDGE

3.1 Background
This Chapter follows the actual step-by-step process of constructing a suspension (N-type) bridge. It
deals with all the necessary information, drawings and formats for the construction of a suspension type
cable bridge.

Figure: F.3.1: Suspension type bridge, Endamayno, Tigray Region

3.2 Presentation of the suspension bridge type


This type of bridge has an upward cambered walkway. The main cables are hung over the towers and
anchored to the main anchorage foundations. The walkway cables (Spanning cables) are anchored to the
tower foundations. The main components of this bridge are: Main cables and Spanning cables, Towers,
Walkway system, Main anchorage foundations and Walkway/Tower foundations.

This type of bridge is selected when the bridge site is comparatively flat terrain and the Suspended type
bridge is not feasible due to insufficient free board. The Suspension type bridge is more expensive and
needs more inputs to design and construct than the Suspended type bridge.

The major structural elements of the suspension bridge presented in this Manual are steel wire ropes of
Ø 13, 26 and/or 32mm which are anchored by gravity soil or rock anchors at either side of the river. The
superstructure is completely unstiffened and thus allows “some” reasonable degree of lateral, vertical
and torsional vibrations.

For economic reasons steel has been used as little as possible, and cement has been applied for
guarantying the durability of the structures. There are only four towers (pylons) the user can choose
from. They are supported by hinged bearings at the bottom. Turnbuckle systems for fine adjusting the
main cables are only provided at one anchor, whereas provisions for later adjusting the spanning cables
turnbuckles are applied at both the walkway and tower foundations.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 115

For reducing vibrations and uncomfortable swinging, the bridge deck has been reinforced by two 2”
G.I. pipes underneath the walkway. These pipes can also be used for carrying drinking or irrigation water
across the river.

Adequate measures have been foreseen to install wind-bracings when needed.

3.3 Technical features and limitations


The Short Span Trail Bridge Standard as presented in this Chapter conforms to mainly Indian Standard,
but also Swiss and German codes and norms have been applied. All its components fulfill the necessary
safety factors by applying the loadings prescribed in the Long Span Trail Bridge standard design.
All the construction materials conform to international specifications. The exposed steel parts are all hot
dip galvanized, and should not be altered unless proven to fulfill standard norms.

For practical, economic and safety reasons the span range for the Short Span Trail Bridge
Standard presented in this Manual is limited to 120 m.

Longer spans are possible but would require special engineering input. As with every standard design,
not all site conditions are covered in this book. It is especially not applicable in unfavourable geological
site conditions. At such sites, as mentioned above for longer spans, an engineer’s input is mandatory.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 116

3.4 Bridge Design


This Chapter along with the Forms and Drawings is intended to provide quick and reliable technical
methods for design and bridge-construction for engineers and overseers.

¡¡ Layout and Sections

Figure F.3.2: A Typical Profile of Suspension Type Bridge

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 117

Figure F.3.3: A Typical Plan of Suspension Type Bridge

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 118

¡¡ Walkway Section

3.5 Basic design concept

3.5.1 Loadings

For designing the bridge structure loadings, the Long Span Trail Bridge standard is followed.
¡¡ Live load
The live load for smaller spans of up to 50m is equivalent to 400 kg/m2 and for longer spans as per
the formula,
50 50
P=300+100· kg/m2or 3 + kN/m2
l l
¡¡ Dead load
The dead load includes the weight of all permanent components of the bridge structure.

Width of Walkway 106 cm


Dead Load (without weight of Main & ~ 75 kg/m or
Spanning Cables) ~ 74 kN/m

¡¡ Wind load
The design wind load is taken as a uniformly distributed load based on a wind speed of 160km/h
acting horizontally to the walkway. This corresponds to a wind pressure of 1.3 kN/m2 acting on
the lateral bridge area of 0.3m2 per meter span. By maintaining a wind coefficient of 1.30 (acc.
to Swiss Standard) the actual lateral design wind load is 0.50 kN (1.3 x 1.3 x 0.3) per meter span.
The Standard Steel Towers and Foundation Structures are sufficient to resist this design wind load.
Wind load also affects the dynamic behavior of the bridge. For minimizing dynamic wind effects on
the bridge, sufficient dead load of the walkway, a steel deck and the most favourable span to sag
ratios have been introduced. Therefore, no lateral stabilizing measures (windguy system) has been
considered in this standard short span suspension bridge design. However, for special cases (spans
more than 120m or extreme windy areas exceeding wind speed of 160km/h) there is a provision
for fixing a windguy system.
¡¡ Snow load
The probability of the occurrence of a full load on a bridge loaded with snow is low. Moreover,
the design live load itself is comparatively high. Therefore, it is considered that the snow load is
already covered by the design live load.
¡¡ Temperature effect
Loading (cable forces) according to temperature effect is negligible in comparison with other
loading conditions. Therefore, temperature effect has not been considered for the standard bridge
design.
¡¡ Seismic load

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 119

There is low probability of a full live load occurring at the time of an earthquake. Moreover, the
design live load itself is comparatively high. Therefore, it is assumed that the seismic load is already
covered by the design live load. A separate loading combination with seismic load has not been
considered.

3.5.2 Construction materials

The construction materials used for the standard trail bridges are wire ropes (cables), steel parts, steel
fixtures and fasteners (thimbles, bulldog grips, nuts and bolts), concrete and stone masonry. Specifications
of these materials are based on the Indian Standard (IS).

¡¡ Steel Wire Ropes (Cables)


For more convenient stock keeping, handling and transportation only three cable diameters have
been used. This also reduces the number of steel parts, logistics, and eventually, bridge costs.

Steel wire ropes should comply with all the requirements of:
• IS 1835 – 1977 Steel wires for ropes
• IS 6594 – 1977 Technical supply conditions for wire ropes and strands
• IS 9282 – 1979 Specification for wire ropes and strands for suspension bridges
• IS 9182 (part II) –1979 Specification for lubricants for wire strands and ropes

Nominal diameters: 13 mm 26 / 32 mm
Rope Construction 7 x 7 (6/1) 7 x 19 (12/6/1)
Elongation non pre-stretched pre-stretched
Lay RHO, right hand ordinary lay
Core WRC, Wire Strand Core
Tensile strength of wire 160 kg/mm2 or 1.57 kN/mm2
Preforming Preformed
Coating Galvanized ‘A’ heavy
Impregnation Non-drying type and non-bituminous lubricant

Nominal Minimum Breaking Permissible load


Weight
Diameter of Metallic Area Load in in
Cable [mm2]
Mm kg/m N/m Tones kN Tones kN

13 0.64 6.3 73 10.5 103 3.5 34


26 2.51 24.6 292 39.3 386 13.1 129
32 3.80 37.3 442 59.6 585 19.9 195
Modulus of Elasticity, E = 110 kN/mm2 = 11 t/ mm2

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 120

Cross-section and Lay of Wire Rope


¡¡ Bulldog Grips
Bulldog Grips are used at the cable terminals to secure the cable ends.

Bulldog grips should conform to IS 2361-1970 specifications for Bulldog Grips. The bridges must
be dropforged and suitably scored to grip a round strand rope of right hand lay having six strands.
Bridges, U-bolt and nuts should be hot dip galvanized with a minimum zinc coating of 40μm. The
size of the bulldog grip should be equal to the size of the cable to be anchored or connected.

Bulldog Grip

¡¡ Thimbles
Thimbles are of open type and should confirm to IS 2315-1978 specifications for Thimbles for
Wire Ropes. They must be forged and hot-dip galvanized with a minimum zinc coating of 40μm.
Thimbles are used at the cable terminals to avoid kinks. The size of the thimbles should be equal
to the size of the cable to be anchored.

Thimble

¡¡ Methods of Applying Bulldog Grips to Wire Ropes

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 121

The number of bulldog grips, gap and overlapping length

Required Number, Gap Overlapping


Diameter of Wire
“n” of Bulldog “G” Length “L”
Rope, mm
Grips cm cm
13 3 10 60
26 5 15 130
32 6 20 180
The bridge of the bulldog grips must be fitted on to the working part of the wire rope and the
U-bolt on to the rope tail. The cable end should be protected from fraying with binding wire.

¡¡ Structural Steel
The steel grade should be of standard quality Fe 410 and the structural steel should comply with
the requirements of:
IS 226 – 1975 Structural Steel
IS 800 – 1984 General Construction in Steel

The tower design is based on the Swiss Standard SIA 161(1979) for Steel Structures.

Mechanical Properties
The steel should have the mechanical properties :
Tensile Strength ft ≥ 410 N/mm2
Yield Stress fy ≥ 250 N/mm2
Permissible Tensile Stress σat = 150 N/mm2
Modulus of Elasticity E = 200 kN/mm2
Elongation δ ≥ 23%
¡¡ Fasteners
Bolts, nuts and washers should be of grade C, property class 4.6 and should comply with the
requirements of:
IS 1363 - 1984 (Part 1) Hexagonal Head Bolts and Nuts
IS 1367 – 1983 Threaded Fasteners

The fasteners should be hot dip galvanized with a minimum zinc coating of 40μm.

¡¡ Reinforcement Steel
The reinforcement steel should be of steel grade Fe415, high yield deformed bars and should
comply with the requirements of:
IS 1786 – 1985 High Strength Deformed Steel Bars for Concrete Reinforcement
IS 456 – 1978 Plain and Reinforced Concrete

Mechanical Properties
The reinforcement steel should have the mechanical properties:
Yield Stress fy ≥ 415 N/mm2
Permissible Tensile Stress σat = 230 N/mm2
Modulus of Elasticity E = 210 kN/mm2

¡¡ Rust Prevention
Rust prevention of steel parts should be done by Hot-Dip Galvanization according to:
IS 2629 - 1966 Recommended Practice for Hot-Dip Galvanization of Iron and Steel
IS 4759 - 1984 Specification for Hot-Dip Zinc Coating on Structural Steel

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 122

Specifications for Hot-Dip Zinc Coating

Minimum Mass of Coating Minimum Thickness of


Products
g/m 2
N/m 2 Coating μm

Structural Steel 610 6.0 80


Threaded work, Nuts & Bolts 300 3.0 40

¡¡ Wire mesh netting


The wire for wire mesh netting should comply with the requirements of:
IS 280 - 1978 Mild Steel Wire for General Engineering Purposes
IS 4826 - 1979 Hot-dipped Galvanizing Coatings on Round Steel Wire

The diameter of the wire should be 12 SWG (2.64mm) and the zinc coating should not be less than
270g/m2. The average tensile strength of the wire should not be less than 380N/mm2.

¡¡ Concrete
The concrete should comply with all the requirements of:
IS 456 – 1978 Plain and Reinforced Concrete
IS 269 – 1989 Ordinary Portland Cement of 53 Grade
IS 383 - 1970 Coarse and Fine Aggregate

The concrete grades used in the standard design are:


Concrete 1:3:6 (M10) for miscellaneous use
Concrete 1:2:4 (M15) for structures
Concrete 1:3:6 mixed with 40% boulder

¡¡ Masonry
The masonry should comply with all the requirements of:
IS 1597 – 1967 Code of Practice for Construction of Stone Masonry
IS 2250 – 1981 Preparation and Use of Masonry Mortars

The stone masonry used in the standard design are:


Chisel Dressed Stone Masonry in 1:4 cement : sand mortar
Hammer Dressed Stone Masonry in 1:6 cement : sand mortar
Dry Stone Masonry.

¡¡ Unit weight of construction materials


The unit weight of the construction materials used in the standard design is given in the table.

Unit Weight,
Materials
kg/m3 kN/m3
Concrete 2200 22.0
Stone Masonry 2100 21.0
Steel 7850 78.5
Soil 1800 18.0
Water 1000 10.0

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 123
Short Span Trail Bridge Standard

3.5.3 3.6.3. STRUCTURAL


Structural ANALYSIS
analysis AND DESIGN
and design

TheThe staticalanalysis
statical analysisisisbased
based on
on the
the calculation
calculationofofforces
forcesand
andstresses in the
stresses structures
in the due to
structures dueexternal
to external
loadings. The calculated forces and stresses are compared with the permissible loads and stresses
loadings. The calculated forces and stresses are compared with the permissible loads and stresses of these of
structures with some safety factors.
these structures with some safety factors.
The Standard Suspension Bridge has been designed as per the statical analysis and basic design
concept.
The Standard Suspension Bridge has been designed as per the statical analysis and basic design concept.
• Bridge Geometry
¡¡ Bridge Geometry

Figure 22.3 The Geometry of the Suspension Bridge


Figure F.3.4: The Geometry of the Suspension Bridge

Where, l = Bridge Span, the horizontal distance between the tower saddles
D = Backstay distance, the horizontal distance between the tower axis and the hinge of the main
cable anchorage or the front of the main cable Foundation
DR = Backstay distance of the main cable on the right bank
DL = Backstay distance of the main cable on the left bank
f = Sag of the Main Cable, the vertical distance from the tower saddle to the lowest point of
the Main Cable
fd = Sag of the Main Cable in a dead load case
ff = Sag of the Main Cable in a full load case
c = Camber of the Spanning Cable, the vertical distance from the spanning cable anchorage to
the highest point of the spanning cable
cd = Camber of the Spanning Cable in a dead load case
βF = Frontstay cable inclination at the tower saddle
βB = Backstay cable inclination at the tower saddle
βf = Cable inclination at full load
β Ff = Frontstay Cable inclination at the tower saddle in a full load case
ht = Tower Height
Fb = Freeboard, the vertical clearance from the high flood level (HFL) to the Spanning Cable
Anchor point
Equation of the bridge geometry:
Total Height of Tower = ht + 0.7m = fd + cd + 1.1m
⎛ 4 f ⎞
Frontstay Cable Inclination = βF = arc ta n ⎜ ⎟
⎝  ⎠
⎛ 4 f f ⎞
Backstay Cable Inclination = β B ≈ β Ff = arc tan ⎜ ⎟ = β f
⎜  ⎟
⎝ ⎠
• Selection of Bridge Geometry
The
¡¡ selection of the
Selection of bridge
Bridgegeometry
Geometry means defining the bridge span "l", Tower height "ht", main cable
" and
sag "fdThe camber "c " of the spanning
selection of dthe bridge geometry cable in a defining
means dead loadthecase. Thespan
bridge relation
“l”, between these “h ”,
Tower height
geometrical
main cable sag “fd” and camber “cd” of the spanning cable in a dead load case. TheThe
parameters of the bridge is given in the above "Equation of the bridge geometry". t
relation
optimum and rational
between these design of a suspension
geometrical parametersbridge largely
of the depends
bridge on the
is given proper
in the aboveselection of theof the
“Equation
bridge bridge
geometry. For this purpose, many combinations of the bridge geometries are calculated
geometry”. The optimum and rational design of a suspension bridge largely depends and the on

133

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 124

the proper selection of the bridge geometry. For this purpose, many combinations of the bridge
geometries are calculated and the final optimal results are presented in the “Table: Selection of
Cables and Standard Towers” of this manual.

Bridge Span:
For the sake of simplicity, the spans have been rounded off to the meter in the standard design. One
suspender at the middle of the span is a must. Hence, the distance from the tower axis to the 1st
suspender will vary as per the odd/even length of the span in meters. The first cross beam is fixed without
a suspender. For further details, refer also to the sub-chapter “Suspenders”, which will be discussed later.

¡¡ Cable Design
A cable hanging between two supports and carrying a uniformly distributed load “q” forms a
parabola. Thus the Main Cables and the Spanning Cables are assumed to be of parabolic geometry.
In the suspension type bridge, the main cables carry all the load. The Spanning Cables carry only
the pretension load. It is assumed that when a bridge is in full load, the tension in the spanning
cable will be zero. The sketch represents the cable geometry and the forces acting on it.

Short Span Trail Bridge Standard

final optimal results are presented in the "Table: Selection of Cables and Standard Towers" of this
manual.

Bridge Span:
For the sake of simplicity, the spans have been rounded off to the meter in the standard design. One
suspender at the middle of the span is a must. Hence, the distance from the tower axis to the 1st suspender
will vary as per the odd/even length of the span in meters. The first cross beam is fixed without a suspender.
For further details, refer also to the sub-chapter "Suspenders", which will be discussed later.
• Cable Design
A cable hanging between two supports and carrying a uniformly distributed load “q” forms a parabola.
The dead Thus
loadthe
sagMain Cables to
fd is chosen and theanSpanning
have Cablesgeometry.
optimal bridge are assumed to be ofsag
The hoisting parabolic geometry.
fh and the full load In the
suspension
sag ff are calculatedtype
as per bridge, the theory
the elastic main cables carry allmethod
by the iteration the load. TheonSpanning
based the chosen Cables carrysag
dead load only the
fd. The pretension
hoisting sag load.
fh isItnecessary
is assumedforthat when
cable a bridge
hoisting is in full
purpose load,
while the tension
erecting in theThe
the bridge. spanning cable
full load sagwill be
zero. The
ff is necessary sketch
for the represents
calculation themaximum
of the cable geometry
tensionand the cable.
in the forcesThe
acting on combination
cable it. is designed
as per the calculated maximum tension.

Bridge Geometry
⇒ Sag of the Main Cable at mid span in Dead Load, fd = 0.075 l to 0.14 l
⇒ Camber of the Spanning Cable at mid span in Dead Load, cd = 0.015 l to 0.03 l
⇒ Cable Inclination at Saddles,
⎛ 4 f ⎞
Frontstay Angle: β F = arctan ⎜ ⎟
⎝ l ⎠
⎛ 4 f f ⎞
Backstay Angle: β B ≈ β Ff = β f = arctan⎜⎜ ⎟⎟
⎝ l ⎠
⎡ 8 ⎛ f d ⎞ 2 ⎤
⇒ Main Cable Length between tower saddles in Dead Load, L md = l ⋅ ⎢1 + ⎜⎜ ⎟ ⎥
3 ⎝ l ⎟⎠ ⎥
⎣⎢ ⎦
2
⎡ 8 ⎛ c ⎞ ⎤
⇒ Spanning Cable Length between tower axis in Dead Load, L sd = l ⋅ ⎢1 + ⎜⎜ d ⎟⎟ ⎥
⎢⎣ 3 ⎝ l ⎠ ⎥⎦
Forces
The fundamental equations for the calculation of the cable forces are:
g l2
⇒ Total Horizontal Tension on the Main Cable in any Load case, H =
8f
H
⇒ Total Maximum Tension on the Main Cable in any Load case, T max =
PART F: TRAIL BRIDGE MANUAL cos β F
gl2
⇒ Total Horizontal Tension on the Main Cable in Full Load case, H=
8f f
⎢⎣ 3 ⎝ l ⎠ ⎥⎦
2
⎡ 8F ⎛- Chapter
c d ⎞ ⎤ 3 - 125
⇒ Spanning Cable Length between tower axis in Dead Load, L sd = l ⋅ ⎢1 + ⎜⎜ ⎟⎟ ⎥
⎢⎣ 3 ⎝ l ⎠ ⎥⎦
Forces
The fundamental equations for the calculation of the cable forces are:
g l2
⇒ Total Horizontal Tension on the Main Cable in any Load case, H =
8f
H
⇒ Total Maximum Tension on the Main Cable in any Load case, T max =
cos β F
gl2
⇒ Total Horizontal Tension on the Main Cable in Full Load case, H=
8f f
H
⇒ Total Maximum Tension on the Main Cable in Full Load case, Tmax =
cos β f
gf l
⇒ Total Horizontal Tension on the Main Cable in Full Load case, V max =
2
The Safety Factor for Cable Breaking Load is taken as fs ≥ 3.

The dead load sag fd is chosen to have an optimal bridge geometry. The hoisting sag fh and the full load sag
Theff Main Cables are
are calculated asselected to withstand
per the elastic theory the maximum
by the iterationtension
methodin based
full load
oncase
the with a safety
chosen dead load sag fd. The
factor of
hoisting sag fh is necessary for cable hoisting purpose while erecting the bridge. The full load sag ff is
fs = TBreak
necessary for the calculation of the maximum / Tmaxin≥the
tension 3 cable. The cable combination is designed as per
the calculated maximum tension.
The Spanning Cables are selected as per the maximum tension Ps in load case [A] or [B] (“Ps” and loading
cases are discussed in sub-chapter “Steel Towers”) with safety of factor, fs = TBreak / TsA/B ≥ 3.

As134
per the above design concept and the optimum bridge geometry, a simplified cable combination
chart has been developed and standardized for the suspension type bridge for spans up to 120m (refer
to “Table: Selection of Cables and Standard Towers” in this Manual). The Main and Spanning Cable
combinations, the dead load sag fd, the hoisting sag fh of the main cables, the camber cd of the spanning
cables, and the cable inclination under full load βf at the tower saddle for a given span is presented in
this chart.

¡¡ Steel Parts
The steel parts of the suspension type bridge are mainly subject to axial tension/compression, bending
stress, shear stress, bond stress and buckling effect. All the steel parts are designed to withstand these
forces and stresses with a safety factor of fs = 1.6.

All the steel parts have been standardized according to the cable combinations and do not depend on
the site condition.

¡¡ Walkway Structure
The walkway structures (Steel deck, Cross beams, Suspenders) are designed for local loadings like all
steel parts with a safety factor of fs = 1.6. A concentrated load P = 150kg (weight of a porter with a load)
positioned along the walkway as well as across the walkway has been considered.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 126

Standard Walkway width (WW) is 106cm.


¡¡ Steel Towers
The Standard Steel Towers are designed as hinged. Different possible loading cases were considered in
order to determine the worst case with respect to buckling and yielding of the most critical elements of
the towers. Two loading cases, i.e. load combination with wind load are found most critical. They are:
Load Case [A] = Dead Load + Full Wind Load
Load Case [B] = Full Load + 1/3 Wind Load
A global wind load of 1.7kN/m2 (170kg/m2) for exposed area, which is equivalent to a wind speed of
160 km/hr has been considered for the design. The wind load is assumed to be a uniformly distributed
horizontal load. Wind loads on different bridge elements are calculated based on the global wind load
and the area exposed to the wind of the respective elements, which are:
on Walkway, Www = 0.50 kN/m (50 kg/m)
on Main Cable, WM = 0.075 kN/m (7.5 kg/m)
on Suspenders, Ws = 0.035 kN/m (3.5 kg/m)
on Tower, Wt = 1.025 kN/m (102.5 kg/m)

The calculations are made under the assumption that all the wind loads are taken by the towers
and ultimately by the walkway and tower foundations. No windguy arrangement is considered in the
calculation.

The simplified model is used for the statical analysis:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 127

With reference to the above Force/Vector diagram, the forces are calculated as:
Forces without wind load Forces with wind load
Wind load, Wb = WM + Ws = 0, PM1 · Sinγ1 + PS1 · Sinα1 = Wb = WMO + WsO
PMO – PSO = gO PMO · Cosγ1 - PS1 · Cosα1 = gO

Where PM and PS are the forces on the main cables and spanning cables respectively.

The statical analysis with the above load combination is mainly to compute the displacement “X” and
the sag “f1” and the corresponding forces on the main cables PM1 and PS1 for both load case [A] and
[B]. The calculation is based on the elastic theory and iteration method. However, while calculating the
stiffness, the constant factors “CSO” of the walkway system, additional stiffness of the steel cross beam/
deck reinforced with pipes are also considered.

Further, as per the calculated “PM1” and “PS1”, the possible maximum vertical load “Vtot” and the
horizontal load “Hw” on the towers are computed. Similarly, the reactions on the tower base “P1”, “P2”
and “PH” are also computed.

Forces on tower

The Tower Capacity Diagram as a function of “Vtot” and “Hw” of the proposed Standard Towers is
presented bellow.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 128

Tower capacity diagram

An appropriate Standard Tower is selected as per the calculated maximum vertical “Vtot” and horizontal
“Hw” loads. The plotted point with the coordinates of “Vtot” and “Hw” in the “Tower Capacity Diagram”
should be below the capacity line of the selected standard tower.

As per the above statical analysis and design concept, a simplified chart for the selection of towers has
been developed and standardized for the suspension type bridge for different bridge spans of up to
120m in the “Table: Selection of Cables and Standard Towers” of this manual.

¡¡ Design of Walkway and Tower Foundations.


The standard design does not have a windguy system. The foundations are designed to take a
wind load satisfying the worst case from the load cases [A] and [B].

The foundations are designed as per the static analysis and the principle of soil and rock mechanics.
The design loads on the Walkway and the Tower Foundations are equivalent to the reactions on
the tower base “P1”, “P2” and “PH” and the tension on the spanning cables “Ts”, which are
calculated for both the load cases [A] and [B].

Overall stability of the foundation:


The mode of failure and respective safety factors for the overall stability of the foundation block are
considered in the design.
¡¡ Sliding FSL ≥ 1.5
¡¡ Toppling B*/2 ≥ B/3 ; L*/2 ≥ L/3
¡¡ Maximum Ground Pressure σmax ≤ σperm , Bearing capacity of soil or rock
¡¡ Ground Shear Failure (Ultimate Bearing Capacity ) in case of soil

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 129

Short Span Trail Bridge Standard


Consider the design concept of the foundation:

Re sistingForce Rv ⋅ tan φsl


Safety factor against Sliding: Fsl = = ≥ 1.5
DrivingForce RH
Where RV = W + (- P1) + P2 - TSV + Eav
RH = TSH + Eah
for the calculation of the Active Earth Pressure Eah and Eav refer to sub-chapter "Main Cable Deadman
Anchor Block in Flat Ground", which will be discussed later in this chapter.

Safety factor against Overturning:


The safety factor against overturning is checked by the eccentricity of the resultant force. The eccentricity
should be within the limit of
B* / 2 ≥ B / 3 and L* / 2 ≥ L / 3

Where B* / 2 = B / 2 - ⏐ex ⏐; ex = ∑ M X / R V ;


L* / 2 = L / 2 - ⏐eY ⏐; ey = ∑ M Y / RV ;

∑MX = TSH ⋅ (H - 0.70) + Eah ⋅ H / 3 - Eav ⋅ B / 2 - Eahp ⋅ hp / 3 + Eavp ⋅ B / 2


∑MY = - P1 ⋅ (C/C1) / 2 + P2 ⋅ (C/C1) / 2 + PH ⋅ H

Safety against Ground Bearing Pressure:


RV
σ max = Z . ≤σ perm , permissible bearing pressure of soil/rock.
B*L*
Where the Z – factor considers the location of the resultant force and is a function of (L*/2)/ L and (B*/2)/B.
Value of Factor, Z
Value of Factor, Z
Z - factor (B* / 2) / B
(L* / 2) / L 0.50 0.45 0.40 0.35 0.30 0.25

138

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 130

Z-factor (B* / 2) /B

(L* / 2) / L 0.50 0.45 0.40 0.35 0.30 0.25

0.5 1 1.17 1.28 1.33 1.33 1.33

0.45 1.17 1.30 1.36 1.39 1.39 1.39

0.4 1.28 1.36 1.41 1.43 1.43 1.43

0.35 1.33 1.39 1.43 1.46 1.47 1.47

0.3 1.33 1.39 1.43 1.47 1.49 1.50

0.25 1.33 1.39 1.43 1.47 1.50 1.50

Nevertheless, the minimum size of the foundation is governed by the geometric requirements to fix the
tower, anchorage/other steel parts and also the topography.

Safety against Ground Shear Failure (Ultimate Bearing Capacity):


In our case, the horizontal force acting on the foundation is relatively greater than the vertical load.
The safety against ground shear failure or bearing capacity depends mainly on the vertical load on the
foundation. Therefore, in a majority of the cases, this failure mode is not the deciding factor in the
foundation design. However, there is comparatively high vertical force acting on the walkway and the
tower foundation, and ground shear failure may be the deciding factor in its design.

The ground shear failure analysis is carried out by Terzaghi’s extended and amplified bearing capacity
formula. Terzaghi’s formula is extended and amplified with different co-efficient considering the foundation
shape and topography of the ground.

Safety against ground shear failure, FBC = P*/P ≥ 2.0 for Terzaghi’s formula,

where, P* - shear resistance of the ground (ultimate bearing capacity ) determined by


Terzaghi’s formula
P - vertical resultant force on the foundation

Terzaghi’s formula for shallow foundation with continuous footing and horizontal ground,

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 131

P* = B [c NC + (ϒ t + q) Nq + 0.5 B ϒ Nϒ],

cohesion factor “c” is not considered since the predominant soil type in Nepal is non-cohesive,

therefore, Terzaghi’s formula above will be,

P* = B [ (ϒ t + q) Nq + 0.5 B ϒ Nϒ]

In our case, safety against ground shear failure, FBC = P*/Rv ≥ 2.0

where, P* - shear resistance of the ground (ultimate bearing capacity ) determined by


Terzaghi’s formula with correction factors for topography and shape of the
foundation,

Rv - vertical resultant force on the foundation

Terzaghi’s formula with correction factors,

P* = B* L* [(ϒ t + q) Nq Sq dq iq b’q gq + 0.5 ϒ B* Nϒ Sγ dγ iγ b’γ gγ]

where, ϒ = unit weight of soil


t = embedded depth in front of the foundation
q = surcharge load, q = 0 in our case
Nq = eΠtanφ tan2(45° + 0.5 φ), co-efficient for embedded depth or surcharge load,
Nϒ ≈ 1.8 (Nq − 1) tan φ, co-efficient for effect of the foundation width,

shape correction factors considering limited length (L*) of continuous footing,


Sq = 1 + (B*/L*) tan φ
Sγ = 1 – 0.4 (B*/L*)

depth correction factors considering the embedded depth (t = hp) of the foundation,
dq = 1 + 0.035 tan φ(1 – Sin φ)2 arctan(t /B*),
dγ = 1

embedded depth in front of the foundation should be guaranteed up to the length of the influence area
L*infl,
L*infl = B* tan(45° + 0.5φ) e0.5 Π tanφ

correction factor for the inclination of the load,


iq = [1 – 0.5 tan(δR - α) =]5

iγ = [1 – (0.7 – 0.0022 α°) tan(δR - α)]5

correction factor for the inclination of the foundation base, α,


b’q = e(– 0.035 α° tan φ)

b’γ = e(– 0.047 α° tan φ)

correction factor for the inclination of the base line (εB), which is approximately equal to the inclination of
the ground slope (ε) but limited to the ground slope up to 15° only,
gq = gγ = (1 – 0.5 tan εB)5, εB ≈ ε

Experience shows that Terzaghi’s amplified and extended formula with correction factors is no more
reliable for foundations on steep slope ground. It gives relatively acceptable results only for a ground
with a slope of less than 15°. Therefore, an overall slope stability analysis is made for slope ground of
more than 15° inclination. The overall slope stability analysis is carried out by using Bishop’s simplified
method of slices.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 132
Short Span Trail Bridge Standard

n
1
∑ [(G i + Vi ) tan φ ']
mαi
F= i =1
n
a
∑ [(G
i =1
i + Vi )sin α i ] + RH
R

⎡ tan φ ' tan α i ⎤


mαi = cosα i ⎢1 + ⎥
⎣ F ⎦

where, in our case


where, in our case Gi = weight of the soil of the given slice "i"
Gi = weight of the soil of the given slice “i”
Vi = RV, resultant vertical force on the foundation, which is applicable only at
Vi = RV, resultant vertical force on the foundation, which is applicable only at one slice and in
other slices, Vi = 0.
one slice and in other slices, Vi = 0.
H = H = RH,force
RH, resultant horizontal resultant horizontal
on the force on the foundation
foundation
φ’ = 1.10 x φ, φ' = 1.10 x φ,
φ = angle of internal friction of the soil
αi = φ =at the
tangential angle angle
slipofcircle
internal friction
of given of “i”
slice the soil
R = Radius of the α i =circle
slip tangential angle at the slip circle of given slice "i"
a = lever arm of “R ”
R H = Radius of the slip circle
F = safety factor, which is taken as 1.5 for the Bishop’s Method
a = lever arm of “RH”
As the factor of safety “F”F appears
= safety
onfactor, which
both the is taken
sides, as 1.5
a process offor the Bishop's
successive Method is required.
approximation
It is necessary to fix the center “o” corresponding to the critical slip circle with minimum safety factor. A
As the program
computer factor ofissafety "F" appears
also available to do on
thisboth the sides, a process of successive approximation is
analysis.
required. It is necessary to fix the center "o" corresponding to the critical slip circle with minimum
As safety
per thefactor.
analysis, it is established
A computer programthat
is ifalso
theavailable
foundations are
to do placed
this behind the slope line with the angle
analysis.
of internal friction “φ”, shear failure in general, will not occur.
As per the analysis, it is established that if the foundations are placed behind the slope line with the
Design
angle of
of the Reinforced
internal Concrete
friction "φ", shear Frame of general,
failure in the Foundation:
will not occur.
The Concrete Frame of the walkway/tower foundation is designed as reinforced concrete columns and
foundation with a combined footing. Consider a typical design of the frame:
Design of the Reinforced Concrete Frame of the Foundation:
The Concrete Frame of the walkway/tower foundation is designed as reinforced concrete columns and
foundation with a combined footing. Consider a typical design of the frame:

PART F: TRAIL BRIDGE MANUAL 141


F - Chapter 3 - 133

The cross sectional sizes of the columns (b x l ) are designed to fulfill the geometric requirement to fix the
tower legs and also to fulfill the criteria of the direct compressive stress on the columns:
Direct compressive stress on column, σac = P1 or P2 / b x l ≤ σac perm., permissible direct compressive
stress.

The size of the foundation (B x L) is also checked to fulfill the criteria of the maximum bearing pressure
at the foundation.

Maximum bearing pressure at the foundation, σmax = (P1 + P2 + W) / L x B ≤ σperm , permissible bearing
pressure of soil/rock.

A balanced section has been assumed for the reinforced concrete design of the frame. The concrete
grade of M10 has been adopted. The structural and loading model is used for the structural analysis.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 134

Consider a typical result of the structural analysis:

The thickness of the foundation slab and the requirement of reinforcement steel is designed as per the
structural analysis (Shear Force and Bending Moment diagram).

Thickness of Foundation Slab:


The thickness of the foundation slab “d” has been designed as per the maximum bending moment, and
Shortchecked forBridge
Span Trail punching shear due to vertical load from the column.
Standard

2 M max
d=
ξ (1 − ξ / 3 ) Bσ c perm

⎛ 2 ⎞
Where ξ = n µ ⎜⎜ 1 + − 1 ⎟⎟ ;
⎝ nµ ⎠
AS
µ= ; Ratio of Area of steel to area of concrete section
dB
ES
n= ; Ratio of modulus of elasticity of steel to concrete;
EC

in longitudinal direction of the foundation slab


M max ( Long .)
AS =
η d σ t perm

PART
in transverse direction of the foundation F: TRAIL BRIDGE MANUAL
slab
M max ( Trans .)
AS =
η d σ t perm
AS
µ= ; Ratio of Area of steel to area of concrete section
dB F - Chapter 3 - 135
ES
n= ; Ratio of modulus of elasticity of steel to concrete;
Required Area of steel: EC

in longitudinal direction of the foundation slab


M max ( Long .)
AS =
η d σ t perm

in transverse direction of the foundation slab


M max ( Trans .)
AS =
η d σ t perm
1
Where η = 1−
σt
3 (1 + )
nσ c
σt - permissible tensile stress of the reinforcement steel
σc - permissible compressive stress of the concrete

• Main Cable Anchor Blocks


¡¡ Main Cable Anchor Blocks
There areThere
mainlyare mainly
three three
basic basic
types typescable
of main of main cable
anchor anchor
blocks blocks depending
depending upon theupon the soil/rock
soil/rock type andtype an
the prevailing topography:
the prevailing topography:
• Main
⇒ Cable
Main Gravity
Cable Anchor
Gravity BlockBlock
Anchor in soil/rock for Slope
in soil/rock Topography;
for Slope Topography
• Main Cable Deadman Anchor Block in Soil for Flat Topography;
• ⇒
Main Main Cable
Cable DrumDeadman
Anchor Anchor
Block in Block
Rock. in Soil for Flat Topography
⇒ Main Cable Drum Anchor Block in Rock
The Main Anchor Blocks are designed to withstand the maximum tension of the main cables based on
the cable structure analysis and considering the soil/rock parameters.
The Main Anchor Blocks are designed to withstand the maximum tension of the main cables based o
the Gravity
Main Cable cable structure
Anchoranalysis
Block inand considering
Soil/Rock the soil/rock
for Slope parameters.
Topography:
This type of anchor block is designed for slope topography. The mode of failure and respective safety
factors
Mainare Cable
considered in the
Gravity design.
Anchor Block in Soil/Rock for Slope Topography:
• Sliding FSL ≥ 1.5;
This type of anchor block is designed for slope topography. The mode of failure and respective safety facto
• Toppling FT ≥ 1.5;
are• considered in the design.
Maximum Ground Pressure σmax < Bearing capacity of soil or rock;
• ⇒ Sliding
Ground Shear Failure (Ultimate Bearing ≥ 1.5 ) in case of soil.
FSL Capacity
⇒ Toppling
Consider the design concept of the anchor block:T
F ≥ 1.5
⇒ Maximum Ground Pressure σ max < Bearing capacity of soil or rock
⇒ Ground Shear Failure (Ultimate Bearing Capacity ) in case of soil

144

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 136

The anchor blocks have been designed to satisfy the safety factor against sliding with the theoretical
basis:
Safety factor against Sliding:

Re sistingForce R 'V ⋅ tan φsl


Fsl = = ≥ 1.5
DrivingForce R' H
Where R'V = RV ⋅ cosα + RH ⋅ sinα; RV = Wce
Re' sistingFor 1 – W2 R +'VEav
⋅ tan- Tv
φsl
RVsistingFor
R'H =Re RFce = Rce
⋅ sinαRe+sistingFor
Hsl⋅ cosα; V R
⋅ tan
H R
φ= '
slV
T⋅Htan
+E =
φ
ah '
≥ 1.5
Fsl = Fsl = = DrivingFor
= ≥ce1.5 ≥ 1.R5 H sl

DrivingForDrivingFor
ce ceAllR ' HRe sisting
R ' H Moments ΣM A+
Safety factor R'V = RV ⋅ cosα + RH ⋅Fsinα;
Whereagainst Overturning: ovt = RV = W1 – W2 + Eav - =Tv ≥ 1.5
Where Where R'V = RR'
V ⋅V cosα
= R+
V ⋅Rcosα
H sinα;
+ R H ⋅ R
sinα;
V = WR1V– W= 2W+1 E

R'H = V sinα + RH ⋅ cosα; RH = TH +Eah
⋅ All
W
av - 2 Driving
Tv
+ E av - Moments
Tv ΣM −
A
R'H =Where
RR'
V ⋅H sinα
= RV+ ⋅Rsinα
HΣM ⋅ cosα;
A
++ RH= ⋅W R cosα;
H = TRH+E2 ah⋅=XT2H++E
1 ⋅ X1 –HW Eavah ⋅ B
-
ΣMA = TV ⋅ (B – XF ) + T All
(Y Re sisting Moments ΣM A+
3 =H ⋅ 1 - ΔH) ++Eah ⋅ (ha+
/3 - Δ=H) ≥ 1.5
All Re sisting All Re Moments
ovt sistingAllMomentsΣM A Moments
ΣM A
Fovt = Fovt = =Driving =≥ 1.5 ≥ 1 .5 ΣM A−
All Driving All Moments
Driving Moments − −
Where ΣMA+ = W1 ⋅ X1 – W2 ⋅ X2 + Eav ⋅ B ΣM A ΣM A
Where Where ΣMA+ =ΣMW1A⋅+ X= 1 W–W ⋅ 2X⋅A1X- –2 =
1ΣM +WT
E2Vav⋅ ⋅X B
2 +–EX
⋅ (B av3)⋅ +B TH ⋅ (Y1 - ΔH) + Eah ⋅ (ha/3 - ΔH)
- -
ΣMA =ΣMTV A⋅ (B=–TXV3⋅) (B + T–HX⋅ 3(Y ) +1 -TΔ H H)
⋅ (Y+1 E- ahΔH)
⋅ (h+a/3Eah
- Δ⋅ H)
(ha/3 - ΔH)

Safety against Ground Shear Failure:


The position of the anchor blocks should be behind the slope line of the internal angle of friction of the
soil to guarantee the safety against Ground Shear failure.

Lateral Pressure at Side Walls:


Further, a curtain wall with cement stone masonry of sufficient thickness has been provided to resist the
lateral pressure resulting0the ⎡ effect
.28Τbulging n 2 due ⎤to tension on the
z deadman beam. The lateral pressure
σ Z of
“σZ” along the length Η ⎢
= the foundation “B” has been n
⎥ ; calculated
= as lateral pressure against a rigid wall as
2
per the elastic theory of soilBmechanics
⎢⎣ 0.16with
+ n ⎦
the (
2 3 ⎥formula: B )
0.28Τ Η ⎡ n2 ⎤
σz
0.12 z
⎡
0.28Τ Η 0.28Τ ΗnZ σ 2⎡= σ⎤m2a x ⎢ ⎤
n 0.1
z ⎥ ; z n=
σZ = σ Z2 = ⎢ ⎢ B⎥ 2; ⎢ 0⎥.16
0.08
0.06
⎣3 n (
; =+ n 2 )
3
n⎥⎦= B
B 2
(
⎢⎣ 0B.16 3
()
+⎢⎣n02.16⎥⎦ +σz n 2 ⎥⎦
0.04
0.02

0.12
0
) B B
B
σz σz σ 0 2 4 6 8 10
m0.1
ax
0.12 0.12 0.08
σ m0.1
ax σ m0.1
ax 0.06
0.08 0.08
0.06 0.06
0.04
0.02 Distribution of lateral pressure σ z along B
0.04 0.04 0 B
0.02 0.02 0 2 4 6 8 10
0 0 B B
0 0 2 2 4 4 6 6 8 8 10 10

Distribution of lateral pressure σ z along B


Distribution
Distribution
of lateralofpressure σ z alongσBz along B
lateral pressure

Main Cable Deadman Anchor Block in Soil for Flat Topography


This type of anchor block is designed for soil in a flat topography. The advantage of earth resistance,
i.e. passive earth pressure in front of the anchor block, is used in the design. This concept results in very
economical anchor block designs.

PART F: TRAIL BRIDGE MANUAL


Main Cable Deadman
economical Anchor
anchor block Block in Soil for Flat Topography
designs.
This type of anchor block is designed for soil in a flat topography. The advantage of
F - earth resistance,
Chapter 3 - 137i.e.
passive earth pressure in front of the anchor block, is used in the design. This concept results in very
economical anchor block designs.

undisturbed

undisturbed

The modes of failure and respective safety factors have been considered in the design.
The modes of failure and respective
⇒ Sliding FSL ≥ 3.5safety factors have been considered in the design.
¡¡ Sliding FSL ≥ 3.5
To the
To prevent prevent
anchortheblock
anchor block
from from movingthetowards
moving the soil in front, afactor
high has
safety factor has been
The modes of failure
applied. and respective safety towards
factors have soil
beeninconsidered
front, a high
in safety
the design. been applied.
¡¡ Toppling FT ≥ 1.5
⇒ Sliding FSL ≥ 3.5
Tofactor
Safety prevent
⇒ the anchor
Toppling
against Fblock
Sliding:T 1.5 moving towards the soil in front, a high safety factor has been
≥ from
applied.
H ult + (W + W E − TV ) tan φ1
⇒ Toppling FT ≥ 1.5 Fsl = ≥ 3.50
TH
Where Hult = Eph – Eah + 2 EoL H ult + (W + W E − TV ) tan φ1
Fsl = ≥ 3.50
hp
2
TH cos 2 Φ 1
E ph = λ ph L γ 1; λ ph = 2
2 2 EoL
Where Hult = Eph – Eah + sin( Φ 1 − δ ) sin( Φ 1 + ε ) ⎤
⎡
2 ⎢1 − ⎥
hp cos cos
⎢⎣ 2
Φδ1 cos ε ⎥⎦
E ph = λ ph L γ 1; λ ph = 2
2 2 ⎡ sin( Φ 1 − δ ) sin( Φ + ε ) ⎤
ha ⎢ 1 − cos 2 Φ 2 1 ⎥
E ah = λ ah L γ 2; λ ah = cos δ cos ε
2 ⎣⎢ ⎡ sin( Φ 2 + δ ) sin( Φ 2 − ψ⎦⎥ ) ⎤ 2
⎢1 + ⎥
coscos Φδ2 cosψ
2
ha ⎢⎣ 2
⎥⎦
E ah = λ ah L γ 2; λ ah = 2
1 2 ⎡ sin( Φ 2 + δ ) sin( Φ 2 − ψ ) ⎤
6
3
[
E OL = λ O γ 1 h p tan φ1 λ ph⎢1++λ ah ]
cos δ cosψ
⎥
⎣⎢ ⎦⎥
λo = 1 - Sinφ1
1
E OL =
Safety factor against 6
λ O γ 1 h p 3 tan φ1
Overturning:
[ λ ph F
ovt
] +
+ λ ah= All Re sisting Moments = ΣM A ≥ 1.5
All Driving Moments−
ΣM A
λo = 1 - Sinφ1
+
Where MA+ = W ⋅ B/2 + WE ⋅ F B/2 +=EphAll
⋅ hp/3 Eav ⋅ BMoments = ΣM A ≥ 1.5
Re+sisting
ovt
All Driving Moments ΣM A−
MA- = TH ⋅ ht + TV ⋅ 0 + Epv ⋅ 0 + Eah ⋅ ha/3

Where MA+ = W E⋅ avB/2


= E+ahWtan (2/3φ )
E ⋅ B/2 +2 Eph ⋅ hp/3 + Eav ⋅ B

MA- = TH E⋅ pv
ht =
+ETph tan (- 1/2φ )
V ⋅ 0 + Epv ⋅ 01 + Eah ⋅ ha/3

Eav = Eah tan (2/3φ2)


Epv = Eph tan (- 1/2φ1)

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 138
Short
ShortSpan
SpanTrail
TrailBridge
BridgeStandard
Standard
¡¡ Main Cable Drum Anchor Block in Hard Rock
An An
RCC RCC
An RCC Concrete Concrete
Concrete Drum
DrumDrum
withwith
with rock
rockrock anchoring
anchoring
anchoring is
isisusedused
used forfor
for foundations
foundations
foundations onon
on sound
sound
sound hard
hard
hard rock.
rock.
rock.TheThe
The sizesofof
sizes
sizes ofthe
the
the
drum drum
and and anchorage
anchorage rods rods
are are designed
designed to to withstand
withstand the the
shearshear forces
forces duedue
to to the
the horizontal
horizontal forcesofofthe
forces the
drum and anchorage rods are designed to withstand the shear forces due to the horizontal forces of the
cables.cables. It isdesigned
It is also also designed to the
to resist resist the forces
uplift uplift forces
on the on
cables. It is also designed to resist the uplift forces on the Drum. Drum.the Drum.

The The required number of anchorNrods, N is calculated as:


Therequired
requirednumber
numberofofanchor
anchorrods,
rods, Nisiscalculated
calculatedas:
as:

TTmax cos
cosββf f
AAstst == max
ττperm
perm

44AAstst TT maxsin
sinββf f
NN== oror NN== max
ππdd 22
σσBBperm
perm
ππddl l

NNisisadopted
adoptedwhichever
whicheverisisgreater
greaterfrom
fromthe
theabove.
above.

MainMain
Main CableCable
CableDrum
DrumDrum Anchor
Anchor
Anchor Block
Block
Block in Fractured
ininFractured
Fractured Hard
Hard
Hard Rock
Rock
Rock or or Soft
orSoft
SoftRockRock
Rock
A Drum anchorage without rock anchoring is used for foundations on fractured hard rock or fresh soft
AADrum
Drum anchorage
anchoragewithout
rock. withoutrock
rockanchoring
anchoringisisused
usedfor
forfoundations
foundationson
onfractured
fracturedhard
hardrock
rockororfresh
freshsoft
softrock.
rock.

The
Thesize
Theofof
size the
size ofDrum
the Drum isisdesigned
the Drum designed totowithstand
is designed withstandthe
theshear
to withstand the failure
shear failure
shear due
duetoto
failure the horizontal
theto
due horizontal forces ofofthe
forcesforces
the horizontal cables.
theof
cables.
the cables.

Determine
Determinethe
theDrum
Determine size
sizeand
the Drum
Drum sizethe
and number
and
the ofofreinforcement
the number
number bars
barsas:
of reinforcement
reinforcement bars as:
as:

⎛⎛ππDD2 2 ⎞⎞
⎜⎜⎜ ⎟⎟⎟ττc c ++AAststττstst ==TTmax cos β
max cos βf f
⎝⎜⎝ 44 ⎠⎟⎠

The
Therequired
requirednumber
numberofofanchor
anchorrods,
rods,NN isiscalculated
calculatedas:
as:

44AAstst
NN==
ππdd2 2
•• Suspenders
Suspenders
¡¡ Suspenders
The
Thesuspenders
The suspenders have been
suspenders have
havebeen
designed for
beendesigned
the
forfor
designed
bridge
thethe
bridgegeometry
geometry
bridge
under
under
geometry
dead load,
deaddead
under i.e.l,l,fi.e.
load,i.e.
load, dfdand
and c . .The
l, fdcddand
Thelayout
cdlayout
. The
ofofthe suspenders
the suspenders is :
is :
layout of the suspenders is :

FFs s== FFs s==

PART F: TRAIL BRIDGE MANUAL


147
147
F - Chapter 3 - 139

Short Span Trail Bridge Standard

The basic equation


The basic for the
equation for calculation of the
the calculation suspender
of the length
suspender is c/cisn c/cn
length vertical center
vertical distance
center between
distance the
between
the main cable to the spanning
main cable to the spanning cable. cable.

4(f d +cd ) 2
c/cn = ⋅ x n + 1.10 m
2
l

Where, Xn = (n -1), in meters

nmax = 0.5 ⋅ l – Fs + 1

Total Suspender Length,

ln = c/cn – 1160mm

Length of standard piece,

ls = 1650mm,

Cutting length of standard piece,

lsc = 1800 mm.

Number of standard pieces,

l n − 350
j n = INTEGER
1650
Length of extra piece,

lr = ln – 1650 jn

Cutting length of extra piece,

lrc = lr + 150 mm

Weight of suspender,

Wn = 1.116 jn + 0.62 ⋅ 10-3 lrc, kg.

Total number of suspenders,


PART F: TRAIL BRIDGE MANUAL
N = (0.5l - Fs) ⋅ 4 + 2

Total weight of the suspenders,


Cutting length of extra piece,
F - Chapter 3 - 140
lrc = lr + 150 mm

Weight of suspender,

Wn = 1.116 jn + 0.62 ⋅ 10-3 lrc, kg.

Total number of suspenders,

N = (0.5l - Fs) ⋅ 4 + 2

Total weight of the suspenders,

n = max
W = 2 W1 + 4 ⋅ ∑W n
n=2

This chapter on Structural Analysis and Design provides the background upon which the pre-calculations
have This based. Itonis Structural
been chapter Analysis and
by far not exhaustive, Design provides
but provides the background
enough information upon
to follow the which the pre-
methodologies
calculations
applied. have been based. It is by far not exhaustive, but provides enough information to follow the
methodologies applied.

3.6 Design of standard suspension bridge

3.6.1 The major components of the suspension type bridge


148

3.6.2 Design procedure

¡¡ Draw the bridge profile from the survey data.

Selection of Bridge Type:


¡¡ According to the bridge profile, assess the possibility of designing a Suspended type bridge. It is
more preferable due to the economic (cost per meter) reasons and easy construction technology,
which is more appropriate for the community bridge building approach.
¡¡ Guideline for selection of a bridge type as per the prevailing topography of the bridge site:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 141

Topography Recommended Bridge Type


The suspended bridge type (raised foundation
blocks on both banks) may be feasible only for
small spans. Most likely only the Suspension
type may be feasible. It is necessary to assess
both the types and adopt whichever is feasible
and more cost effective.

The Suspended bridge type (raised foundation


blocks on the lower bank) may be feasible only
for small spans. Most likely only the Suspension
type may be feasible. It is necessary to assess
both the types and adopt whichever is feasible
and more cost effective.

Both the Suspended and Suspension types are


feasible. It is necessary to assess both the types
and adopt whichever is more cost effective.
The Suspended type is preferable.

The Suspended type is most appropriate.


If the Suspended type is not feasible, the
Suspension type with one tower may also be
feasible. The present handbook does not deal
with this. Refer to the Volume A of LSTB (SBD)
Manual for the design. The Suspension type
with both towers is not feasible.

The Suspended type is most appropriate.

¡¡ If both the Suspended and Suspension types are feasible, make a cost (total cost) comparison and
adopt the more cost effective one.
¡¡ If the cost is nearly equal for both the bridge types, choose the Suspended type.
¡¡ Select the Suspension type only, if the Suspended type is not feasible for the given topography or
it is more economical than the Suspended type.
¡¡
Steps for Designing a Suspension Type Bridge:
¡¡ Fix the position of the Walkway and Tower foundations in the bridge profile.
¡¡ Determine the span of the bridge.
¡¡ Select the Bridge Geometry, i.e., Sag fd, Camber cd and Tower Height ht as per the span.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 142

¡¡ Select the Main Cables, Spanning Cables and Tower Type as per the span.
¡¡ Design the Walkway & Tower Foundations.
¡¡ Fix the position of the Main Cable Anchor Blocks and determine the Backstay Distances.
¡¡ Design the Main Anchor Blocks.
¡¡ Calculate the lengths of the Suspenders.
¡¡ Transfer the bridge data in the bridge profile and prepare the General Arrangement Drawing.
¡¡ Compile and fill in the Standard Steel & Construction Design Drawings.
¡¡ Calculate the quantities of works and prepare a Cost Estimate.

3.6.3 Designing the position of the walkway & tower foundations

Fix the position of the bridge foundations and the actual span of the bridge in the bridge profile. This
bridge profile will be the basis for the layout of the bridge at the construction site. Fulfill the criteria while
fixing the position of the bridge foundations.

Criteria for fixing the Bridge Foundations


¡¡ The Bridge Foundations should be placed behind the line of angle of internal friction (safe slope
line) of the soil or rock. This angle in general is 35° for soil and 60° for rock.
¡¡ The Bridge Foundations should be placed at least 3 meters back from the soil slope and 1.5 meters
back from the rock slope.
¡¡ If the bank is flat and prone to flooding or excessive bank erosion, the foundations should be
placed sufficiently back from the bank edge to ensure the long term safety of the bridge.
¡¡ The top of the Walkway and the Tower Foundation should be at the same level on both the banks.
¡¡ The height of the Walkway and the Tower Foundation should be as low as possible. However, it
should not be less than 1.0m from the ground level which gives a 0.55m clearance between the
spanning cable and the ground. For slope topography (both rock and soil), the Walkway and Tower
Foundation height H has been fixed at 1.0m.
¡¡ The freeboard Fb, between the lowest part of the bridge, i.e., the anchorage point of the spanning
cable and the highest flood level, should generally be not less than 5.0 m

Criteria for fixing the location of the Bridge Foundation

Procedure for fixing the Bridge Foundations


According to the above criteria, draw the bridge profile as per the steps.
¡¡ Draw the bridge axis profile on an A3 size paper in the scale 1:200 (for up to 50m span) or 1:
400 (for spans above 50m) with all the details like axis points A and B, HFL, WL and the tentative
location of the Walkway and Tower Foundations (if available in the survey data) on both banks.
¡¡ Fix the position of the walkway and the tower foundations on both the banks as per the procedure:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 143

Case-1: When the Tentative Position of the Walkway and the Tower Foundations has been fixed
during the Survey.
Step 1: Fix the Front of the Walkway and the Tower Foundations on one bank.
Mark the front of the foundations as fixed during the survey. Check the position of the front of the
foundation on the lower bank or the bank which has a limited defined area for positioning the foundation.

Draw the safe slope line. The front of the foundation should be behind this slope line and also sufficiently
back from the bank/slope edge. If the minimum required distance from the bank/slope edge is not
sufficient or out of the safe slope line, shift its position backward.

If the bank is flat and the bank height is low and prone to flooding or excessive bank erosion, the
foundation should be placed sufficiently back from the bank edge to ensure a long term safety of the
bridge.

Step 2: Fix the Elevation of the Spanning Cable Anchor point.


Mark the elevation line of the spanning cable anchor point (Es). Es = G.L. + 0.55m. Check the available
freeboard Fb. Fb = Es – HFL, which should not be less than 5.0m. If the freeboard is not enough rise the
Es. In case of flat topography, raise the Es = G.L. + 1.55 or 2.55 or 3.55 (max.) till sufficient freeboard
is achieved. In case of slope topography shift its position backward till sufficient freeboard is achieved.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 144

Step 3: Fix the Front of the Walkway and Tower Foundations on the other bank.
Once the elevation of the spanning cable anchor point Es has been established on one bank, draw
this elevation line towards the other bank. Draw the safe slope line. Mark the front of the foundation
according to the Es line behind the safe slope line maintaining the minimum required distance from the
bank/slope edge.

Step 4: Determine the Bridge Span, l.


Mark the center line of the towers on both the banks. The center line of the tower is 1.25m for rock and
1.75m for soil back from the front of the foundation. Calculate the bridge span which is the distance
between the center lines of the Towers. Determine the final span round to the nearest meter.

Step 5: Finalize the Bridge Profile.


Finalize the bridge profile with the final span l (span should be fixed round to the meter) and the elevation
of the spanning cable anchor points Es.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 145

Case-2: When the Position of the Walkway and the Tower Foundations has not been fixed during
the Survey.

Step 1: Fix the Front of the Walkway & Tower Foundations on one bank.

Fix the position of the front of the tower on the lower bank or the bank which has a limited defined area
for positioning the foundation.

Draw the safe slope line. The front of the foundation should be behind this slope line and also sufficiently
back from the bank/slope edge.

If the bank is flat and the bank height is low and prone to the flooding or excessive bank erosion, the
foundation should be placed sufficiently back from the bank edge to ensure the long term safety of the
bridge.

Step 2: Fix the Freeboard Level.

Mark the minimum freeboard level. The minimum freeboard from the highest flood level is 5.0 meters.
The minimum freeboard level will be the lowest possible position (Elevation, Es) of the spanning cable
anchor point.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 146

Step 3: Fix the Elevation of the Spanning Cable Anchor point.


Mark the elevation line of the spanning cable anchor point Es at the minimum freeboard level. Check the
height from the ground level (Es - G.L.), which should be 0.55 or 1.55 or 2.55 or 3.55 m (max). Adjust Es
accordingly.

Step 4: Fix the Front of the Walkway and the Tower Foundations on the other bank.
Same as Step 3 of Case-1.

Step 5: Fix the Bridge Span, l.


Same as Step 4 of Case-1.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 147

Step 6: Finalize the Bridge Profile.


Finalize the bridge profile with the final span l and the elevation of the spanning cable anchor points Es.

Short Span Trail Bridge Standard

3.7.4. CABLE DESIGN & STANDARD TOWER SELECTION


Designing the cables for a Suspension bridge involves the selection of the required number and the diameter
of the Main Cables and the Spanning Cables for the given bridge span.

3.6.4 Cable
The designtable
standard & standard tower selection
for the Selection of Cables and Standard Towers and the Bridge Geometry has been
developed for spans of up to 120m.
Designing the cables for a Suspension bridge involves the selection of the required number and the
diameter of the Main Cables and the Spanning Cables for the given bridge span.
To design the cables and to select the type of standard tower:
The standard
• table forthe
Determine thespan
Selection
of theof Cables
bridge andthe
from Standard Towers and the Bridge Geometry has been
bridge profile.
developed for spans of up to 120m.
• Select the number and diameter of the Main Cables, the diameter of the Spanning Cables and
To design the
the cables
type ofand to select
Standard the type
Tower fromofthe
standard tower:
Selection of Cables and Standard Towers according to
¡¡ Determine the span
the bridge span. of the bridge from the bridge profile.
¡¡ Select the number and diameter of the Main Cables, the diameter of the Spanning Cables and the
typeCable
Design the of Standard Tower
Structure andfrom thethe
Select Selection
Standardof Tower
Cables and Standard Towers according to the bridge
span.
I. Survey Data & Calculation of Freeboard
Design the Cable Structure and Select the Standard Tower
I. Survey Data & Calculation of Freeboard

1. Span of the Bridge (should be rounded off to the nearest meter)  = ................... m

2. Elevation of the spanning cable at anchor point Es = ................... m

3. Highest Flood Level HFL = ................... m

4. Freeboard (min. 5.00m) Fb = Es - HFL = Fb = ................... m

PART
(If the freeboard is less than F: TRAIL
5.00m, BRIDGE
try either MANUAL
to raise the elevation of the spanning cable anchor point
by increasing the height of the Walkway & Tower Foundation or to adjust the span.)
F - Chapter 3 - 148

1. Span of the Bridge (should be rounded off to the nearest meter)  = ................... m

2. Elevation of the spanning cable at anchor point Es = ................... m

3. Highest Flood Level HFL = ................... m

4. Freeboard (min. 5.00m) Fb = Es - HFL = Fb = ................... m

(If the freeboard is less than 5.00m, try either to raise the elevation of the spanning cable anchor point
by increasing the height of the Walkway & Tower Foundation or to adjust the span.)

II. Selection of Cables and Standard Towers (Pylon)


Select the Main and Spanning Cable combinations, the standard tower height ht and the Bridge Geometry
according to the span from the table. The bridge geometry refers to the main cable dead load sag fd,
the dead load camber cd of the spanning cable and the main cable inclination at the tower saddle ßf in
the full load case.

156

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 149

Table F.3.1: Selection of Cables and Standard Tower

Tower Cables Dead


Hoisting Camber Full load
Span Height/ Nos. & diameter Load
Sag Backstay
(Type) Sag
Angle
l Main Spanning Windguy fh cd ßf
m ht mm mm mm fd
m m deg.
m m
30.0 4.20 4.13 0.90 29.83
31.0 4.15 4.07 0.95 28.83
32.0 4.10 4.02 1.00 27.88
33.0 4.10 4.01 1.00 27.24
34.0 4.08 3.99 1.02 26.55
35.0 4.05 3.95 1.05 25.84
36.0 4.22 4.12 0.88 26.14
2 ø 26

2 ø 26
37.0 4.19 4.09 0.91 25.49
38.0 4.16 4.05 0.94 24.88
39.0 4.13 4.01 0.97 24.31
40.0 4.10 3.97 1.00 23.77
41.0 4.10 3.96 1.00 23.39
42.0 4.10 3.96 1.00 23.03
5.50
(1)

43.0 4.30 4.16 0.80 23.48


44.0 4.40 4.26 0.70 23.53
45.0 4.40 4.25 0.70 23.20
Not Required

46.0 3.92 3.77 1.18 20.45


47.0 3.90 3.74 1.20 20.10
48.0 4.05 3.89 1.05 20.39
49.0 4.05 3.88 1.05 20.13
50.0 4.10 3.93 1.00 20.07
51.0 4.10 3.92 1.00 19.83
52.0 4.10 3.91 1.00 19.61
53.0 4.10 3.90 1.00 19.40
54.0 4.10 3.88 1.00 19.20
2 ø 32

2 ø 32

55.0 4.25 4.04 0.85 19.46


56.0 5.47 5.30 1.48 22.82
57.0 5.44 5.26 1.51 22.46
58.0 5.41 5.22 1.54 22.12
59.0 5.38 5.18 1.57 21.79
60.0 5.35 5.13 1.60 21.48
7.35
(2)

61.0 5.35 5.12 1.60 21.26


62.0 5.45 5.22 1.50 21.32
63.0 5.65 5.42 1.30 21.65
64.0 5.85 5.63 1.10 21.97
65.0 5.95 5.72 1.00 22.01

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 150

Tower Cables Dead


Hoisting Camber Full load
Span Height/ Nos. & diameter Load
Sag Backstay
(Type) Sag
Angle
l Main Spanning Windguy fh cd ßf
m ht mm mm mm fd
m m deg.
m m
66.0 7.02 6.81 1.78 24.48
67.0 6.99 6.77 1.81 24.15
68.0 6.96 6.73 1.84 23.83
69.0 6.93 6.69 1.87 23.54
2 ø 32
70.0 7.20 6.97 1.60 23.97
9.20
(3)

Not Required
71.0 7.40 7.17 1.40 24.22
72.0 7.60 7.37 1.20 24.47
73.0 7.70 7.46 1.10 24.47
74.0 7.65 7.40 1.15 24.14
75.0 7.65 7.40 1.15 23.93
76.0 8.57 8.36 2.08 25.38
77.0 8.54 8.33 2.11 25.07
78.0 8.51 8.29 2.14 24.77
79.0 8.48 8.25 2.17 24.48
80.0 8.45 8.22 2.20 24.20
81.0 8.42 8.18 2.23 23.93
82.0 8.39 8.14 2.26 23.67
2Ø 32

83.0 8.36 8.10 2.29 23.41


84.0 8.33 8.06 2.32 23.16
4 ø 26

85.0 8.30 8.02 2.35 22.92


86.0 8.27 7.98 2.38 22.68
26 (Optional only)

87.0 8.24 7.94 2.41 22.46


11.05
(4)

88.0 8.21 7.90 2.44 22.24


89.0 8.35 8.04 2.30 22.34
90.0 8.55 8.24 2.10 22.56
91.0 8.85 8.54 1.80 22.96
92.0 9.05 8.74 1.60 23.16
93.0 9.25 8.95 1.40 23.36
94.0 9.25 8.94 1.40 23.20
∅Ø

95.0 9.20 8.87 1.45 22.96


96.0 7.98 7.68 2.67 20.00
97.0 8.05 7.75 2.60 20.00
4 ø 32

98.0 8.13 7.82 2.52 20.00


99.0 8.20 7.89 2.45 20.00
100.0 8.28 7.96 2.37 20.00

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 151

Tower Cables Dead


Hoisting Camber Full load
Span Height/ Nos. & diameter Load
Sag Backstay
(Type) Sag
Angle
l Main Spanning Windguy fh cd ßf
m ht mm mm mm fd
m m deg.
m m
101.0 8.35 8.03 2.30 20.00
102.0 8.40 8.07 2.25 20.00
103.0 8.50 8.17 2.15 20.00
104.0 8.57 8.23 2.08 20.00
105.0 8.65 8.31 2.00 20.00
106.0 8.71 8.36 1.94 20.00

ø 26 (Optional only)
107.0 8.80 8.45 1.85 20.00
108.0 8.85 8.49 1.80 20.00
109.0 8.94 8.58 1.71 20.00
4 ø 32

2 ø 32
11.05

110.0 9.00 8.63 1.65 20.00


111.0 9.00 8.62 1.65 19.88
112.0 8.95 8.55 1.70 19.70
113.0 8.95 8.54 1.70 19.60
114.0 8.95 8.53 1.70 19.50
115.0 8.90 8.46 1.75 19.34
116.0 8.90 8.45 1.75 19.24
117.0 8.85 8.38 1.80 19.09
118.0 8.85 8.37 1.80 19.00
119.0 8.85 8.36 1.80 18.92
120.0 8.85 8.34 1.80 18.83

III. Fixing the Back Stay Lengths (DL and DR) and the Main Cable Anchorage Elevations
(EL and ER)

Fix the position of the Main Anchorage Foundation Blocks on both the banks in the bridge profile:
¡¡ Draw the centerline of the towers on both the banks in the bridge profile and mark the tower top
(saddle) with the elevation Et = Es + 0.70 + ht.
¡¡ From the tower top, draw a line at the angle ßf towards the ground level. Fix the position of
the main cable anchor point along this line by considering the topography and the geological
condition of the bridge site.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 152

In flat topography (up to ground slope angle 15°):

In slope topography:

Measure the DL or DR backstay distance with a scale in the bridge profile.


¡¡ Calculate the Main Cable Anchor Elevation (ER and EL) as per the formula.
ER = Et – DR x tan ßf, EL = Et – DL x tan ßf

IV. Calculation of Cable Lengths

dia Backstay Length [m] Cutting Length*


Type of Cable Nos.
(mm) DL DR [m/pc]

Main Cable ……… ……… ……… ………


Spanning Cable ……… 2 0 0
Handrail Cable 13 2 0 0
Fixation Cable 13 2 0 0
Windguy 1 (U/S) 0 0
Cable** 26
(Optional) 1 (D/S) 0 0

Windties**
13 1 0 0
(Optional)

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 153

* Cutting Length:
1.08 x Span + 1.15 (DL+DR) + Anchorage Length on the Right Bank
Main Cable =
+ Anchorage Length at Left Bank.
Calculate the Anchorage Lengths only after selecting of the Main Cable Anchor Blocks.
Spanning Cable = 1.03 x Span + 3.5 m.
Handrail Cable = Span
Fixation Cable = Span
** Windguy Cable = Refer to Design of Wind guy Arrangement
** Windties = Refer to Design of Windguy Arrangement

V. Calculation of (fh) Hoisting Sag


This calculation has to be made after the tower erection is completed.

1. Actual Span measured in the field l =.................. m


Hoisting Sag (from “Table: Selection of Cables and Standard Tower” as
2. fh =.................. m
per actual span)
3. Tower Height ht =.................. m
4. Marking Point on Tower (from steel tower base) at hoisting sag elevation = ht – fh =.................. m

3.6.5 Design of the main cable anchor

The standard designs for the Main Cable Anchor Blocks have been developed for all possible cases for
spans of up to 120m. The Standard Anchor Blocks have been developed by the design concept and
statical analysis.

The design of the Main Cable Anchor Blocks is mainly to select the standard anchor block types suitable
to the given cable combination, geology and topography of the site. Fill in the required data in the
corresponding selected drawings.

There are basically six typical designs of the Main Anchor Blocks, depending on the flat or hill slope
topography, and the soil or rock type.

Figure F.3.5: Main Cable Deadman Anchor Block with Turnbuckle on Flat Topography

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 154

Figure F.3.6: Main cable anchor block with direct cable connection on flat topography

Figure F.3.7: Main cable gravity anchor block with turnbuckle on hill slope topography

Figure F.3.8: Main cable gravity anchor block with direct cable connection on hill slope

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 155

Figure F.3.9: Main cable drum anchor block in hard rock

Figure F.3.10: Main cable drum anchor block in fractured hard rock or soft rock

To select the main anchor block type:


¡¡ Define the number and the diameter of the main cables.
¡¡ Define the topography of the ground where the anchor block will be placed as whether flat or
slope. The topography is defined as flat if the ground slope is less than 15°, and slope if the ground
slope is more than 15°.
¡¡ Define the soil or rock type from Survey Form and Checklist
¡¡ Select the anchor block type and the corresponding drawing from the selection tables according
to the above design data.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 156

Design the Main Anchor Blocks

(1). Design Data

ƒƒ Main Cables: Nos. …..…Ø………..mm

Right Bank Condition


Geology: Soil

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 15° slope) (more than 15° slope)

What is the Soil Type? Gravelly Sandy Silty


If Rock, what is the Rock Hard Rock Hard Rock Soft Rock
Type? (only a few fractures) (highly fractured)
Left Bank Condition
Geology: Soil

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 15° slope) (more than 15° slope)

What is the Soil Type? Gravelly Sandy Silty


If Rock, what is the Rock Hard Rock Hard Rock Soft Rock
Type? (only a few fractures) (highly fractured)

(2). Selection of Anchor Types


Select the appropriate Anchor Types for the Right and Left Banks according to the above design data.
Except for the Drum Rock Anchors, all Main Cable Anchors are either designed with Turnbuckle or for
Direct Cable Connection. One Main Cable Anchor should always have a Turnbuckle for fine adjustment
during sag setting, whereas for economic reasons, the cable anchor on the other bank should always have
a direct cable connection.

Rules for selecting the anchor blocks:


¡¡ Select an Anchor Block with a turnbuckle for one bank and an Anchor Block with a direct cable
connection for the other bank. Always choose the more convenient bank for the Main Cable
Anchor with turnbuckle.
¡¡ If one bank is soil and the other bank is rock, always select a Drum Anchor for the rocky bank.
¡¡ If both banks are rocky, select an Anchor Block for one bank and a Drum Anchor for the other bank.

Procedure for selection:


¡¡ According to the Soil type and Slope of the ground or the Rock type, refer to the respective tables
for selection of the Anchor Types.
for Soil and Flat Ground (Cable Connection with Turnbuckle) : Table F.3.2
for Soil and Flat Ground (Direct Cable Connection) : Table F.3.3
for Soil and Hill Slope (Cable Connection with Turnbuckle) : Table F.3.4
for Soil and Hill Slope (Direct Cable Connection) : Table F.3.5
for Hard, Soft or Fractured Rock (Cable Connection with Turnbuckle) : Table F.3.6
for Hard Rock (Direct Cable Connection ) : Table F.3.7
for Fractured Hard Rock or Soft Rock (Direct Cable Connection) : Table F.3.8

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 157

¡¡ Match the design data in the table:


The number and diameter of the Main Cables  if Soil, Soil type  select Anchor type and
corresponding. Drawing No. for the right bank and the left bank respectively .
¡¡ Fill in the required data in the selected drawings.

(3). Anchor Type Selection Tables


¡¡ In Soil and Flat Ground

Table F.3.2: Selection of the main cable deaman anchor in soil & flat ground
(Cable connection with turnbuckle)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All DA 1 42/1Ncon
2x32 All DA 2 42/3Ncon
4x26 All DA 3 44/1Ncon
4x32 All DA 4 44/3Ncon

Table F.3.3: Selection of the main cable deadman anchor in soil & flat ground
(Direct cable connection)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All DA 5 42/2Ncon
2x32 All DA 6 42/4Ncon
4x26 All DA 7 44/2Ncon
4x32 All DA 8 44/4Ncon
¡¡ In Soil and Hill Slope

Table F.3.4: Selection of the main cable anchor block in soil & hill slope
(Cable connection with turnbuckle)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All AB 1 42/5Ncon
2x32 All AB 2 42/7Ncon
4x26 All AB 3 44/5Ncon
Gravelly or Sandy AB 4 44/7Ncon
4x32
Silty AB 5 44/9Ncon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 158

Table F.3.5: Selection of the main cable anchor block in soil & hill slope
(Direct cable connection)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All AB - 6 42/6Ncon
2x32 All AB - 7 42/8Ncon
4x26 All AB - 8 44/6Ncon
Gravelly or Sandy AB - 9 44/8Ncon
4x32
Silty AB - 10 44/10Ncon
¡¡ In Hard, Soft or Fractured Rock

Table F.3.6: Selection of the Main Cable Anchor Block in Hard, Soft or Fractured Rock
(Cable Connection with Turn Buckle)

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 AB - 11 45Ncon
4x26 or 4x32 AB - 12 46Ncon
¡¡ In Hard Rock

Table F.3.7: Selection of the Main Cable Drum Anchor in Hard Rock
(Direct Cable Connection)

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 DR-1 47Ncon
4x26 or 4x32 DR-2 48Ncon
¡¡ In Fractured Hard Rock or Soft Rock

Table F.3.8: Selection of the Main Cable Drum Anchor in Fractured Hard Rock or Soft Rock
(Direct Cable Connection)

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 DR-3 49Ncon
4x26 or 4x32 DR-4 50Ncon

3.6.6 Walkway and tower foundation design

The standard designs of the Walkway and Tower Foundations have been developed for all possible cases
for spans of up to 120m. The standard foundation designs have been developed by the design concept
and statical analysis.

The design of the Walkway and the Tower Foundations is mainly to select the standard foundation types
as per the cable combinations, geology and topography of the site. Fill in the required data in the
corresponding selected drawings.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 159

Illustration of a typical design:

A. Design Data

ƒƒ Windguy Cable: …..nos. Ø…….mm

Right Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

Right Bank / Downstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

B. Selection of Walkway & Tower Foundations


Select the appropriate Walkway & Tower Foundations for the Right and Left Banks according to the above
design data.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 160

Procedure for selection:


¡¡ According to the Soil or Rock type, refer to the respective table for selection of the Walkway and
Tower Foundations.

In Soil : Table F.3.9


In Hard Rock : Table F.3.10
In Fractured Hard Rock or Soft Rock : Table F.3.11

¡¡ If the bank is rock, always take the foundation block height H as 1.0m,
If the bank is soil and in hill slope, always take the foundation block height H as 1.0m.
If the bank is soil on flat ground, but the foundation block height H as needs to be raised, always
take the foundation block height as 1.0 or 2.0 or 3.0 or 4.0m (maximum). H is limited to a maximum
of 4.0m.
¡¡ Match the design data in the table: Number and Diameter of Main Cables  Foundation Block
Height  if Soil, Soil type  select Foundation type and corresponding Drawing No. for right bank
and left bank respectively.
¡¡ Fill in the required data in the selected drawings.

C. Walkway & Tower Foundations Selection Tables

¡¡ In Soil
Table F.3.9: Selection of walkway & tower foundations in soil

Foundation
Block Height
Main Cable Soil Type Foundation Type Drawing No.
(H)
m
1.0 All TF-1 92/1Ncon
2.0 All TF-2 92/2Ncon
2x26 and 2x32mm
3.0 All TF-3 92/3Ncon
4.0 All TF-4 92/4Ncon
1.0 All TF-6 94/1Ncon
2.0 All TF-7 94/2Ncon
Gravelly/Sandy TF-8 94/3Ncon
4x26mm 3.0
Silty TF-9 94/4Ncon
Gravelly/Sandy TF-10 94/5Ncon
4.0
Silty TF-11 94/6Ncon
1.0 All TF-6 94/1Ncon
2.0 All TF-7 94/2Ncon
Gravelly/Sandy TF-8 94/3Ncon
4x32mm 3.0
Silty TF-9 94/4Ncon
Gravelly/Sandy TF-10 94/5Ncon
4.0
Silty TF-11 94/6Ncon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 161

¡¡ In Hard Rock

Table F.3.10: Selection of walkway & tower foundations in hard rock

Foundation Block
Main Cable Height (H) Foundation Type Drawing No.
m
2x26 or 2x32mm 1.0 TF-5 92/5Ncon
4x26 or 4x32mm 1.0 TF-12 94/7Ncon

¡¡ In Fractured Hard Rock or Soft Rock

Table F.3.11: Selection of walkway & tower foundations in fractured hard rock or soft

Foundation Block
Main Cable Height (H) Foundation Type Drawing No.
m
2x26 or 2x32mm 1.0 TF-1 92/1Ncon
4x26 or 4x32mm 1.0 TF-6 94/1Ncon

3.6.7 Suspender list

Suspenders are designed for the bridge geometry under dead load, i.e. l, fd and cd . The layout of the
suspenders and its geometrical calculation is based on the design calculations done in this manual, as per
the said suspender layout and the geometrical calculation for suspender lengths, a computer program
“Suspender Design” has been developed.

(a). Calculation of Suspender List


Calculate the Suspender List using the computer program “SSTB Design” which is found in another
folder.

The calculation is based on the fixed geometry of the bridge as per the table: Selection of Cables and
Standard Towers. The only input data required for the calculation is the span, l.

(b). Procedure for the Suspender List calculation


¡¡ Open the computer program SSTB Design > “Suspender Design”
¡¡ Enter the input data: Bridge Name, Bridge Span and width of Walkway and Tower Foundation
¡¡ Print result.

Attach the printout of the Suspender List with the drawing of the suspenders.

3.6.8 List of drawings

Select the required Steel Drawings and Construction Drawings according to the selected Main Cables,
Towers, Main Cable Anchorage Blocks and Walkway / Tower foundations.

In addition, prepare a General Arrangement drawing for individual bridge designs

3.6.9 Other structures

Besides the bridge structure, some other adjacent structures may be required for overall bridge stability
and for safety measures. These are the structures.
(1). Windguy Arrangement
(2). Retaining Structures
(3). Slope Protection works

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 162

(4). River Bank Protection

(1). Windguy Arrangement (Optional)


The design maximum wind speed is calculated assuming 160km/h, and has favourable nature of the wind
flow. Theoretical calculations show that there is no need for a windguy arrangement for Suspension type
bridges with spans of up to 120m. Therefore, no Windguy Arrangement is necessary in this standard
suspension bridge. All the towers and foundation structures are designed to withstand the maximum
design wind pressure without a windguy arrangement. However, during high wind time (the frequency
and duration of such times are very low) the comfort while crossing the bridge will be significantly reduced
and it may even be risky. Therefore, no traffic should be allowed to cross the bridge during such windy
times.

For special cases, i.e., when the bridge span is more than 120m or in extremely windy areas with wind
speeds in excess of 160km/hr and unfavourable nature of the wind flow, a windguy system is mandatory.

¡¡ Design of Windguy Arrangement


The design of the Windguy Arrangement is to
• Select the diameter of the windguy cable and windties cable;
• Calculate the geometry of the windguy arrangement;
• Select the windguy cable anchor blocks.

A general layout of the Windguy Arrangement.

Figure F.3.11: Layout of windguy arrangement

A computer program “Windguy Design” has been developed for the calculation. The calculation is
based on the layout of the Windguy Arrangement as per the above and the fixed geometry of the bridge
in dead load case as per the Table: Selection of Cables and Standard Towers section of this manual..

Calculate the Windguy Arrangement using the computer program “SSTB Design” which is found in a
separate folder of this Manual.

¡¡ Procedure for Windguy Arrangement Design


• Open computer program SSTB Design > “Windguy Design”.
• Give input data: Bridge Name, Bridge Span and Elevation of Steel Tower Base.

The calculation procedure:


¡¡ Give the tentative of the first windtie position BL and BR as per the topography. The program will
adjust to the nearest actual values.
¡¡ Give the level of the windguy anchorage on the right bank HL and the left bank HR. HL and HR
should be at or below the level of the spanning cable at the anchorage point (Es).

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 163

¡¡ Give (select) the windguy anchor position on the right bank (or the left bank), i.e. CR and DR or (CL
and DL). Accordingly, the windties listing will be from the right bank to the left bank or vise versa.
¡¡ Input the value of VR (distance up to the vertex of the windguy cable) if the selected windguy
anchor position is on the right bank or VL if the selected windguy anchor position is on the left
bank.
¡¡ Check the span and sag ratio of the windguy cable, l/b. It should be not less than 8 and not more
than 10. If this condition is not met, input a new V till the l/b is within the recommended limit.
¡¡ Check the lowest point of the windguy cable “Δhlp”. It should not be negative (-). If the Δhlp is
negative, input new HL and HR with a lower value.
¡¡ Input value of DLo (or DRo) = 0. The result will show CLo (or CRo), and horizontal angle αL (or αR).
¡¡ Draw a straight line with the given CLo (or CRo), and horizontal angle αL (or αR). Fix the windguy
anchor position on the left bank (or the right bank) along this line. The anchorage block should be
placed on safe ground and at the optimum foundation location so that it has sufficient embedded
depth and also so that deep excavation can be avoided.
¡¡ After fixing the anchorage position, measure the actual CL (or CR ). The result will give the DL (or
DR).
¡¡ Draw longitudinal sections along the windguy cables at the foundations and determine the exact
position of the front of the foundations (DR, CR and DL, CL) and the windguy cable elevations (HR,
HL).
¡¡ If the final position of the foundations does not match the design as above, repeat the calculation
process from the second bullet till all the foundations are located at the optimum positions.
¡¡ Print out the results. The results will give:
• Position of the anchorage blocks,
• Diameter of the windguy cable,
• c/c inclined distance (length) of the windties,
• Cutting length of the windties,
• Horizontal c/c distance of the windties and
• Inclined distance Dw (interval between the windties).
¡¡ Calculate the Windguy Arrangement for both upstream and downstream.

Attach the printout of the results with the drawing of the Windguy Clamp and also transfer it to the
General Arrangement Drawing.

¡¡ Windguy anchorage foundation blocks design (optional)


The design concept and the statical analysis are the same as for the main anchor blocks. Accordingly, the
standard anchor blocks have been developed. Basically, there are four typical standard anchor blocks.

Figure F.3.12: Gravity anchor block in soil

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 164

Figure F.3.13: Gravity anchor in rock

Figure F.3.14: Drum anchor block in hard rock

Figure F.3.15: Drum Anchor Block in Fractured Hard Rock & Soft Rock

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 165

¡¡ Design Data
Fill in the Design Data from Survey Form and Checklist.

ƒƒ Windguy Cable: …..nos. Ø…….mm

Right Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

Right Bank / Downstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 166

ƒƒ Windguy Cable: …..nos. Ø…….mm

Left Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

Left Bank / Downstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Block Height from the Ground (data
1.0m 2.0m 3.0m 4.0m
from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Block Height from the Ground (data
from bridge profile): 1.0m in case of Rock

¡¡ Selection of Standard Windguy Cable Anchor Blocks


Select appropriate Windguy Cable Anchor Blocks for U/S and D/S for the Right and Left Banks
according to the above design data. are the rules for selection of the anchor blocks.
• If the bank is rock, always take the block height as 1.0m.
• If the bank is soil on a hill slope, always take the block height as 1.0m.
• If one bank is soil and the other bank is rock, always select a Drum Anchor for the rocky bank.
• If both banks are rocky, select an Anchor Block for one bank and a Drum Anchor for the other
bank.
• If the bank is soil but on flat ground and the anchor block height needs to be raised, always
take the foundation block height as 1.0, or 2.0, or 3.0, or 4.0 m (maximum).

Procedure for selection:


According to the Soil or Rock type, refer to respective table for selection of Windguy Cable Anchor
Blocks.

In Soil : Table F.3.12


In all types of Rock (Anchor Block) : Table F.3.13
In Hard Rock (Drum Anchor) : Table F.3.14
In Fractured Hard Rock or Soft Rock : Table F.3.15
(Drum Anchor)

In the tables match the design data: Diameter of Windguy Cable  Soil or Rock type  Block Height
 select Anchor Block type and corresponding Drawing No.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 167

¡¡ Fill in the required data in the selected drawings.

Selection Tables for Windguy Anchor Blocks

¡¡ In Soil

Table F.3.12: Selection of windguy cable anchor block in soil

Windguy Cable Anchor Block


Soil Type Anchor Type Drawing No.
[mm] Height
1 All WAB - 1 55/1Acon
2 All WAB – 2 55/2Acon
26
3 All WAB – 3 55/3Acon
4 All WAB – 4 55/4Acon
¡¡ In Hard, Fractured or Soft Rock

Table F.3.13: Selection of windguy cable anchor block in all types of rock

Windguy Cable
Anchor Type Drawing No.
[mm]
26 WAB - 5 56Acon
¡¡ In Hard Rock

Table F.3.14: Selection of windguy cable drum anchor in hard rock

Windguy Cable
Anchor Type Drawing No.
[mm]
26 WDR - 1 57Acon
¡¡ In Fractured or Soft Rock

Table F.3.15: Selection of windguy cable drum anchor in fractured or soft rock

Windguy Cable
Anchor Type Drawing No.
[mm]
26 WDR - 2 58Acon

(2). Retaining structures


Retaining structures are necessary to retain the earth (soil, fractured rock and weathered soft rock) behind
the anchorage blocks of the bridge. There are many options for retaining structures. But for trail bridges,
the most feasible are retaining walls. Retaining walls can be of gabion boxes, rubble masonry and dry
stone masonry. For the short-span trail bridges, dry stone retaining walls or breast walls are preferable
since they require only local materials.

The choice between retaining wall and breast wall depends on different factors, such as the available
space behind the blocks, the required height of the protection, soil conditions etc.

Retaining walls are used when the earth to be retained is loose soil with a great protection height. For the
design of the retaining wall, use the table.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 168

Retaining Walls

Banded Dry
Type Dry Stone
Stone / Masonry

Section

Top width, Wt 0.6 - 1.0 m 0.6 - 1.0 m


Base width, Wb 0.5 - 0.7 H 0.6 - 0.65 H
Front batter varies varies
Back batter varies vertical
Inward dip of foundation, n 1:3 1:3
Foundation depth ≥ 0.50 m ≥ 0.50 - 1 m
Range of height, H 1-6m 6-8m
Hill slope angle, < 35º 20º

Breast walls are used when the earth to be retained is fractured or weathered rock, or compact soil with
temporarily unstable nature. For designing breast walls, use the table.

Breast Walls

Banded Dry
Type Dry Stone
Stone / Masonry

Section

Top width, Wt 0.50 m 0.5 m


Base width, Wb 0.29H 0.3H 0.33H > 0.5H
Back batter 3:1 4:1 5:1 3:1
Inward dip of foundation, n 1:3 1:4 1:5 1:3
Foundation depth ≥ 0.50 m ≥ 0.50 m ≥ 0.50 m ≥ 0.50 m
Range of height, H <6m <4m <3m 3-8m
Hill stope angle, a 35 - 60º 35 - 60º

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 169

(3). Slope Protection Measures


Slope protection measures depend on the factors influencing slope instability. It is recommended that
a bridge site be selected, where there are no slope instability features. However, it is often necessary to
drain out the surface runoff and seepage water from the slope as a slope protection measure.

Water should be collected as close as possible to its origin and safely channeled to a nearby watercourse.
The surface drainage can be catch drain on the slope or drainage around the anchorage foundation or a
combination of both.

The choice depends on the position of the anchorage foundation and the profile of the natural terrain.

The drain should be open type. The design shows the typical sections of the drain.

To avoid self-scouring, the drain outlet should be protected.

In seepage area, sub-surface drainage is required around the anchorage foundation. Consider a typical
layout design of the sub-surface drain.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 170

A typical design of the sub-surface drain.

(4). Bio-Engineering
Surface drainage alone may not be sufficient to protect unstable slopes. The most effective method for
stabilizing such slopes is bio-engineering in combination with light civil structures such as catch drains,
check dams, cascades, etc. This is a cheap and easy method. The main concept of this method is to
grow trees and plants such as shrubs or grasses. Deep rooted and fast growing trees and plants are most
suitable for this purpose. The proper selection of plant types is most important, and it should be based
upon local experience. Some of the vegetation measures are:
¡¡ Planted grass lines: contour/horizontal or down slope/vertical or random planting;
¡¡ Grass seeding;
¡¡ Turfing;
¡¡ Shrub and tree planting;
¡¡ Shrub and tree seeding;
¡¡ Fascines (bundles of live branches are laid in shallow trenches).

(5). River Bank Protection


River protection works are of a temporary nature and costly. They require frequent maintenance to keep
the structures functional. Therefore, avoid bridge sites which require river protection works as far as
possible.

3.6.10 General arrangement drawing

General Arrangement (GA) drawing shows the overall plan and profile of the bridge. The GA should
reflect the major components of the bridge and its geometry for dead load case including the elevations
of the tower tops, cable anchor points, the foundations and general dimensions. The GA is required for
an overall view of the designed bridge and also for layout of the bridge before construction.

Draw the GA on the same bridge profile as already done in this manual and mention the data on the
plan and the profile of the bridge:
¡¡ Span l, tower height ht dead load sag fd, dead load camber cd, and backstay angle βf
¡¡ Backstay distance DR and DL

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 171

¡¡ Distance from the axis points A and B (which were fixed during the survey) and the center of the
towers;
¡¡ Main cable elevations at the tower saddles (top of the tower), elevation of the lowest point of the
main cable;
¡¡ Spanning cable elevation at the anchor point;
¡¡ All bridge foundation levels;
¡¡ The LWL, HFL and available freeboard Fb;
¡¡ Overall dimensions of the bridge structures and their elevations.

The completed GA should be sufficient for the layout of the bridge. An example of the GA drawing of
a bridge.

3.7 Design example

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 172

3.8 Bridge standard drawing

3.8.1 Introduction & overview of drawings

The Bridge Standard Drawings represent the centerpiece of the Short Trail Bridge Standard. They are
composed as a unit component system and are categorized into two groups: Construction Drawings
and Steel Drawings

Both drawing groups are linked with each other and depending on the bridge design the required
drawings are to be selected.

Legend for the Drawing Numbers and Suffixes:

Drawing No. Suffix Bridge or Drawing Type

STEEL 08 A For All Bridge Types


DRAWINGS 07 N For Suspension Bridge Types

56 Ncon construction drawings


Anchor Types for Main Cables
CONSTRUCTION

DA 5 Deadman Anchor Type 5


DRAWINGS

AB 7 Anchor Block Type 7


DR 3 Drum Rock Type 3
Anchor Types for Windguy Cables
WAB 2 Windguy Anchor Block Type 2
WDR 1 Windguy Drum Rock Anchor Type 1

Walkway and Tower Foundations


TF 11 Tower Foundation Type 11

3.8.2 Construction drawings

Construction drawings for suspension bridges consist of three different drawing groups.

These are:
¡¡ Main Cable Anchor Drawings
¡¡ Walkway and Tower Foundation Drawings
¡¡ Windguy Cable Anchor Drawings

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 173

Table F.3.16: Construction Drawings for Main Cable Anchors

(in Different Soil & Rock Conditions)

Main Cables
Anchor Turnbuckle or Direct Drawing
Ground Condition c/c1 Ø
Type Nos Cable Connection No.
[m] [mm]
With Turnbuckle 42/1Ncon

DEADMAN ANCHORS
26
Direct Cable Connection 42/2Ncon
2 2.50
FLAT GROUND

"DA" TYPES
With Turnbuckle 42/3Ncon
32
Direct Cable Connection 42/4Ncon
Gravelly, Sandy or Silty
With Turnbuckle 44/1Ncon
26
All Soil Types:

Direct Cable Connection 44/2Ncon


4 3.50
With Turnbuckle 44/3Ncon
32
Direct Cable Connection 44/4Ncon
SOIL

With Turnbuckle 42/5Ncon


2 2.50
Direct Cable Connection 42/6Ncon
26
With Turnbuckle 44/5Ncon
4 3.50
Direct Cable Connection 44/6Ncon
GRAVITY BLOCK ANCHORS
HILL SLOPE

With Turnbuckle 42/7Ncon


2 2.50
Direct Cable Connection 42/8Ncon
"AB" TYPES

Gravelly 32
With Turnbuckle 44/7Ncon
or Sandy 4 3.50
Direct Cable Connection 44/8Ncon
Soil

Silty With Turnbuckle 44/9Ncon


4 3.50 32
Soil Direct Cable Connection 44/10Ncon

26 or
2 2.50 With Turnbuckle 45Ncon
Soft or 32
Fractured Rock 26 or
4 3.50 With Turnbuckle 46Ncon
32
26 or
DRUM ROCK ANCHORS

2 2.50 Direct Cable Connection 47Ncon


ROCK

Hard 32
"DR" TYPES

Rock 26 or
4 3.50 Direct Cable Connection 48Ncon
32
26 or
Soft or 2 2.50 Direct Cable Connection 49Ncon
32
Fractured
Rock 26 or
4 3.50 Direct Cable Connection 50Ncon
32

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 174

Table F.3.17: Construction drawings for walkway and tower foundations

(in Different Soil and Rock Conditions)

Main Cables Block


Ground Foundation Drawing
Height
Condition Nos Ø [mm] Type No.
[m]

2 1.0 TF 1 92/1Ncon

Gravelly, Sandy or
All Soil Types: 2 2.0 TF 2 92/2Ncon

Silty 2 3.0 TF 3 92/3Ncon


2 4.0 TF 4 92/4Ncon
SOIL

4 1.0 TF 6 94/1Ncon

26 or 32 mm
4 2.0 TF 7 94/2Ncon

Gravelly or 4 3.0 TF 8 94/3Ncon


Sandy 4 4.0 TF 10 94/5Ncon
4 3.0 TF 9 94/4Ncon
Silty
4 4.0 TF 11 94/6Ncon
Soft or 2 TF 1 92/1Ncon
Fractured
ROCK

Rock 4 TF 6 94/1Ncon
1.0
2 TF 5 92/5Ncon
Hard Rock
4 TF 12 94/7Ncon

Table F.3.17: Construction Drawings for Windguy Cable Anchors

(in Different Soil and Rock Conditions)

Windguy Cable Block Height Foundation


Ground Condition Drawing No.
Ø [mm] [m] Type
1.0 WAB 1 55/1Ncon
Gravelly, Sandy
All Soil Types:

One Cable Ø 26 mm

2.0 WAB 2 55/2Ncon


or Silty
SOIL

3.0 WAB 3 55/3Ncon

4.0 WAB 4 55/4Ncon

All Rock Types 1.0 WAB 5 56Ncon


Hard Rock WDR 1 57Ncon
ROCK

Fractured or Drum Type


WDR 2 58Ncon
Soft Rock

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 175

Table F.3.19: Summary of All Construction Drawings

Drawing
Drawing Titles
Nos.
FITTING
Walkway Fitting for Suspension Bridge for 106 cm Walkway Width 19Ncon
DETAILS
DEADMAN ANCHORS
IN FLAT GROUND 2 Cables Ø 26mm, c/c1 = 2.50m, with Turn Buckle, all Soil Types 42/1Ncon
2 Cables Ø 32mm, c/c1 = 2.50m, with Turn Buckle, all Soil Types 42/3Ncon
4 Cables Ø 26mm, c/c1 = 3.50m, with Turn Buckle, all Soil Types 44/1Ncon
4 Cables Ø 32mm, c/c1 = 3.50m, with Turn Buckle, all Soil Types 44/3Ncon
2 Cables Ø 26mm, c/c1 = 2.50m, Direct Cable Connection, all Soil Types 42/2Ncon
2 Cables Ø 32mm, c/c1 = 2.50m, Direct Cable Connection, all Soil Types 42/4Ncon
4 Cables Ø 26mm, c/c1 = 3.50m, Direct Cable Connection, all Soil Types 44/2Ncon
4 Cables Ø 32mm, c/c1 = 3.50m, Direct Cable Connection, all Soil Types 44/4Ncon
MAIN CABLE ANCHORS

2 Cables Ø 26mm, c/c1 = 2.50m, with Turn Buckle, all Soil Types 42/5Ncon
GRAVITY BLOCK ANCHORS

2 Cables Ø 2mm, c/c1 = 2.50m, with Turn Buckle, all Soil Types 42/7Ncon
4 Cables Ø 26mm, c/c1 = 3.50m, with Turn Buckle, all Soil Types 44/5Ncon
IN HILL SLOPES

4 Cables Ø 32mm, c/c1 = 3.50m, with Turn Buckle, gravelly or sandy Soil 44/7Ncon
4 Cables Ø 32mm, c/c1 = 3.50m, with Turn Buckle, silty Soil 44/9Ncon
2 Cables Ø 26mm, c/c1 = 2.50m, Direct Connection, all Soil Types 42/6Ncon
2 Cables Ø 32mm, c/c1 = 2.50m, Direct Connection, all Soil Types 42/8Ncon
4 Cables Ø 26mm, c/c1 = 3.50m, Direct Connection, all Soil Types 44/6Ncon
4 Cables Ø 32mm, c/c1 = 3.50m, Direct Connection, gravelly or sandy Soil 44/8Ncon
4 Cables Ø 32mm, c/c1 = 3.50m, Direct Cable Connection, silty Soil 44/10Ncon
2 Cables Ø 26 or 32mm, c/c1 = 2.50m, with Turn Buckle, all Rock Types 45Ncon
4 Cables Ø 26 or 32mm, c/c1 = 3.50m, with Turn Buckle, all Rock Types 46Ncon
DRUM ROCK
ANCHORS

2 Cables Ø 26 or 32mm, c/c1 = 2.50m, Direct Connection, in hard Rock 47Ncon


4 Cables Ø 26 or 32mm, c/c1 = 3.50m, Direct Connection, in hard Rock 48Ncon
2 Cables Ø 26 or 32mm, c/c1 = 2.50m, Direct Connection, fractured/soft Rock 49Ncon
4 Cables Ø 26 or 32mm, c/c1 = 3.50m, Direct Connection, fractured/soft Rock 50Ncon

2 Main Cables c/c1 = 2.50m, Block Height 1.0m, all Soil Types & soft Rock 92/1Ncon
WWALKWAY AND TOWER

2 Main Cables c/c1 = 2.50m, Block Height 2.0m, all Soil Types 92/2Ncon
2 Main Cables c/c1 = 2.50m, Block Height 3.0m, all Soil Types 92/3Ncon
FOUNDATIONS

2 Main Cables c/c1 = 2.50m, Block Height 4.0m, all Soil Types 92/4Ncon
2 Main Cables c/c1 = 2.50m, Block Height 1.0m, in hard Rock 92/5Ncon
4 Main Cables c/c1 = 3.50m, Block Height 1.0m, all Soil Types & soft Rock 94/1Ncon
4 Main Cables c/c1 = 3.50m, Block Height 2.0m, all Soil Types 94/2Ncon
4 Main Cables c/c1 = 3.50m, Block Height 3.0m, in gravelly or sandy Soil 94/3Ncon
4 Main Cables c/c1 = 3.50m, Block Height 3.0m, in silty Soil 94/4Ncon
4 Main Cables c/c1 = 3.50m, Block Height 4.0m, in gravelly or sandy Soil 94/5Ncon
4 Main Cables c/c1 = 3.50m, Block Height 4.0m, in silty Soil 94/6Ncon
4 Main Cables c/c1 = 3.50m, Block Height 1.0m, in hard Rock 94/7Ncon
1 Cable Ø 26mm, Block Height 1.0m, for all Soil Types 55/1Acon
1 Cable Ø 26mm, Block Height 2.0m, for all Soil Types 55/2Acon
OPTIONAL
ANCHORS
WINDGUY
CA BLE

1 Cable Ø 26mm, Block Height 3.0m, for all Soil Types 55/3Acon
1 Cable Ø 26mm, Block Height 4.0m, for all Soil Types 55/4Acon
1 Cable Ø 26mm, Block Height 1.0m, for all Rock Types 56Acon
1 Cable Ø 26mm, Drum Type in hard Rock 57Acon
1 Cable Ø 26mm, Drum Type in fractured or soft Rock 58Acon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 176

IV. Steel Drawings

Table F.3.20: Summary of All Steel Drawings

Drawing
Drawing Titles
Nos.

WALKWAY
Crossbeam for Walkway Width = 106 cm 07N
CROSS BEAM

Steel deck Standard Panel, length = 198 cm / width = 34 cm 08A


STEEL DECK
Steel deck Special Panel, length = 223 cm / width = 34 cm 10A

WINDGUY Windties Cable Clamps for Windguys Cable ø 26 or 32 mm 11A


PARTS Windguys Cable Anchorage for one Cable End ø 26 or 32 mm 50A

Suspenders for 2 Main Cables ø 26 or 32mm 31N


SUSPENDERS
Suspenders for 4 Main Cables ø 26 or 32mm 32N

MAIN CABLE Main Cable Anchorage for 2 Main Cables ø 26 or 32mm 42N
ANCHORS Main Cable Anchorage for 4 Main Cables ø 26 or 32mm 44N

WALKWAY
Walkway and Tower Foundation, c/c1 = 2.50m, c/c2 = 383mm 92N
& TOWER
Walkway and Tower Foundation, c/c1 = 3.50m, c/c2 = 383mm 94N
FOUNDATION

Guide for selecting Towers, Main Cables and Spanning Cables 99N
Assembly and Layout for Tower No. 1, Height 5.50 m 141N
Assembly and Layout for Tower No. 2, Height 7.35 m 142N
Assembly and Layout for Tower No. 3, Height 9.20 m 143N
Assembly and Layout for Tower No. 4, Height 11.05 m 144N
TOWERS Base Element for Tower Nos. 1, 2, 3 or 4 100N
Intermediate Element for Tower Nos. 1, 2 or 3 109N
Intermediate Element for Tower No. 4 110N
Top Element for Tower Nos. 1, 2 or 3 119N
Top Element for Tower No 4 121N
Saddle for Tower Nos. 1, 2, 3 or 4 135N

3.8.3 Concept of the standard drawings

Steel Drawings

Each drawing provides the necessary information and specifications for manufacturing the desired steel
parts. The required quantity of steel parts has to be filled in the space provided in the title of the drawing.

Specific items such as reinforcement rods and flats for Main Cable Anchorages have to be filled in the
empty spaces in the material lists.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 177

Example:

From the total weights of each drawing, the grand total for each steel category has to be added up as
follows:
A : Means the entire weight of the steel including galvanization to be transported to the site.
B: This is the total structural steel, raw or untreated (u+g). This includes steel profiles, plates and flats
but not reinforcement bars and other steel items.
C: This is the weight of the Nuts, Bolts and Washers (galvanized weight).
D : This is the weight of the Bulldog Grips or Thimbles and other items, if required (galvanized weight).
R: This is the raw weight of the Reinforcement Steel or Plain Rods. They are never galvanized.
The Total Transportation Weight A = B + C + D + R
g: The small g indicates the weight of the structural steel to be galvanized. This weight is part or can
be the sum of the total structural steel (B), but it is not an additional weight.

The above distinctions have been made for quotation purposes because the price per kg (or piece) of the
different items varies greatly. Reinforcement steel is usually much cheaper than structural steel; and nuts
and bolts are much more expensive than structural steel, etc. The weight of the steel to be galvanized is
necessary to obtain the price for galvanization separately, excluding the cost of the steel itself.

The steel drawings are usually not necessary at the construction site, but for assembly and identification
of the steel parts, a copy of each steel drawing should be available. Also for maintenance at a later stage,
copies of the steel drawings are useful.

Construction Drawings
The construction drawings are the actual site drawings. A complete set is absolutely necessary at the site.
¡¡ Walkway fitting drawing for 106cm walkway width
¡¡ Main cable anchorages
For the Main Cable Anchorages, there are three main categories of drawings

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 178

• Deadman Anchors for Flat Ground of up to 15o slope


• Gravity Block Anchors for Hill Slopes of more than 15o
• Drum Rock Anchors for Fractured or Soft Rock or Hard Rock
With these three groups of Main Cable Anchorages, the optimal kind of anchor can be applied.
The selection depends upon the prevailing Soil/Rock condition on the respective river banks and
the number and diameter of the Main Cables.
¡¡ Walkway and Tower Foundations
Depending on the Soil/Rock conditions the number of Main Cables and the required Block Height,
Walkway and Tower Foundations are to be selected. There are five options for the two Main Cables
with c/c1 = 2.50m, and seven options for the four Main Cables with c/c1 = 3.50m.
¡¡ Windguy Cable Anchors
There are two categories of anchor types for anchoring the windguy cables.
• Gravity Block Anchors for different Soil/Rock conditions with variable Block Heights (five
types)
• rum Rock Anchors for fractured or soft Rock or for hard Rock (two types)
All drawing types are complete designs and fulfill their respective parameters selected in the
design form. Also, in all drawings, the necessary quantities of construction materials have already
been calculated as far as possible. They have to be filled in their respective tables given in the cost
estimate Form No. 3 or 4.
It is absolutely necessary to fill in the Elevations and Cable diameters as indicated in the drawings.
The levels are to be determined in the topographic profile of the survey, whereas the cable
diameter can be taken from the design form.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 179

Examples of Construction Drawings:

Main Cable Anchor Block:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 180

3.8.4 Relationship between construction and steel drawings

Each Construction Drawing has related Steel Drawings. Their respective related drawing numbers are
mentioned on the drawing itself, and also the respective steel parts numbers are indicated on the
construction drawing for easy reference.

The Construction Drawings “Walkway Fitting”, Nos. 19Ncon show the elements of the superstructure
of the bridge. The related Steel Drawings are the Steel Crossbeam, Steel deck, Suspender and Tower
Drawings.

Crossbeam for
Suspension Bridge
Drawing Nos. : 07N

Steeldeck Panels
Construction Drawing

Drawing Nos. : 08A & 10A

Steel Drawings
Walkway Fitting
Drawing Nos. : 19Ncon

Suspender for
Suspension Bridge
Drawing Nos. : 31N or 32N

Assembly & Layout


of Tower
Drawing Nos. : 141N …144N

The Steel Drawings for the cable anchorages Nos. 42N, 44N and 50A are related to the corresponding
Construction Drawings Nos. 42/1Ncon…44/1Ncon…and 55/1Acon.

The Steel Drawings for the walkway and tower foundations Nos. 92N, and 94N are related to the
corresponding Construction Drawings Nos. 92/1Ncon…and 94/1Ncon.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 181

3.9 Construction

3.9.1 Bridge layout

The Bridge Layout is to fix the bridge position and foundations at the site as per the design.

Procedure for General Bridge Layout (refer to General Arrangement ‘GA’ Drawing):
¡¡ Find the existing pegs and Bench Marks.
¡¡ Measure the horizontal distance between the axis pegs A (L) and B(R), and compare with the
measurement given in the General Arrangement.
¡¡ Check the elevations of the axis pegs A (L) and B(R), and compare with the elevations given in the
GA.

¡¡ If the horizontal distance between the axis pegs A (L) and B(R) and their elevations are not similar
to the measurements given in the GA, readjust the design according to the actual measurements.
¡¡ If the horizontal distance between the axis pegs A (L) and B(R) and their elevations are identical to
the measurements given in the GA, fix the position of all the foundation blocks.

Procedure for Detailed Foundation Layout:


¡¡ Align the centerline of the bridge by joining the permanent points with mason threads or by
ranging between the axis pegs ‘A’ and ‘B’ .

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 182

¡¡ Mark the front of the tower foundation on the bridge centerline (peg 1) with reference to the axis
peg. The distance between the front of the tower foundation and the axis peg is given in the GA.

¡¡ Check the location of the front of the tower to ensure that it is at a sufficient distance (minimum 3m
for soil slope and 1.5m for rock slope) from the bank edge.
¡¡ Measure the length of the foundation from peg 1 and fix peg 2. Set up two additional centerline
pegs at a safe distance for the excavation works (pegs 3 and 4).

¡¡ Draw an offset line (right angle) through pegs 1 and 2 by the 3-4-5 method. Starting from peg 1,
set out pegs 5, 6, 7 and 8 for the reference line of the front edge, and from peg 2 for pegs 9, 10,
11 and 12 for the back edge of the tower foundation.

¡¡ Determine the center point of the tower foundation and measure the distance between the tower
axis and the front of the main cable anchor and fix peg 17. Draw an offset line through peg 17 as
a reference line of the front edge. Set up one additional centerline peg at a safe distance behind
and proceed as above.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 183

¡¡ Determine the reference line at the downstream edge with the help of pegs 5 and 9. For the
upstream edge, use pegs 6 and 10.

¡¡ Fix the elevation line (datum level), and indicate the depth of the excavation work for the tower
foundation and the main cable anchor as per the elevations indicated in the GA and Anchorage
Drawings.

Setting out Windguy Cable Anchorages

In the General Arrangement, the location of the wind guy cable anchorages is usually given in reference
to the tower and bridge centerline: D1, d1, D2, d2.
The correct direction is given by the angles β1, β2.

Plan:
D1, D2 = Distances measured from the bridge
axis to the front (centerline) of the
windguy cable anchorage block.
d1 d2 = Distances measured from the tower
centerline to the front of the windguy
cable anchorage block.
β1, β2 = Angle between the windguy cable
and the direction of the bridge
centerline.

Procedure:
¡¡ Measures the distances d1 and d2 from point Q along the bridge centerline, peg out points O, P.
¡¡ Set up the theodolite at O and P and measure the distances D1 and D2 perpendicular to the
centerline of the bridge, peg out points R and S.
¡¡ Set up the theodolite at R and S and set out reference lines R - T, S - U.
PRT = β2 + 90o (β2 + 1009)
OSU = β1 + 90o (β1 + 1009)
Peg out points T and U. The reference lines R - T and S - U are the centerlines of the windguy cable
anchorages.
¡¡ Check if the ground levels at points S, U, R and T correspond approximately to the layout given in
the General Arrangement.
If the ground levels are correct:
¡¡ Set out the windguy cable anchorage blocks in the same way as for the Main Anchor Foundation.

Setting out Windguy Cable Anchorages in Case of Obstacles


If it is not possible to set out the windguy cable anchorages, according to the above procedure due to
obstacles (rock, trees, etc.), it can also be done by using other points like R’ and S’.

PART F: TRAIL BRIDGE MANUAL


If the ground levels are correct:
• Set out the windguy cable anchorage blocks in the same way as for the Main Anchor Foundation.
F - Chapter 3 - 184
Setting out Windguy Cable Anchorages in Case of Obstacles

If it is not possible to set out the windguy cable anchorages, according to the above procedure due to
Plan:
obstacles (rock, trees, etc.), it can also be done by using other points like R' and S'.

Plan:

191

Procedure:
¡¡ Choose points P’ and O’ and measure the distance k from the tower centerline to P’ and O’.
¡¡ Set up the theodolite at O’, and measure the distance D’1 and D’2 perpendicular to the centerline
of the bridge and determine R’ and S’.
¡¡ Set up the theodolite at R’ and S’ and set out reference lines R’ - R and S’ - S.
Measure the distances r, s and peg out points R, S.

Notes:
¡¡ The points O’, P’ can also be at the river bed if it will facilitate setting out the reference lines.
¡¡ Reference lines for the windguy anchorages are to be set out in the same way as employed for
the main anchorages.

A bench mark should be set up for checking the elevation.

Setting out Windguy Cable Anchorages in Case of Incorrect Survey

In some instances, the elevation of the ground near the windguy anchorages may not coincide with the
General Arrangement. In such cases, the layout of the anchorage block position can be revised at the site.

1. Prepare a new profile drawing along the reference lines R - T and S - U and indicate the windguy cable
elevation according to the General Arrangement (first phase).
2. Relocate the windguy cable anchorage block in such a way that its position in relation to the ground
is similar to the layout given in the General Arrangement (second phase).
3. Draw the new location of the anchorage block in the profile drawing and transmit the data to the
design office which will prepare a revised design.
4. Make sure that the windguy cable is still long enough.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 185

3.9.2 Foundation excavation

In Soil:
The foundations should be excavated with slopes to provide stability in the cut slope. The cut slope in soil
should generally not exceed 3:1 (V:H). The foundations should be excavated stagewise. Trenches should
be excavated vertically with sheeting, or must be banked with slopes which afford the necessary stability.

All safety requirements for the protection of personnel during excavation must be met.

Slope Pitch
¡¡ In well consolidated stable ground, maximum slope pitch 3:1 (3m vertical, 1m horizontal);
¡¡ In moderately consolidated but stable soil, maximum slope pitch 2:1 (2m vertical, 1m horizontal);
¡¡ In non-cohesive ground, maximum slope pitch 1:1 (1m vertical, 1m horizontal).

If slope stability is impaired by unfavourable strata morphology, artesian water, intermediate friction
layers, vibration, etc. the slope pitch must be reduced.

The most important point to be borne in mind during excavation is the fact that almost
every bridge foundation bed is inclined. It is not allowed to excavate horizontally and
form the incline with fill material!

To ensure that the foundation bed is clean and undisturbed, the bottom 10cm should be excavated only
shortly before the concrete is poured.

During all excavation stages, the excavation depth should be accurately maintained. For this, establish an
elevation line (datum level) and measure the foundation depths with fixed sticks.

All the excavated soil should be safely disposed of without damaging the existing vegetation down hill,
thus not affecting the environment.
In Rock:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 186

Rock excavation is necessary to prepare the platform for a drum anchorage or for a rock block anchorage.
The rock should be excavated manually without blasting.

Excavation in rock is done by first drilling holes to weaken the rock, and then using crowbars to break it
up and dig out the pieces. The cutting can be carried out by forming steps.

3.9.3 Cement works

(a). Composition and Mixtures


Cement concrete is a mixture of the four components:
¡¡ Cement;
Ordinary Portland Cement commonly used for general construction works.
¡¡ Sand;
¡¡ Gravel;
¡¡ Water.

Cement is very sensitive to humidity and moisture; therefore it should never be stored for a long time. In
the rainy season, cement bags have to be packed in additional sealed plastic bags plus additional nylon
bags for protecting the cement against water and damage.

Sand should be clean, sharp, angular, hard and durable. Sand must be well washed and cleaned of mud
or any organic material before use. A well-graded sand should be used for cement works. All or most of
the sand should pass through a 3 mm sieve or mesh wire. However, it should not be too fine, only 15% of
the sand at the most can be smaller than 150 microns, which is like dust.

Gravel should be clean, hard, angular and non-porous. Usually riverside gravel makes the best aggregate
for preparing concrete. The corn size of gravel should be smaller than 40 mm (1½ inches) but bigger than
5 mm.

Water from rivers or lakes is usually suitable for making cement mixtures. Do not use water from ponds
or swamps, it may contain a lot of organic materials.

The main characteristic of any cement work is given by the mix proportions of their components:
¡¡ Cement Mortar = A mix of Cement and Sand
¡¡ Cement Concrete = A mix of Cement, Sand and Gravel

Of course, Water is added in both cases, but the mix proportions of cement, sand and gravel give the
main characteristics of any cemented work.

Mixing the above components thoroughly is of utmost importance. Hand mixing should be done on
a clean watertight platform. Cement and sand should first be mixed dry, and then gravel added. Now
the whole mixture should be turned over three times dry. Then mixing should take place for at least five
minutes by slowly sprinkling water until the concrete is of a uniform color.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 187

Table F.3.21: Quantities for Various Types of Cement Works

Dry required quantities for one cubic meter


Mixed proportions
wet:
Type of
Cement Stones
Cement
Work Sand Gravel or
Cement : Sand : Gravel bags kg
[m3] [m3] Boulders
@ 50 kg
[m3]
1 : 1 – 20.4 1020 0.71 – –
1 : 2 – 13.6 680 0.95 – –
Cement
1 : 3 – 10.2 510 1.05 – –
Mortars
1 : 4 – 7.6 380 1.05 – –
1 : 6 – 5.0 250 1.05 – –
Cement
Plaster
1 : 4 – 0.18 9 0.024 – –
(20 mm
1 : 6 – 0.12 6 0.024 – –
includes
12% waste)
uncoursed 2.66 133 0.37 1.1
1 : 4 –
stone
Cement 1 : 6 –
masonry 1.75 87.5 0.37 1.1
Stone
Masonries coursed 2.28 114 0.32 1.1
1 : 4 –
stone
1 : 6 –
masonry 1.50 75 0.32 1.1
Cement 1 : 4 : 8 3.4 170 0.47 0.94 –
Concretes 1 : 3 : 6 (M10) 4.4 220 0.47 0.89 –
(plain or 1 : 2 : 4 (M15) 6.4 320 0.45 0.85 –
reinforced) 1 : 1½ : 3 (M20) 8 400 0.42 0.84 –

"Plum" 1 : 3 : 6
2.64 132 0.28 0.54 0.50
Concrete with 50% boulders
Source: Indian Practical Civil Engineers’ Handbook, Section 20

The amount of Water should be about half the volume of cement. One 50-kg bag of cement has a
volume of approximately 35 liters, which is equal to approximately two kerosene tins.

Concrete and Mortar should be placed in its final position within one hour. After placing, it should be
well compacted with the help of rods in order to remove any air pockets. For concrete of high quality,
good compaction is essential. This may mean extra work during placing, but on no account should more
water be added for reducing the compacting work. Concreting should never be done if it is raining.

Curing means keeping completed cement works wet until the setting process is completed. If concrete
works are not kept continuously wet during the setting process, cement mortar, cement stone masonry
work and especially concrete do not develop their full strength. Curing should be done for at least 28
days.

For increasing the strength of concrete, ribbed Tor-Steel bars are added which makes Reinforced Cement
Concrete or RCC.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 188

(b). Tower and Walkway Foundations


The tower and walkway foundations are built for providing a safe base for the towers and for the firm
anchoring of the spanning cables, fixation cables and handrail cables. This structure also stands for the
bridge entrance.

For this structure, the proper and precise placing of the reinforcement bars is very important. The bars
usually come pre-bent from the workshop, but can also be bent at the site if required.

D1 = 15 x d
D2 = 6 x d, for rod ø ≤ 20mm
8 x d, for rod ø > 20mm
D3 = 4 x d, for rod ø ≤ 16mm

Reinforcement steel is fixed with binding wire, so that no displacement can happen during concreting.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 189

(c ). Cable Anchors in Soil


Cable Anchors in Soil or Soil Anchors are constructed either as:
¡¡ Deadman Anchor or;
¡¡ Gravity Block Anchor.

For both Soil Anchor Types (Deadman and Block Anchors), ready-made design drawings are provided for
Anchors with Turnbuckles or Anchors with Direct Cable Connection. One Cable Anchor should always
have a Turnbuckle which allows fine adjustment of the cables during erection time. For economic reasons,
the cable anchor on the other bank should always have a direct cable connection.

Always choose the more convenient bank for the Main Cable Anchor with Turnbuckle.

For the Main Cable Anchor with Direct Cable Connection the Main Cables have to be fixed during the
construction of the particular Anchor Type. The Main Cables, therefore have to be ready during the time
of construction. Make sure that the Main Cables are deposited at a safe distance from the bank of the
river during construction time.

For both the Soil - Anchor Types (Deadman and Block Anchor), ready-made design drawings are provided
for Anchor with Turnbuckles or Anchors with direct cable connections. These drawings contain all the
information about assembly and location of each steel part.

Example of a Deadman Anchor without Turnbuckle:

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 190

Example of a Gravity Block Anchor with Turnbuckle:

Nos. & Spacing of Bulldog Grips


Cable φ mm Nos. G
13 3 10cm
26 5 15m
32 6 20m

(d). Constructing Drum Anchorages in Rock


There are two types of Drum Anchorages in Rock:
¡¡ Drum Anchor in Hard Rock
¡¡ Drum Anchor in Soft or Highly Fractured Rock

Drum Anchorages in Hard Rock are made by drilling holes of 32mm diameter ( = diameter of crowbar)
into the rock. Clean the boreholes of dust and debris by flushing them with water.

Fill the holes with cement mortar 1:1 before the anchor rods are inserted.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 191

Drum Anchorages in Soft or Fractured Rock are not done by drilling holes, but by excavating a round
pit. The Anchor Reinforcement has to be placed into the pit and is fixed with the help of stirrups. The
excavated pit is then filled and well compacted with concrete 1 : 2 : 4 up to the ground level. At this
stage, the anchor rods should protrude (stick out) approx. 50 cm. The formwork for the drum is made
using a chitra (bamboo mat) lined with a plastic sheet bound together with binding wire.

(e). Water Management and Backfilling

The life expectancy of a bridge largely depends on proper water management.

Any water seepage encountered during excavation should be intercepted as close as possible to its
origin, and channeled safely to a nearby watercourse. Especially vulnerable is the place behind the
Deadman Beam! If in doubt, or in case of unusual humidity or water seepage, provide a drain behind
the Deadman Beam with a side outlet. Sometimes, water seepage occurs during the rainy season only.
Inquire with the local people.

Divert surface water and provide drainage channels as necessary. Do not hamper existing irrigation
channels, rather improve and adjust them with some cement works. Discuss solutions with the local
people and decide on the spot.

As a general rule divert water as far away from the bridge foundations as possible.

For managing surface water well, also fill in the gaps around completed anchor blocks well above the
existing surface. Back filling prevents surface water from flushing out the excavations.

(f). Finishing Work

Provide finishing structures like retaining walls, staircases, small trail improvements, adjustments to
nearby houses, etc. if they add functional value to the bridge. Never do cement pointing or other non-
functional works. Also check the vegetation and plant life in the vicinity of the bridge. Plant some new
trees if possible, especially if some had to be cut down in order to build the bridge.

3.9.4 Bridge erection

As soon as the anchor blocks and tower foundations are completed, the bridge erection works can be
started. Bridge erection and fitting works are somewhat difficult and dangerous, and require especially
skilled laborers who will not suffer from giddiness. Because of this somewhat risky work, the necessary

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 192

safety precautions should strictly be followed and the respective responsibilities should be clarified
before starting the work.

(a). Erection of Towers

The towers must be temporarily fixed during erection, because they rest as a line load on the base
plate which acts like a hinge. In order to avoid serious accidents during erection, the towers must be
temporarily fixed at their base. For this purpose, temporary side struts have been provided at the tower
and walkway foundation with an additional 8 angles for each tower (see position B-5 of Base Element
Drawing No. 100N). The angles are supplied with holes at one end only, the exact position of the
required hole at the other end must be marked at the site when erecting the base element.

First fix both the Base Elements at the bottom and put them in an exactly vertical position. Fix the
temporary angles at the bottom of each side, then mark the required position of the hole at the other
end of the angle. Also give a position number to each angle so that they won’t get mixed up. Have one
hole of diameter 17 mm drilled at the site or at the nearest workshop. For tower erection, refer to the
respective Assembly and Layout Drawings.
¡¡ Check Steel parts for labels and numbers put by the manufacturer.
¡¡ Use steel-cones for easy fitting works and tighten the nuts and bolts fully only after the next
diagonal bracing has been put in place. Each tower consists of the parts :
Type of element Part Nos. in Steel Drawing
Base Element B
Intermediate Element I
Top Element T
Saddle S

Retighten all Nuts and Bolts firmly after fixing the Elements.

After the bridge has been erected, the cable clamps on the top of the pylon must be firmly
tightened, and then all the side struts (angles) must be removed.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 193

Figure F.3.16: Assembly & Layout of Tower No. 4

(b). Hoisting of Main Cables and Sag Setting


Usually, the main cables are pulled across the river with the help of nylon ropes. In case of a deep or
turbulent river, attach an empty airtight plastic can (jerry can) at the end of the cable. This will prevent
the cable-end from getting stuck between stones and rocks lying on the riverbed.

Make sure that the respective Main Cables are pulled on either side of the Tower and Walkway Foundations.
Fix them temporarily at the respective Turnbuckle at the Main Cable Anchor.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 194

The hoisting sag setting of the Main Cables is one of the most important tasks during the erection
of the bridge.

The towers should stand exactly vertical, the saddle cover plates are loose, and the temporary side struts
are fixed. With this arrangement, the main cables can slide over the saddles when the bridge is being
erected and the cables become longer; and the towers remain in vertical position in dead load.

With a leveling instrument, the exact hoisting sag is fixed in the way:
¡¡ Mark the elevation of the hoisting sag on both the towers with permanent paint.
¡¡ Now set up the leveling instrument on the tower foundation so that its line of sight matches with
the mark on the tower across the river. Setting up the leveling instrument at the prescribed hoisting
sag elevation has to be done by trial and error, and may require several attempts. Make use of the
three adjustment wheels of the leveling instrument when the eyesight is close to the mark.
¡¡ Pull the Main Cables until they reach a level of about 20 cm higher than the hoisting sag.
¡¡ Clamp the cables around the thimbles at the cross bar of the Turnbuckle of the main cable
anchorage. Make sure that the crossbar is in the middle position of the threaded anchor bars when
clamping the main cables, secured with two nuts in the front and one in the back
¡¡ The Main Cables should be left in this “over pulled” position for at least 12 hours so that some
relaxation can take place.
¡¡ Now move the Turnbuckles to achieve the exact sag setting. For compensating elongations due
to change in air temperature, recheck the hoisting sag at different times of the day and make
the necessary adjustments. It is recommended to adjust the final sag setting during the hot day
after noon, when the cables have accumulated maximum heat, i.e., during maximum elongation
condition.
¡¡ The hoisting sags of all the Main Cables must be identical at any point of time.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 195

Also check the sags from time to time when the fitting works are going on. Different elongations may
take place due to dissimilar hidden cable relaxations when the tension increases. Adjust possible sag
differences with the help of the turnbuckles at the main cable anchor so that the Main Cables are always
parallel and compare the dead load sag with the pre-calculated values.

(c). Hoisting of Spanning Cables


Fix the Spanning Cables at the Turnbuckles of the Tower and Walkway Anchorage on one river bank.
Make sure that the crossbar of the turnbuckles are at the outermost position secured with two nuts each
so that more tension can be applied when all the fitting work is completed.

Pull the cables across the river and secure them at the corresponding turnbuckles on the other bank
(crossbar at the outermost position).

It is not necessary to achieve the sag corresponding to the required dead load camber, since this requires
very high pulling forces. Just make sure that both the Spanning Cables are hanging approximately
parallel and are high enough over the highest water-level of the river.

It is much easier to adjust the spanning cables when the suspender is being fitted .

(d). Fitting Suspenders and Center Row of Steel Deck


Fitting the suspenders and walkway elements are the most difficult and daring job. Adequate safety
precautions should be strictly followed and the respective responsibilities should be clarified.

The suspender fitting work should start from both the towers and proceed towards the center of the
bridge. This procedure is easier and has more advantages than starting the fitting work from the center.
However, in order to achieve a proper symmetry of the suspenders, the central suspender must be fitted
first. The only disadvantage will arise when finishing the fitting works at the middle of the bridge. Due to
inaccuracies, the remaining spacing at the center of the bridge might be either too long or too short. For
minimizing this imprecision, the required distances to the towers and the center have to be rechecked
after fitting 10 suspenders.

Preparation for Suspender Fitting Works:


¡¡ Lay out all the suspenders in sequence on the ground.
¡¡ Prepare all crossbeams, J-hooks and steel deck.
¡¡ Prepare two fitter platforms, one for the main cables and one for the spanning cables, and two
gauged sticks of exactly 1.00m length.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 196

Assembly:
First fit the top portion on the main cables, then hang the platform under this by using either steel bars
or cables.

Fitting the central suspender:


With the help of the fitting platform, the suspender in the center has to fitted first. Determine and
reconfirm the center with a tape and level instrument, then fit the first suspender-pair at the center of the
bridge. To avoid excessive load on the center suspender during erection time, bind all cables (spanning
and main cables) together a security rope.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 197

The security rope supports the suspender in the middle during erection work.

Tighten the spanning cable to some extent; now the cables are ready to be fixed to the suspenders.

Sketches and Procedures for Fitting Operations:

¡¡ Start the fitting work from both sides of the bridge and work towards the center of the bridge;
¡¡ Fix one cable car on top of the main cables and one on top of the spanning cables;
¡¡ Fix the first two suspenders to the main cables at the prescribed distance from the tower;
¡¡ Lift the spanning cable until the suspenders can be connected with the threaded rod of the walkway
crossbeam. Note: The first crossbeam at the bridge entrance is fitted without a suspender (see
Drw. No. 19Ncon).
¡¡ In order that the suspenders are fixed exactly 1m apart, use gauged sticks for exact fitting;
¡¡ Re-adjust the spanning cables from both the banks as the suspenders are being fitted;

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 198

¡¡ After fitting ten pairs of suspenders, check the distances to the tower and to the center;
¡¡ Adjust only inaccuracies by moving the crossbeams;
¡¡ Gradually start fitting the center row only with standard steel deck panels (refer Drawing No. 19
Ncon).

When the center is reached, there will be some extra length of spanning cable. For adjusting this, pull
the spanning cables from both the banks with the tirefor machine through the loose J-hooks. Make sure
that the middle row of the steel deck is fitted when doing this work.

¡¡ When all the suspenders have been fixed, tighten the spanning cables with the cable pulling
machine as much as possible before fitting the 2” G.I. pipes, the crossbeams and before fitting the
rest of the steel deck panels.
¡¡ Fix the handrail cables by pulling them through the suspender-rings just above the suspender
turnbuckle, and secure them to the handrail posts by winding the cable end twice around the post.

Fitting the 2” G.I. Pipes :


Two 2” G.I. pipes have to be mounted from below to the steel deck cross beams. This provides additional
vertical but also lateral stability to the entire walkway. These pipes can also be used for transferring water
across the river as per local requirement.

The G.I. pipes have to be fitted before the outer rows of the steel decks are mounted in the way:
¡¡ Lay two pipes of 6m length end to end on the ground and join them together firmly. Use a 2” die
set and jute threads to make the joint water tight.
¡¡ In the same way, also fix half of the “union” at each end of the 12m piece.
¡¡ Now carry the 12m pipe to the bridge, pass it through the suspenders by securing it with nylon
ropes until the entire 12m piece is on the outside of the suspenders.
¡¡ Now bring the pipe into proper position underneath the walkway, and secure it immediately with
the U-clamps and join it with the “union”.
¡¡ In case a union coincides with a crossbeam, cut the pipe and make a new thread with the die set.

This work requires special attention. While passing the pipe outside the suspenders, several
workers are necessary and sufficient ropes are required to secure the pipe at all times.

(e). Finishing the Walkway Steel Deck

Now the remaining steel deck panels can be fitted. Fit them in a staggered way. Adjust the end of the
walkway with “Special Panels”; and cut off any extra length with a hacksaw as required.

¡¡ If wood is used for the bridge deck, the planks should be 2 meters long and at least 4 cm thick.
They should also be fitted in a staggered way. Use washers below the Bolt Heads. The distance
between the Crossbeams is 1 meter.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 199

Note:
1. Make sure that all U-clamps which hold the G.I. pipes have been tightened firmly before fitting the steel deck.

(f). Fitting Wind Bracings


Wind bracings might be required to protect the superstructure from excessive wind forces.
¡¡ Cut the windtie cable according to the calculations resulting from the design form;
¡¡ Mark each cable at 60 cm from the end;
¡¡ Fit a thimble and bulldog grips so that the “60 cm mark” is at the top of the thimble;

¡¡ Now mark each windtie cable at 2.0 meter from the other cable end.

The windtie cables are now ready to be fixed to the windguy cable. For this, fix the cut windtie cables at
intervals of 5.0 m to the windguy cable; take starting measurements from the design form or the general
arrangement.

Make sure that the windtie cable clamps and bulldog grips are firmly tight, and that the nuts are locked
against the threaded plates of the windtie cable clamp, because these connections are no more accessible
after hoisting.
Hoisting Windguy Cables :
¡¡ Place the wind guy cable in position along the bridge deck outside the suspenders;

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 200

¡¡ Pass the respective windties through the walkway beam and clamp them loosely at the “2.0m
Mark”;
Note: The steel deck in the center row at this location must be removed for this purpose.
¡¡ Fix one bulldog grip with the windtie from the other side just outside the crossbeam;
¡¡ After fixing all the windties, attach both cable-ends of the windguy cable to the respective windguy
anchors (crossbar at the outermost position). Simultaneously apply tension to both the windguy
cables to avoid any unsymmetrical lateral load to the bridge, then tighten all clamps at the “2.0m
Mark”.

Adjusting the Windguy / Tie - Cables :


¡¡ The windties can be adjusted by loosening the bulldog grips at both sides of the crossbeam;
¡¡ Loosen also one bulldog grip outside the crossbeam, but make sure that the grip on the other side
is tight !
¡¡ Now allow the first windtie - pair to balance off any unsystematic forces;
¡¡ Then grasp the cable under tension (only one cable) from the middle of the crossbeam and lift it
as required;
¡¡ Use a crowbar and wooden logs to hold this position.
¡¡ Now push the loose cable-end of the second windtie through the loose bulldog grip outside the
crossbeam and clamp it to the cable under tension;
¡¡ Now do the same with the other end, and finally again clamp both the cables to the crossbeam.

This procedure should be repeated until all the windties are tight. The adjustment length is not limited
and can also be done again at a later stage.

(g). Finishing Erection Works


Now there remains the fitting of the fixation cable, the wire mesh and some final adjustments.

Like the handrail cable, the fixation cable is to be passed through the rings at the bottom of the suspenders.
Anchor the fixation cables to the spanning cable anchor.

The fencing is woven on the spot between the handrail and the fixation cables with gabion wire (12 SWG
ø 2.64 mm)

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 201

Finally the whole bridge should be aligned with the help of the turnbuckles at the suspenders so
that the walkway forms a smooth parabola. If a wind bracing system (windguys and windties) is in place,
recheck whether all the cables are tight and align the walkway so that it is perfectly straight.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 202

Suspension Trail Bridge Design Form

Bridge Design
1. Cables and Standard Tower Selection
2. Main Cable Anchor Blocks
3. Walkway & Tower Foundations
4. Windguy Cable Anchor Blocks
5. Suspender List
6. Windguy Arrangement
7. Bridge Standard Drawings

Bridge No. & Name:______________________________________

River Name:______________________________________________

District Name:____________________________________________

Designed by:_____________________________________________

Approved by:____________________________________________

Date:____________________________________________________

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 203

Bridge No: ………… Bridge Name:……………… Woreda:……………………

Span: ………………. Walkway Width : ……….

1. Cables and Standard Tower Selection for Suspension Bridge Type

A. Survey Data & Calculation of Freeboard

Bridge Geometry

1. Span of the Bridge (should be rounded off to the nearest meter) l = ________

2. Elevation of the Spanning Cable Anchor Point Es = ________

3. Highest Flood Level HFL = ________

4. Freeboard (min. 3.00m) Fb = Es - HFL = Fb =

(If the freeboard is less than 5.00m, try either to raise the elevation of the spanning cable anchor
point by increasing the height of the Walkway and Tower Foundation, or to adjust the span to gain the
required elevation.)

B. Selection of Cables and Standard Tower (Pylon)


¡¡ Select the Main and Spanning Cable combinations, standard tower type and tower height ht and
the Bridge Geometry according to the span from the table. The Bridge Geometry refers to main
cable dead load sag fd, dead load camber cd of spanning cable and the main cable inclination at
tower saddle βf in full load case.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 204

Table F.3.22: Selection of Cables & Standard Towers

Tower Cables Dead Full load


Hoisting
Span Height/ Nos. & diameter Load Camber Backstay
Sag
l (Type) Sag cd Angle
Main Spanning Windguy fh
m ht fd m ßf
mm mm mm m
m m deg.
30.0 4.20 4.13 0.90 29.83
31.0 4.15 4.07 0.95 28.83
32.0 4.10 4.02 1.00 27.88
33.0 4.10 4.01 1.00 27.24
34.0 4.08 3.99 1.02 26.55
35.0 4.05 3.95 1.05 25.84
36.0 4.22 4.12 0.88 26.14
2 ø 26

2 ø 26
37.0 4.19 4.09 0.91 25.49
38.0 4.16 4.05 0.94 24.88
39.0 4.13 4.01 0.97 24.31
40.0 4.10 3.97 1.00 23.77
41.0 4.10 3.96 1.00 23.39
42.0 4.10 3.96 1.00 23.03
5.50
(1)

43.0 4.30 4.16 0.80 23.48


44.0 4.40 4.26 0.70 23.53
45.0 4.40 4.25 0.70 23.20
Not Required

46.0 3.92 3.77 1.18 20.45


47.0 3.90 3.74 1.20 20.10
48.0 4.05 3.89 1.05 20.39
49.0 4.05 3.88 1.05 20.13
50.0 4.10 3.93 1.00 20.07
51.0 4.10 3.92 1.00 19.83
52.0 4.10 3.91 1.00 19.61
53.0 4.10 3.90 1.00 19.40
54.0 4.10 3.88 1.00 19.20
2 ø 32

2 ø 32

55.0 4.25 4.04 0.85 19.46


56.0 5.47 5.30 1.48 22.82
57.0 5.44 5.26 1.51 22.46
58.0 5.41 5.22 1.54 22.12
59.0 5.38 5.18 1.57 21.79
60.0 5.35 5.13 1.60 21.48
7.35
(2)

61.0 5.35 5.12 1.60 21.26


62.0 5.45 5.22 1.50 21.32
63.0 5.65 5.42 1.30 21.65
64.0 5.85 5.63 1.10 21.97
65.0 5.95 5.72 1.00 22.01

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 205

Tower Cables Dead Full load


Hoisting
Span Height/ Nos. & diameter Load Camber Backstay
Sag
l (Type) Sag cd Angle
Main Spanning Windguy fh
m ht fd m ßf
mm mm mm m
m m deg.
66.0 7.02 6.81 1.78 24.48
67.0 6.99 6.77 1.81 24.15
68.0 6.96 6.73 1.84 23.83
69.0 6.93 6.69 1.87 23.54

2 ø 32
70.0 7.20 6.97 1.60 23.97
9.20
(3)

Not Required
71.0 7.40 7.17 1.40 24.22
72.0 7.60 7.37 1.20 24.47
73.0 7.70 7.46 1.10 24.47
74.0 7.65 7.40 1.15 24.14
75.0 7.65 7.40 1.15 23.93
76.0 8.57 8.36 2.08 25.38
77.0 8.54 8.33 2.11 25.07
78.0 8.51 8.29 2.14 24.77
79.0 8.48 8.25 2.17 24.48
80.0 8.45 8.22 2.20 24.20
81.0 8.42 8.18 2.23 23.93
82.0 8.39 8.14 2.26 23.67
2Ø 32

83.0 8.36 8.10 2.29 23.41


84.0 8.33 8.06 2.32 23.16
4 ø 26

85.0 8.30 8.02 2.35 22.92


86.0 8.27 7.98 2.38 22.68
Ø 26 (Optional only)

87.0 8.24 7.94 2.41 22.46


11.05
(4)

88.0 8.21 7.90 2.44 22.24


89.0 8.35 8.04 2.30 22.34
90.0 8.55 8.24 2.10 22.56
91.0 8.85 8.54 1.80 22.96
92.0 9.05 8.74 1.60 23.16
93.0 9.25 8.95 1.40 23.36
94.0 9.25 8.94 1.40 23.20
95.0 9.20 8.87 1.45 22.96
96.0 7.98 7.68 2.67 20.00
97.0 8.05 7.75 2.60 20.00
4 ø 32

98.0 8.13 7.82 2.52 20.00


99.0 8.20 7.89 2.45 20.00
100.0 8.28 7.96 2.37 20.00

PART F: TRAIL BRIDGE MANUAL


Short Span Trail Bridge Standard

F - Chapter 3 - 206

(…continued)
Span Tower
Tower Cables
Cables Dead
Dead Hoisting Camber Fullload
Full load
Height/(Type)
Hoisting
Span Height/ Nos.
Nos. and Diameter
& diameter Load Sag
Load Sag
Sag
Camber Backstay
Backstay
l (Type) Sag cd Angle
Angle
f
ml ht ht Main
Main Spanning Windguy
Spanning Windguy fdfd fm
h
h m
cd ßßf f
m mm mm mm
m m mm mm mm mm m m deg.
deg.
101.0
101.0 8.35 8.35 8.03
8.03 2.302.30 20.0020.00
102.0
102.0 8.40 8.40 8.07
8.07 2.252.25 20.0020.00
103.0
103.0 8.50 8.50 8.17
8.17 2.152.15 20.0020.00
104.0 8.57 8.23 2.08 20.00
104.0 8.57 8.23 2.08 20.00
105.0 8.65 8.31 2.00 20.00
105.0 8.65 8.31 2.00 20.00
106.0 8.71 8.36 1.94 20.00
106.0 8.71 8.36 1.94 20.00

only)
107.0 8.80 8.45 1.85 20.00
107.0 8.80 8.45 1.85 20.00

(Optional only)
108.0 8.85 8.49 1.80 20.00
108.0 8.85 8.49 1.80 20.00
109.0 8.94 8.58 1.71 20.00

ø 26(Optional
4 ø 32

32
109.0 8.94 9.00 8.58 1.71 20.00
11.05

4 ø 32

22øø 32
11.05

110.0 8.63 1.65 20.00


(4)

110.0
111.0 9.00 9.00 8.63
8.62 1.651.65 19.8820.00
111.0
112.0 9.00 8.95 8.62
8.55 1.701.65 19.7019.88
112.0
113.0 8.95 8.95 8.55
8.54 1.701.70 19.6019.70
ø 26

113.0
114.0 8.95 8.95 8.54
8.53 1.701.70 19.5019.60
114.0
115.0 8.95 8.90 8.53
8.46 1.751.70 19.3419.50
115.0
116.0 8.90 8.90 8.46
8.45 1.751.75 19.2419.34
117.0
116.0 8.90 8.85 8.38
8.45 1.801.75 19.0919.24
118.0
117.0 8.85 8.85 8.37
8.38 1.801.80 19.0019.09
119.0
118.0 8.85 8.85 8.36
8.37 1.801.80 18.9219.00
120.0
119.0 8.85 8.85 8.34
8.36 1.80
1.80 18.83
18.92
120.0 8.85 8.34 1.80 18.83
The above Cable Combinations and Towers have been calculated for the specifications and bridge
geometry.
The above Cable Combinations and Towers have been calculated for the specifications and bridge geometry.
Cables: construction 7 x 19, wire strand core, 160 kg/mm2 (1.57 kN/mm2) tensile strength of wire
Safety Factor: minimum 3
Live load 5000 50 or
p = (300 + ) kg/m 2 or (3 + ) kN/m 2
span span
p = 400 kg/m2 (4 kN/m2) if the span is 50.0 m or less

Backstay Angle βf ≅ ______________


Camber cd = ________________
Safety factor for Tower = ________________
Example: Span = 115m
⇒ selected cable combination: Main Cables : nos. 4∅ 32mm
Spanning Cables : nos. 2∅ 32mm
Windguy Cable : not applicable
Dead Load Sag, fd = 8.9m
Dead Load Camber, Cd = 1.75m
Backstay Angle, β f = 19.34°
⇒ selected Tower: Standard Tower Height, ht = 11.05m / Type 4

213

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 207

Selected Cable combination, Bridge Geometry and Standard Towers from Table 3.17

MC Main Cables: nos._____________


SC Spanning Cables: nos._____________
HRC Handrail Cables: nos._____________
FC Fixation Cables: nos._____________
WGC Windguy Cables (optional): ________________
Dead Load Sag, fd =_____________
Dead Load Camber, cd =_____________
Standard Tower Type: ________________
Tower Height, ht =_____________

C. Fixing the backstay lengths (dl and dr) and main cable anchorage elevations (El and Er)

¡¡ Fix the position of the Main Anchorage Foundation Blocks and their Cable Anchorage Elevations
on the right bank and the left bank in the bridge profile considering the topography and geological
condition of the ground
¡¡ Measure the backstay length DL and DR with a scale in the bridge profile.
¡¡ Calculate the Main Cable Anchorage Elevation.
Calculated Backstay Length (DL and DR) and
Main Cable Anchorage Elevations, (EL and ER):

Et = Es + 0.70 + ht =__________
βf (from Table 1.1) =__________
DL (measured from the bridge profile) =__________
DR (measured from the bridge profile) =__________
EL = Et - (DL x Tan βf) =__________
ER = Et - (DR x Tan βf) =__________

D. Calculation of Cable Lengths

Backstay Length
[m] Cutting Length*
Type of Cable dia (mm) Nos
[m/pc]
DL DR
Main Cable
Spanning Cable
Handrail Cable
Fixation Cable

Windguy Cable** 1 (U/S)


(Optional) 1 (D/S)
Windties**
(Optional)

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 208

* Cutting Length:
Main Cable = 1.08 x Span + 1.15 (DL+DR) + Anchorage Length on the Right Bank +
Anchorage Length on the Left Bank.
Calculate the Anchorage Lengths only after selection of the Main Cable Anchor Blocks.

Spanning Cable = 1.025 x Span + 3.5m.


Handrail Cable = Span
Fixation Cable = Span

** Windguy Cable = Refer to Design of Windguy Arrangement


** Windties = Refer to Design of Windguy Arrangement

E. Calculation of Hoisting Sag


This calculation has to be made after the tower erection work has been completed.

1. Actual Span measured in the field l =....… m


2. Hoisting Sag (from Table No 1.1 as per actual span) fh =.…... m
3. Tower Height ht =…..… m
4. Marking Point on Tower (from steel tower base) at hoisting sag elevation = ht – fh = ….... m

2. Main Cable Anchor Design


A. Design Data

Main Cables:

Right Bank Condition


Geology: Soil

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 15° slope) (more than 15° slope)

What is the Soil Type? Gravelly Sandy Silty


If Rock, what is the Rock Hard Rock Hard Rock Soft Rock
Type? (only a few fractures) (highly fractured)
Left Bank Condition
Geology: Soil

If Soil, how is the Ground Flat or Hill Slope


Surface? (up to 15° slope) (more than 15° slope)

What is the Soil Type? Gravelly Sandy Silty


If Rock, what is the Rock Hard Rock Hard Rock Soft Rock
Type? (only a few fractures) (highly fractured)

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 209

B. Selection of Anchor Types


Select the appropriate Anchor Types for the Right and Left Banks according to the above design data. s
are the rules for selecting the anchor blocks:
¡¡ One Main Cable Anchor should always have turnbuckle for fine adjustment during sag setting,
whereas for economic reasons, the main cable anchor on the other bank should always have a
direct cable connection.
¡¡ Always select an Anchor Block with turnbuckle for one bank and an Anchor Block with direct cable
connection for the other bank.
¡¡ If one bank is soil and the other bank is rock, always select a Drum Anchor for the rocky bank.
¡¡ If both banks are rocky, select an Anchor Block for one bank and a Drum Anchor for the other bank.

Procedure for selection:


¡¡ According to the Soil type and Slope of ground or the Rock type, refer to the respective tables for
the selection of the Anchor Types.
In Soil and Flat Ground (Cable Connection with Turnbuckle) : Table F.3.23
In Soil and Flat Ground (Direct Cable Connection) : Table F.3.24
In Soil and Hill Slope (Cable Connection with Turnbuckle) : Table F.3.25
In Soil and Hill Slope (Direct Cable Connection) : Table F.3.26
In Hard, Soft or Fractured Rock (Cable Connection with Turnbuckle) : Table F.3.27
In Hard Rock (Direct Cable Connection) : Table F.3.28
In Fractured Hard Rock or Soft Rock (Direct Cable Connection) : Table F.3.29
¡¡ In the table match the design data: Number and Diameter of Main Cables  if Soil, Soil type
select Anchor type and corresponding Drawing No. for right bank and left bank respectively.

Anchorage Type Selection Tables


¡¡ In Soil and Flat Ground (Ground slope up to 15° )

Table F.3.23: Selection of Main Cable Deadman Anchor in Soil & Flat Ground
(Cable Connection with Turnbuckle)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All DA 1 42/1Ncon
2x32 All DA 2 42/3Ncon
4x26 All DA 3 44/1Ncon
4x32 All DA 4 44/3Ncon

Table F.3.24: Selection of Main Cable Deadman Anchor in Soil & Flat Ground
(Direct Cable Connection)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All DA 5 42/2Ncon
2x32 All DA 6 42/4Ncon
4x26 All DA 7 44/2Ncon
4x32 All DA 8 44/4Ncon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 210

¡¡ In Soil and Hill Slope

Table F.3.25: Selection of Main Cable Deadman Anchor in Soil & Hill Slope
(Cable Connection with Turnbuckle)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All AB 1 42/5Ncon
2x32 All AB 2 42/7Ncon
4x26 All AB 3 44/5Ncon
Gravelly or Sandy AB 4 44/7Ncon
4x32
Silty AB 5 44/9Ncon

Table F.3.26: Selection of Main Cable Anchor Block in Soil & Hill Slope
(Direct Cable Connection)

Main Cable Foundation Soil Type Anchor Type Drawing No


2x26 All AB 6 42/6Ncon
2x32 All AB 7 42/8Ncon
4x26 All AB 8 44/6Ncon
Gravelly or Sandy AB 9 44/8Ncon
4x32
Silty AB 10 44/10Ncon

¡¡ In Hard, Soft or Fractured Rock

Table F.3.27: Selection of Main Cable Anchor Block in Hard, Soft or Fractured Rock
(Cable Connection with Turnbbuckle)

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 AB - 11 45Ncon
4x26 or 4x32 AB - 12 46Ncon

¡¡ In Hard Rock

Table F.3.28: Selection of Main Cable Drum Anchor in Hard Rock

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 DR-1 47Ncon
4x26 or 4x32 DR-2 48Ncon
¡¡ In Fractured Hard Rock or Soft Rock

Table F.3.29: Selection of Main Cable Drum Anchor in Fractured Hard/Soft Rock

Main Cable, mm Anchor Type Drawing No


2x26 or 2x32 DR-3 49Ncon
4x26 or 4x32 DR-4 50Ncon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 211

==> Selected Main Cable Anchor type and corresponding Drawings from the Table above:

Left Bank: Anchor Type _____________ Drawing No_______________


Right Bank: Anchor Type___________ Drawing No_______________

2. Walkway and Tower Foundation Design


A. Design Data

Fill in the Design Data from Survey Form and Checklist

ƒƒ Main Cable: Nos

ƒƒ Spanning Cable: Nos


Right Bank / Condition
Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Foundation Block Height (H) from the
1.0m 2.0m 3.0m 4.0m
Ground (data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Foundation Block Height (H) from the
Ground (data from bridge profile): 1.0m in case of Rock

Left Bank / Condition


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Foundation Block Height (H) from the
1.0m 2.0m 3.0m 4.0m
Ground (data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Foundation Block Height (H) from the
Ground (data from bridge profile): 1.0m in case of Rock

B. Selection of Walkway and Tower Foundations


Select the appropriate Walkway and Tower Foundations for the Right and the Left Banks according to the
above design data.

Procedure for selection:


¡¡ According to the Soil or Rock type, refer to the respective tables for the selection of the Walkway
and Tower Foundations.
In Soil : Table F.3.30
In Hard Rock : Table F.3.31
In Fractured Hard Rock or Soft Rock : Table F.3.32
¡¡ If the bank is rock, always take the foundation block height H as 1.0m,
If the bank is soil on a hill slope, always take the foundation block height H as 1.0m.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 212

If the bank is soil on flat ground but the foundation block height H needs to be raised, always take the
foundation block height as 1.0, or 2.0, or 3.0, or 4.0m (maximum).
¡¡ In the table, match the design data: Number and Diameter of Main Cables  Foundation Block
Height, H  if Soil, Soil type  select Foundation type and corresponding Drawing No. for right
bank and left bank respectively.

Walkway and Tower Foundations Selection Tables


¡¡ In Soil

Table F.3.30: Selection of Walkway & Tower Foundation in Soil

Foundation
Block Height
Main Cable Soil Type Foundation Type Drawing No
(H)
m
1.0 All TF-1 92/1Ncon
2.0 All TF-2 92/2Ncon
2x26 and 2x32mm
3.0 All TF-3 92/3Ncon
4.0 All TF-4 92/4Ncon
1.0 All TF-6 94/1Ncon
2.0 All TF-7 94/2Ncon
Gravelly/Sandy TF-8 94/3Ncon
4x26mm 3.0
Silty TF-9 94/4Ncon
Gravelly/Sandy TF-10 94/5Ncon
4.0
Silty TF-11 94/6Ncon
1.0 All TF-6 94/1Ncon
2.0 All TF-7 94/2Ncon
Gravelly/Sandy TF-8 94/3Ncon
4x32mm 3.0
Silty TF-9 94/4Ncon
Gravelly/Sandy TF-10 94/5Ncon
4.0
Silty TF-11 94/6Ncon

¡¡ In Hard Rock

Table F.3.31: Selection of Walkway & Tower Foundation in Hard Rock

Foundation Block Height


Main Cable (H) Foundation Type Drawing No
M
2x26 or 2x32mm 1.0 TF-5 92/5Ncon
4x26 or 4x32mm 1.0 TF-12 94/7Ncon

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 213

¡¡ In Fractured Hard Rock or Soft Rock

Table F.3.32: Selection of Walkway & Tower Foundation in Fractured Hard Rock or Soft Rock

Foundation Block Height


Main Cable (H) Foundation Type Drawing No
M
2x26 or 2x32mm 1.0 TF-1 92/1Ncon
4x26 or 4x32mm 1.0 TF-6 94/1Ncon

¡¡ Selected Walkway and Tower Foundations and corresponding drawings from the table above:

Right Bank: Foundation Type _________ Drawing No ________________


Left Bank: Foundation Type __________Drawing No________________

4. Windguy Anchorage Foundation Blocks Design (Optional)


A. Design Data
Fill in the Design Data from Survey Form and Checklist.

Windguy Cable: …..nos Ø …….mm

Right Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Anchor Block Height from Ground
1.0m 2.0m 3.0m 4.0m
(data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Anchor Block Height from Ground
(data from bridge profile): 1.0m in case of Rock

Right Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Anchor Block Height from Ground
1.0m 2.0m 3.0m 4.0m
(data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Anchor Block Height from Ground
(data from bridge profile): 1.0m in case of Rock

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 214

Left Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Anchor Block Height from Ground
1.0m 2.0m 3.0m 4.0m
(data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Anchor Block Height from Ground
(data from bridge profile): 1.0m in case of Rock

Left Bank / Upstream


Geology: Soil

If Soil, what is the Soil Type?


Gravelly Sandy Silty
Anchor Block Height from Ground
1.0m 2.0m 3.0m 4.0m
(data from bridge profile):
Hard Rock Hard Rock Soft Rock
If Rock, what is the Rock Type?
(only a few fractures) (highly fractures)
Anchor Block Height from Ground
(data from bridge profile): 1.0m in case of Rock

B. Selection of Windguy Cable Anchor Blocks (Optional only)


Provision of a windguy cable arrangement is optional.
Select appropriate Windguy Cable Anchor Blocks for the Right and Left Banks according to the
above design data. s are the rules for selection of the anchor blocks.
¡¡ If the bank is rock, always take block height as 1.0m.
¡¡ If the bank is soil in hill slope, always take block height as 1.0m.
¡¡ If one bank is soil and the other bank is rock, always select a Drum Anchor for the rocky bank.
¡¡ If both banks are rocky, select an Anchor Block for one bank and a Drum Anchor for the other bank.
¡¡ If the bank is soil but on flat ground and the Anchor Block Height needs to be raised, always take
the foundation block height as 1.0, or 2.0, or 3.0, or 4.0m (maximum).

Procedure for selection:


¡¡ According to the Soil or Rock type, refer to the respective tables for selection of the Windguy
Cable Anchor Blocks:

In Soil : Table F.3.33


In all types of Rock (Anchor Block) : Table F.3.34
In Hard Rock (Drum Anchor) : Table F.3.35
In Fractured Hard Rock or Soft Rock (Drum Anchor) : Table F.3.36

¡¡ In the table, match the design data: Diameter of Windguy Cable  Soil or Rock type  if soil
Anchorage Block Height  select Anchor Block type and corresponding Drawing No. for right
bank upstream and downstream and for left bank upstream and downstream respectively.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 215

Select the appropriate Windguy Cable Anchor blocks from the table.

¡¡ In Soil

Table F.3.33: Selection of Windguy Cable Anchor Block in Soil

Windguy Cable Anchor Block


Soil Type Anchor Type Drawing No
[mm] Height
1 All WAB - 1 55/1Acon
2 All WAB – 2 55/2Acon
26
3 All WAB – 3 55/3Acon
4 All WAB – 4 55/4Acon

¡¡ In Hard, Fractured or Soft Rock

Table F.3.34: Selection of Windguy Cable Anchor Block in All Types of Rocks

Windguy Cable
Anchor Type Drawing No
[mm]
26 WAB - 5 56Acon

¡¡ In Hard Rock

Table F.3.35: Selection of Windguy Cable Drum Anchor in Hard Rock

Windguy Cable
Anchor Type Drawing No
[mm]
26 WDR - 1 57Acon

¡¡ In Fractured or Soft Rock

Table F.3.36: Selection of Windguy Cable Drum Anchor in Fractured or Soft Rock

Windguy Cable
Anchor Type Drawing No
[mm]
26 WDR - 2 58Acon

¡¡ Selected Windguy Cable Anchor and corresponding Drawings from the table above:

Right Bank, Upstream: Anchor Type………..., Drawing No………………


Downstream: Anchor Type………..., Drawing No………………
Left Bank, Upstream: Anchor Type………..., Drawing No………………
Downstream: Anchor Type………..., Drawing No………………

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 216

5. Suspender List
Calculate the Suspender List using the computer program “Suspender Design” which is included in the
CD-ROM (available in separate form).

The calculation is based on the fixed geometry of the bridge and the design of the suspender.

Figure F.3.17: Typical Design of Suspender

The only input data required for the calculation is the span, l. The procedure for the Suspender List
calculation is as follows:
¡¡ Open the computer program SSTB Design > “Suspender Design”.
¡¡ Enter the input data: Bridge Name, Bridge Span and width “B” of the Walkway and Tower
Foundation.
¡¡ Print result.
Attach the printout of the Suspender List with the drawing of the suspenders (Drawing No 31N or 32N).

6. Design of Wind guy Arrangement


For special cases, i.e., when the bridge span is more than 120m or in extremely windy areas with wind
speeds in excess of 160 km/hr and unfavourable nature of the wind flow, a windguy system is mandatory.

For calculation of the Windguy Arrangement use the computer program “Windguy Design” which is
included in the CD-ROM which is available in separate form.

The calculation is based on the fixed geometry of the bridge.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 217

Layout of Windguy Arrangement

The calculation procedure is as:


¡¡ Open the computer program SSTB Design > “Windguy Design”.
¡¡ Enter the general input data: Bridge Name, Bridge Span and elevation of Steel Tower Base.
¡¡ Enter the tentative position of the first windtie BL and BR as per the topography. The program will
adjust to the nearest actual values.
¡¡ Give the level of the windguy anchorage on the right bank HL and the left bank HR. HL and HR
should be at or below the level of the spanning cable at the anchorage point (Es).
¡¡ Give (select) the windguy anchor position on the right bank (or the left bank), i.e. CR and DR or (CL
and DL). Accordingly, the windties listing will be from right bank to left bank or vise versa.
¡¡ Input the value VR (distance up to the vertex of the windguy cable) if the selected windguy anchor
position is on the right bank or VL if the selected windguy anchor position is on the left bank.
¡¡ Check the span and sag ratio of the wind guy cable, l/b. It should be not less than 8 and not more
than 10. If this condition is not met, input new V till l/b is within the recommended limit.
¡¡ Check the lowest point of the windguy cable Δhlp. It should not be negative (-). If Δhlp is negative,
input new HL and HR with a lower value.
¡¡ Input the value of DLo (or DRo) = 0. The result will show CLo (or CRo), and the horizontal angle αL (or
αR).
¡¡ Draw a straight line with the given CLo (or CRo), and horizontal angle αL (or αR). Fix the windguy
anchor position on the left bank (or the right bank) along this line. The anchorage block should be
placed on safe ground and at the optimum foundation location so that it has sufficient embedded
depth and also so that deep excavation can be avoided.
¡¡ After fixing the anchorage position, measure the actual CL (or CR). The result will give DL (or DR)
¡¡ Draw longitudinal sections along the windguy cables at the foundations and determine the exact
position of the front of the foundations (DR , CR and DL, CL) and the windguy cable elevations (HR,
HL).
¡¡ If the final position of the foundations does not match the design as above, repeat the calculation
process from the second bullet till all the foundations are located at the optimum positions.
¡¡ Calculate the Windguy Arrangement for both upstream and downstream.
¡¡ Attach the printout of the results (Windguy Arrangement) with the drawing of the Windtie cable
Clamps (Drawing No. 11A) and also transfer it to the General Arrangement Drawing.

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 218

7. Bridge Standard Drawings


Select the required Steel Drawings and Construction Drawings from the design as above and the drawing
list.

Steel Drawings

Required
Drawing Title Drawing No.
Drawing
Walkway and Anchors
Steel Crossbeams
Suspenders
Main Cable Anchor With Turn
Buckle
Main Cable Anchor With Direct
Cable Connection
Walkway and tower Foundation

Steel Deck

Towers
Assembly and Layout
Base Element
Intermediate Element
Top Element
Saddle

Optional
Windtie Cable Clamps for
One Cable Ø 26 or 32mm
Windguy Parts
Windguy Cable Anchorage for
one Cable Ø 26 or 32mm

Construction Drawings

Drawing Title Drawing No Required Drawing


Walkway Fitting
LB:
Main Cable Anchors
RB:
RB
Walkway and Tower Foundations
LB

PART F: TRAIL BRIDGE MANUAL


F - Chapter 3 - 219

Designed by: …………………………………. Date: ……………….

Cable hoisted by: ………………………………… Date: ………………..

PART F: TRAIL BRIDGE MANUAL


FEDERAL DEMOCRATIC REPUBLIC OF

ETHIOPIAN ROADS AUTHORITY

IT Y
E TH

OR
IO

IA
NR UT
H
P

OADS A

DESIGN MANUAL FOR LOW VOLUME ROADS


ROAD MAINTENANCE BOOKLET - PART G
FINAL DRAFT, AUGUST 2011
Part G
ROAD MAINTENANCE BOOKLET
G-i

G TABLE OF CONTENTS
Table of Contents........................................................................................... i

Foreword....................................................................................................... ii

Acknowledgements...................................................................................... ii

The Aims of this Booklet............................................................................... 1

Road Features............................................................................................... 2

Basic Access.................................................................................................. 4

The Purpose of Maintenance........................................................................ 9

Regular Maintenance.................................................................................. 10

Occasional Maintenance (Periodic)............................................................. 13

Road Maintenance Tools............................................................................. 14

Maintenance Activities................................................................................ 15

Drainage..................................................................................................... 23

Road Surface............................................................................................... 51

Priorities...................................................................................................... 91

Work Options.............................................................................................. 94

Planning and Productivity........................................................................... 99

Further Advice and Assistance.................................................................. 102

Terminology.............................................................................................. 104

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - ii

G FOREWORD
This first edition of this Road Maintenance Booklet has been prepared
for circulation and discussion with the aim of gathering comments and
contributions from stakeholders and potential users.

It is intended that the document will be refined and will be suitable for
translation into a number of local languages to facilitate application at
Wereda, Kebele and community level.

Please send any comments and contributions:-


Deputy Director General
Ethiopian Roads Authority
Ras Abebe Ageray Street
P O Box 1770 Addis Ababa
e-mail: bekelenegussie@ethionet.et

G ACKNOWLEDGEMENTS
Some material for this document has been taken from the World Road
Association (PIARC) International Road Maintenance Handbook. Additional
images have been provided by Intech Asset Management. This Booklet
has been prepared with the assistance of Intech Asset Management under
the direction of Ti-UP with funding of the UK Department for International
Development.

© Ethiopian Road Authority, 2011 (English language version)

Part or all of this document may be copied, adapted or translated, provided


the source is acknowledged.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 1 - 1

1. THE AIMS OF THIS BOOKLET


This Booklet has been developed by the Ethiopian Road Authority as one
of a number of initiatives to help rapidly expand, develop and maintain the
rural road network to provide greatly improved road access and lower the
transport costs for rural communities.

The Booklet specifically sets out guidance for Weredas, Kebeles and all
local communities on how to maintain their own road access using the
limited resources available to them. It also advises how it may be possible
to mobilise outside resources to enhance the impact of their own initiatives.

By focussing on the use and mobilisation of available local resources, such


as a range of materials and local labour and skills, it is entirely possible to
build and maintain durable all-weather road access suitable for all traffic
from pedestrians and animal transport up to buses and trucks.

The Booklet advises:


¡¡ What needs to be done to achieve all-year Basic Road Access;
¡¡ How to identify the main problems/defects and solve them;
¡¡ How to make the most of local materials and skills;
¡¡ How to maintain the road access at low cost;
¡¡ How to make priorities
¡¡ How to organise and plan the work
¡¡ Where to obtain further advice and outside assistance.

The Maintenance activities and codes are according to the ERA Maintenance
Technical Specifications and proposed amendments.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 2 - 2

2. ROAD FEATURES
Road Cross Section (imagine a vertical slice through the road)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 2 - 3

Drainage Features

The Terminology Section (13) provides the


explanation for each road term or feature.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 3 - 4

3. BASIC ACCESS
Basic Access is achieved to provide year-round passage to routes by turning
them from weather-dependent tracks into proper roads. A proper road can
be formed from the natural soil (Engineered Earth Road) in many locations.

The main features of a road are:-


¡¡ A camber to shed rainwater to each side of the road
¡¡ Side drains (ditches), mitre drains (turnout drains), drifts and culverts
to manage the water collected from the road surface and to discharge
it carefully to avoid erosion or other problems.

This usually means that the road surface needs to be slightly higher than the
ground at the road side.

Most natural soils can be built into an (Engineered Natural Surface) Earth
Road. However, for route sections with weak soils, or if traffic increases to
more than about 50 motor vehicles per day, or on steep hills, it may be
necessary to improve the road surface with gravel or various types of paving.
This can be achieved at relatively low cost by applying a Spot Improvement
approach to improve these limited problem sections, using local labour and

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 3 - 5

materials. Other documents such as the ERA LVR Design Manual (2011)
describe how such improvements can be designed and constructed.

Most routes can be built to Engineered Natural Surface (ENS) – Earth road
standard for most of their length. If in doubt about soil suitability, seek advice
from the regional road authority. The Spot Improvements at problem
sections of the route can be selected from the following list of options:
¡¡ Drift
¡¡ Culvert
¡¡ S-01: Engineered Natural Surface (ENS)
¡¡ S-02: Natural gravel
¡¡ S-03: Waterbound/Drybound Macadam
¡¡ S-04: Hand Packed Stone
¡¡ S-05: Stone Setts or Pavé
¡¡ S-06: Mortared Stone
¡¡ S-07: Dressed stone/cobble stone
¡¡ S-08: Fired Clay Brick, Unmortared/mortared joints
¡¡ S-09: Bituminous Sand Seal
¡¡ S-10: Bituminous Slurry Seal
¡¡ S-11: Bituminous Chip Seal
¡¡ S-12: Bituminous Cape Seal
¡¡ S-13: Bituminous Otta Seal
¡¡ S-14: Non-reinforced concrete
¡¡ S-15: Ultra-thin reinforced concrete

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 3 - 6

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 3 - 7

The choice of spot improvement should be based on the location features


and the materials and skills available locally. Great care should be used in
using gravel as a road surface in some circumstances. It is unlikely that it will
be suitable due to high costs of replacing the surface material that will be
lost due to rainfall or traffic, or dust nuisance in the locations; where:
¡¡ Traffic is more than 200 motor vehicles per day
¡¡ Annual Rainfall is greater than 2,000mm
¡¡ Slope of road surface is more than 6%
¡¡ Through community settlements
¡¡ The haul distance from the quarry/pit to the road site is more than
10km
¡¡ The road section experiences flooding
¡¡ The gravel is of poor quality.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 3 - 8

Despite initial low construction costs, it is important to appreciate that under


a Basic Access and Spot Improvement strategy it is essential to arrange
the necessary Regular maintenance of the ENS and any gravel surface and
drainage, and Occasional maintenance of the improved surface sections, to
preserve the initial construction investment.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 4 - 9

4. THE PURPOSE OF MAINTENANCE


From the moment that a road is constructed or upgraded, it will deteriorate
due to the effects of weather and traffic. Maintenance is required to be
carried out from time to time to restore its condition to be close to its as-
constructed state. If maintenance is not carried out the road will continue
to deteriorate making passage increasingly difficult, uncomfortable and
expensive to road users. The road may even become impassable for part
or all of the year.

It is convenient to view Maintenance as correcting Defects.

In practical terms it is useful to identify and quantify the Defects, and then
arrange the necessary Maintenance to be carried out. In this booklet defects
and Maintenance activities are grouped into the following:

Regular Maintenance
¡¡ Roadside activities
¡¡ Drainage
¡¡ Road surface
• Earth Road
• Gravel Road

Occasional (Periodic) Maintenance


¡¡ Road surface
• Gravel Road
• Paved Road

From time to time, other activities not covered by this booklet may be
required. Advice should be obtained from the regional road authorities for
any problem or road aspect not covered in this booklet.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 5 - 10

5. REGULAR MAINTENANCE
These are the maintenance activities that are likely to be required somewhere
on a road link every year. Most of the tasks may be carried out manually.
Mechanised alternatives are available for some tasks as indicated.
Regular Maintenance is conveniently divided into three main groups of
activities that are often carried out on a seasonal basis:
¡¡ Roadside activities
¡¡ Drainage
¡¡ Road surface
• Earth Road
• Gravel Road

Roadside Activities
Defect Maintenance Activity (Page No.)
1. Trees and bushes growing on
131 Brush clearing (P15)
roadside
2. Shoulder uneven or eroded, or 132 Shoulder Rehabilitation
does not drain properly (manual) (P19)
133 Plant grass and water
3. Shoulder erosion
(P21)
1. Grass on shoulder or in drain
134 Cut grass (P15)
requires cutting
2. Shoulder uneven or eroded, or 240 Shoulder Blading
does not drain properly (minor) (mechanised) (P19)
2. Shoulder uneven or eroded, or 241 Shoulder Rehabilitation
does not drain properly (major) (mechanised) (P19)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 5 - 11

Drainage

Defect Maintenance Activity (Page No.)


4. Culvert silted/obstructed 121 Culvert Cleaning (P23)
122 Ditch Clearing (Manual)
5. Ditch silted
(P26)
6. Ditch or slope eroded 123 Repair Erosion Damage
(minor) (Selected Fill) (P34)
7. Ditch or slope eroded 124 Repair Erosion Damage
(major) (rockfill) (P35)
7. Slope eroded (major) 129 Wattling (P35)
125 Mortared Masonry Repair
8. Mortared Masonry damaged
(P40)
9. Dry Masonry damaged 126 Dry Masonry Repair (P43)
127 Gabion Structure Repair
10. Gabion structure damaged
(P45)
128 Build wooden/stone scour
11. Erosion in ditch
check (P48)
230 Ditch clearing (mechanised)
5. Ditch silted
(P30)

Road Surface
Defects and maintenance requirements will depend on the road surface
type.

Earth Road
Defect Maintenance Activity (Page No.)
12. Road surface potholed, rutted
112 Reshape & Compact Earth
or uneven, and does not drain
Road Camber (P51)
to edge

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 5 - 12

Gravel Road

Defect Maintenance Activity (Page No.)


13. Road Surface potholed 110 Spot Repair Selected Material
(P59)
13. Road Surface potholed 111 Spot Repair Crushed
Aggregate (P59)
14. Road Surface rutted or uneven, 220 Blade Gravel Road (light) (P63)
and does not drain to edge
(Minor: <3cm)
15. Road Surface rutted or uneven, 221 Blade Gravel Road (heavy)
and does not drain to edge (P68)
(Major: >3cm)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 6 - 13

6. Occasional Maintenance (Periodic)


These are the maintenance activities that may be required somewhere on a
gravel or paved road section or link after a period of a number of years. The
category of repair depends on the type of road surface constructed. All of
the Occasional Maintenance tasks may be carried out manually with the aid
of simple tools or equipment.

Gravel or Paved Road

Defect Maintenance Activity (Page No.)


16. Gravel layer too thin 317 Gravel Resurfacing (Selected
Material) (P77)
318 Gravel Resurfacing (Crushed
Aggregate) (P77)
17. Paved road pothole or surface 113a Spot /pothole Repair
defect (Macadam) (P84)
113b Spot /pothole Repair
(Stone setts)
113c Spot /pothole Repair
(Mortared stone)
113d Spot /pothole
Repair(Dressed stone)
113e Spot /pothole
Repair(Emulsion chip seal)
113f Spot /pothole
Repair(Emulsion sand seal)
113g Spot /pothole
Repair(Emulsion gravel/
slurry seal)
113h Spot /pothole Repair(Un-
mortared brick)
113i Spot /pothole Repair(Mortar
jointed brick)
113j Spot /pothole Repair(Non-
reinforced concrete)
217 Pothole Reinstatement (cold
mix)
219 Pothole (Base Failure
Repair)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 7 - 14

7. ROAD MAINTENANCE TOOLS


Road Maintenance activities require a range of simple and inexpensive
tools and control aids. However construction quality tools are preferable to
agricultural quality.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 15

8. MAINTENANCE ACTIVITIES
Defect 1: Grass, weeds, bushes or trees have been allowed to grow
unchecked at the side of the road.

Development, if neglected:
¡¡ Drainage ditches cannot be cleaned;
¡¡ Surface water can pond at the edge of the road and weaken the road
surface;
¡¡ Silt can accumulate at the edge of the road;
¡¡ The visibility for road users is reduced, with increased risk of accidents
with persons or animals;
¡¡ Increased fire hazard in the dry season.

Maintenance Activity
¡¡ 134 Grass Cutting
¡¡ 131 Bush Clearing

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 16

These two activities may be required individually or together. With the


exception of arid areas, these are Regular activities, involving control of
grass, weeds, bush and trees where these are not controlled by animal
grazing. They may be required to be carried out at least once a year after
the rainy season, or more often where the climate causes vegetation to grow
rapidly.

¡¡ Grass Cutting
Grass and weeds should be cut at least once a year after vegetation
reaches full growth or according to local experience. The vegetation
should be trimmed by hand. Sickles, scythes, slashers, bushknives, or
similar handtools will be required.
¡¡ Bush control & Trees
Any bushes on the road shoulders or drains should be cut down.
Dead or leaning trees within the right-of-way which may fall on the
roadway or block the drainage system, or block sight lines should
be removed. The felling of trees, or the removal of large branches at
heights of more than 2 metres above ground level can be hazardous.
This work should only be carried out under expert supervision or by
experienced workers. Trees should be felled using one- or two-man
saws or axes. Ladders should be used for climbing trees, and ropes
should be used to restrain trees and control felling. Traffic should be
halted when the tree is finally toppled. All debris should be removed
and disposed of safely.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 17

¡¡ Disposal of debris
All cuttings and debris should be disposed of safely so that there is
no risk of drains being blocked or fire hazard.
¡¡ Herbicides
Herbicides (weed-killer) are chemical agents intended to destroy or
reduce vegetation growth. It is not recommended that herbicides
or any chemical methods be used to control roadside vegetation.
Some reasons are:
• Herbicides can cause pollution of crops, rivers and streams
and drinking water supplies;
• Herbicides are often dangerous to health;
• Herbicides are expensive, and must often be imported;
• Herbicides do not always produce satisfactory results.
¡¡ Burning
Do not burn roadside vegetation to control its growth or the debris
from Bush Clearing activities. The results may be more harmful than
desired:
• The fire could spread and destroy valuable vegetation (trees,
grass crops), and traffic signs;
• Vegetation may grow faster after burning;
• Smoke and flames blowing across the road are dangerous for
traffic.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 18

Defect 2: Shoulder eroded, mis-shaped or does not drain away from


roadway

Development, if neglected:
¡¡ Hazard to road users, increased risk of accidents;
¡¡ Obstruction of water flow from roadway;
¡¡ Inadequate support for the road surface;
¡¡ Water collects and softens/weakens the shoulder and pavement;
¡¡ The edge of the pavement will break when vehicle wheels run over it;
¡¡ The roadside ditch may become blocked by the excess material.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 19

Maintenance Activity
¡¡ 132 Shoulder Rehabilitation (manual);
¡¡ 241 Shoulder Rehabilitation (mechanised);
¡¡ 240 Shoulder Blading (mechanised).
a) Rehabilitation Manual Method (132)
The low surfaces and all high material should be loosened with a pick axe or
mattock. The shoulder should be reshaped to slightly above the final level
and the correct crossfall using a shovel and rake. Any low spots should be
topped up with fresh material of the same type and quality as the existing
shoulder. The crossfall of the uncompacted material should be checked with
a camber board. Excess material should be spread over the embankment
slope or transported by wheelbarrow to a convenient and safe dumping
site. Material should not be deposited on the roadway or in the drainage
ditch. If the material is dry it should be sprinkled with water. The shoulder
is then compacted with hand rammers or a hand roller. The compacted
surface should butt smoothly onto the roadway. Check the finished crossfall
with the camber board and repeat the reshaping if necessary. Brush all loose
material and debris from the roadway.

b) Rehabilitation Mechanised Method (241)


The existing surface of the shoulder should be scarified with the tines of
a motor or towed grader. This will loosen the raised areas and allow the
loosened material to key into any existing low areas. The shoulders should
be reshaped to slightly above the final level and the correct crossfall using a
number of passes of the motor or towed grader blade. Care must be taken
not to damage the edge of the roadway with the blade. Any low spots
should be topped up with fresh material of the same type as the existing
shoulder. The cross fall of the uncompacted material should be checked
with a camber board. Excess material and vegetation should be graded to
the embankment side slope. In cuttings, excess material and vegetation
should be graded into a windrow for removal by wheelbarrow, tractor and

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 20

trailer or truck. Material should not be deposited on the roadway or into the
drainage ditch.

If the material is dry it should be sprinkled with water. The shoulder is then
compacted using a self-propelled, towed or pedestrian roller. The compacted
surface should butt smoothly onto the roadway. Check the finished crossfall
with the camber board and repeat the reshaping if necessary. Brush all loose
material and debris from the roadway.

c) Shoulder Blading (mechanized) (240)


This Regular maintenance activity may be carried out if no additional
material is required to be added to the shoulder. The shoulder material
should contain sufficient moisture to enable the reshaped material to be
compacted by the grading equipment or a roller. It is therefore ideally
carried out in the rain season. Otherwise, water should be added to ensure
a more durable surface finish.

Defect 3: Existing roadside surface requires protection from erosion (this


activity may be required as a follow up from Maintenance Activity 123)
Development, if neglected:
On some steep slopes or erodible soils surface scour may occur if vegetation
cover is not established. This could damage to the roadway, shoulders,
drainage system or earthworks.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 21

Maintenance Activity
¡¡ 133 Plant grass and water.
a) Seeding
Grass seeding will only be successful if climate and soil conditions are
favourable. The best advice can be provided by the local department of
agriculture on:
¡¡ Topsoil required;
¡¡ Seed type, rate of spread;
¡¡ Fertilizer types, rate of spread;
¡¡ Most favourable season and weather for seeding;
¡¡ Other preparatory treatment of the soil (for example mixing-in
ground limestone).

Typical procedure:
¡¡ Loosen the soil to a depth of 10 cm in the area to be seeded using
rakes or similar tools;
¡¡ Spread the topsoil to a depth of at least 5 cm;
¡¡ Water the area to be seeded;
¡¡ Apply fertilizer at the specified rate;
¡¡ Apply ground limestone at the specified rate and mix-in;
¡¡ Apply seeds by hand at the specified rate;
¡¡ Lightly roll the seeded area within 24 hours using hand roller, only if
the soil does not adhere to the roller;
¡¡ The seeded area should be watered as required until the grass has
taken hold.

b) Turfing (grass sodding)


This method is suitable when climate and soil conditions are favourable
and when fresh grass sods (soil clumps containing grass and its roots) are
available.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 8 - 22

The general procedure is:


¡¡ Prepare the area to be turfed to required levels and slopes;
¡¡ Where no topsoil is present, haul suitable topsoil to site and spread
evenly to a depth of not less than 5cm;
¡¡ Water as required;
¡¡ Cover the area with freshly cut sods without weeds. Sods are to have
thickly matted roots which should not have dried out. Tamp sods
with tamper or use hand roller. On slope use stakes to hold sods in
position;
¡¡ Water the turves at intervals until the grass takes hold.

Other patterns of sodding are:


¡¡ Spot sodding (sods spaced about 50cm in holes deep enough to
take sod and about 5cm topsoil);
¡¡ Trench sodding. Lay sods on 5cm topsoil bed in parallel trenches;
¡¡ Trench spacing about 50cm along contour or x-shaped pattern.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 23

9. DRAINAGE
Defect 4: Culvert silted or obstructed with debris

Development, if neglected:
The intended waterway opening will be so reduced that flood water cannot
flow as intended. Flood water will back-up or pond on the upstream side of
the culvert and may eventually over-flow the road embankment. The road is
then in danger of being washed away.
Maintenance Activity
¡¡ 121 Culvert Cleaning/
In order to function properly, a culvert must retain the full opening over its
complete length. In addition, the upstream approaches and the downstream
area must be free of obstructions. Floating debris (tree branches, bushes,
etc.) carried by water is a great danger to culverts. The debris may completely
block the culvert inlet. The following Regular Maintenance activities may be
required:
¡¡ If debris racks are already provided, these as well as the culvert
opening should be freed of all accumulated obstructions.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 24

¡¡ Clear any sand, silt or debris from inside the culvert. Sanding or silting
of culverts, especially those with openings smaller than 1 metre, can
be a particular problem. These culverts can be cleaned by pulling a
cable or rope through, to which is attached any suitable object (e.g.
a bucket). Alternatively a long handled trowel and spike can be used.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 25

If the silting problem continues despite regular clearing, it may be


necessary to reconstruct the culvert at a higher level or enlarge it.

Material and debris from the culvert must be spread or dumped


where they cannot cause an obstruction to water flow, preferably on
the downstream side of the culvert, well away from the watercourse.

This Maintenance task is best carried out before the rains and after
any heavy rainstorm.

Defect 5: Ditch silted


Ditch partially or fully blocked by vegetation growth, bushes, fallen trees,
debris, loose silt, loose rocks.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 26

Development, if neglected:
Concentration of flood flow causing erosion and possible overtopping and
damage to roadway, paving or shoulders.
Maintenance Activity
¡¡ 122 Ditch Clearing (Manual);
¡¡ 230 Ditch clearing (Mechanised).

a) Manual Method (122)


This is a Regular Maintenance activity. The object is to remove all soil,
high vegetation, materials and objects from the ditch which could possibly
interfere with water flow or cause an eventual blockage of the ditch. This
can include for example, rocks, loose silt and sand, weeds, trees, bushes,
including their roots, etc. Dispose of these materials well away from the
roadside so that water flow will not be impeded and they will not fall or
wash back into the ditch. NO soil material or debris should be placed on
the roadway. On unlined ditches a short grass cover can help to stabilise the
bottom and sides of the ditch. Therefore where a side ditch is established to
the correct depth and profile with grass cover and no erosion, it is advisable
to merely cut the grass short. This will leave the roots in place to bind the
drain surface together.

At some locations it may be necessary to RESHAPE/REGRADE/DEEPEN the


ditch.

It is advisable to adopt a trapezoidal ditch shape when using labour


methods. The excavation using a hoe/mattock and shovel is easier than
for a V-shaped ditch. An added advantage is that the flat invert causes less
concentration of water than a V-ditch.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 27

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 28

A ditch & slope template should be used to obtain the correct drain shape.
¡¡ Using the template a 50 cm wide slot should be excavated to the
correct ditch shape every 10 metres along the drain. The slots act as
a guide for excavating the ditch to the correct shape;
¡¡ In flat areas, the gradient of the ditch should be checked using
ranging rods and profiles or similar methods, to ensure that water
will not pond. The levels at adjacent slots should be checked using
a line and level or abney level, and the level of the slot adjusted if
necessary.
¡¡ Excavate all surplus material between the slots and to the correct
shape with the aid of stringlines stretched between the slots. If
necessary the intermediate invert levels can be checked using a
traveller sighted between the ranging rod profiles.
¡¡ Material excavated from the drain must be removed and spread well
clear of the drain so that it cannot later fall or wash back into the
ditch.
¡¡ The shape can be checked during the excavation activity using the
ditch template.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 29

When excavating a completely new ditch it is preferable to split the task into
two operations:
i) Cut the central rectangular shape and check with a template
(INVERT);
ii) Cut the slopes and check with the full ditch & slope template
(SLOPES).

The alignment or route of the drain should be set out using stringlines and
pegs. The ranging rods and profiles should be set up at the start and outfall
of the ditch. Intermediate profiles may be required on long ditches. The
levels of intermediate slots can be determined using the traveller.

This Maintenance task is best carried out before the rains and after any
heavy rainstorm.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 30

b) Mechanised Method (230)


This activity is suggested where long sections of V- shaped ditches are to
be maintained and cleaned and where high daily outputs are possible. The
activity may be carried out by a motor or towed grader. The grader should
always work by cutting in the direction of water flow in the ditch.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 31

Case 1:
When the grader can operate only on the shoulder and in the ditch, but not
beyond the ditch:
¡¡ Start by grading the outside slope of the ditch, using the blade to
windrow the soil to the bottom of the ditch between the rear wheels.
(This can be repeated to obtain the desired depth of ditch. This part
of the task can also be done manually);
¡¡ The next blade pass(es) are to clean the invert of the ditch by
removing the windrow to the top of the ditch at road shoulder;
¡¡ The final pass is required to move the windrow material away from
the shoulder ditch edge.
If the recovered material quality is inferior to that of the road surface, THE
MATERIAL MUST BE REMOVED FROM THE SITE.
¡¡ On completion, the ditch should generally have a depth of 50 cm
(minimum), which can be checked with a ranging rod and tape/rule;
¡¡ If necessary the grade of the ditch invert can be checked using the
methods described in a) Manual Method (122).

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 32

Case 2:
When the grader can operate beyond the ditch. Reverse the operations
shown previously:
¡¡ Grade the inside slope, windrowing material to the bottom of the
ditch. Repeat as necessary to achieve the desired depth of ditch;
¡¡ Remove the windrow material to the top of the outside slope;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 33

¡¡ Move windrow away from ditch edge and spread the material so that
it will not wash back into the ditch;
¡¡ On completion, the ditch should have a depth of 50 cm (minimum),
which can be checked with a ranging rod and tape/rule;
¡¡ If necessary the grade of the ditch invert can be checked using the
methods as described in a) Manual Method (122).

Defect 6: Ditch or slope eroded (minor)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 34

Development, if neglected:
Damage to drainage system, roadway, structures, paving or shoulders.
Maintenance Activity
¡¡ 123 Repair Erosion Damage (Selected Fill).

This activity may be sufficient for minor erosion damage to a ditch. However,
reconstructing the ditch profile with selected fill material alone may not be
sufficient to prevent the defect recurring within a short time. Loose stones
or boulders should be removed. The defective section of ditch should be
cut back to firm material and fresh material placed in layers not exceeding
15cm thickness and compacted with a hand rammer. If the material is dry
it should be sprinkled with water before compaction. The added material
should be trimmed back to the correct ditch profile and checked with the
ditch template. Dispose of the excess materials well away from the roadside
so that water flow will not be impeded and they will not fall or wash back
into the ditch.

Check Dams (scour checks) may need to be installed to prevent recurrence.


Similar minor repairs may be carried out to eroded slopes.

It is likely that additional measures will be required such as:


¡¡ 33 Plant grass and water;
¡¡ 124 Repair Erosion Damage (rockfill);
¡¡ 128 Build wooden/stone Check Dam;
¡¡ 135 Wattling.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 35

Defect 7: Ditch or slope eroded (major)

Development, if neglected:
Damage to drainage system, roadway, structures, earthworks paving or
shoulders.
Maintenance Activity
¡¡ 124 Repair Erosion Damage (rockfill);
¡¡ 129 Wattling.

A number of activities are possible to repair erosion damage to ditches and


slopes. However it is important to try to determine the cause of the erosion
so that the repair will minimize the risk of it recurring. It is advisable to obtain
engineer’s advice where erosion is extensive.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 36

Ditches (Open Drains)


Drain sections are often laid at a steep gradient or on sharp bends without
erosion protection along or at the outfall of the drain. The following options
should be considered:
¡¡ Repair the ditch with rockfill lining
¡¡ Repair the ditch with timber lining;
¡¡ Provide masonry lining;
¡¡ Regrade/Realign ditch;
¡¡ Provide relief ditch or culvert.
Slopes

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 37

Slips or slope erosion/ instability are usually caused by adverse ground


conditions or ground/surface water or both. The remedial works should be
specified by an Engineer after an inspection of the site and the necessary
investigations. Works are likely to be expensive and it is important to ensure
an appropriate solution is adopted.

Dealing with slips and unstable slopes is hazardous and particular care
should be taken to safeguard manpower, equipment and the road users.
The cheapest solution (if appropriate) is expected to be Wattling. Other, but
more expensive, remedial options are:
¡¡ Counterfort drains;
¡¡ Stone pitching the slope;
¡¡ Reducing slope angle;
¡¡ Clearing slip material;
¡¡ Surcharging the slope;
¡¡ Gabions;
¡¡ Cribwork;
¡¡ Masonry retaining wall;
¡¡ Concrete retaining wall.

Wattling (129)
This activity may be suitable after repairing an eroded slope with Maintenance
Activity: 123 Repair Erosion Damage (Selected Fill)

Wattling will help to resist surface water erosion of a slope. Wattles are
bundles of plant stems up to 3 m long, tied together and laid in shallow
trenches, staked into position on contour lines (lines of the same height),
or x - form lines. As with turfing and seeding, a favourable climate and soil
conditions are essential for the successful use of wattling.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 38

Wattling helps to stabilize slopes, reduce surface erosion and provides a


bench on which grass can become established. Plant stems which root
easily are preferred. Advice on suitable plants and planting time should be
obtained from the local department of agriculture.

Typical procedure:
¡¡ Cut wattling stems at suitable source and transport them to site
immediately. Stems should not be allowed to dry out;
¡¡ Tie bundles of stems 15 - 20 cm diameter, alternating the ends;
¡¡ Excavate a trench in the slope along the desired line. The trench
should be deep enough to accommodate tied wattling stems (this
work can be completed beforehand);
¡¡ Place wattling stems in trench and use stakes to fix them in position.
Overlap bundles and stake through the overlaps;
¡¡ Cover the wattling with topsoil and tamp them firmly in place;
¡¡ Watering may be necessary until the roots take hold.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 39

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 40

Defect 8: Mortared Masonry damaged

Development, if neglected:
Further damage to structure or roadway, slope or structural failure.
Maintenance Activity
¡¡ 125 Mortared Masonry Repair

This activity should only be carried out on masonry structures in reasonably


good condition. It is recommended that the structure be completely rebuild
if it is in danger of collapse.
¡¡ Clean and rake out defective joints of weak mortar, soil and vegetation
using compressed air or a water spray, hammer and chisel;
¡¡ At locations where the joint has to be completely renewed, the stone
or brick should be eased out of place temporarily until a new mortar
bed is placed;
¡¡ Replace missing stones with sound pieces;
¡¡ Use templates and stringlines if necessary to ensure the correct
shape and incline of the face of the mortared masonry work;
¡¡ Dampen the joint surfaces where fresh mortar has to be applied;
¡¡ Mix a mortar of cement and sand as required or specified (normally 1
cement: 4 sand) and add only enough water to permit mortar to be
well mixed and applied;
¡¡ Apply fresh mortar to joint, filling all space available, compacting
with a suitable wooden rammer. Do not use mortar which has fallen
on the ground;
¡¡ Smooth joints with a suitable tool (a piece of rubber or plastic water
hose, or bent reinforcing steel);
¡¡ The final mortar surface should be inset slightly from the stone/brick
surface to achieve a tidy finish;
¡¡ In dry or windy weather conditions, mortar can dry out too quickly.
Prevent this by sprinkling water on joints after the mortar has set and

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 41

until mortar has completely hardened. Alternatively cover the work


area with wet jute sacks or similar;
¡¡ Clean visible stone or brick surfaces which have been stained
by mortar or cement-water in the process of the work so that the
finished work will present a neat appearance;
¡¡ Remove surplus materials and leave the site in a clean and tidy
condition.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 42

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 43

Defect 9: Dry Masonry damaged

Development, if neglected:
Further damage to structure or roadway, slope or structural failure.
Maintenance Activity
¡¡ 126 Dry Masonry Repair.

Try to use the established local dry stone construction techniques.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 44

This activity should only be carried out on dry masonry structures in


reasonably good condition. It is recommended that the structure shall be
reconstructed if it has settled or is in danger of collapse.

¡¡ Carefully take down the defective areas of dry stone masonry,


stacking the stone for re-use;
¡¡ Clean and rake out defective joints of soil and vegetation using
hammer and chisel, and brush;
¡¡ Re-build the dry stone work using the salvaged stones and carefully
selecting each stone to ensure good bonding horizontally and
through the width of the stonework. Use smaller stones to wedge
the larger ones where necessary;
¡¡ Add new stones if necessary;
¡¡ Use templates and stringlines if necessary to ensure the correct
shape and incline of the face of the dry masonry work;
¡¡ Pack the spaces between stones with soil or gravel;
¡¡ Remove surplus materials and leave the site in a clean and tidy
condition.

Defect 10: Gabion structure damaged


Development, if neglected:
Further damage to structure or roadway, slope or structural failure.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 45

Maintenance Activity
¡¡ 27 Gabion Structure Repair.

Gabions are usually made of zinc coated steel baskets, although they may
also be made from welded mesh sheets, galvanised chainlink fencing and
woven wire depending on the circumstances and locally available materials.
The baskets are hand-filled with rock and stones between 12 and 30 cm size.

In this way they attain great stability, but will allow minor settlement. Repairs
may be required due to bulging or breaking of the basket due to foundation
or backing movement, or settlement of the stone within the basket. Gabions
are designed to allow some settlement. Repairs should aim to ensure that
the stone continues to be contained. Repairs will normally consist of opening
the baskets, re-packing the stone inside, topping up stone if necessary and
re-securing the lid of the gabion. It may be necessary to weave new cage
material over broken or deformed areas, and any suitable steel mesh or
woven sheets can be used for this.

Where a gabion box is required to be replaced or added, the procedure for


building a new gabion box should be used as follows.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 46

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 47

The gabion baskets are normally supplied folded flat complete with
tying wire so that the transport volume is minimised. Foundations should
be excavated level and cleaned as for a conventional structure, with any
unsuitable material removed and replaced with good soil, stone or gravel,
and compacted. The baskets should be erected in their final position.

Cages should be woven together using 3 mm binding wire securing all


edges every 15 cms with a double loop. The binding wire should be drawn
tight with a pair of heavy duty pliers and secured with multiple twists (1 and
2). The centre gabion only should be filled initially to act as an anchorage.
The connected baskets should be stretched and staked with wires and pegs
to achieve the required shape (3). Filling should be carried out by hand
using hard durable stones not larger than 250 mm and not smaller than the
size of the mesh. The best size range is 125 to 200 mm. The stones should
be tightly packed with a minimum of voids. Boxes of 1 metre height should
be filled to 1/3 height. Horizontal bracing wires should then be fitted and
tensioned with a windlass to keep the vertical faces even and free of bulges
(4 and 5). Further bracing should be fixed after filling to 2/3 height. 500
mm height boxes should be braced at mid height only. 250/330 mm deep
gabions do not require internal bracing. The stones should be carefully
packed to about 3 to 5 cms above the top of the box walls to allow for
settlement. Smaller material can be used to fill the voids on the top face, but
excessive use of small stones should be avoided. The lids are then closed
and stretched tightly over the stones, (carefully) using crowbars if necessary
(6). The corners should be temporarily secured to ensure that the mesh
covers the whole area of the box. The lid should then be securely woven to
the tops of the walls removing stones if necessary to prevent the lid from
being overstretched.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 48

Defect 11: Erosion in ditch

Development, if neglected:
Damage to drainage system, roadway, structures, paving or shoulders.
Maintenance Activity

¡¡ 128 Build wooden/stone scour check (Check Dams).

Unlined ditches may suffer from scour of the invert and sides. Simple repairs
may be achieved by filling the affected areas with soil and trimming to the
correct profile, and turfing where climatic conditions are favourable. The
turves will probably need to be pegged in place to retain them, and watered
until established. Simple Check Dams (scour checks) may be constructed of
wood or stones. Larger ones may be constructed of stone masonry, brick
or concrete. They reduce the speed and erosion force of the water. They
also hold back the silt carried by the water flow to provide a series of gently
sloping sections of ditch separated by steps.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 49

The Check Dams must not be too high otherwise water will be forced onto
the surrounding ground, the shoulder or the roadway. The Check Dam
construction should therefore be controlled with the aid of a template.
Check Dams should not be constructed on ditches with gradients of less
than 4%. This will encourage too much silting of the drain and could lead
to road damage. The gradient of the side drain should be checked with an
abney level or line and level to determine the requirements for scour
checks (spacing guidance in the ERA LVR Design Manual).

After the basic Check Dam has been constructed, an apron should be built
immediately downstream either using stones or grass turves pinned to the
ditch invert with wooden pegs. The apron will help resist the forces of the
water flowing over the Check Dam. Grass sods should be placed against
the upstream face of the Check Dam, to prevent water seeping through the

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 9 - 50

Check Dam and to encourage the silting behind the Check Dam. The long
term objective is to establish complete grass cover over the silted Check
Dams to stabilise them.

Well constructed Check Dams will allow the water to gently cascade over
(and not through) the checks, removing energy from the water and reducing
erosion power.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 51

10. ROAD SURFACE


Whenever works are carried out on the road
surface warning signs should be placed before
each end of the work site.

EARTH ROAD
Defect 12: Road surface potholed, rutted or
uneven, and does not drain to edge

Development, if neglected:
Road becomes waterlogged or impassable.
Maintenance Activity
¡¡ 112 Reshape & Compact Earth Road Camber.

This activity is carried out using labour, basic hand tools and control aids.

The Method comprises the following steps:


¡¡ Setting out;
¡¡ Excavation of ditch and slope;
¡¡ Excavation of backslope;
¡¡ Camber formation and final compaction.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 52

These steps are shown on this page.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 53

Setting out
¡¡ The PROFILE method of setting out enables a smooth vertical
alignment to be re-established on a severely deteriorated road
surface;
¡¡ The alignment will consist of straight gradients and vertical curves;
¡¡ The centre line of the road is pegged every 10 metres;
¡¡ A ranging rod is fixed at each 10 metre peg;
¡¡ Each ranging rod is fitted with a profile board. The profile board can
slide up and down the ranging rod and be clamped at any height.

Setting out is arranged in sections of 60 to 100 metres, which approximate


to either a straight gradient or vertical curve on the road line.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 54

Check that the amount of earthworks at each centre line (finished level) peg
is acceptable, or repeat the procedure using different assumptions.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 55

Once the centre line level pegs are fixed, set out the pegs for the edge of
the roadway and both sides of the ditch using the tape measure, camber
board and spirit level for the required road cross section.

Pegs should be driven in to the required finished cross section level, or a


fixed height above.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 56

EXCAVATE DITCH AND SLOPE


¡¡ Material is excavated from the ditch and slope area and used to form
the camber until the required shape of ditch and slope is achieved;
¡¡ Check shape with the ditch and slope template, and spirit level;
¡¡ If too much material is excavated, discard the surplus material well
beyond the side drain;
¡¡ If the filling placed is greater than 15 cm deep, then it is preferable
to spread and compact the fill material with rakes and hand rammers
or a hand/animal drawn roller in 15-20 cm layers.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 57

Excavate backslope
If insufficient material is excavated to form the camber, dig additional
material from the backslope or from beyond the side drain.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 58

Camber formation and final compaction


¡¡ Continue adding material to the camber to achieve the required
profile after compaction;
¡¡ Stringlines stretched directly and diagonally across the running
surface between the setting out pegs can be used to check the
shape;
¡¡ Compact the fill material to the final profile, preferably using a hand
or animal drawn roller;
¡¡ If a roller is not available, use hand rammers or the tyres of any
vehicle to uniformly compact the soil across the roadway width.

The shaped and compacted earth road surface is a suitable foundation on


which one of the surface options can be constructed directly.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 59

Gravel Road
Defect 13: Road Surface potholed

Development, if neglected:
Gravel surface loss increases. Road becomes very rough, slowing and
damaging traffic, and may become waterlogged or impassable.
Maintenance Activity
¡¡ 110 Spot Repair – Selected Material;
¡¡ 111 Spot Repair – Crushed Aggregate.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 60

Potholes and ruts should be repaired with materials similar to the surrounding
surface. This can be either selected gravel material (110) or crushed stone
aggregate (111) with sufficient fines to bind the material together.

¡¡ Loose material and standing water is brushed from the pothole or


rut to be patched;
¡¡ Large or deep potholes should have their sides cut back to be
vertical and to reach sound material;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 61

¡¡ The moisture content of the material can be checked quickly by


squeezing it in the hand. If the material is wet enough to stick
together, it is suitable for use. If water runs out of the material, it is
too wet and should not be used;
¡¡ If the material is dry, the area to be patched should be sprinkled with
water and water should also be added to the patching material;

¡¡ The area is filled with gravel to a depth of about 10 centimetres;


¡¡ If the material is dry, it should be sprinkled with water to help
compaction;
¡¡ The layer is then compacted using the roller or hand rammer;
¡¡ In this way the thickness of the patch is built up in layers;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 62

¡¡ Finally, the patched area is filled evenly with the gravel to


approximately 3 centimetres above the level of the surface and is
spread and raked to the correct shape;
¡¡ 3 centimetres is approximately the thickness of a rake handle;
¡¡ The patch is then compacted using the roller or hand rammer to give
a surface which is only slightly above the level of the surrounding
road to allow for further traffic consolidation;
¡¡ Both large or small areas to be patched are repaired in the same way,
the rammer is used for the smaller potholes. The roller if available is

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 63

used for larger areas, although the hand rammers will still be required
for the corners and short edges;
¡¡ Patching work started must not be left unfinished overnight. At night
the site should be made safe for traffic and all signs and obstacles
removed from the road.

Defect 14: Road Surface rutted or uneven, and does not drain to edge
(Minor: <3cm)

Development, if neglected:
Gravel surface loss increases. Road becomes very rough, slowing and
damaging traffic.
Maintenance Activity
¡¡ 220 Blade Gravel Road (light)

Light grading may be carried out with a motor grader or a tractor towed
grader to correct minor defects on the gravel road surface such as
corrugations, shallow ruts and flat camber. The task may also be achieved
using labour with handtools.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 64

a) Light reshaping, Manual Method


¡¡ The surface material may be loosened, trimmed and reshaped with
a pickaxe, hoe or mattock and rakes to form the required camber
and crossfall;
¡¡ The shape is checked with the camber board and spirit level;
¡¡ If gravel stockpiles are provided, any local depressions are filled with
material transported in a wheelbarrow, pannier or other device;
¡¡ The loose material is compacted with the hand rammer;
¡¡ Pegs and string lines can be used to help to achieve the correct
shape and camber.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 65

b) Light reshaping, Mechanised Method


The motor grader or tractor towed grader is used to draw the surface
material back to the crown of the roadway. Normally only 4 passes will be
required to achieve this minor reshaping. It is best carried out during the
rains when there is sufficient moisture in the material for reconsolidation
under traffic, so that expensive watering and compaction operations will
not be required.

Minor corrugations can be dealt with by using a low cost drag towed by a
tractor or other vehicle.

Drags can be made from old tyres or various arrangements of discarded


steel sections.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 66

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 67

¡¡ The tractor tows the drag at up to 5 km/hour depending on the type


of drag and on the type and condition of the road surface;
¡¡ The length of pass should be as long as possible;
¡¡ The number of passes needed will depend upon the conditions and
the width of the road;
¡¡ The equipment should work in the same direction as the traffic flow;
¡¡ DO NOT drive too fast or the drag will jump over the surface
irregularities and raise a lot of dust, it will also cause a hazard to
traffic.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 68

Defect 15: Road Surface rutted or uneven, and does not drain to edge
(Major: >3cm)

Development, if neglected:
Road becomes very rough, slowing and damaging traffic. Water ponds on
road surface. Gravel surface loss increases and danger of total gravel layer
loss and road being impassable.
Maintenance Activity
¡¡ 221 Blade Gravel Road (heavy)

Heavy grading may be carried out with a motor grader or a tractor towed
grader. However the task will also require towed or self-propelled watering
and compaction equipment. The task may also be achieved using labour
and hand tools by adapting the methods of Maintenance activity 112.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 69

Preparation
Patching (Activity 110 or 111) of large potholes or depressions should be
carried out in advance of the grading. Areas of standing water should be
drained. This preparation will ease the work and make the resulting surface
last longer.

Scarifying
Using a motor or tractor towed grader it may be necessary to scarify the
existing surface to cut to the bottom of any surface defects and loosen the
material for reshaping.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 70

Machine Attendants
These help direct traffic and grader turning, and remove large stones and
other unwanted material from the path of the grader.

Grading
The grader works on one
side of the road at a time and
works in passes about 200
metres long to convenient
and safe turning points.
Heavy Grading will require
additional passes to achieve
the required camber. Work
should be completed on one
side of the road at a time. An
even number of passes should
be used to avoid a flat finished
crown. Normally initial cutting
passes are required to bring
material in from the edges of
the road. Spreading passes
redistribute the material away
from the crown. The initial
passes cut to the bottom of
the surface irregularity and
deposit a windrow just beyond
the centre line.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 71

Watering
The towed or self-propelled water tanker sprays the windrow with water,
if required. The windrow is spread back across the road depositing all the
material to give the correct camber. A second application of water may be
required to obtain the correct moisture content for compaction.
Cambering
The aim should be to develop a proper crown on the road. The road should
be cambered to fall away from the crown at a rate of about 6 to 7 cm for
each metre from the centre of the road before compaction. This should
achieve a crossfall of about 4 to 6 cm per metre (4 to 6%) after compaction.
If there is insufficient camber, water will not drain easily from the surface
of the road, potholes will form and the road will deteriorate quickly. This is
particularly important on gradients, where the rain water tends to run along
the road forming erosion channels.

Do not make a final pass down the centre of the road with the grader blade
horizontal. This flattens the centre of the road and causes water to pond
leading to rapid deterioration of the surface. Do not leave a windrow on the
road overnight as this is a danger to traffic.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 72

Compaction
When towed, self-propelled compaction plant is being used, it must follow
close up behind the grader, but only on sections where grading has been
completed. Usually about eight passes of a roller will be needed to achieve
full compaction, working towards the centre of the road. Shoulders are
treated as part of the running surface.

Junctions and Bends


Graders must not park up near junctions or bends where they will be a
danger to traffic.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 73

Check the Camber


Camber should be checked with a camber board at about 100 metre
intervals along the road. To use the camber board place it on its edge across
the road with the shorter end pointing towards the centre line. Check the
level bubble. If it is central, the camber is correct. If it is not central, the
camber is either too steep or too flat and further grading and compaction
are required.

Superelevation
On bends the surface must be straight (at 4-6%) from shoulder to shoulder
with the outer shoulder higher. This is called superelevation. This is because
any crown on a bend can be very dangerous to traffic. The superelevation
must be retained for the complete length of the bend.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 74

On the transition at each end of the bend into the straight sections, the
superelevation should be gradually reduced until the normal cross section
shape with about 1 in 25 to 1 in 17 (4-6%) crossfall is obtained again.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 75

Structures
The shape of the road must
be maintained over culverts
to avoid a hump. Material
should be brought in if
necessary from either side
of the culvert to maintain
a cover to the top of the
culvert of at least 3/4 culvert
diameter.

Bridge decks should be


kept free from gravel. Loose
material should be swept
away by the attendants. It is
important to have smooth
approaches to the bridge.
They should be smoothed out
using the back of the blade
with the grader working in
reverse, or by hand.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 76

Blade position

For most grading work, the cutting blade


is set to be vertical.

For cutting hard surfaces, the cutting


blade should be set back at the top to
give the most effective cutting angle.
Scarifying passes should also be made
before cutting.

For spreading, the cutting blade should


be set forward at the top.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 77

OCCASIONAL MAINTENANCE – GRAVEL ROAD


Defect 16: Gravel layer too thin

Development, if neglected:
Road becomes very rough, slowing and damaging traffic. Water ponds on
road surface. Gravel surface loss increases and danger of total gravel layer
loss and road being impassable.

Maintenance Activity
¡¡ 317 Gravel Resurfacing (Selected Material);
¡¡ 318 Gravel Resurfacing (Crushed Aggregate).

Gravel surfaces wear down due to the wasting effects of traffic and weather.
Loss rates can be up to 5cm thickness each year or more even on a Low
Traffic Volume Road. Re-gravelling will be required when (or before) the
residual thickness of gravel reduces to about 5 – 8 cm, otherwise there is a
danger of vehicle wheels ‘punching’ through to the weaker material below.
This would result in mixing and effectively the loss of the gravel layer.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 78

Great care should be taken in locating and selecting suitable gravel


material. It should be obtained from a recognized approved source and
meet materials specification requirements. This can be either selected
gravel material, or crushed stone aggregate with sufficient fine material to
bind the material together.

Gravel or crushed stone should not contain any pieces larger than 3cm,
as this will seriously affect performance. ‘Oversize’ pieces should be hand-
picked or ‘screened’ out. Due to the high cost of re-gravelling, technical
advice should be obtained on sources and material suitability. It is unlikely
that re-gravelling will be cost-effective if the material has to be hauled more
than 10km and other types of road surface may be more economical.

Diversion?
Wherever possible, before the re-gravelling work starts, a diversion should
be opened up adjacent to the road. If traffic is diverted from the work site, it
will enable the job to be carried out more efficiently and safely.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 79

Quarry or Borrow Pit


Before the regravelling work starts, gravel should be stockpiled at the quarry
or borrow pit. It may also be helpful to start hauling the material to site.
Plan the quarry excavations and stockpiles so that:
¡¡ The quarry can be fully exploited with removal of the maximum
amount of gravel;
¡¡ The overburden is stockpiled so that it will not hinder future
extension, and that it can be used to reinstate the quarry;
¡¡ The best material is taken, where gravel quality is variable within the
quarry;
¡¡ Material is stockpiled to minimise segregation;
¡¡ Environmental damage by poor drainage and erosion is minimised
both during and after exploitation of the quarry.
The quarry layout should:
¡¡ Permit efficient excavation and stockpiling of gravel;
¡¡ Allow the trucks, tractor and trailers or other haulage vehicles to
enter and leave without obstructions;
¡¡ Repair the quarry access road, if necessary, to ensure safe passage
of haulage vehicles.

Site Preparation
Traffic warning signs should be place at either end of the re-gravelling site.
The existing road surface must be graded-off or reshaped by hand to
provide a firm regular surface on which to work. Where possible, the edges
should be ‘boxed’ to provide lateral support for the new gravel. The graded/
reshaped surface should be watered and compacted. The camber should
be checked with a camber board and the road level should fall 4 to 6 cm for
each one metre width of road (4-6 %).

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 80

The road drainage system should be checked and repaired if necessary


(see Drainage defects and activities), otherwise the performance of the new
gravel surface will be affected.

At the quarry or borrow pit, the bulldozer or excavation labourers should


have stockpiled sufficient gravel for the work. The excavating and stockpiling
of gravel should create low, broad heaps to prevent segregation of the
coarser material.

Gravelling operations
When the initial grading/shaping of the road surface is complete, the loader
or the quarry labour should start to load the tippers or trailers with gravel for
transport to the re-gravelling site.

The supervisor at the quarry should ensure that gravel is taken from the
correct stockpiles and that the trucks/trailers are loaded correctly. Tippers or
tractor trailers should always circulate continuously between the quarry and

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 81

the site. Loading resources should be adjusted to try to keep the haulage
equipment working continuously. Dumping should start at the far end of the
site so that the heaps of gravel do not impede tippers or other haulage
vehicles delivering later loads.
¡¡ Material should be
dumped on one side of the
road only;
¡¡ Loads should be placed
at the correct spacing
as instructed by the
supervisor, necessary to
give the required thickness
of gravel over the
complete road width after
compaction;
¡¡ If the road is not closed,
material should be
dumped on the shoulder,
or dumped and spread
immediately by labour;
¡¡ The tankers or towed
bowsers should have
filled up with water using
the pump and then have
driven to the site;
¡¡ Initially the existing road
surface is sprayed with
water;
¡¡ Spreading of the gravel can start when there is a working length of
about 200 metres of dumped material if using a motor or towed
grader. If spreading is by labour, the gravel can be spread as soon
as it is dumped, or even unloaded by labour if non-tipping haulage
equipment is used;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 82

¡¡ The material is alternately spread by the grader/labour and watered


with the tanker/bowser until its moisture content is correct for
compaction;
¡¡ The amount of water to be added must be determined by moisture
content tests on site or by the supervisor;
¡¡ The tankers/bowsers circulate continuously between the site and the
source of water;
¡¡ The new material is now graded or spread by labour to produce a
camber of 4 to 6 cm for each one metre width of road (4 to 6 %).
Guide pegs and stringlines should used if labour spreading is used;

¡¡ The camber should now be


checked with the camber
board at approximately 100
metre intervals along the
road for machine spreading
and every 10 metres if
labour is used;
¡¡ To use the camber board,
place it on its edge across
the road with the shorter
end pointing towards the
centre line. Check the level
bubble;
¡¡ If it is central, the camber is
correct;
¡¡ If it is not central, the
camber is either too steep
or flat and further grading/
manual reshaping, and
compaction are required;
¡¡ When the correct camber
has been achieved,
compaction can start using a self propelled or towed roller, or a
pedestrian vibrating roller for labour works;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 83

¡¡ Water should not be added during rolling as the material may stick
to the wheels or drums;
¡¡ Rolling should start at the edge of the road and work towards the
middle. The roller should aim to progress from section to section at
the same rate as the grader or labour operations;

¡¡ Typically about eight passes of the roller will be needed to achieve


full compaction;
¡¡ It is possible to re-gravel without the use of water and compaction,
but it is difficult to achieve satisfactory results, and subsequent
gravel material loss from the surface will be faster. The watering and
compaction help to preserve the investment in the gravel.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 84

OCCASIONAL MAINTENANCE – PAVED ROAD


Defect 17: Paved road pothole or surface defect

Development, if neglected:
Road becomes very rough, slowing and damaging traffic. Water ponds on
road surface, speeding the deterioration and increasing risk of accidents.
Road user costs increase substantially. Road may become impassable.
Maintenance Activity
Depending on the type of paved road surface:

113a Spot /pothole Repair (Macadam)


113b Spot /pothole Repair (Stone setts)
113c Spot /pothole Repair (Mortared stone)
113d Spot /pothole Repair (Dressed stone)
113e Spot /pothole Repair (Emulsion chip seal)
113f Spot /pothole Repair (Emulsion sand seal)
113g Spot /pothole Repair (Emulsion gravel/slurry seal)
113h Spot /pothole Repair (Un-mortared brick)
113i Spot /pothole Repair (Mortar jointed brick)
113j Spot /pothole Repair (Non- reinforced concrete)
217 Pothole Reinstatement (cold mix)
219 Pothole (Base Failure Repair)

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 85

Although well constructed paved roads or sections should give many


years of trouble-free service, from time to time defects can be expected to
develop in any surface, such as:
¡¡ Cracking;
¡¡ Rutting;
¡¡ Potholes;
¡¡ Edge break.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 86

These defects are normally limited in extent and can be repaired using
labour, suitable hand tools and limited materials. Equipment is not normally
required. Any work on the road surface should be signed either side of the
repairs to warn road users and for the safety of those carrying out the work.

For all of the paved road surface types, the repair techniques are very similar,
and consist of:
¡¡ Marking out the area to be repaired;
¡¡ Excavation of the area to be repaired;
¡¡ Backfilling the hole with new material.
Marking out the area to be repaired
The area to be treated is marked out with chalk by
drawing a rectangle around the defects.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 87

Excavation of the area to be repaired


It is necessary to:
¡¡ Remove all loose or damaged material from within the marked
out area of the road surface back to a firm, sound material.
Sledgehammers, crowbars, hammers and chisels may be required;
¡¡ Increase the depth of the hole until firm, dry material is found and
then trim the walls of the hole so that they are vertical. If water or
excessive moisture is present, then arrangements must be made to
drain it away from the pavement foundation;
¡¡ Trim the bottom of the hole such that it is flat, horizontal and free
from loose material then compact it with a hand rammer.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 88

Backfilling the hole with new material


The repair will depend on the type of surface. Specifications and
requirements on each material are contained in the ERA Specifications. The
same specifications and standards should be applied to the repair.

The hole is filled with a selected material to match the existing surrounding
good surface and base materials. This can consist of new material, or in the
case of e.g. stone paving, recycled undamaged pieces.

The material is placed in the hole and compacted in one or more layers of
regular thickness depending on the depth and materials involved.

Generally, the last layer, prior to compaction, must have an excess thickness
of about 1/5 the depth of the final layer, in order to allow for settlement on
compaction.

Compaction is continued depending on the size of the excavation, using the


vibrating roller, plate compactor or with a hand rammer, until the surface is
level.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 89

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 10 - 90

Porous repairs will require a seal coat to prevent penetration of water.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 11 - 91

11. PRIORITIES
With limited resources, it is usually necessary to set priorities for carrying
out maintenance work, which should always take priority over any route
upgrading or improvement works. Protect what you have before extending
your assets and liabilities!

Furthermore, maintenance is most effective when applied to ‘maintainable’


routes, that is, those with a road camber and drainage system already
established. This is preferable to trying to work on un-drained tracks and
sunken road sections, which will consume a lot of resources with limited
impact.

The usual questions are:


¡¡ Which Route?, and
¡¡ Which maintenance activity?
Route Priorities
Within the community, the routes with the higher maintenance priorities
should be agreed. For simplicity and clarity it is best to divide all routes into
2 or 3 priority groups based on the following suggested criteria:
¡¡ Strategic inter-community or main road links;
¡¡ Is it maintainable? That is, does it already have camber and working
drainage system?;
¡¡ Traffic: high (e.g. more than 25 motor vehicles/day) or low;
¡¡ Population served;
¡¡ Value of volume of crops extracted each year;
¡¡ Serving markets, educational or health facilities.

It is useful to list the routes as an Inventory of assets to be managed, as an


aid to planning and management, for example:

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 11 - 92

Example inventory

Reasons for priority

Days impassable
motor vehicles
Daily traffic in
Length (km)

last year
Route

Priority A

Main road 65 (July


6.0km gravel Main Access 0
to A 2010)

3.5km earth
50 (April School and
A to Kebele B with spot 0
2010) teff exports
paving

Priority B

20 (January
A to Kebele C 4.5km earth Horticultural 2
2010)

12 (October
Kebele B to C 5.0km earth Dispensary 5
2009)

Pottery,
<10 (July
Kebele C to D 7.0km earth sunken 15
2010)
sections

It is beneficial for this information to be displayed at prominent community


locations.
Maintenance Activity Priorities
Where it is possible to arrange route maintenance on a number of occasions
each year, the following seasonal priorities should be made for each group
of maintenance activities:
BEFORE RAINS: Drainage
DURING RAINS: Drainage & Road Surface unpaved road sections
AFTER RAINS: Road Surface paved sections & Roadside activities &
Occasional maintenance

In this way the vital road drainage system is prepared for the rainy periods
and is kept functioning through the rains. The earth and gravel surfaces are

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 11 - 93

most effectively maintained during the rains where there is moisture in the
materials to help consolidate them after reshaping. Paved road surfaces will
usually develop any defects during the rains and are best repaired when the
drier weather comes.

Some variations to these general priorities may be applied due to local


conditions.

If a route becomes blocked or totally impassable, this is no longer a


maintenance problem. It will usually require the reasons to be investigated
and additional resources required to re-open it. Assistance or advice may be
required from regional authorities.
UPGRADING OR SPOT IMPROVEMENT PRIORITIES
Each year an assessment should be made of any desirable spot improvements
that should be made if resources are available. These could include any of
the options listed on Page 5.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 12 - 94

12. WORK OPTIONS


There are a number of ways that maintenance work can be organised
depending on the financial and human resources available, and ‘in-house’
capacity of the organising authority.

In practical terms, the maintenance of Low traffic Volume Roads (LVR) will be
principally carried out by labour methods with possible occasional support
of intermediate or heavy equipment. The last option is usually too expensive
to mobilise and inefficient for remote rural road works.

The main work organization options are detailed in this section with their
typical advantages and disadvantages. The Works Options are:
¡¡ Option 1- Small Contractor (Private);
¡¡ Option 2 - Force Account;
¡¡ Option 3- Community Group;
¡¡ Option 4 - Length Person or Family Contract;
¡¡ Option 5 - Compulsory/Voluntary Labour;
¡¡ Option 6 - Hire-in equipment;
¡¡ Option 7- Large Contractor Based System.

Option 1- Small Contractor (Private)


The small enterprises will be based in the region or local Weredas. They may
be general or building contractors with established contracting experience
in earthworks, masonry and concrete. They would be expected particularly
to make use of local labour, and may have access to light equipment such as
a concrete mixer or tractor. This implementation option can be suitable for
All Maintenance activities.
Advantages:
¡¡ Overheads lower than big contractor;
¡¡ Low mobilization and demobilization costs;
¡¡ Experience of the enterprise;
¡¡ Available range of building and maintenance skills;
¡¡ Local enterprise committed to the community;
¡¡ Good prospects for local employment and money being injected
into the commune.
Disadvantages:
¡¡ Time, resources and costs involved with preparing and managing
the contract;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 12 - 95

¡¡ Market for maintenance works currently not developed so prices may


be distorted (guideline unit costs should be available from regional
authorities);
¡¡ Small contractor may have to hire in some equipment;
¡¡ May initially require some training/ mentoring, or a higher level of
supervision than large contractors;
¡¡ May have difficulty in obtaining credit for purchases, or financing
cash flow;
¡¡ Insufficient funds currently available to pay for this approach for all
maintenance work (but may be suitable for selected works – see also
Option 6);
¡¡ Risks of disputes over interpretation of contract responsibilities.
Option 2 – Force Account
This option makes use of a permanently employed and equipped workforce
to carry out the maintenance work such as regional road management units.
This implementation option can be suitable for All Maintenance activities.
Advantages:
¡¡ Direct response to all maintenance needs;
¡¡ Rapid mobilization when funds are available;
¡¡ Retain skills and experience within organization, familiarity with the
network, standards, work methods;
¡¡ Minimum of works documentation requirements;
¡¡ Dealings/ disputes with outside parties minimized.
Disadvantages:
¡¡ In some cases no budget or funds are currently available for this
option;
¡¡ Difficulties in equipment procurement & the lowest initial purchase
cost policy can hinder the standardization and efficiency;
¡¡ Poor mobility of the workforce around the network unless transport
is provided (at considerable cost);
¡¡ Paid labour and equipment may be standing if no funds available
for works;
¡¡ Low efficiency and poor management/use of available resources,
poor cost-awareness;
¡¡ Little pressure to try new solutions/ technologies;
¡¡ High mobilization and demobilization costs if sourced from regional
level.
Option 3 – Community Group
The use of a group of persons based within the community and organized
specifically to carry out the maintenance works under an agreement or
contract with the commune authority. This can be for a single route, or a

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 12 - 96

number of routes served by the community. This approach differs from


the Length person or Family contract approach only in that the number of
persons expected to be involved would be greater, and that consequently
work would probably be concentrated at a particular time or times of the
year. This implementation option can be particularly suitable for the Regular
Maintenance activities
Advantages:
¡¡ Low cost compared to most other forms of contract (due to low
overheads, low mobilization and demobilization costs, absence of
profit component, and by local participation);
¡¡ Can be cash or in-kind payment according to commune circumstances
¡¡ Simple contract/agreement required;
¡¡ Direct response to Regular maintenance needs – Rapid mobilization,
or planned seasonal inputs;
¡¡ Retain skills and experience within the commune, familiarity with the
network and any problem sections;
¡¡ Close control of the works personnel;
¡¡ Pride of ‘ownership’ for the network;
¡¡ No dealings/disputes with parties outside of the commune;
¡¡ Employment and money/resources recycled within the commune;
¡¡ Employment can be targeted at poor or disadvantaged persons in
the commune.
Disadvantages:
¡¡ Possibly insufficient cash funds available to pay for this approach in
poor communes;
¡¡ Possible difficulties in controlling output and quality;
¡¡ Not suitable in areas of dispersed or low population density;
¡¡ No equipment capability;
¡¡ May not have access to construction quality hand tools;
¡¡ May initially require some training/ mentoring.
Option 4 – Length Person or Family Contract
A contract or agreement is drawn up for an individual or family to carry
out specified routine maintenance activities on a section of road, at certain
times of the year, for a payment in cash or in-kind for work on a full or part
time basis.

Usually a labourer is appointed for a distinct section of road close to his/


her home, typically 1 to 1.5 km in length. He or she is provided with all
the necessary hand tools to carry out all the routine maintenance activities
as instructed by the commune authority. An advantage is that regular
maintenance of the entire road can be arranged at all times and one person
can be made fully responsible of a road section. A disadvantage is that

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 12 - 97

supervision has to mobile frequently to ensure that performance does not


deteriorate.
Advantages:
¡¡ Low cost compared to most other forms of contract (due to low
overheads, low mobilization and demobilization costs, absence of
profit component, and by local participation);
¡¡ Can be cash or in-kind payment according to commune circumstances
¡¡ Simple contract/agreement required;
¡¡ Flexible approach to seasonal needs. Rapid mobilization;
¡¡ Pride of ‘ownership’ for the network;
¡¡ No dealings/disputes with parties outside of the commune;
¡¡ Employment and money/resources recycled within the commune;
¡¡ Employment can be targeted at poor or disadvantaged persons in
the commune.
Disadvantages:
¡¡ Possibly insufficient cash funds available to pay for this approach in
poor communes;
¡¡ Possible difficulties in controlling output and quality;
¡¡ Not suitable in areas of dispersed or low population density;
¡¡ No equipment capability;
¡¡ May not have access to construction quality hand tools;
¡¡ May initially require some training/ mentoring;
¡¡ System will degenerate if supervisor is not continuously mobile and
effective in management.
Option 5 – Compulsory/Voluntary Labour
The use of local (commune) labour to carry out maintenance works on the
roads is one of the options for maintaining community roads. The approach
can be suitable for Regular Maintenance activities. If the whole community
can be persuaded to attend a ‘maintenance day’ once or twice a year with
their hand tools, there will be sufficient labour resources to carry out the
necessary maintenance work under the guidance of a trained supervisor.
This is the cheapest way to maintain a LVR and involves no taxation or levy
to the community. Everybody contributes and benefits equitably. Wealthier
inhabitants, traders or other well-wishers could contribute hand tools,
equipment hire or food to create a community occasion.
Advantages:
¡¡ No financing or cash accounting involved;
¡¡ In richer communities, individuals can elect to pay cash instead. This
can provide funding for materials, handtools and equipment hire, or
even paid labour;
¡¡ Minimum of works documentation requirements;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 12 - 98

¡¡ Direct response to all maintenance needs;


¡¡ Rapid mobilization;
¡¡ Retain skills & experience within commune;
¡¡ Direct supervision of works;
¡¡ Pride of ‘ownership’ for the network;
¡¡ No dealings/disputes with outside parties.
Disadvantages
¡¡ All persons contribute equally, whether rich or poor;
¡¡ Can be a severe burden on the community’s poorest persons;
¡¡ Difficulties in controlling output and quality;
¡¡ Not suitable for work during the agricultural ‘high’ season;
¡¡ Not suitable in areas of dispersed or low population density;
¡¡ Few prospects for PAID commune employment or money being
injected into the commune;
¡¡ No equipment capability;
¡¡ May not have access to construction quality hand tools;
¡¡ May initially require some supervisor and ‘gang leader’ training/
mentoring.
Option 6 – Hire-in equipment
This is an option to supplement the other options to provide equipment for
specific operations such as towed grading or haulage. The funding could
be provided by the commune authority if available, or a benevolent trader,
farmer or other well-wisher.
Option 7 – Large Contractor Based System
The employment of a large equipment-based contractor may be considered.
They are usually based in Addis Ababa or regional centres and their
overheads, mobilization and demobilization costs and profit components
would be very high. Therefore, the costs would be extremely high and
unlikely to be affordable by the commune. In cases where a contractor
is funded externally to construct a road, this contractor may be engaged
as part of the contract to maintain the road for some years after the final
construction acceptance is finalized.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 13 - 99

13. PLANNING AND PRODUCTIVITY


It is important to plan the maintenance works according to the defect repair
needs, the priorities and the resources available to carry out the works.
The following labour resource requirements were developed from research in
East Africa and can be used as an outline planning guide.

Outline regular maintenance planning for LVRs

Number of Person-days of work expected/km/year

Number of Gravel Gravel Earth Earth


“wet” months <50vpd >50vpd <50vpd >50vpd

4 40 45 45 52

8 68 75 79 88

12 100 107 115 125


NOTES:
i. Number of “wet” months per year are with rainfall >25mm.
ii. Estimates assume ‘maintainable’ road with proper camber and drainage
system and gradients <6%. Not applicable for problem soils such as
‘black cotton’.
ii. Does not include Occasional works such as re-gravelling.

With good record keeping a similar table can be developed for each
commune. Equipment inputs may be required for materials haulage and
towed grading. Earth and gravel roads require reshaping/grading typically
between 1 and 4 times per year. These estimates will help to make resource
and cost estimates for each road each year.
Productivity Targets
To plan and manage maintenance works it is useful to have productivity
Standards, Norms or Targets. These need to be flexible considering the
variable nature of LVR maintenance works, experience of the supervisor
and workforce, and whether the work is carried out on a paid or voluntary
basis. Development of local Norms or Targets can take considerable time to
achieve. The following Targets were developed from research in East Africa
and practice in Southern Africa, and can serve as a reference point. The
standards were developed under close supervision conditions, with a well
trained workforce. They represent the best productivities that can be achieved
with a well organised and managed workforce. They should therefore be as
targets to be worked towards. It is expected that under normal conditions
60 – 80% of the productivity standards should be achieved. Good record
keeping can allow local productivity standards to be developed over time.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


PRODUCTIVITY TARGETS
TASK DIFFICULTY
PER PERSON-DAY
G - Chapter 13 - 100

NOTES
MAINTENANCE
CODE UNIT 1 2 3 4
ACTIVITY
131 Bush dearing - light m/day 425 260 190 - Difficulty = Width of bush
1) up to 1.0m
131 Bush dearing - dense m/day 275 225 175 - 2) 1.0 to 2.0m
3) Over 2.0m
Difficulty = Depth of erosion
Shoulder Rehabilitation 1) up to 1.0m
132 m/day 100 80 65 -
(manual) 2) 1.0 to 2.0m
3) Over 2.0m
Difficulty = Planting width
1) up to 1.0m
133 Plant grass and water m/day 100 80 65 -
2) 1.0 to 2.0m
3) Over 2.0m
Wet areas 425 260 190 Difficulty = Width of grass cutting
134 Cut grass - light m/day -
Dry areas 310 230 170 1) up to 1.0m
2) 1.0 to 2.0m
134 Cut grass - dense m/day 310 240 175 - 3) Over 2.0m
Difficulty = 4 passes more than 3cm of cut.
220 Blade Gravel Road (light) route km/day 10 - - -
Excludes mobilisation & demobilisation

PART G: WEREDA ROAD MAINTENANCE BOOKLET


Difficulty = 8 passes more than 3cm of cut.
221 Blade Gravel Road (heavy) route km/day 4260 - - -
Excludes mobilisation & demobilisation

Difficulty = Hauling distance from stockpile


110 Spot Repair - Selected Wheel barrows/
25 18 13 8 1) up to 100m 2) 100 to 200m
111 Material/Aggregate day
3) Over 200m
PRODUCTIVITY TARGETS
TASK DIFFICULTY
PER PERSON-DAY
NOTES
MAINTENANCE
CODE UNIT 1 2 3 4
ACTIVITY
Difficulty = Type of reshaping over road width
Reshape & Compact Earth
112 route m/day 70 50 - - 1) (Light) up to 75mm cut
Road Camber
2) (Heavy) over 75mm cut
Difficulty = Silt Depth in Culvert barrel
1) up to 1/4 2)1/4 to 1/2
121 Culvert Cleaning As shown 4 Culverts/day
3)1/2 to 3/4 3) 1/2 to 3/4
4) over 3/4 Tasks for 600dia. Culverts 7m long
Difficulty = Silt Depth
Ditch Clearing - Manual
122 m/day 55 40 25 - 1) Up to 10cm 2) 10 to 20cm
(Culvert outfall)
3) Over 20cm
Difficulty = Silt depth
Ditch Clearing - Manual
122 m/day 60 45 30 - 1) Up to 10cm 2) 10 to 15m
(Turnout drains)
3) Over 15m
Wet areas 65 45 30 Difficulty = Silt depth
Ditch Clearing - Manual (side
122 m/day Dry soft soil 55 40 30 - 1) Up to 10cm 2) 10 to 15cm
drains)
Dry hard soil 30 23 18 3) Over 15m
Difficulty = Silt depth
Repair Erosion Damage Wet areas 100 80 60
123 m/day - 1) Up to 10cm 2) 10 to 15cm

PART G: WEREDA ROAD MAINTENANCE BOOKLET


(Selected Fill in drains) Dry areas 100 50 23
3) Over 15cm
Difficulty = Depth of erosion
1) Up to 15cm 2) 15 to 30cm
126 Dry Masonry Repair m/day 7 4 - -
3) Over 30cm
Task for linear metres of dry stone wall
Build wooden/stone scour Difficulty = Type of scour check
128 No/day 5 7 - -
check 1) wood 2) stone
G - Chapter 13 - 101
G - Chapter 14 - 102

14. FURTHER ADVICE AND ASSISTANCE


Documentation
The following documents and publications may be accessed for further
information. Many of these documents may be accessed or downloaded
free of charge from ERA or www.gtkp.com .
1. Ethiopian Road Authority, LVR Design Manual, 2011;
2. Ethiopian Road Authority, Maintenance Technical Specifications;
3. Ethiopian Road Authority, Intermediate Technology Roadworks
Equipment, November 2010 ;
4. ILO, Guide to Tools and Equipment for Labour Based Road
Construction, 1981;
5. Intech Associates & MOPW Kenya, Road Maintenance Manual, 1992;
6. I&D Consult & MOPW Kenya, Headmans Handbook for Maintenance,
1991;
7. Transport and Road Research Laboratory, Overseas Road Note 7,
Volume 2, Bridge Inspector’s Handbook, 1988;
8. World Road Association (PIARC), International Road Maintenance
Handbook, 4 Volumes, 1994 and revisions.
Knowledge Sources
1. Knowledge Centre, Ethiopian Roads Authority, Ras Abebe Ageray
Street, P O Box 1770 Addis Ababa;
2. Global Transport Knowledge Partnership, www.gtkp.com,
info@gtkp.com;
3. Transport Research Laboratory;
4. CSIR, South Africa;
5. Expertise;
6. Knowledge Centre, Ethiopian Roads Authority, Ras Abebe Ageray
Street, P O Box 1770 Addis Ababa;
7. Regional Road Authorities.
Financial or other support
It is appreciated that communities and local authorities have very limited
resources and funding available for improving or maintaining rural access
roads. However, regional, national and international organisations may be
interested to help with partial funding for cost-effective, well-targeted rural
transport development initiatives which can demonstrate rural development
or poverty reduction benefits.

The following organisations have previously supported rural transport


initiatives. Details of their headquarter and local representative offices may
be obtained from the internet.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 14 - 103

¡¡ African Development Bank;


¡¡ Arab Bank for Economic Development in Africa;
¡¡ CARE;
¡¡ DANIDA;
¡¡ DFID, UK;
¡¡ DGIS (Netherlands);
¡¡ FINNIDA (Finland);
¡¡ Helvetas (Switzerland);
¡¡ Irish Aid;
¡¡ Islamic Development Bank;
¡¡ Kuwait Fund;
¡¡ NORAD (Norway);
¡¡ OPEC Fund for International Development;
¡¡ Saudi Fund for Development;
¡¡ Swiss Development Cooperation;
¡¡ SIDA;
¡¡ USAID (USA);
¡¡ World Bank.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 104

15. TERMINOLOGY
¡¡ Abney Level - Small hand held slope measuring and levelling
equipment.
¡¡ Aggregate - Hard mineral elements of construction material
mixtures, for example: sand, gravel, crushed rock.
¡¡ Aggregate Brooming - Using a broom to spread chippings on a
surface.
¡¡ Alligator Cracks - See Crazing.
¡¡ Apron - The flat invert of the culvert inlet or outlet.
¡¡ Asphalt - Another word for bitumen. Sometimes used to describe
plant mixed bituminous materials.
¡¡ Asphaltic Concrete - A high quality manufactured mixture of
bitumen and aggregates. Expensive and usually only used on main
roads.
¡¡ Attendant or Lengthman - A person contracted or appointed to
maintain a section of road. Can be male or female and the term
‘Attendant’ or ‘Lengthman’ assumes either sex.
¡¡ Basin - A structure at a culvert inlet or outlet to contain turbulence
and prevent erosion.
¡¡ Berm - A low ridge or bund of soil to collect or redirect surface water.
¡¡ Bituminous Slurry (Slurry-Seal) - Mixture, usually of fine-grained
aggregates, water, bituminous binder (emulsion), cement, and
sometimes an additive, for a road surface seal.
¡¡ Bituminous Binder (Asphalt) - A petroleum oil based or natural
product used to bind or coat aggregates for road pavements.
¡¡ Bleeding - Defect: Excess binder on the surface of the pavement.
¡¡ Blinding -
a) A layer of lean concrete, usually 5 to 10 cm thick, placed on soil
to seal it and provide a clean and level working surface to build the
foundations of a wall, or any other structure.
b) An application of fine material e.g. sand, to fill voids in the surface
of a pavement or earthworks layer.
¡¡ Block Cracking - Defect: Interconnected cracks forming a series of
large blocks usually with sharp corners or angles.
¡¡ Brick (clay) - A hard durable block of material formed from burning
(firing) clay at high temperature.
¡¡ Bridge - A structure usually with a span of 5 metres or more, providing
a means of crossing above water, a railway or another obstruction,
whether natural or artificial. A bridge consists of abutments, deck
and sometimes wingwalls and piers, or may be an arch.
¡¡ Camber - The road surface is normally shaped to fall away from
the centre line to either side. The camber is necessary to shed rain
water and reduce the risk of passing vehicles colliding. The slope

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 105

of the camber is called the crossfall. On sharp bends the road


surface should fall directly from the outside of the bend to the inside
(superelevation).
¡¡ Camber Board - Apparatus for checking the crossfall of the road
camber, or the shoulder.
¡¡ Cape Seal - A road surface layer formed by slurry seal laid on top of
a bituminous chip seal.
¡¡ Carriageway - The road pavement or bridge deck surface on which
vehicles travel.
¡¡ Cascade - A drainage channel with a series of steps, sometimes with
intermediate silt traps or ponds, to take water down a steep slope.
¡¡ Catchpit - A manhole or open structure with a sump to collect silt.
¡¡ Catchwater Drain - See Cutoff.
¡¡ Causeway or Vented Drift - Low level structure constructed across
streams or rivers with openings to permit water to pass below road
level. The causeway may become submerged in flood conditions.
¡¡ Check Dams – (see also Scour Checks) Small checks in a ditch or
drain to reduce water velocity and reduce the possibility of erosion.
¡¡ Chippings - Clean, strong, durable pieces of stone made by crushing
or napping rock. The chippings are usually screened to obtain
material in a small size range.
¡¡ Chip Seal - A surface layer formed by stone chippings laid onto a
bituminous seal coat.
¡¡ Chute - An inclined pipe, drain or channel constructed in or on a
slope.
¡¡ Cobble Stone (Dressed stone) - Cubic pieces of stone larger than
setts, usually shaped by hand and built into a road surface layer or
surface protection.
¡¡ Coffer Dam - A temporary dam built above the ground to give
access to an area which is normally, or has a risk of being, submerged
or waterlogged. Cofferdams may be constructed of soil, sandbags
or sheetpiles.
¡¡ Compaction - Reduction in bulk of fill or other material by rolling or
tamping.
¡¡ Counterfort Drain - A drain running down a slope and excavated
into it. The excavation is partly or completely filled with free draining
material to allow ground water to escape.
¡¡ Cracking - Defect: Narrow breaks in a surfacing or pavement material
caused by overloading, fatigue or weakness of the material.
¡¡ Crazing (Alligator Cracks) - Defect: Interconnecting network of
cracks in the road surfacing.
¡¡ Cribwork - Timber or reinforced concrete beams laid in an
interlocking grid, and filled with soil to form a retaining wall.
¡¡ Cut-off/Catchwater Drain - A ditch constructed uphill from a cutting
face to intercept surface water flowing towards the road.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 106

¡¡ Debris Rack or Grill - Grill, grid or post structure located near a


culvert entrance to hold back floating debris too large to pass
through the culvert.
¡¡ Deck - The part of a bridge that spans between abutments or pier
¡¡ supports, and carries the road traffic.
¡¡ Depression - Defect: Localised low areas of limited size in the
pavement surface or in any other surface.
¡¡ Ditch (Open Drain) - A long narrow excavation designed or intended
to collect and drain off surface water.
¡¡ Drag - An apparatus towed behind a vehicle or piece of equipment
to remove minor irregularities and redistribute loose surface material.
¡¡ Drainage - Interception and removal of ground water and surface
water by artificial or natural means.
¡¡ Drainage Pipe - An underground pipe to carry water.
¡¡ Dressed Stone - See Cobble Stone.
¡¡ Drift or Ford - A stream or river crossing at bed level over which the
stream or river water can flow.
¡¡ Earth Road - See ENS.
¡¡ Edge Cracking - Defect: Longitudinal cracking near the edge of the
pavement.
¡¡ Embankment - Constructed earthworks below the pavement raising
the road above the surrounding natural ground level.
¡¡ ENS (Engineered Natural Surface) - An earth road built from the
soil in place at the road location, and provided with a camber and
drainage system
¡¡ Excess Aggregate - Defect: Aggregate particles not coated with
binder after application of binder.
¡¡ Flow Spreader - A structure designed to disperse the flow at the
outfall of a ditch or drain to minimise the risk of erosion down stream.
¡¡ Fog Seal - A very light film of binder sprayed onto a road to bind or
enrich the surface.
¡¡ Ford - See Drift
¡¡ Formation - The shaped surface of the earthworks, or subgrade,
before constructing the pavement layers.
¡¡ Fretting - Defect: The loss of chippings from the surface seal or
premix layer due to poor bond between the aggregate and the seal
or binder.
¡¡ Gabion - Stone-filled wire or steel mesh cage. Gabions are often
used as retaining walls or river bank scour protection structures.
¡¡ Glazing - Defect: Wear or embedment of chippings in the surfacing
giving a smooth, shiny appearance.
¡¡ Hand Packed Stone - A layer of large, angular broken stones laid by
hand with smaller stones or gravel rammed into the spaces between
stones to form a road surface layer.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 107

¡¡ Incremental paving - Road surface comprising small blocks such as


shaped stone (setts) or bricks, jointed with sand or mortar.
¡¡ Invert - The lowest point of the internal cross-section of a ditch or
culvert.
¡¡ Layby - An area adjacent to the road for the temporary parking of
vehicles.
¡¡ Lengthman - See Attendant.
¡¡ Loss of Surface Aggregate - Defect: Removal of aggregate from
a surface dressing, or from surfacings with coated aggregate, or
concrete.
¡¡ Macadam - A mixture of broken or crushed stone of various sizes
(usually less than 3cm) laid to form a road surface layer.
¡¡ Manhole - Accessible pit with a cover forming part of the drainage
system and permitting inspection and maintenance of underground
drainage pipes.
¡¡ Margins - The right of way or land area maintained or owned by the
road authority.
¡¡ Mitre Drain (Turn Out Drain) - leads water away from the Side
Drains to the adjoining land.
¡¡ Occasional Maintenance - Operations that are occasionally required
on a section of road after a period of a number of years.
¡¡ Open Drain (Ditch) - A long narrow excavation designed or intended
to collect and drain off surface water.
¡¡ Ottaseal - A surface layer formed by rolling natural gravel into a soft
bituminous seal coat.
¡¡ Outfall - Discharge end of a ditch or culvert.
¡¡ Parapet - The protective edge, barrier, wall or railing at the edge of
a bridge deck.
¡¡ Pass - A single longitudinal traverse made by a grader, roller or other
piece of equipment working on the road.
¡¡ Patching - The execution of minor local repairs to the pavement and
¡¡ shoulders.
¡¡ Pavé - See Sett
¡¡ Paved Road - For the purpose of this booklet, a paved road is a road
with a Stone, Bituminous, Brick or Concrete surfacing.
¡¡ Pavement - The constructed layers of the road on which the vehicles
travel.
¡¡ Permeable Soils - Soils through which water will drain easily e.g.
sandy soils. Clays are generally impermeable except when cracked
or fissured (e.g. ‘Black Cotton’ soil in dry weather).
¡¡ Plumbing - Using a calibrated line, with a weight attached to the
bottom, to measure the depth of water (e.g. for checking erosion by
a structure).
¡¡ Profile - An adjustable board attached to a ranging rod for setting
out.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 108

¡¡ Ravelling - See Stripping.


¡¡ Regular Maintenance - Operations required to be carried out once
or more per year on a section of road. These operations are typically
small scale or simple, but widely dispersed, and require un-skilled or
trained manpower.
¡¡ Reinforced Concrete - A mixture of coarse and fine stone aggregate
bound with cement and water and reinforced with steel roads for
added strength.
¡¡ Riprap - Stones, usually between 5 to 50 kg, used to protect the
banks or bed of a river or watercourse from scour.
¡¡ Road Base and Subbase - Pavement courses between surfacing and
subgrade.
¡¡ Road Maintenance - Suitable regular and occasional activities to
keep pavement, shoulders, slopes, drainage facilities and all other
structures and property within the road margins as near as possible
to their as constructed or renewed condition. Maintenance includes
minor repairs and improvements to eliminate the cause of defects
and avoid excessive repetition of maintenance efforts.
¡¡ Roadway - The portion within the road margins, including shoulders,
for vehicular use.
¡¡ Sanding - Spreading course sand onto a bituminous road surface
that is bleeding.
¡¡ Sand Seal - A surface layer formed by sand laid onto a bituminous
seal coat.
¡¡ Scarifying - The systematic disruption and loosening of the top of a
road or layer surface by mechanical or other means.
¡¡ Scour - Defect: Erosion of a channel bed area by water in motion,
producing a deepening or widening of the channel.
¡¡ Scour Checks – (see also Check Dams) Small checks in a ditch or
drain to reduce water velocity and reduce the possibility of erosion.
¡¡ Scuppers - Drainage pipes or outlets in a bridge deck.
¡¡ Sett (Pavé) - A small piece of hard stone trimmed by hand to a size
of about 10cm cube used as a paving unit.
¡¡ Shoulder - Paved or unpaved part of the roadway next to the outer
edge of the pavement. The shoulder provides side support for the
pavement and allows vehicles to stop or pass in an emergency.
¡¡ Slip - Defect: Slope material sliding downhill because of instability,
water penetration or flow.
¡¡ Slope - A natural or artificially constructed soil surface at an angle to
the horizontal.
¡¡ Slot - A sample cross section of the road or drain constructed as a
guide for following earthworks or reshaping.
¡¡ Slurry Seal - A mixture usually containing fine graded aggregates,
water;

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 109

¡¡ bitumen emulsion, cement and sometimes an additive, spread on


the road surface by a specially equipped machine, or by hand.
¡¡ Sods - Turf but with more soil attached (usually more than 10 cms).
¡¡ Soffit - The highest point in the internal cross-section of a culvert, or
the underside of a bridge deck.
¡¡ Special Maintenance - Certain serious, unforeseen situations
necessitating remedial action to be taken as soon as possible, e.g.
flood damage, major slips. Consult the regional authorities regarding
these.
¡¡ Spray Lance - Apparatus permitting hand-application of bituminous
binder at a desired rate of spread through a nozzle.
¡¡ Squeegee - A small wooden or metal board with a handle for
spreading bituminous mixtures by hand.
¡¡ Streaking - Defect: Alternate lean and heavy lines of bitumen
running parallel to the pavement centre line, caused by blocked or
incorrectly set spray nozzles.
¡¡ Stringer - Longitudinal beam in a bridge deck or structure.
¡¡ Stripping (Ravelling) - Defect: The loss of surface seal from the
pavement due to poor bond between the seal and the lower
pavement layer.
¡¡ Subbase -See Road Base.
¡¡ Subgrade - Upper layer of the natural or imported soil (free of
unsuitable material) which supports the pavement.
¡¡ Sub-Soil Drainage - See Underdrainage.
¡¡ Surface Dressing - A sprayed or hand applied film of bitumen
followed by the application of a layer of stone chippings, which is
then rolled.
¡¡ Surface Treatment - Construction of a protective surface layer e.g.
by spray application of a bituminous binder, blinded with coated or
uncoated aggregate.
¡¡ Surfacing - Top layer of the pavement. Consists of wearing course,
and sometimes a base course or binder course.
¡¡ Tar Binder - A binder made from processing coal.
¡¡ Template - A thin board or timber pattern used to check the shape
of an excavation.
¡¡ Traffic Lane - The portion of the carriageway usually defined by road
markings for the movement of a single line of vehicles.
¡¡ Transverse Joint - Joint normal to, or at an angle to, the road centre
line.
¡¡ Traveller - A rod or pole of fixed length (e.g. 1 metre) used for
sighting between profile boards for setting out levels and grades.
¡¡ Turf - A grass turf is formed by excavating an area of live grass and
lifting the grass complete with about 5 cms of topsoil and roots still
attached.
¡¡ Turn Out Drain - See Mitre Drain.

PART G: WEREDA ROAD MAINTENANCE BOOKLET


G - Chapter 15 - 110

¡¡ Underdrainage (Sub-Soil Drainage) - System of pervious pipes or


free draining material, designed to collect and carry water in the
ground.
¡¡ Unpaved Road - For the purpose of this booklet an unpaved road is
a road with a soil or gravel surface.
¡¡ Vented Drift - See Causeway.
¡¡ Weephole - Opening provided in retaining walls or bridge abutments
to permit drainage of water in the filter layer or soil layer behind
the structure. They prevent water pressure building up behind the
structure.
¡¡ Windrow - A ridge of material formed by the spillage from the end of
the machine blade or continuous heap of material formed by labour.
¡¡ Wingwall - Retaining wall at a bridge abutment to retain and protect
the embankment fill behind the abutment.
¡¡ 2WD - Two Wheel Drive vehicle or equipment.
¡¡ 4WD - Four Wheel Drive vehicle or equipment.

PART G: WEREDA ROAD MAINTENANCE BOOKLET

You might also like