You are on page 1of 729

Maintenance of Road Drainage Infrastructure

Road Drainage Manual

A guide to the Planning, Design, Operation and

A guide to the Planning, Design, Operation and Maintenance of


Road Drainage Infrastructure

Road Drainage Manual

Road
Drainage
Manual
A Guide to the Planning, Design,
Operation and Maintenance of Road
Drainage Infrastructure

2nd Edition, March 2010

Prepared and issued by the


Engineering & Technology Division

Department of Transport and Main Roads


Road Drainage Manual

COPYRIGHT
State of Queensland (Department of Transport and Main Roads) 2010
Copyright protects this publication. Except for the purposes permitted by and subject to the conditions
prescribed under the Copyright Act, reproduction by any means (including electronic, mechanical,
photocopying, microcopying or otherwise) is prohibited without the prior written permission of the
Queensland Department of Transport and Main Roads. Enquiries regarding such permission should be
directed to the Road & Delivery Performance Branch, Queensland Department of Transport and Main
Roads

First published 2002

as Road Drainage Design Manual

Second edition 2010

Published by Queensland Department of Transport


and Main Roads, Brisbane, Australia.

Reprinted with
corrections 2003

FEEDBACK
Your feedback is welcomed. Please send to <mr.techdocs@tmr.qld.gov.au>.

TRADEMARKS ACKNOWLEDGEMENT
Terms mentioned in this document that are known or understood to be trademarks, whether registered
or not, have been identified. Where trademarks have been confirmed as registered in Australia, this
has been indicated by the addition of the symbol, otherwise the symbol is used. While all care
has been taken to identify trademarks, users should rely on their own inquiries to determine trademark
ownership. Use of a term in this document as a trademark should not be regarded as affecting the
validity of any trademark.

DISCLAIMER
This publication has been created for use in the design, construction, maintenance and operation of
road transport infrastructure in Queensland by or on behalf of the State of Queensland.
Where the publication is used in other than the departments infrastructure projects, the State of
Queensland and the department give no warranties as to the completeness, accuracy or adequacy of the
publication or any parts of it and accepts no responsibility or liability upon any basis whatever for
anything contained in or omitted from the publication or for the consequences of the use or misuse of
the publication or any parts of it.
If the publication or any part of it forms part of a written contract between the State of Queensland and
a contractor, this disclaimer applies subject to the express terms of that contract.

March 2010
ii

Department of Transport and Main Roads


Road Drainage Manual

FOREWORD
This Road Drainage Manual (2nd Edition) provides guidance in the planning, design, operation
and maintenance of road drainage infrastructure in all urban and rural environments for the
Department of Transport and Main Roads, Queensland.
Road drainage is an important element of the road environment and therefore must be
considered in all planning, design, construction and maintenance projects. The appropriate
management of stormwater about the road environment is vital to user safety, longevity of the
road asset and protection of the environment. This demands a holistic and multi-disciplinary
approach to the road drainage planning and design activity in order to satisfy both hydraulic and
environmental requirements in determining an appropriate drainage solution at a cost acceptable
to the community.
The information within this guide is sufficient to undertake normal daily work. Competent
planners, designers and engineers shall apply the requirements within this document in an
intelligent way. However, it is expected that when situations arise that cannot be resolved
using this guide, the subject will be examined in more detail using relevant experts. Noncompliance with the requirements and intent of this manual must be considered a design
exception and will require appropriate technical certification and approval by Department of
Transport and Main Roads, Queensland.

Ian Reeves
General Manager (Engineering and Technology)
24 March 2010

March 2010
iii

Department of Transport and Main Roads


Road Drainage Manual

IMPORTANT INFORMATION
The requirements of this document represent Technical Policy of the department and contain
Technical Standards. Compliance with the departments Technical Standards is mandatory for
all applications for the design, construction, maintenance and operation of road transport
infrastructure in Queensland by or on behalf of the State of Queensland.
This document will be reviewed from time to time as the need arises and in response to
improvement suggestions by users. Please send your comments and suggestions to the feedback
email provided previously.
This document is approved for use by:

24 March 2010
D. A. Hicks
Chair of Steering Committee and Director (Road Planning & Design)

24 March 2010
A. D. Krosch
A/Executive Director (Design, Environment & Stewardship)
And the General Manager (Engineering & Technology) by his signature of the Foreword.

March 2010
iv

Department of Transport and Main Roads


Road Drainage Manual

TABLE OF CONTENTS
As of March 2010
Chapter

Issue
Date

Acknowledgements

Mar 10

Framework

Mar 10

General Design Requirements

Mar 10

Strategic Planning & Development Control

Mar 10

Data Collection
Appendix 4A Data Collection Checklist
Hydrology
Appendix 5A Pipe Flow Charts
Appendix 5B - Worked Examples

Mar 10
Mar 10
Mar 10
Mar 10
Mar 10

Approach to Drainage Design

Mar 10

Environmental Consideration and Design

Mar 10

Open Channel Design


Appendix 8A Vegetal Retardance Curves
Appendix 8B - Worked Examples
Culvert Design
Appendix 9A Design Nomographs
Appendix 9B Design Form
Appendix 9C - Worked Examples
Floodway Design
Appendix 10A - Worked Examples
Road Surface and Sub-Surface Drainage Design
Appendix 11A - Worked Examples

Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10

Basins

Mar 10

Erosion and Sediment Control


Appendix 13A Preparing an ESCP
Appendix 13B Fact Sheets for ESC Measures
Appendix 13C Inspection of ESCs
Appendix 13D Revegetation Guidelines
Operation, Maintenance and Remediation
Appendix 14A Inspection and Treatment of Metal Culverts

Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10
Mar 10

4
5

10
11
12

13

14

Glossary

Mar 10

References

Mar 10

March 2010
v

Department of Transport and Main Roads


Road Drainage Manual

March 2010
vi

Department of Transport and Main Roads


Road Drainage Manual

ACKNOWLEDGEMENTS
The Content Owners of this new edition wish to acknowledge the effort of the various authors
(both internal and external to the department) and members of the steering committee
involved with the first release. Much of their work exists in this edition.
The on-going development and maintenance of the manual is the responsibility of the Road
Engineering Standards Section (E&T), under the guidance of the following steering
committee:
Member

Role

Mr D Hicks
Director (Road Planning &
Design)

Committee Chair and


Departmental Management Representative

Mr M Whitehead
Principal Engineer (Road
Design Standards)

Manual Manager and


Departmental Technical Specialist (Design)

Dr W Weeks
Director (Hydraulics)

Departmental Technical Specialist


(Hydrology/Hydraulics)

Mr R Stone
Director (Environment and
Heritage)

Departmental Technical Specialist (Environment)

Ms K Mahony
Principal Environmental
Officer (Southern)

Departmental Regional (Toowoomba) Representative


(Environment)

Mr T Smith
Manager (Design
Services)

Departmental Regional (Townsville) Representative


(Design)

Ms T Graham
Aurecon

Industry Representative (Hydrology/Hydraulics)

Mr P Woods
Aurecon

Industry Representative (Design)

Key technical authors / contributors to the development of this edition are: Mr Mike
Whitehead, Dr William Weeks, Mr Juan Carlos Delgado, Mr Wayne Huntley, Mr John
Blurton, the Urban Stormwater Quality Management Working Group (USQMWG) and
Environmental Soil Solutions Australia.
The department would also like to acknowledge the Department of Environment and
Resource Management for their approval to reference and use parts of the Queensland Urban
Drainage Manual (QUDM) issued in 2008.

March 2010
vii

Department of Transport and Main Roads


Road Drainage Manual

March 2010
viii

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

Chapter 1
Framework

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

Chapter 1 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

Table of Contents
1.1

1.2

1.3

Introduction

1.5

1.6

1-1

1.1.1

Overview of 2nd Edition of the Road Drainage Manual

1-1

1.1.2

Applicability of this Manual

1-2

1.1.3

Importance of Road Drainage

1-2

1.1.4

Reference Documents and Manuals

1-3

1.1.5

Structure of this Manual

1-4

Limits of Manual

1-5

1.2.1

General

1-5

1.2.2

Hydrology / Hydraulics

1-5

1.2.3

Environmental Design

1-7

1.2.4

Catchment Hydrology in Rural Areas

1-7

1.2.5

Catchment Hydrology in Urban Areas

1-8

Road Infrastructure Delivery


1.3.1

1.4

1-8

Introduction

1-8

Legislation and Policy

1-9

1.4.1

Introduction

1-9

1.4.2

Relevant Legislation and Other Authorities

1-9

1.4.3

Key Aspects of Road Drainage Policy

1-10

1.4.4

Community Engagement

1-10

1.4.5

Impacts on Landowners

1-11

Guiding Principles

1-12

1.5.1

Drainage Infrastructure and Land Use

1-12

1.5.2

Assessment of Future Development

1-13

1.5.3

General Hydraulic and Environmental Consideration

1-13

1.5.4

Economic Considerations

1-13

1.5.5

Maintenance

1-14

Drainage Issues

1-14

1.6.1

Planning the Drainage System

1-14

1.6.2

Urban Drainage

1-14

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

1.6.3

1.7

Climate Change

March 2010

iv

Rural Drainage Systems

Chapter 1
Framework

1-15

1-16

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

Chapter 1
Framework
1.1 Introduction
1.1.1

Overview of 2nd Edition of


the Road Drainage Manual

In 2002, the Queensland Department of


Main Roads released the Road Drainage
Design Manual (RDDM) that documented a
new approach to the planning, design,
construction and maintenance of drainage
infrastructure that established a multidisciplinary approach to the provision of
drainage infrastructure.
In 2006, a review of the RDDM was
initiated to:
reflect the changes to the planning
and design framework introduced in
the departments Preconstruction
Processes Manual;
provide consistency with the
departments Road Planning and
Design Manual;
address comments and suggestions
for improvements to the existing
manual published in 2002;
better align the departments
drainage design practice to accepted
industry practice.
This edition has a new title, Road Drainage
Manual - A guide to the Planning, Design,
Operation and Maintenance of Road
Drainage Infrastructure (RDM), to better
reflect the departments whole-of-life
approach to the provision of drainage
infrastructure. It is a new manual that
replaces the previous edition of the manual

and is a guide to those involved in the


planning,
design,
operation
and
maintenance of road drainage infrastructure
for small, simple rural and urban
catchments.
It does not include the
hydrology / hydraulics for major waterways
and/or floodplains, complex catchments or
significant drainage structures such as
bridges.
The principles of a full multi-disciplined
approach to the provision of effective road
drainage infrastructure that were established
in the previous edition have been
maintained. The sizing and location of
drainage structures are developed by taking
into
account
relevant
hydraulic,
environmental, safety and maintenance
requirements.
Effects of climate change, provided for in
the previous edition of this manual, have
been
reviewed,
but
no
new
recommendations or changes have been
made. Ongoing review of this issue is
required and will continue.
Information on computer aided design of
drainage has been included to reflect the
departments decision to adopt the software
platform, 12d Model, to embrace an
integrated electronic modelling system that
provides for seamless interoperability
through survey, planning, design, tendering,
construction, as-constructed and archiving.
Project electronic modelling and archiving
and/or drawing storage provides a common
basis for accessing data on a consistent and
reliable basis appropriate for use by
different personnel over the whole of life of

March 2010

1-1

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

the drainage infrastructure.


Drawing
storage is controlled by Planroom located
within the Geospatial Technologies Section,
Engineering & Technology Division.

1.1.2

Applicability of this
Manual

This manual represents the departments


policy with respect to the planning, design,
operation and maintenance of road drainage
infrastructure and must be applied on all
road infrastructure projects for which the
department is responsible. As such, the
manual applies equally to all personnel,
departmental or not, that are involved in the
drainage aspects of departmental projects,
including:
Departmental management;
Departmental and/or consulting
project (preconstruction / design /
construction
/
maintenance)
managers;

System Manager Framework and therefore


provides guidance in relation to the
sustainable planning, design, construction,
operation and maintenance of most road
drainage infrastructure in both urban and
rural environments throughout Queensland.
Limitations to the application and use of
this manual are outlined in Section 1.2 of
this chapter.
The manual is to be used by appropriately
qualified and experienced personnel or
others who have received appropriate
training and are supervised by qualified and
experienced personnel.
The manual also integrates best practice and
environmental management techniques into
the provision of road drainage. It includes
technical governance requirements for the
selection, design and construction of
appropriate drainage structures that satisfy
hydraulic requirements whilst minimising
the potential for environmental and asset
harm.

Planners (strategic / project / town);

1.1.3

Development control officers;


Civil
designers
(engineers
technologists / designers);

Surveyors;
Environmental scientists / engineers;
Geologists / geotechnical engineers;
and
Construction
personnel.

maintenance

This manual facilitates the development and


implementation of drainage solutions for
state controlled roads and roads within
Queensland that are part of the National
Land Transport Network (Auslink). It
operates from Phase 3, Corridor Planning
and Stewardship, to Phase 6, Program
Finalisation, of the departments Road
March 2010

1-2

Importance of Road
Drainage

It is recognised that a road requires a


drainage system to deal with stormwater
runoff. Therefore, the drainage system
becomes an important and integral
consideration in the planning and design of
road infrastructure.
In order to provide an appropriate and
economic drainage system, all road
projects, irrespective of location, size, cost
or complexity, must consider and address
the following aspects:
provision of an acceptable level of
flood immunity and accessibility;
impact of flooding of public and
private property;

Department of Transport and Main Roads


Road Drainage Manual

conveyance of stormwater through


the road reserve at a development
and environmental cost that is
acceptable to the community as a
whole;
protection of the roadway asset;
safety of all road users;
pollutant discharge from the road
reserve to receiving waters;
land degradation caused by erosion
and sedimentation during road
construction,
operation
and
maintenance;
any impact on habitats for terrestrial
and aquatic flora and fauna;
any impact on the movement of
terrestrial and aquatic fauna.
This requirement particularly applies to
projects where it is proposed to keep the
existing drainage infrastructure.
The
original design intent must be reviewed and
understood, and then the existing system
needs to be assessed against the above
aspects for performance, adequacy and
continued durability. Design must ensure
that the original intent is restored,
deficiencies corrected and modifications /
changes appropriately considered and
detailed.

1.1.4

Reference Documents and


Manuals

This manual is underpinned by an


integrated, multi-disciplinary approach.
That is, the user of this manual must focus
at all times on the consideration of
hydrologic,
hydraulic,
environmental,
surrounding land use, maintenance and
safety factors. Typically, these factors
should be considered concurrently.

Chapter 1
Framework

In using this manual, reference will need to


be made to other departmental documents
developed to assist in the planning, design,
construction and maintenance of drainage
infrastructure on state-controlled roads in
Queensland.
These reference documents include the:
Road Planning and Design Manual A guide to QLD practice
Preconstruction Processes Manual
Guidelines for Strategic
Network Planning

Road

Project Cost Estimating Manual


Road
Project
Processes Manual

Environmental

Environmental Legislation Register


Road Landscape Manual
Soils Manual
Roads in the Wet Tropics Manual
Road Traffic Noise Management:
Code of Practice.
It is important to note that the departments
Road Planning & Design Manual - A guide
to QLD practice (RPDM) is its main road
design reference, however in the interest of
national uniformity, the RPDM primarily
references Austroads Guide to Road
Design (GRD) series.
Drainage is an integral component of road
infrastructure. Drainage design cannot be
separated from the overall road planning
and design process. Drainage must always
be considered in conjunction with the
geometric design of a road. Although this
manual has been published as a separate
document, it has a complementary
relationship the RPDM and therefore
planners and designers in particular, must

March 2010

1-3

Department of Transport and Main Roads


Road Drainage Manual

use this manual in conjunction with the


RPDM.

1.1.5

Structure of this Manual

This manual has been designed so that at


any stage of drainage infrastructure
delivery, users can refer to the chapter or
section that is applicable to their needs.
However, reference to other chapters will
be required. To facilitate this, references
have been made to other, relevant sections
of the manual, other manuals, guidelines

Chapter 1
Framework

and policies where possible. Additionally,


checklists and flow charts are provided
throughout the manual to assist the user to
proceed from section to section of the
manual.
This edition has been structured into parts
that reflect the planning and design process.
These parts are then further divided into
chapters of like purposes. Table 1.1.5
outlines the structure of this manual.

Table 1.1.5 - Structure of Road Drainage Manual


Chapter

Purpose of Chapter(s)
Sets the context and limitations of this
manual
Defines the general design requirements
(including considerations, controls, criteria
& standards) for planning and design.
Defines the base drainage requirements
for strategic planning (road route
strategies & road link plans) and
development applications.
Defines strategic and project data.
Establishes importance of data collection
and retrieval. Describes the data
collection process and identifies sources
of data.
Describes the processes for determining
and analysing hydrologic data to quantify
specific design criteria.
Describes the design methodology and
process to be followed for drainage
components.
Provides guidance for understanding
environmental assessments and the
consideration of requirements in design.

Framework

Design Requirements

Strategic Planning & Development


Control

Data Collection

Hydrology

Approach to Drainage Design

Environmental Consideration and


Design

8
9
10

Open Channel Design


Culvert Design
Floodway Design

11

Road Surface and Sub-surface Drainage Design


Provides hydraulic design methods for
Basins
design of various basins
Provides the requirements for drainage in
Erosion and Sediment Control
construction.
Emphasises the importance of
Operation, Maintenance and
performance monitoring & maintenance of
Remediation
drainage infrastructure.

12
13
14

March 2010

1-4

Title of Chapter

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

o Bureau of Meteorology (BoM)


website

1.2 Limits of Manual


1.2.1

Hydraulic design:

General

The following sections outline the


limitations placed on civil designers in road
drainage design and the use of this manual.
Furthermore, it should be noted that this
manual is not a complete drainage guide
covering all aspects of road drainage. It is a
guide for small, simple rural and urban
catchments only and does not include in
detail, the hydrology / hydraulics for major
waterways and/or floodplains, complex
catchments
or
significant
drainage
structures (e.g. bridges).

1.2.2

Hydrology / Hydraulics

This section details the requirements placed


on internal and consulting engineering
services regarding hydrologic assessment
and hydraulic design.
1.2.2.1 Departmental
Designers

Civil

Departmental civil designers engaging in


drainage design must be suitably qualified,
have demonstrated capability in drainage
design and should also have attended the
mandated departmental technical training
courses in road drainage.

o Hydraulics package within 12d


Model
o PC DRAIN
o DRAINS
o CULVERT
(Departmental
software) (refer Section 1.2.2.5)
o CulvertMaster & FlowMaster
Multi-purpose:
o Drainage Design Assistant (DDA)
Basic Tools
Use of any other software package must be
approved in writing by Principal Engineer
(Road Design Standards), Road Planning &
Design Section, Engineering & Technology
Division. Refer also to Sections 1.2.2.5,
1.2.2.6 and 1.2.2.7.
1.2.2.3 Prequalification
Consultant Engineers
Designers

of
and

consultants
must
be
Engineering
prequalified with the department in the
prequalification category of Hydraulic
Design (HD) before providing any drainage
advice or undertaking any drainage design
or review on behalf of the department.

1.2.2.2 Approved Software for use


by
Departmental
Civil
Designers

1.2.2.4 Approved Software for use


by Prequalified Consultants

The following software packages are


deemed approved for use on projects by
departmental civil designers:

The following software packages are


approved for use on departmental projects
by prequalified consultants:

Hydrology:
o RAIN (Departmental software)

Hydrology:
o RAIN (Departmental software)
o RORB

March 2010

1-5

Department of Transport and Main Roads


Road Drainage Manual

o RAFTS

o URBS
o WBNM
o Bureau of Meteorology (BoM)
website
Hydraulic design:
o Hydraulics package within 12d
Model
o HEC-RAS
o MIKE 11
o PC DRAIN
o DRAINS
o EXTRAN, SWMM, UDD
o CULVERT
(Departmental
software) (refer Section 1.2.2.5)
o CulvertMaster & FlowMaster
Hydraulic design - 2D Modelling:
o SOBEK, DELFT-FLS
o MIKE FLOOD, MIKE 21
o TUFLOW
o SMS
Multi-purpose:
o Drainage Design Assistant (DDA)
Basic Tools
Water quality:

Chapter 1
Framework

1.2.2.5 Mandated Software for use


by all Civil Designers
The department may require the use of
specific software on projects to suit internal
processes and systems. This requirement
will be specified within this section of this
manual, released Planners and Designers
Instructions and/or relevant project
documentation.
All departmental designers and prequalified
consultants are required to undertake the 3D
modelling and quantity calculation of cross
drainage (culvert) infrastructure using the
departments CULVERT software. Using
this program ensures that departmental
processes and practices are followed for:
drawing drainage cross sections (e.g.
location, skew, invert heights, culvert
component details, bedding and
backfill material quantities, etc.)
when used within 12d Model; and
producing a culvert electronic model
that allows splicing into and storage
with the project electronic model,
when used within 12d Model.
The hydraulic design of standard culverts
should also use CULVERT, however other
approved hydraulic design software can be
used. Where other software or manual
methods are used, CULVERT should be
used as a check.

o MUSIC
o AQUALM
The department may allow pre-qualified
consultants to use other software packages,
but use of any other software package must
be approved in writing by Director
(Hydraulics),
Hydraulics
Section,
Engineering & Technology Division.
Refer also to Sections 1.2.2.5, 1.2.2.6 and
1.2.2.7.
March 2010

1-6

1.2.2.6 Other
Tools

Computer

Based

Further to Sections 1.2.2.2 and 1.2.2.4, any


spreadsheet or computer based tool
developed and then used to assist with
drainage design must be checked and tested
for applicability, accuracy of results and
compliance to current standards and
methodologies as prescribed in this manual

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

/ departmental Standard Specifications and


Standard Drawings. Certification of design
is deemed to cover use of these
spreadsheets / tools.

Examples of these non-standard structural


drainage elements include:
all bridges;

1.2.2.7 Currency of Software


While it is desirable that all software
packages used, as listed in Sections 1.2.2.2
and 1.2.2.4, be the latest version, it is not a
requirement. However it is a requirement
that project workings must clearly state the
software package and version being used.
It is also a requirement that users check /
test software to be used for compliance to
current standards and methodologies as
prescribed in this manual / departmental
Standard Specifications and Standard
Drawings. There is often a time lag
between the release of updated manuals /
standards / specifications and updates to
software packages. Where software is not
compliant and difference is minor (major
differences would prohibit usage), any
output must be adjusted accordingly and
adjustments recorded and checked.
Again, certification of design is deemed to
cover this requirement.
1.2.2.8 Structural
Design
Drainage Elements

consultant, design must be approved by the


departments Bridge Design Section.

of

any proposal to replace an existing


bridge with a culvert;
any culvert installation using nonstandard components.

1.2.3

Environmental Design

While
this
manual
promotes
environmentally sustainable drainage, it
does not give guidance on all relevant
environmental design considerations and
criteria. Project specific details as defined
in environmental approvals, licences and
permits will need to be incorporated. Refer
also Chapter 2.

1.2.4

Catchment Hydrology in
Rural Areas

The Rational Method has been adopted to


determine the runoff from small, simple
rural catchments up to 25 km2 in area (refer
Chapter 5). The method is not applicable
for complex catchments, irrespective of
size. Complex catchments include:
multiple streams;

This manual does not cover the structural


design of drainage elements and structures.
While there are standard designs for many
drainage elements, non-standard drainage
elements and structures must be referred to
the departments Bridge Design Section
located within the Engineering &
Technology
Division
or
suitably
prequalified consultants. Where design is
conducted by a suitably prequalified

branched catchments;
mixed land use catchments;
situations where a catchment may be
inundated by another catchment;
situations where the catchment may
overflow into an adjacent catchment;

March 2010

1-7

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

Table 1.2.4 - Designers for Rural Catchment Hydrology


Catchment Size

Who decides if catchment/design is


simple or complex?

Hydraulics Section or
consultant
Simple Departmental
Region or District
Complex - Hydraulics Section
or consultant
Simple Departmental
Region or District
Complex - Hydraulics Section
or consultant

Over 25 km2
10 km2 to 25 km2

Hydraulics Section or consultant

Less than 10 km2

Departmental Region or District

Permissible Designer

Note: The Region/District may elect to refer simple designs to consultants or Hydraulics Section.

catchments with significant storage


capacity (dam, swamp and major
retention / detention basin); and
irrigated land.
Table 1.2.4 details both decision points and
permissible designers.
In this table,
Hydraulics Section refers to the section
located
within
the
departments
Engineering & Technology Division and
consultants refers to suitably prequalified
consultants.

1.2.5

Catchment Hydrology in
Urban Areas

The Rational Method has been adopted to


determine the runoff from simple urban
catchments up to 1 km2 in area and where
the flow path is relatively simple (refer
Chapter 5). Relatively simple is defined as
a channel / flow path with limited
floodplain storage or piped drainage
systems and/or with no detention / retention
basins or similar flood storages (natural or
artificial) in the catchment.
Designs and/or reviews of designs
involving more complex catchment
characteristics are beyond the capability of
Rational Method and more detailed

March 2010

1-8

computer modelling is generally required.


In these cases, project must be referred to
the departments Hydraulics Section or
suitably prequalified consultants. Complex
urban catchments include:
retention / detention basins or other
flood storage areas;
mainly pipe flow
overland flow;

with

limited

complex flow path(s) and/or channel


system.

1.3 Road Infrastructure


Delivery
1.3.1

Introduction

The departments delivery process for road


infrastructure is described and detailed in
the Preconstruction Processes Manual.
Delivery is undertaken using the OnQ
project management methodology.
Drainage requirements are normally
established during the identification of the
project solution options and selection
process.
Actual information and data inputs for
drainage projects are usually enhanced as

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

investigations are progressively completed


and the solutions developed through an
iterative process. This can occur over
several years and may involve various
studies that are commissioned during the
project solution identification process. The
outputs from these studies usually identify a
range of issues to be addressed by the
designer. The issues are included in the
Design Development Report Small
Projects (Form M4211) or the Design
Development Report Large Projects
(Form M4212) with the designer including
statements on how these issues were
considered and incorporated into the design.

Torres Strait Islander


Heritage Act 2003;

All drainage risks are to be recorded in the


Risk Management Record (Form M4213)
with all risks being mitigated or removed as
the design evolves. Remaining risks at the
end of the design development process are
to be costed and the total cost included in
the project estimate contingency amount.

Transport Infrastructure Act 1994;

1.4 Legislation and Policy


1.4.1

Introduction

The department has obligations under State


and Federal legislation.
The department also provides direction
through a number of key corporate
documents including the departments
Strategic Plan, Road Network Strategy,
Roads Implementation Program and
Environmental Management Policy and
Strategy.

1.4.2

Relevant Legislation and


Other Authorities

The following Queensland Legislative Acts


are most pertinent to road drainage:
Aboriginal Cultural Heritage Act
2003;

Cultural

Nature Conservation Act 1992;


Land Act 1994;
Coastal Protection and Management
Act 1995;
Environmental Protection Act 1994;
Fisheries Act 1994;
Sustainable Planning Act 2009;
Soil Conservation Act 1986;
State Development and Public Works
Organisation Act 1971;

Water Act 2000;


Workplace Health and Safety Act
1995.
The following Commonwealth Legislative
Act also applies to road drainage on
Queensland roads:
Environment
Protection
and
Biodiversity Conservation Act 1999
(Commonwealth Act).
The following departments and authorities
are considered key stakeholders with
respect to road drainage and who also
administer some of the above listed Acts;
Department of Environment and
Resource Management (previously
Queensland
Environmental
Protection
Agency
(EPA)
&
Queensland Department of Natural
Resources and Water (NR&W));
Department
of
Employment,
Economic
Development
and
Innovation (previously Queensland
Department of Primary Industry and
Fisheries);

March 2010

1-9

Department of Transport and Main Roads


Road Drainage Manual

Local authorities / Councils


The Queensland Government has also
constituted numerous Statutory Authorities
across the state to perform specific
functions with relation to water resources in
local geographic areas.
These include:
Drainage boards;
Water boards;
Water authorities;
River(s) improvement trusts;
River(s) commissions;
Advisory councils;
Water management authorities;
Basin commission;
Referral panels;
Water supply authorities; and
Sun Water
The Workplace Health and Safety Act 1995
and the Workplace Health and Safety Act
Regulation 1997 set out the requirements
for safety requirements in workplaces.
Designers of drainage infrastructure need to
consider each installation as a workplace
during maintenance operations and
incorporate
provisions
that
permit
maintenance work to be completed in an
appropriate way that manages exposure to
risk in accordance with amendments to the
Workplace Health and Safety Act issued in
June 2007 (which addresses safety in
design and obligations of designers).
It is recommended that project teams
include or have access to personal who are
current in their knowledge, understanding
and application of the above legislation and
the functions and responsibilities of local
authorities etc.

March 2010

1-10

Chapter 1
Framework

1.4.3

Key Aspects of Road


Drainage Policy

Areas of hydraulic design that may require


considerations include:
changes to afflux from drainage
works that cause adverse impacts on
neighbouring property;
diversion of runoff to a different
point of discharge than that occurring
naturally;
concentration of runoff into a culvert
or open channel which is not a
watercourse
and
where
the
concentration is not dissipated by the
time the downstream flow reaches a
property or development boundary;
changes to stream morphology; and
outlet works on drainage structures
which do not sufficiently dissipate
energy, prevent scour or limit
siltation.
All of these situations should be avoided. If
avoidance is not possible, the relevant
stakeholders must be consulted. For a
discussion of the wider legal issues relevant
to the design of stormwater and drainage,
refer to the 2008 edition of the Queensland
Urban Drainage Manual (QUDM), as
published by the Department of Natural
Resources and Water (which became
Department of Environment and Resource
Management in March 2009).

1.4.4

Community Engagement

The department has a strong commitment to


engaging all stakeholders, including other
levels of government, industry and the
public. This is outlined in the departments
Community Engagement Policy, Standards,
Principles and Guidelines.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

consultation programme should


gather as much relevant data on
flooding as possible.
This
information can include detailed data
such as historical flood levels or
more general descriptions of flow
patterns or how often a road is
closed.

Community engagement can assist in


integrating economic, social, environmental
and engineering objectives of road
infrastructure delivery. With respect to
road drainage, community engagement is
both:
a process to improve departmental
understanding
of
stakeholder
expectations; and
an avenue for stakeholders to be
advised of constraints, technical
issues and possible options.
Public requests to reduce standards can be
considered, but not adopted at the expense
of policy, safety and sustainable
engineering.
The degree and timing of community
engagement regarding drainage will vary
between projects. However, the department
emphasises the need to commence
community engagement as early as possible
before decisions are made.
Flooding and drainage issues are important
to the community and effort is required to
gather and understand these concerns.
Members of the public may also provide
local data / knowledge for use in the
environmental assessment or design of a
project.
Flooding and drainage are relevant in
several parts of community engagement
programmes, with the main aspects as
follows:
Data collection.
Information
gathered from the community and
other stakeholders, particularly those
who have long term first hand
knowledge of the area, is often very
valuable and helps to establish
parameters for hydraulic analysis. In
the early phases of a project, the

Assessment of flood criteria. The


community and stakeholders may
have opinions on the level of service
that would be acceptable. While this
is only one input to this question, it is
valuable and can assist in
determining the most appropriate
level to be adopted.
Assessment of impacts.
The
community who may be affected by a
road project (say in areas where
afflux occurs) can be consulted to
discuss the individual levels of
impact.

1.4.5

Impacts on Landowners

Road drainage may have impacts on


neighbouring landowners and these
landowners may be concerned if there are
adverse impacts.
There is even the
potential for legal action in some situations,
so these impacts must be carefully
analysed. In some cases, adverse impacts
are unavoidable and appropriate mitigation
or compensation may be required.
Adverse impacts on landowners include:
(a) Afflux, or changes in flood levels on
the property. In particular, roads may
increase upstream flood levels.
(b) Changes in distribution of flood flows.
The road drainage may divert flow and
this could be a problem, if additional
flow is diverted on to a property or if

March 2010

1-11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

water is diverted away from a farm


dam for example.

statutory authorities should be undertaken


as applicable.

(c) Changes in flow velocity.


Flow
velocity is increased through bridges
and culverts and this could cause scour
damage.

Roads are a major land use and do influence


natural, built, commercial and social
environments. These influences result from
the design and selection of drainage
infrastructure and associated works and
designers, constructors and maintenance
personnel need to be cognisant of the
influence of their designs, construction
practices and procedures.

(d) Concentration of flow. Culverts may


concentrate flow that was previously
widespread sheet flow and this could
cause scour or other problems.
(e) Scour. Concentrated and high velocity
flows near culverts / bridges may cause
damage to neighbouring properties.

Inappropriate road drainage infrastructure


for example, can change the characteristics
of a waterway by altering:

(f) Water quality. Runoff from the road


surface and high velocity flows could
cause downstream water quality
impacts.

flooding patterns, including flow


distribution;

(g) Duration of inundation. This can be an


issue with agricultural land and some
urban areas, if the road increases the
time of inundation.

peak water levels;

1.5 Guiding Principles

erosion and sedimentation patterns;

This section introduces guiding principles


that should be part of the various
consideration processes involved in the
planning, designing, constructing and
maintaining road drainage infrastructure.

fauna transfer; and

1.5.1

Drainage Infrastructure
and Land Use

Changes in land use may create changes to


existing drainage patterns with associated
environmental, commercial or social
consequences.
The management of
drainage resulting from different land uses
or changes to land uses is usually
undertaken by a primary authority with
responsibility for the catchment which is
usually a local authority. Reference to
relevant government departments and

March 2010

1-12

flood heights;

water velocities, especially through


bridges and culverts;
duration of inundation;

terrestrial and aquatic fauna habitats.


These changes may create impacts that lead
to in service issues that require additional
unplanned investment to address:
flood mitigation works;
erosion and sedimentation problems;
reductions
valuations;

in

adjoining

salinity issues; and


increases in pollutant levels.

land

Department of Transport and Main Roads


Road Drainage Manual

1.5.2

Assessment of Future
Development

The department needs to plan and design


drainage works based on assumptions of
future catchment and floodplain conditions.
This can be difficult but nonetheless
important. Changes in conditions can occur
during the life of drainage infrastructure.
These changes can affect the runoff to the
road or changed floodplain / channel
conditions may affect flood levels adjacent
to the road.
The department has a responsibility to
provide advice on any development
applications that affect the department's
infrastructure, and can ensure that suitable
mitigation measures are applied before
approval can be provided.
Mitigation
measures include detention storages,
channel works or other similar measures.
As well as being effective initially, some
mitigation measures may not be effective in
the long term and comments need to be
made on these. For example, localised
dredging may be filled by later sediment
movement or cleared channel vegetation
may regrow. As well, cumulative impacts
may need to be considered, where
individual developments may have minor
impacts that are acceptable, but the impact
of a number of similar developments may
be unacceptable when all have occurred.
It is important that the department maintain
close review of any potential impacts on
drainage and ensure that these are managed
appropriately.

1.5.3

General Hydraulic and


Environmental
Consideration

This section provides an overview of


general hydraulic and environmental

Chapter 1
Framework

considerations
infrastructure.

relating

to

drainage

A road embankment may form an


obstruction to a natural waterway, bushland
corridor or floodplain if there are
insufficient or unsuitable bridge and/or
culvert
openings
in
the
road.
Considerations of the following may
therefore be required:
an unacceptably high increase in
upstream flood level (afflux) may
have adverse impacts on property,
infrastructure or the environment;
an increase in stream velocities
through the structures may initiate or
cause continuing erosion downstream
of the road;
movement of fish may be affected by
higher velocities or changes in the
channel invert level and/or extended
culvert lengths;
restricted flow may cause increased
time of inundation of the land
upstream of the road. In some cases
the road may be designed as part of a
large retention basin, and in this
situation peak discharges in floods
are reduced; and
roads may restrict the movement of
fauna from one side of the road to the
other. This will be identified in
environmental studies and drainage
structures may then be designed to
allow for this movement (e.g. one
culvert cell with a higher invert
level).

1.5.4

Economic Considerations

The
provision
of
road
drainage
infrastructure is typically a consideration of
the desired level(s) of flood immunity.

March 2010

1-13

Department of Transport and Main Roads


Road Drainage Manual

Solutions to deliver the preferred levels of


immunity may be delivered at initial
construction, staged construction may be
used to achieve higher levels over time, or
the lower levels may be acceptable if
alternative routes are available.
Where budgets restrict construction of
infrastructure that delivers the desired level
of flood immunity, a risk analysis that
details the costs of providing infrastructure
versus those of not providing drainage
infrastructure to the nominated standard is
necessary.
While budget may affect some aspects of
road design, other factors such as safety
cannot be compromised.

1.5.5

Maintenance

The provision of any type of road drainage


infrastructure should not be undertaken
without due consideration of maintenance
practices. Key issues include:
the need to provide adequate access
for maintenance purposes. A lack of
access will lead to a corresponding
lack of maintenance, and hence
possible failure of the drainage
infrastructure with respect to
performance;
consideration of the cost and
frequency of maintenance activities;
an understanding of what equipment
will be needed to undertake the
maintenance work;
assessment
of
maintenance
equipment usage and that it could
lead to failures;
assessment of the maintenance
activity and possible special safety
considerations.

March 2010

1-14

Chapter 1
Framework

Specific guidelines relating to the


consideration of maintenance issues are
provided at several locations within this
manual, but in particular, Chapter 14.

1.6 Drainage Issues


This section outlines key drainage issues
that planners and designers need to be
aware of and incorporate into their thinking
during the planning, designing, constructing
and
maintaining
road
drainage
infrastructure.

1.6.1

Planning the Drainage


System

Planning of the drainage system will likely


involve input from professionals in many
fields, and requires detailed consideration
of
environmental,
recreational,
transportation, traffic, social, economic, and
infrastructure needs. Because of the high
costs associated with drainage works, the
drainage aspects of a project should be
placed high in the hierarchy of the planning
process.
Drainage is a grade related service (i.e.
relies on gravity) and therefore can often
impose a significant constraint within the
planning process.

1.6.2

Urban Drainage

Extracts from QUDM (NR&W 2008) are


included in this section to enhance
understanding of urban drainage systems,
although this is beyond the scope of this
manual.
QUDM is not a stand-alone planning and
design
guideline
for
stormwater
management. It must be used in
coordination with other recognised manuals
dealing with specialised stormwater topics.

Department of Transport and Main Roads


Road Drainage Manual

The list of broad stormwater objectives of


QUDM (NR&W 2008) (as listed in Section
1.03 of that manual) is:
(a) Protect
and/or
enhance
downstream
environments,
including
recognised
social,
environmental
and
economic
values, by appropriately managing
the quality and quantity of
stormwater runoff.
(b) Limit flooding of public and private
property
to
acceptable
or
designated levels.
(c) Ensure
stormwater
and
its
associated drainage systems are
planned, designed and managed
with appropriate consideration and
protection of community health and
safety
standards,
including
potential impacts on pedestrian and
vehicular traffic.
(d) Adopt and promote water
sensitive
design
principles,
including appropriately managing
stormwater as an integral part of
the total water cycle, protecting
natural features and ecological
processes within urban waterways,
and optimising opportunities to use
rainwater/stormwater
as
a
resource.
(e) Appropriately integrate stormwater
systems into the natural and built
environments while optimising the
potential
uses
of
drainage
corridors.
(f) Ensure stormwater is managed at a
social,
environmental
and
economic cost that is acceptable to
the community as a whole and that
the levels of service and the
contributions to costs are equitable.

Chapter 1
Framework

(g) Enhance community awareness of,


and
participation
in,
the
appropriate
management
of
stormwater.
All of the objectives may not be relevant in
all circumstances and individual objectives
may be expanded to highlight site-specific
issues.

1.6.3

Rural Drainage Systems

Many departmental drainage structures in


rural areas are located on channels (both
natural and artificial) and include farm
drains, natural gullies, creeks and rivers.
Most of the guidance in this manual, and
many other manuals, is concerned with
these watercourses, and appropriate
hydraulic design is needed for each. These
design considerations include selection of
appropriate structures, sizes and locations
as well as assessment of afflux, flood
immunity and scour. In some cases, future
catchment development may need to be
considered. This is especially important if
urbanisation is possible in the future.
However in purely rural catchments,
specific consideration may need to be
provided for special circumstances, such as
soil conservation works, farm dams or laser
levelling of farmland.
As well, there are particular drainage
problems with extensive flat floodplains
where shallow overland flow occurs. There
is a risk that this overland flow will be
concentrated by drainage structures and
sufficient structures must be placed in
suitable locations to limit this concentration
and to spread the flow. Also in these sheet
flow situations, even small structures such
as levees can affect flow patterns
significantly and these changes in flow

March 2010

1-15

Department of Transport and Main Roads


Road Drainage Manual

patterns must be considered to ensure that


the structures are located correctly.
In some agricultural areas, the duration of
flood inundation is important, therefore the
design of roads and road drainage must
consider this aspect.

1.7 Climate Change


The department needs to consider the
possible impacts of climate change on many
aspects during the planning and design
process. Part of this consideration is the
possible impact on road drainage. This is
important because of the value and critical
significance of road drainage infrastructure.
If a significant change in the risk is
identified, the department will update
relevant procedures.
With respect to recommended allowances
for increases or possible changes to rainfall
intensities, the department takes guidance
from Australian Rainfall and Runoff, A
Guide to Flood Estimation (AR&R),
published by the Institution of Engineers
Australia. AR&R is continually monitored
by the department and any changes in
procedures that become apparent will be
incorporated into this manual.
The previous edition of this manual
recommended that designs in coastal
regions should make an allowance for an
increase in sea level of 300 mm for the
effects of climate change. This allowance
had considered the consensus of opinion
concerning sea level rise and could be
regarded as a conservative approach. Many
local authorities in coastal regions of
Queensland also make a similar allowance.
Selection of an approach for design ocean
levels allows for a number of conditions
such as the tides, storm surge and the risk
of coincidental occurrence of floods and

March 2010

1-16

Chapter 1
Framework

high ocean levels.


Generally the
combination of these conditions in design
procedures leads to a conservative ocean
level for design. However the adoption of
an allowance for sea level rise has been
accepted without much debate.
An allowance for a change in rainfall
intensity for application in design for
bridges and other drainage structures is
more difficult.
The Queensland
Government publication ClimateSmart
Adaptation 2007-12 suggests that bridge
design should provide for an increase in
rainfall intensity. However, AR&R is the
major reference in this field and almost
universally accepted for design rainfall
estimation throughout Australia.
The
department therefore adopts the rainfall
intensities as recommended by AR&R.
The design rainfall intensities (data and
procedures for estimating design rainfall)
were prepared by the Bureau of
Meteorology and published in the 1987
edition of AR&R.
Any update of rainfall intensities /
estimation procedures, will rely on the
statistical analysis of recorded rainfall data
and historical records. Review of AR&R is
currently underway and possible changes in
rainfall intensities are being researched /
studied.
Because of the potential significance of
impacts, the department will continue to
monitor
relevant
literature
and
recommendations and if changes are
appropriate / required, will update this
manual to allow for climate change
impacts.
This means that there is no recommendation
for incorporation of a change in rainfall
intensity for road drainage designs included
in this edition of the manual.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 1
Framework

However because of the uncertainty in


estimates of possible changes in rainfall
intensity related to climate change, the
department requires the assessment of
potential impacts using the procedures
included in the governments Climate
Change
Impact
Statement
(CCIS)
procedure.
Consideration of rainfall
intensity is only one portion of the CCIS,
but is the only part mentioned here. This is
a risk based procedure essentially involving
a sensitivity test, where the impacts of
higher rainfall intensities are tested. These
implications are then reviewed and
appropriate mitigation or management
measures are developed. Climate modelling
does not give a clear indication of changes
in rainfall intensity, so the sensitivity test is
the only feasible method of analysis.

While this procedure is adopted, the general


drainage analysis does not assume any
changes in rainfall intensity.

March 2010

1-17

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Chapter 2
General Design
Requirements

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Chapter 2 Amendments Mar 2010

Revision Register
Issue/
Rev
No.
1

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Table of Contents
2.1

Introduction

2-1

2.1.1

Definitions

2-1

2.1.2

Determining and Understanding Requirements

2-1

2.1.3

Immunity Criteria

2-2

2.1.4

Design Life / Service Life

2-2

2.2

Road Locality

2-3

2.3

Design Considerations

2-4

2.3.1

Identifying Design Considerations

2-4

2.3.2

Geometric Considerations

2-4

2.3.3

Geographic Considerations

2-6

2.3.4

Environmental Considerations

2-8

2.3.5

Selection of Drainage Infrastructure

2-9

2.3.6

Maintenance Considerations

2-13

2.3.7

Safety Considerations

2-14

2.3.8

Staged Construction of Roads

2-15

2.4

Design Controls

2-17

2.5

Design Criteria

2-17

2.5.1

Introduction

2-17

2.5.2

General Hydraulic Criteria

2-18

2.5.3

Cross Drainage Criteria

2-26

2.5.4

Longitudinal Drainage Criteria

2-27

2.5.5

Road Surface Drainage Criteria

2-29

2.5.6

Immunity Criteria for the National Land Transport Network

2-29

2.5.7

Immunity Criteria for State Controlled Roads - Rural Catchments

2-29

2.5.8

Immunity Criteria for State Controlled Roads - Urban Catchments

2-29

2.5.9

Environmental Criteria

2-30

2.6

Water Sensitive Urban Design

2-31

2.7

Extreme Rainfall Events

2-32

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

2.8

2.7.1

Erodible Soil Environments

2-32

2.7.2

Excessive Flooding

2-33

Self Cleaning Sections

2 2.

March 2010

iv

Chapter 2
General Design Requirements

2-33

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Chapter 2
General Design Requirements
2.1

Introduction

The purpose of this chapter is to introduce


and discuss a number of general design
requirements
for
road
drainage
infrastructure. The requirements presented
cover a range of topics. More specific
design requirements are contained in
relevant chapters.
The intention is that this chapter should be
referenced first to establish general and
some specific drainage requirements for a
project. Topic specific chapters, such as
Chapters 7, 8, 9 and so on should be then
referenced as applicable / required.

2.1.1

Definitions

The term design requirements, encompasses


all design:
considerations;
controls;
criteria; and
standards ... that must be included in
or be part of the design process.
Design considerations encompass all
aspects, issues, functionality, expectations,
demands, constraints, risk and cost, that
need to be appropriately addressed, or to be
taken into account, in order to satisfy design
criteria and determine trade-offs. Refer
Section 2.3 for further discussion regarding
design considerations.
Design controls are aspects of the road
environment or project that cannot be
changed, or are extremely difficult to

change, and therefore place some restriction


or control on the design. Refer Section 2.4
for further discussion regarding design
considerations.
Design criteria set the expected level of
achievement or conformance for relevant
design parameters or design inputs. The
design criteria ensure that the end result can
be judged and defended. An example of a
design criterion with respect to road
drainage would be the average recurrence
interval for design for a particular project or
structure. Section 2.5 presents a number of
fundamental design criteria while more
specific criteria will be found in relevant,
topic specific chapters.
Design standards however, set approved or
prescribed values or limits for specific
elements of design or set procedures and/or
guides that must be followed. A design
standard with respect to road drainage
would be the use of the Rational Method to
determine the runoff from a small rural
catchment. Design standards are presented
throughout this manual.
Both design criteria and design standards
set the mandatory limits designers must
work within and/or achieve.

2.1.2

Determining and
Understanding
Requirements

To assist in determining and understanding


the design requirements for a road project,
the Road Planning & Design Manual
(RPDM) describes a Hierarchy of Roads,
and this provides an overview of the road

March 2010

2-1

Department of Transport and Main Roads


Road Drainage Manual

network and defines the roads that the


Department of Transport and Main Roads
has stewardship for in Queensland. It
develops the philosophy that the road
hierarchy provides a useful planning tool
for many decision making activities
concerning the road network.
This
hierarchy permits the department to set the
design requirements for projects on roads
according to their function within the
network.
The department also sets the design
criteria and design standards for road
links that will deliver a planned level of
service. Therefore, the overall standard of
drainage infrastructure for a specific road
project needs to reflect and be consistent
with the function of that road within the
road hierarchy.
In order to satisfy the relevant design
criteria for a project and to determine
appropriate trade-offs, a number of aspects,
issues,
functional
requirements,
expectations, demands, constraints, risk
elements and costings need to be
appropriately addressed, or taken into
account. These design considerations,
with respect to drainage infrastructure, are
progressively identified and developed to
address
geographic,
hydraulic,
environmental, cultural, native title and
land use issues.
During pre-project planning, target drainage
criteria for the road are established and
documented in relevant Road Route
Strategies and Road Link Plans. The
respective design criteria are then
progressively addressed during the Design
Development Phase. Excessive cost to
provide drainage at the specified Average
Recurrence Interval (ARI) may result in a
review of the criteria in order to achieve a
fit for purpose outcome or may result in

March 2010

2-2

Chapter 2
General Design Requirements

the project being delivered in stages over


multiple years to achieve the desired
outcome.

2.1.3

Immunity Criteria

The immunity criteria discussed within this


manual and set in the design brief / contract
documentation relates to individual
drainage components (such as cross or
longitudinal drainage) and not the road
project, section or link. Furthermore, by
setting the immunity criteria for various
drainage components on a project does not
imply that the road inherits the same
immunity level(s). It is extremely difficult
to assess immunity and set criteria for a
road.
Refer Chapter 3 (particularly Section
3.2.3.2) for a more detailed discussion
regarding this issue.

2.1.4

Design Life / Service Life

It is important to define, in general terms,


the difference between design life and
service life.
The design life of a component or system
of components is the period of time during
which the item is expected, by its designers
or as required by specification, to work or
perform its intended function within
specified design parameters / operating
conditions. In other words, the design life
is the life expectancy of the item under
normal / specified operating conditions.
With respect to road drainage, operating
conditions can include:
environmental / atmospheric
geographic conditions;

foundation, bedding and support /


cover conditions; and
traffic and loading.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

For example, the department specifies the


design life for a new structural component
(such as a culvert) as 100 years. Therefore,
it is expected that the culvert will last 100
years before replacement or major repair is
expected.

2.2

The service life of a component or system


of components is the period of time over
which the item actually provides adequate
or satisfactory performance before repair or
replacement is required.

2.2.1.1 Road Environment

If the operating conditions over the life of


the component or system remain within the
original design parameters, theoretically,
the design life will equal service life.
However, if the operating conditions are or
move outside of the original design
parameters, service life will be less than
design life.
In some situations, this
reduction in time can be considerable,
leading to premature failure.
In relation to drainage infrastructure,
designers should be mindful of these two
terms and ensure, where possible, that the
designed drainage components or system
are appropriately selected for the
anticipated operating conditions.
For
example, concrete culvert components will
have a reduced service life if standard
culvert components are used in:
streams carrying abrasive material
(such course gravel or small stones);
or
aggressive environments such as sea
water; or
situations where they are subject to
excessive peak loads (generally from
overloaded vehicular traffic).

Road Locality

There are two major environments or zones


potentially affected by drainage and are
defined as the road environment and the
external environment.

The road environment is the zone which the


Department of Transport and Main Roads
has responsibility for and therefore is under
its control. It is defined as the road corridor
as defined by property boundaries (also
known as road reserve). It is important to
note that not all boundaries are clearly
defined, particularly within large western
properties and reserves. In these situations,
the road reserve is usually based about the
existing road centreline and planners and
designers need to further investigate to
establish applicable boundaries.
2.2.1.2 External Environment
The external environment is the zone
outside of the road corridor which may
include sensitive areas such as wetlands
rainforest, sand dunes, waterways or private
properties or other infrastructure (e.g.
railways). The external environment may
extend for some distance from the road
environment and is not the responsibility of
the department. However the department
needs to liaise or work with relevant
stakeholders and authorities with respect to
any proposed project as drainage work
within the road environment may affect the
external environment both upstream and
downstream of the project.

March 2010

2-3

Department of Transport and Main Roads


Road Drainage Manual

2.3
2.3.1

Design
Considerations

(a) Geometric

Identifying Design
Considerations

(c) Environmental

The construction of new or upgraded road


drainage infrastructure may lead to changes
in the existing road and external
environments. Problems associated with
erosion and sedimentation, flooding
(changes in peak water levels) and water
quality are of concern to the department,
adjacent land owners, road users and the
local community. The occurrence of these
problems, particularly after a project is
completed, can be costly to remedy and
may lead to reduced amenity.
Effective project planning covering both the
road and external environments plays a
major role in minimising the potential for
adverse impacts.
The planning and design of road drainage
infrastructure can be quite complicated and
involves the consideration of a diverse set
of data in order to develop the most
appropriate drainage solution for a project.
Collaborative planning by a group of
professionals with complementary skills is
often a productive way to identify all
aspects, issues, functional requirements,
expectations, demands, constraints, risks
and possible costs to be considered in a
project.
The department has identified a generic set
of considerations that address drainage
issues across Queensland.
In order to develop the most appropriate
drainage solution, the project team for each
project, must select applicable drainage
considerations
from
the
following
categories:

March 2010

2-4

Chapter 2
General Design Requirements

(b) Geographic

(d) Crossing Type


(e) Maintenance
(f) Safety
(g) Staged Construction of Roads
It should be noted that identified design
considerations may present several options
when being addressed. It is possible that
upon further consideration or review, some
design considerations may no longer be part
of a project while others develop into key
design controls.

2.3.2

Geometric Considerations

There are two aspects of geometrics that


must be considered in the drainage design
for a road project.
Some parts or
components of these aspects may in turn
become design controls. The first aspect
deals with the geometry of the stream or
creek and the second aspect deals with the
geometry of the road-stream crossing.
2.3.2.1 Stream Geometrics
It is important to determine the geometry of
the stream or flow path, in particular:
stream longitudinal alignment;
stream gradient; and
channel shape.
Stream alignment refers to the natural
meanders of the stream or creek channel.
While most streams have only one
alignment for all flows, it is possible to
have the situation where the alignment for a
low flow differs from the alignment for a
high flow in the same stream.
This
situation must be identified and considered

Department of Transport and Main Roads


Road Drainage Manual

when designing the road-stream crossing.


It is possible to alter the alignment of
existing streams to improve the hydraulic
performance of the road-stream crossing,
however it is preferable to maintain or
preserve the existing stream alignment as
changes will affect the existing flow
parameters (velocity, depth of flow and
energy). Furthermore, it is important to
note that licences are required from the
Department of Environment and Resource
Management to change the alignment of
any defined watercourse. Experience has
shown that the process of obtaining relevant
licence maybe difficult.
Stream gradient refers to the vertical
alignment of the stream or creek and
changes to gradient will also affect flow
parameters. Gradient has a significant
influence on flow velocity and velocity in
turn has a significant affect on sediment
transport and scour potential.
Channel shape needs to be considered as it
will tend to dictate the size and
configuration of drainage structures.
Altering the channel shape to accommodate
a drainage structure will affect flow
parameters and could increase the risk of
erosion. It is preferable to maintain or
preserve the existing channel shape as close
as possible and culvert structures should be
designed to fit the shape of the stream.
Some channels may not contain all of the
design storm runoff and overtopping of the
banks will occur.
Multiple culvert
installations for the one catchment will be
required and in this instance, specialist
advice / design will be required.
Lastly, designers must include an
understanding of stream morphology when
considering stream geometrics. Streams are
dynamic and can change over time. This
aspect is important.

Chapter 2
General Design Requirements

2.3.2.2 Road Geometrics


As identified in Chapter 1, drainage is an
integral component of road infrastructure
and therefore drainage design cannot be
undertaken in isolation from the geometric
design of the road.
In the design of the road-stream crossing, it
is important to consider the skew angle
between the road alignment and drainage
structure. Keeping the skew angle as small
as possible (or eliminating it altogether)
reduces costs and construction difficulty
and is therefore the most desirable option.
Given that it is highly recommended to
preserve
stream
alignment,
this
consideration however does not imply any
priority of drainage over road alignment
and high skew angles may be unavoidable
at times.
The design of the vertical alignment should
be undertaken in conjunction with the
design of the drainage system. An initial
vertical alignment design would be used to
undertake the initial drainage design of
various structures. It may then be necessary
to adjust the vertical alignment in order to
achieve the most efficient and effective
drainage design (considering allowable
headwater levels, afflux and minimum
cover requirements for structures). In this
instance, the requirements for drainage may
become a design control on the vertical
alignment. However, the drainage designer
needs to be aware that constraints placed on
vertical alignment would make it a design
control on the drainage system and force
the design to change.
Furthermore, vertical alignment together
with cross-sectional crossfall of the road
also affects longitudinal drainage channels
(such as table drains) and therefore must be
designed considering minimum grade
requirements for flows and minimising
March 2010

2-5

Department of Transport and Main Roads


Road Drainage Manual

steeper grades where


velocities could result.

higher

Chapter 2
General Design Requirements

erosive

Another important aspect related to the


geometric design of roads is stormwater
runoff from the road surface. This aspect is
critical as water flow (and depth) on the
road surface relates to aquaplaning.
Surface flows are as a result of the
geometric road design (combination of
horizontal, vertical, cross section, crossfall
and superelevation elements) and therefore
any identified problems should be solved
and mitigated through amended geometric
road design.
A drainage solution to
aquaplaning should be only considered as a
last resort option. If a drainage solution is
required, specialist advice is highly
recommended in the development /
assessment of design options.
Lastly, where the possibility of stormwater
crossing over the road exists (whether
intentional or unintentional), adequate
stopping sight distance must be provided
and this factor could affect the vertical
alignment design (refer Road Planning &
Design Manual).

2.3.3

Geographic
Considerations

Geographic conditions play a significant


role in the determination of what type of
drainage structure and/or controls may be
adopted at a given location. Structures and
controls that are appropriate in one part of
the state may not be suitable in other areas.
This is also true for the prevention of
erosion.
This section discusses some key issues for
different situations and regions across the
state.
Most of the departments roads are located
in rural regions, so standard practices for

March 2010

2-6

the planning and design of road drainage


should address most of the issues that will
arise in these areas.
However, it is
important to note that these issues can also
apply also in urban regions. The design of
drainage systems in all regions should
ensure that the road level and associated
drainage infrastructure is adequate to
provide the specified level of flood
immunity.
Furthermore, the drainage
structures should be sized to ensure that the
flow velocities and afflux are acceptable.
Specific issues to be addressed include:
awareness of local drainage and
management plans;
ensuring property and crops will not
be affected by an increase in water
levels or duration of inundation;
concentration of flow on floodplains
should be minimised because of the
risk of scour;
maintaining free drainage, and not to
create ponding of low flows;
changes to flow patterns; and
consideration of seasonal variations
in hydraulic roughness linked to
changes in vegetation cover.
Urban regions have similar issues to rural
areas, but may also present other
constraints. Constraints may be present in
the form of adjacent infrastructure
(including businesses and housing) or a
limit in available space. Because of the
more intense level of development, afflux is
usually of more concern in urban areas than
in most rural locations. In addition, local
authorities may have prepared catchment or
stormwater management plans, which will
affect the future management of stormwater
and watercourses in the area. The drainage

Department of Transport and Main Roads


Road Drainage Manual

design should be compatible with these


plans wherever they exist.
Considerations in urban regions include:
provision for higher peak flows
arising from uncontrolled upstream
development (many local authorities
now require flow increases to be
mitigated);
assessment of the requirements of
any catchment management plan or
stormwater
management
plan
prepared for the watercourse;
need for pollution control measures;
interaction
of
road
drainage
provisions with existing services;
minimisation of ground disturbance
during construction as urban
environments often have limited
space for large control measures such
as sediment basins; and
consideration and control of afflux
effects. There is often a requirement
that negligible afflux increases be
generated upstream / downstream of
the proposed drainage structure.
With respect to possible changed
water levels, it is important that each
case is assessed fully in keeping with
a risk management approach.
Coastal regions provide unique conditions
and hence may other require special
considerations. Conditions include regular
tidal inundation, a corrosive environment
and sandy soils (i.e. soils with little
cohesion). Coastal environments are also
often highly sensitive to pollution.
Considerations in coastal regions include
ensuring that:
legal requirements with respect to the
protection of marine environments

Chapter 2
General Design Requirements

(e.g. protection of fish habitats and


marine plants) are met;
natural flow systems (e.g. tidal
exchange) are properly assessed and
will therefore not be compromised;
corrosion resistant materials are used;
potential Acid Sulphate Soils
(typically below 5 metres AHD) are
identified
and
managed
appropriately;
storm tide influence upon peak water
levels and greenhouse related sea
level rise is considered;
designs allow for the presence of
highly erodible or mobile materials
such as sand;
consideration is given to directing
drainage to natural channels or
swales rather than to hard structures;
and
bi-directional flows (due to tidal
effects)
are
identified
and
appropriately considered.
Design of road drainage in flat terrain is
often difficult for several reasons,
including:
(a) Flow velocities in flat areas are usually
low so larger structures are needed to
convey the flow.
(b) Flow may be widespread and/or
shallow so minor obstructions to the
flow may divert flow significantly.
These minor obstructions include
levees and other floodplain works.
Even the road itself may cause major
diversions.
(c) It is often difficult to determine the
catchment areas accurately because of
minimal relief in terrain and the

March 2010

2-7

Department of Transport and Main Roads


Road Drainage Manual

presence of minor obstructions as


discussed in (b) above.

(d) Poorly defined flow paths also mean


that it is sometimes difficult to place
culverts in the most suitable locations.
(e) In flat terrain, the impacts of the road
on flood levels may extend for
significant distances upstream of the
road. Where afflux is a concern, this
impact may often be critical.
(f) There is usually an increased risk of
erosion at culvert outlets because flow
will be concentrated by drainage
structures, particularly where there are
poorly defined flow paths and/or most
flow occurs across the floodplain.
In mountainous or steep terrain, the most
common factor influencing design is the
gradient of the natural ground. Issues for
consideration where topography is steep
include:
control of velocities in roadside
drains and culvert outlets;
collection and discharge of water
from the upslope side of the road to
the downslope side;
prevention of erosion at outlets onto
steep areas; and
need for small scale drop structures,
weirs or drop manholes.
Locations subject to inundation by water,
such as floodplains or coastal areas, either
by tidal flow or backwater, require careful
consideration
of
how
drainage
infrastructure will operate under a range of
water levels. The presence of high and low
water levels requires significantly different
approaches:

March 2010

2-8

Chapter 2
General Design Requirements

when downstream water levels are


high, the hydraulic capacity of a
structure may be limited; and
when downstream water levels are
low, high velocities can result,
thereby maximising the potential for
erosion to occur.
It is therefore very important that both cases
are considered during the design of
drainage infrastructure.
Regular inundation (i.e. changing water
levels) can also accelerate the erosion
process, through the saturation of banks,
which may then fail as water levels reduce.
In coastal regions, the effects of tidal
influence and climate change allowance
(refer Chapter 1) on areas of inundation will
need to be assessed.

2.3.4

Environmental
Considerations

Drainage has the potential of causing


environmental harm.
Therefore it is
important that environmental impacts are
assessed and mitigated (as appropriate) as
part of the development and operation of a
road drainage system.
The risk of scour / erosion and sediment
movement caused by the concentration of
flows that typically occurs with drainage
structures is of particular concern. Causal
factors, including changes in flood flow
patterns and changes in peak water levels
should also be checked. In some instances,
a new road embankment could lead to long
term ponding of water which in turn could
have adverse environmental impacts.
Environmental considerations will vary
significantly from project to project, and
hence it is not practical to list all potential
issues in this section. However, there are

Department of Transport and Main Roads


Road Drainage Manual

two types of environmental consideration


for which details have been provided.
These are; the provision for fauna passage
and the maintenance of water quality. In
many projects, it will be important to
ensure that the design of drainage
infrastructure adequately caters for the
existence of fauna, and also for the
maintenance (or improvement) of the
quality of stormwater runoff.
Chapter 4 describes the role of the
environmental assessment (process and
documentation) in obtaining and analysing
data for the purposes of identifying
potential environmental considerations for a
projects drainage design.
Careful
review
of
any
relevant
environmental assessment documentation,
including any recommended management
strategies, needs to be undertaken, as some
of these strategies may become design
requirements or criteria. The recommended
management strategies are generally based
on the requirements of relevant legislation,
policy, codes, guidelines and current best
practice within the industry.
Streams are usually riparian corridors.
Therefore, they can provide a corridor for
fauna movement as well as providing a
habitat for terrestrial and aquatic fauna. All
drainage works on the road should
minimise any adverse impact on these
important corridors.
Further to the above discussion, there are
two important documents that should also
be
referenced
when
considering
environmental issues.
The first is
Australian Runoff Quality A guide to
Water Sensitive Urban Design (2006),
published by Engineers Australia and the
second is Water Sensitive Urban Design

Chapter 2
General Design Requirements

Technical Design Guidelines for South East


Queensland (2006), published by Moreton
Bay
Waterways
and
Catchments
Partnership.

2.3.5

Selection of Drainage
Infrastructure

Determining the type of structure for any


crossing is an important consideration and
there are a number of factors that need to be
addressed in this process. It may be
necessary to assess several options of
different crossing type and size, in order to
appropriately meet the design requirements
and objectives. There are three main types
of cross drainage structures used on roads
and each has particular advantages and
disadvantages. The three types are bridges,
culverts and floodways as shown in Figure
2.3.5.
2.3.5.1 Relevant Factors
The relevant factors that need to be
considered in selecting infrastructure are
grouped into hydraulic and other factors.
The hydraulic factors include:
Flood
discharge.
Defined
waterways with a large discharge are
more suited to a bridge because of
the larger waterway area. The large
discharge will also generally occur in
rivers or large creeks, where a bridge
is more appropriate and cost
effective. Depending on location and
importance of road, in flat terrain
where the waterways are less defined
and road embankment is typically
low, a floodway may be the better
option.

March 2010

2-9

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Floodway

Bridge

Culvert

Figure 2.3.5 - Primary Drainage Infrastructure Types

Stream channel conditions and


topography.
Similarly with the
consideration of discharge, the shape
and size of the channel and the
catchment will also indicate a bridge,
culvert or floodway. Large and well
defined channels will be better suited
to a bridge, while less well defined,
smaller channels will be more suited
to a culvert, especially where
multiple openings are required (such
as on floodplains). Floodways again
could be considered, particularly in
flat terrain / low embankment
situations.
Afflux constraints.
The most
suitable structure may be indicated

March 2010

2-10

by the amount of flow that can pass


through / over the structure with
acceptable afflux. The location and
extent of afflux needs to be
considered in detail and the
alternatives assessed to minimise
afflux.
Debris properties.
Culverts will
normally have a smaller waterway
area and present a greater obstruction
to the flow. They are more prone to
collection of debris. If there is a
large amount of debris conveyed by a
stream, a bridge or larger culvert may
be more suitable.
Scour risk. Scour can be affected by
the size and type of waterway

Department of Transport and Main Roads


Road Drainage Manual

structure. If a structure concentrates


flow significantly, risk of scour may
be increased, so structures that spread
the flow may be favoured in these
locations.
This is especially
important for drainage in floodplains
where the flow paths may not be well
defined.
Other relevant factors that need to be
considered include:
Road alignment. Sometimes, the
alignment of the road is well defined
and this may not be the best
arrangement for drainage. This may
sometimes occur where land tenure
needs to be considered and the
alignment follows streams rather than
crossing at a zero skew. In these
cases, the sizing and locating of
drainage structures must be carefully
considered.

Chapter 2
General Design Requirements

of the most suitable / appropriate structure


for a particular crossing. These are listed in
Table 2.3.5.2.
2.3.5.3 Culvert Types
Selection of the culvert type is important in
some applications. The choice is between
the following predominate types:
pipes (any material type);
box culverts including slab link
culverts;
slab deck culverts (cast in-situ); and
multi-plate arches.
Arches and cast in-situ culverts are not
common and if being considered, specialist
advice must be obtained including from the
departments Bridge Design Section located
within the Engineering & Technology
Division.

Level of serviceability. This includes


the required flood immunity or
trafficability and the type of structure
that will be best for meeting this
requirement.

There are two issues of particular concern


for selecting the type of culvert. The first is
related to the waterway area at low flow
depths and secondly, the extent to which the
culvert spreads the flow.

Navigation.
Structures crossing
rivers where boat traffic needs to be
considered must allow for specified
clearances for this traffic.

Box culverts and slab deck culverts provide


for a greater waterway area at shallow
depths, while pipes need to flow at a greater
depth before the maximum flow capacity is
reached. The use of pipes however does
tend to spread the flow to a greater extent,
which is often desirable for consideration of
concentration of flow and risk of scour.

Soil conditions.
Particular soil
conditions, such as marine mud or
acid sulphate soils for example may
be a problem and this can affect the
selection of drainage structures.
Fauna and fish movement. This is
an important consideration in many
locations.
2.3.5.2 Bridge, Culvert or Floodway
There are a number of factors and issues
that need to be considered in the selection

March 2010

2-11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Table 2.3.5.2 - General Selection Factors

Structure

Bridges

Culverts

Floodways

March 2010

2-12

Advantages

Waterway area generally


increases with increased deck
height

Provides greatest flood


immunity

Large flow capacity

Fewer problems with debris

Deck widening does not affect


capacity

Less disturbance to riparian


environment about waterway

Less impact on fauna / fish


passage

Disadvantages

Higher design, construction and


maintenance costs

More structural maintenance


required

Spill slopes can be affected by


erosion (potential for costly batter
protection requirements
particularly for higher / exposed
approach embankments)

Pier and abutment can be


affected by scour

Increased buoyancy, drag and


impact risks

Susceptible to stream / channel


migration

Simplest structure to design /


construct

Generally require higher levels of


maintenance

Generally most cost effective


option

Most susceptible to failure

Can accommodate future


changes to road geometry

Higher siltation / debris risk


(blockage)

Less structural maintenance

Increased environmental impacts


(fauna / fish passage)

Can spread flows

Potential for scour at outlet

Subject to abrasion

Future extension may reduce


capacity

Potential for separation at joints

Potential for failure by piping


(leading to failure of embankment)

Allow water flow over road


immunity and safety issues

Increased disruption to traffic due


to overtopping

Can have higher construction


costs than culverts

Batter slopes can be affected by


erosion / scour (particularly for
higher embankments)

Generally have costly batter


protection requirements

Susceptible to stream / channel


migration

Can have environmental impacts


(fauna / fish passage)

Potential for failure of


embankment (depending on
provided protection)

Generally, simple to design

May offer environmental


advantages over culverts and
bridges, since they will tend to
spread flows more widely

Typically have low


embankments

Risk of scour to waterway and


surrounding land is reduced

Department of Transport and Main Roads


Road Drainage Manual

A further consideration for pipe culverts is


material type. There are several different
material types available:
reinforced concrete;
corrugated steel (plate or rolled);
polyethylene and polypropylene; and
fibre reinforced.
Reinforced concrete pipes are the most
common type used by the department. This
is primarily due to the products
availability,
strength,
serviceability,
durability and overall cost (design,
construction and operation).
The selection of steel rather than concrete
for the culvert material can be suitable as
these culverts are generally quicker and
easier to construct as well as easy (and
potentially cheaper) to handle and transport.
Designers should note that traditional steel
culverts are no longer used by the
department as they do not meet the required
100 years design life. To achieve this
requirement, steel culvert must have added
protective coatings (refer Section 9.2.6.8).
Steel culverts with protective coatings need
to be product approved by the departments
Bridge Design Section.
Polyethylene and polypropylene pipes are
limited to smaller diameter sizes, but can
provide an acceptable alternative in some
circumstances.
These pipes however
cannot be used in locations subject to traffic
loads unless they meet the requirements
within MRTS29 (TMR 2010c) and are
product approved by the departments
Bridge Design Section.
Fibre reinforced pipes are also limited to
smaller diameter sizes. These pipes have
some flexibility within the walls of the
culvert and tolerate construction loads and
low cover installations better than

Chapter 2
General Design Requirements

reinforced concrete pipes. Fibre reinforced


pipes must meet the requirements within
MRTS26 (TMR 2010c) and be product
approved by the departments Bridge
Design Section.
Considerations in selecting culvert type and
material type are product and installation
costs, availability (including transport of
product to site), constructability, site
conditions, environmental requirements,
product longevity and serviceability and
will also influence the decision as to which
structure type and material type is most
appropriate for a given site.
Many of these aspects and issues will be
discussed and/or addressed throughout the
remaining chapters of this manual.

2.3.6

Maintenance
Considerations

The provision for maintenance is an integral


component of the planning and design
phases of road drainage.
Adequate
maintenance is necessary for the proper
operation of the drainage system. The lack
of maintenance is one of the most common
causes of failure of drainage systems (and
erosion and sediment controls). This may
be attributed to reasons such as a significant
reduction in hydraulic or storage capacity
(e.g. blockage by debris or sediment).
Specific details on maintenance procedures
and requirements for road drainage systems
are provided in Chapter 14 of this manual.
To enable maintenance to be properly and
safely undertaken during road construction
and operation, consideration must be given
at the design stage to the requirement of the
Workplace Health and Safety Act 1995 to
make a safe maintenance workplace.

March 2010

2-13

Department of Transport and Main Roads


Road Drainage Manual

2.3.7

Safety Considerations

An integral aspect of the detailed design of


all road drainage systems is the underlying
consideration of safety.
Some of the safety issues that require
consideration as part of the road drainage
design process, excluding workplace health
and safety issues, are described below.
Reference should also be made to Chapter
12 of Queensland Urban Drainage Manual
(QUDM) (NR&W 2008).
Maintenance Access - Safe access
needs to be provided to all drainage
structures that require either ongoing
(i.e. mowing of drains) or occasional
(i.e. removal of debris) maintenance.
This access is required for vehicles
and maintenance crews depending on
the type of maintenance that will be
undertaken. Safe access to erosion
and sediment control devices during
the construction phase should also be
allowed.
Human Safety - Where long culverts
potentially
provide
a
hazard
(particularly in urban areas) to human
safety, preventative measures should
be considered.
Safety measures
include fencing, swing gates and
grates at culvert inlets. Any safety
device needs to ensure that it
prevents both access to the culvert
and trapping of a human against the
grate. The effect of any proposed
human safety measure on culvert
capacity and efficiency needs to be
checked.
Traffic Safety - Projecting culvert
ends have the potential to act as
obstructions to out of control
vehicles. Where there are no safety
barriers; culvert ends should be

March 2010

2-14

Chapter 2
General Design Requirements

designed to not present an


obstruction. If obstructions from
projecting culverts or head walls are
unavoidable then safety barriers
should be considered.
Floodway Safety - The main issue
associated with safety at floodways is
adequate sight distance for drivers to
ensure vehicles can stop before
entering the floodwaters. Preferably,
the floodway longitudinal profile
should be horizontal so that the same
depth of water exists over the entire
floodway length.
The floodway
length should be limited and on a
straight stretch of road where
possible. Adequate permanent and
temporary signing must be erected.
As flood water recedes, silt and
debris can be left on the road surface
of a floodway and this can be a
hazard to road users. The department
should
inspect
each
affected
floodway as soon as possible after a
flood event and clear the surface if
required.
Energy Dissipators - Energy
dissipation is necessary due to high
flow velocities. Dissipation devices
usually consist of large obstructions
to the flow and result in a high
degree of turbulence. For these
reasons, energy dissipation structures
should be avoided in urban areas
where possible. Otherwise access
should be limited by appropriate
fencing. Reference should also be
made to Chapter 12 of QUDM
(NR&W 2008).
Energy dissipators are also very costly to
build and maintain and changes to the
design, such as flattening of channel

Department of Transport and Main Roads


Road Drainage Manual

gradient to reduce high velocities, is


preferred.

2.3.8

Staged Construction of
Roads

Whole of life considerations dictate that the


design of a road takes proper account of
both expected and potential changes that
will or may occur as traffic grows and the
surrounding land use develops or changes.
This aspect is most important for projects
that are planned to be built in stages over a
period of time. Making allowance in
current designs for these changes ensures
that future enhancements can be
accommodated in a cost effective, efficient
and safe manner.
Other future benefits include:
reduced disruption to road users and
adjacent property owners;
early / coincident resumption of
properties; and
reduced environmental impacts.
If designs ignore the requirements for future
upgrading, future projects will be more
difficult and very much more expensive to
implement than they would have been if
appropriate provisions were included in the
original design.
The fundamental elements to be addressed
in designs to allow for future upgrades
include:
carriageway or formation widening
(e.g. for an additional through,
auxiliary and/or overtaking lane, for
a noise barrier, for a safety barrier);
duplication of carriageways; and
intersection and interchange changes
or upgrades.

Chapter 2
General Design Requirements

Providing for the future cross-section and


ultimate road configuration when designing
drainage
systems
requires
careful
consideration of the various components of
the drainage system.
2.3.8.1 Cross Drainage
Aspects of cross drainage that require
special
consideration
for
staged
construction include:
hydraulic efficiency and capacity of
the culvert in its initial (short) and
ultimate (long / extended) forms;
possible change in culvert operation
(inlet control / outlet control) and
subsequent outlet velocity changes;
potential variation in afflux and/or
allowable headwater changes;
positioning of culvert inlets and
outlets (within the stream);
changes to the inlet / outlet of
adjacent culverts (in the same stream)
where these are located within the
median of a dual carriageway and
where future widening will be within
the median (e.g. culverts may
become connected);
environmental considerations (e.g.
scour prevention measures, fish or
animal passage);
resumptions (e.g. land required to
accommodate future culvert inlets
and
outlets,
allowance
for
maintenance access); and
cover over future culvert extensions
due to carriageway widening (on the
outside of the formation and/or in the
median).

March 2010

2-15

Department of Transport and Main Roads


Road Drainage Manual

2.3.8.2 Longitudinal Drainage

Aspects of longitudinal drainage that


require special consideration for staged
construction include:
drainage of the ultimate median
which must be provided for with:
height of pipes and inlets designed to
fit the initial and ultimate shapes of
the median and carriageway;
designed capacity and hydraulic
operation suitable for the initial and
ultimate configurations;
conversion of an open channel within
the median to an underground piped
system and the requirements for
outlets;
road safety impacts with drainage
inlets structures within the median.
drainage connections to bridges
(including any pollutant control
devices) may need to be designed for
the ultimate configuration (e.g. need
to cope with additional surface runoff
from a widened structure);
resumptions (e.g. land required to
accommodate catch drains, diversion
drains or channels, maintenance
access, sedimentation basins);
environmental considerations (e.g.
size and location of sedimentation
basins).
2.3.8.3 Surface Drainage
Aspects of surface drainage that require
special
consideration
for
staged
construction include:
aquaplaning (e.g. pavement widening
may create a problem where before
there was none, the application of

March 2010

2-16

Chapter 2
General Design Requirements

superelevation in the initial stage


may need to suit the ultimate stage);
use of crowned multi-lane one-way
carriageways to reduce aquaplaning
will impact on drainage design (e.g. a
third lane added to the median inside
of a two-lane carriageway may be
crowned and so drain towards the
median); and
addition of kerbing / kerb and
channelling in the future (e.g.
channelisation of an unchannelised
intersection, when widening a twolane carriageway to three lanes).
2.3.8.4 Sub-surface Drainage
Aspects of sub-surface drainage that require
special
consideration
for
staged
construction include:
location and capacity of sub-soil
drains;
location of outlets and cleanout
points to allow for ultimate shape;
and
changes to the water table and
groundwater flows.
2.3.8.5 Medians and Obstructions
In divided roads where the ultimate median
has a concrete safety barrier and the median
width is at or near the absolute minimum,
the ultimate median drainage system will
require the use of drop inlets to an
underground drainage which can be located
beneath the barrier itself.
The location of obstructions or immovable
features such as bridge piers and abutments
must be carefully considered to enable the
future stage development of the cross
section of the road to be implemented
without major change to these features.

Department of Transport and Main Roads


Road Drainage Manual

Preserving the required above ground


horizontal and vertical clearances to these
features is essential in this process as well
as providing underground clearances from
footings or abutments to the underground
stormwater drainage system.

2.4

Design Controls

Design controls are aspects of the road


environment or elements of the project that
cannot be changed, or are extremely
difficult or costly to change. These aspects
and elements therefore place restrictions
and controls on the design. Design controls
can either place a direct restriction on a
project or at least influence the
development of design options, becoming
design considerations.
One example of a design control with
respect to drainage may be the width of the
road reserve.
Where resumptions are
undesirable, the existing right-of-way could
limit the available space for drainage
infrastructure and therefore control what
can be done.
Another example may be the location of the
horizontal alignment / centreline. While the
design of the horizontal alignment should
consider drainage elements, there are many
reasons why the location of the horizontal
alignment may be fixed. This could then
directly restrict or influence the drainage
design.
While it is possible, vertical
alignment should rarely be a design control
over drainage design as the both elements
need to be developed holistically in order to
achieve an appropriate design solution.

Chapter 2
General Design Requirements

2.5

Design Criteria

2.5.1

Introduction

Drainage infrastructure for a road project is


planned and designed to provide a standard
or level of drainage immunity that conforms
to good engineering practice and that also
meets
government
and
community
expectations. It is conventionally specified
based on an Average Recurrence Interval
(ARI). This is defined as the average
interval in years between exceedances of a
specified event (i.e. rainfall or discharge)
and is written as ARI x years. The ARI is
really a probability rather than an actual
period between occurrences.
Within a project, the design criteria will
vary in accordance with road type and
whether the design relates to cross drainage,
surface drainage, urban drainage, or
construction phase drainage, including
erosion and sediment control. Guidance as
to suitable ARIs for each of these situations
is provided in the following sections.
Further to the use of ARI, the Annual
Exceedance Probability (AEP) is also used
in relation to flood flows. The AEP is the
probability of exceedance of a given
discharge within a period of one year. It
can be considered as the reciprocal of the
average recurrence interval in years
expressed as a percentage, for example, 50
years ARI is equivalent to 2% AEP or 1:50
AEP.
For further discussion on ARI and AEP,
reference should be made to Book 1,
Section 1.3 of Australian Rainfall and
Runoff, A Guide to Flood Estimation, Vol 1
(AR&R) (IEAust 2001).

March 2010

2-17

Department of Transport and Main Roads


Road Drainage Manual

2.5.2

Chapter 2
General Design Requirements

General Hydraulic Criteria

2.5.2.2 Flow Velocities


The flow velocity is a critical parameter
used in design of drainage structures. It is
the velocity of the flow of the water in the
flow path. The flow velocity can be
calculated for a particular location in a
stream cross section or it can be an average
over a portion or the whole of the cross
section.

Hydraulic criteria include the following:


(a) Design discharge

(b) Flow velocities


(c) Permissible velocities
(d) Flood and stream gradient
(e) Fauna passage requirements

Flow velocity can be calculated using


Mannings Equation, by a hydraulic model
or it can be measured during an actual flood
event.

(f) Fish passage requirements


(g) Erosion and sediment control
(h) Permissible afflux
(i) Tailwater
potential

levels

and

Backwater

(j) Pollution control


(k) Road
closure
AATOS/AATOC

periods

(l) Inundation of adjacent land


(m) Maintenance of flow patterns
Establishing the hydraulic criteria requires
an understanding of the hydrologic and
hydraulic conditions of the site or project.
2.5.2.1 Design Discharge
The design discharge is the flow rate of the
defined probability (or Average Recurrence
Interval) for the required drainage works.
Usually the design discharge is used to
provide the size of the drainage structure
and the level of the road. The design
discharge is expressed as a flow rate,
usually as cubic metres per second (m3/s).
Usually the discharge is calculated directly
by a hydrology procedure, such as the
Rational Method for the drainage structure
and this discharge is used directly.
In more complex situations, the design
discharge is calculated while accounting for
attenuation or diversions.

March 2010

2-18

Flow velocities are usually calculated


initially for the natural channel, without any
drainage works. This velocity indicates the
natural conditions which can be used as a
basis for the consideration of the drainage
works.
Flow velocities can then be
calculated for the conditions with the
addition of the proposed infrastructure.
Flow velocity in a flow path depends on the
slope and geometry of the flow path as well
as the channel roughness and the amount of
flow. It is often very variable across a cross
section and along a reach of a stream.
2.5.2.3 Permissible Velocities
When designing a drainage structure or
channel, the flow velocity is an important
input to the design process. This is because
excessive flow velocities will cause scour.
The risk of scour depends on the gradient
(slope) and geometry of the channel, the
soil conditions and the vegetation cover.
When the velocity of the flow increases
beyond a limit, the risk of scour will
increase. In the design, the permissible
flow velocities need to be defined to help in
the design process.

Department of Transport and Main Roads


Road Drainage Manual

The process used is as follows:


The drainage structure (culvert,
bridge, floodway or channel) is
designed, based on the best available
information.
The design flow velocity for the
preliminary design is calculated.
The maximum permissible flow
velocity is compared to the calculated
design velocity.
The design may be modified to meet
this limit, by increasing the waterway
area or reducing the slope for
example.
If this is impossible because of
constraints, appropriate mitigation
measures will be needed.
The permissible velocities depend on the
material of the channel bed as well as the
type of soil, channel gradient and shape as
well as vegetation cover. Permissible flow
velocities are listed in tables that can be
found in Chapter 8.
While the permissible flow velocities are
set mainly to counter the risk of scour, the
permissible flow velocity may also depend
on other environmental factors, such as the
allowance for fish passage.
2.5.2.4 Flood and Stream Gradient
Flood
and
stream
gradients
are
considerations in drainage designs, since
these affect stream discharges (hydrology)
and flow velocities and flood levels
(hydraulics).
As discussed in Chapter 8, there are three
different gradients or slopes that are
relevant in road drainage design:
Energy gradient; the profile of the
energy line in a flood. While this

Chapter 2
General Design Requirements

slope is not easily measured, it is the


gradient used in the hydraulic
calculations. It is usually estimated
for use in calculations.
Water surface slope; the profile of
the surface of the water. This is the
slope measured by observing a series
of flood levels along the waterway.
In open channels, the water surface
slope is also the Hydraulic Grade
Line (HGL).
Bed slope; the profile or slope of the
bed of the channel. This slope can be
measured from survey data or
topographic maps.
While not
directly used in the hydraulic
analysis, for reasonably uniform
channels, the bed slope can be used
to approximate the water surface
slope / energy gradient.
Another term used is ground or catchment
slope and the value is used in some
hydrology procedures. The value is a
representative slope for the whole
catchment.
Higher gradients lead to greater flow
velocities, which result in lower flood
levels, but increased risk of scour.
2.5.2.5 Fauna Passage
Requirements
When a road is built it tends to fragment
habitat and lead to greater risk to fauna that
cross the road.
Since the drainage
structures cross under the road, these can
potentially be used to provide a safe means
for fauna to cross between habitats. To
provide this, the drainage structures may
need to be modified to make this passage
easier. Both terrestrial and aquatic fauna
(especially fish) need to be considered.

March 2010

2-19

Department of Transport and Main Roads


Road Drainage Manual

Fish passage is an especially important case


and this is described in Section 2.5.2.6.

When considering the requirements for


fauna transfer, several issues must be
considered as follows:
Consult with environmental experts
to confirm that the best information
possible is being incorporated.
Identify the relevant environmental
issues.
Identify fauna types that may use the
transfer.
Determine
requirements.

the

appropriate

Usually the culvert can operate effectively


as a drainage structure while also providing
a means for fauna to cross the road.
Particular considerations for the fauna
transfer are as follows:
Normally it is important to supply a
dry passageway so that the fauna can
move through the culvert without
getting wet.
This can be provided by dividing the
culvert into wet and dry cells, with
the inverts of some cells kept higher
than others. When larger floods
occur, the whole culvert set will
operate.
A low flow channel can provide the
same benefit.
Fencing may be needed to direct the
fauna towards the culvert, but
hydraulic
issues
should
be
considered.
The culvert should not provide
habitat for the fauna, since this
habitat will be removed when a flood
occurs.

March 2010

2-20

Chapter 2
General Design Requirements

Lighting may be needed in long


culverts so that fauna can enter the
culvert is not discouraged by the
dark.
Vegetation is required at the entrance
and exit of the culvert to provide
cover in otherwise exposed areas.
2.5.2.6 Fish Passage Requirements
A waterway barrier is any structure that
limits fish movement along the waterway.
Examples of waterway barriers include:
dams;
weirs;
bridges;
culverts;
tidal barriers;
fords;
causeways / floodways;
silt curtains; and
any other barriers that restrict fish
movement.
Constructing
or
reconstructing
(or
significantly repairing) any waterway
barrier across a (freshwater or tidal)
waterway requires a development approval
under the Integrated Planning Act 1997 and
the Fisheries Act 1994.
Under this
legislation the application will be refused
unless movement of fish across the
waterway barrier is adequately provided.
An exemption can be given if no fish or
habitat exists above the barrier.
The definition of a waterway under the
Fisheries Act 1994 includes a river, creek,
stream, watercourse or inlet of the sea. This
definition does not distinguish between
freshwater and tidal waters and incorporates

Department of Transport and Main Roads


Road Drainage Manual

both permanent and ephemeral flowing


waterways.
The definition includes:
any artificially modified or improved
waterway; and
channels that connect water bodies to
waterways during times of flow,
along which fish would be expected
to move.
It does not include isolated water bodies
such as lakes or some wetlands (e.g. springfed wetlands).
There is no indication in the legislation that
there should be an upstream limit to the
application of waterway and hence the
requirement for a development approval for
waterway barrier works. However, an
upstream limit is relevant, as there is
usually little to be gained in terms of fish
movement from consideration of waterway
barrier works on natural drains and gullies.
To determine whether a site is above or
below the upstream limit, the following
guidelines are provided:
1. Defined bed and banks
The bed and banks of the waterway
need to be continuous rather than
isolated and broken sections of a
depression.
2. An extended, if non-permanent, period
of flow
Flow must continue for a reasonable
period after rain ceases and have
some reliability commensurate with
rainfall.
3. Flow adequacy
The flow needs to be sufficient to
sustain basic ecological processes
and to maintain biodiversity within
the feature.

Chapter 2
General Design Requirements

If the crossing being assessed is determined


to be above the upstream limit, design can
progress normally, however, if a crossing is
below the upstream limit (that is, all 3
conditions met), a structure catering for fish
passage has to be designed and approved as
part of the development approval process.
Further guidance and information can be
obtained
from
the
Fish
Habitat
Management Operational Policy (FHMOP
008)

Waterway
barrier
works
development approvals (2009) issued by
Queensland Fisheries (part of Department
of Employment, Economic Development
and Innovation). Document is available <
http://www.dpi.qld.gov.au/documents/Fishe
ries_Habitats/FHMOP008-Fish-HabManage.pdf >, correct as of January 2010.
Development applications, as required
under the Act, must be submitted to
Queensland Fisheries for approval.
The department is currently working with
Queensland
Fisheries
to
develop
requirements and self-assessable codes for
the design, construction and operation of
drainage structures.
Updates regarding
requirements will be issued either as a
Planners and Designers Instruction or
directly as an update to this manual.
2.5.2.7 Erosion and Sediment
Control
One of the most important environmental
concerns for road drainage is erosion and
sediment control.
This should be
considered in all situations, and appropriate
assessment and mitigation measures must
be supplied. Scour at drainage structures
can be a serious environmental problem as
well as providing a risk of structure failure
and possible road embankment failure.

March 2010

2-21

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Control of scour at culverts and channels


needs to consider the permissible flow
velocities noted in Section 2.5.2.3 and
Chapter 8, which indicates the velocity
limits where scour begins to become a
problem. While these are good guidelines,
each individual situation needs to be
considered on its own merits, since there
may be a large variation for different
situations.

The point of maximum afflux occurs


immediately upstream of the road and then
dissipates while moving further upstream.
There is a point where the afflux drops to
zero and the influence of the bridge on
flood levels disappears. In flat terrain, this
point may be a considerable distance
upstream, but in steep country with high
flow velocities, the afflux may extend only
a very short distance.

Where necessary, erosion control measures


will be needed and these are described in
later sections of this manual.

The afflux also reaches a maximum at the


point of overtopping of the road. Smaller
floods will be conveyed easily through the
structure, while larger floods may
eventually drown out the structure. For
very large floods, there may be no impact
on flood levels, where the structure is
submerged to a significant depth.

2.5.2.8 Permissible Afflux


Afflux is the increase in peak water levels
produced by the introduction of a culvert or
bridge and is the comparison between the
water levels for the existing conditions and
the proposed conditions once the road has
been built.
Afflux is defined for a
particular location and will vary across the
floodplain or along the length of a channel.
The allowable afflux is often a controlling
factor in design of drainage structures and
can be a serious community concern.
While the department must assess the afflux
expected during the planning and design
process, local authorities will often specify
the requirements that they expect in a
region.
Afflux is usually caused by a constriction in
a flow path, from the construction of a
culvert, bridge or floodway. However in
some cases, especially in flat terrain and
where flow may be diverted from one
catchment to another, it could be caused by
a redistribution of flow. Afflux can also be
negative, that is a reduction in flood level,
downstream of a constriction or where flow
is diverted away from a stream or creek.

March 2010

2-22

Afflux needs to be considered in all


drainage designs. During the planning
phase, any properties, infrastructure or
other feature upstream of the crossing must
be reviewed. These structures then need to
be considered in the design and the impact
on flood levels at each of these must be
included in the design process. If there is
nothing that could be adversely impacted by
an increase in flood levels, afflux does not
necessarily form a part of the design. In
this case, the maximum permissible flow
velocity through the structure is the critical
factor.
The allowable afflux will vary for
individual locations. In some particularly
sensitive areas, no afflux may be the
appropriate limit. This would be in areas
where there are already flood prone
properties and even a small increase in level
could cause a significant increase in
damage. In some locations, a small amount
of afflux may be acceptable. In regions
where upstream development does not
provide a control, the flow velocity and/or

Department of Transport and Main Roads


Road Drainage Manual

allowable headwater requirement generally


set the limit. In this instance, the afflux is
often of the order of 250 mm, though higher
afflux may be possible in some situations.
Afflux is reduced usually by increasing the
waterway area of the drainage structure, but
it can also be reduced by channel works or
other mitigation measures.
Reducing the afflux may lead to higher
costs for drainage infrastructure and it may
be impossible to reduce the afflux at some
sensitive locations, even with extensive
mitigation measures. In these cases, careful
assessment of the hydraulics and potential
damage is needed and this should be
followed by consultation with affected
property owners to develop an acceptable
result.
2.5.2.9 Tailwater Levels and
Backwater Potential
Tailwater is important for drainage design,
as it sets the water level at the outlet of a
drainage structure. It therefore can control
the hydraulic performance of the structure.
Tailwater levels must be calculated as part
of the hydraulic design for all drainage
structures. There are a number of situations
required for the calculation of tailwater, as
follows:
Normal stream depth. In this case
the tailwater level is defined by the
normal water level in the downstream
channel, and this depends on the
conditions of the stream or creek.
These conditions are the slope,
channel geometry and stream
roughness. The tailwater level is
calculated
using
Mannings
Equation, backwater analysis or a
stream rating curve.

Chapter 2
General Design Requirements

If there is a downstream confluence


(junction) with another stream, the
tailwater level may be held at a
higher level than would naturally be
the case. In this case, the flow is at a
lower velocity and the water levels
are higher, which means that the
culvert will not operate as efficiently
as it would if the downstream water
level was lower. This is especially
the situation if the road crosses a
tributary just before this tributary
joins a major stream. Two cases
need to be analysed.
Firstly,
assuming a major flood in the
downstream catchment of the major
stream. This may result in a higher
flood level in the tributary, which
may be critical for the design.
Secondly, assuming normal to low
flows in major stream, a local
catchment flood in tributary may
result in lower flood levels but a
critical case for the consideration of
velocities through the structure.
Similarly to the tributary situation, a
downstream lake or dam can affect
the tailwater level. In this case, the
stream flows into a lake, natural or
artificial, and this body of water
holds up the flood levels and thereby
increases the tailwater level. This
increase can occur over time, giving
a dynamic tailwater.
Also, another infrastructure crossing
or artificial constriction downstream
of proposed crossing can affect
tailwater levels.
If the road crossing is close to the
ocean or an estuary, the tailwater
level may be controlled by the level
of the ocean. In this case, the water
level will depend on tidal levels as

March 2010

2-23

Department of Transport and Main Roads


Road Drainage Manual

well as possible effects from storm


tides or waves. The assessment of an
appropriate tidal tailwater level for
design of drainage structures is a
difficult problem. A major issue is
the risk of occurrence of a particular
tide at the same time as a major
flood. Analysis of a range of tidal
levels may be of value as for the
consideration of a downstream
tributary. If a high tide is analysed,
this may give the critical event for
flood levels on the road, but the flow
velocities will be low. On the other
hand, analysis of a lower tide will
give lower flood levels, but the flow
velocities may be critical for the
design.

2.5.2.10 Pollution Control


While roads may make up a relatively small
proportion of the catchment area, they can
contribute a relatively high proportion of
contaminants which are washed into
streams, creeks and other receiving waters.
The contaminants include a range of
materials, especially sediment, metals, oils
and greases, rubber and gross pollutants.
The export of these contaminants may need
to be mitigated by measures provided as
part of the drainage system for the roads.
The Environmental Protection Act 1994
identifies the objective to protect
Queenslands waters while allowing for
development
that
is
ecologically
sustainable.
This purpose is achieved
within a framework that includes:
identifying environmental values for
Queensland
waters
(aquatic
ecosystems, water for drinking, water
supply, water for agriculture,
industry and recreational use);

Chapter 2
General Design Requirements

deciding and stating water quality


guidelines
and
water
quality
objectives to enhance or protect the
environmental values.
The Department of Environment and
Resource
Management,
formerly
Queensland Environmental Protection
Agency (EPA), published the Queensland
Water Quality Guidelines 2006 (QWQG),
based on the Australian Water Quality
Guidelines for Freshwater and Marine
Waters 2000 developed and published by
the Australian and New Zealand
Environment and Conservation Council
(ANZECC).
The QWQG provides
technical guidelines for the protection of
aquatic ecosystems.
These guidelines
initially focus on the protection of aquatic
ecosystems across three geographic regions
for which regional data is available;
South-east;
Central; and
Wet tropics.
For other regions of Queensland, the
national guidelines are adopted until local
guidelines are developed.
Most environmental protection agencies
tend to rely on the ANZECC guidelines.
These guidelines generally are used as the
default framework for setting water policy
objectives for managing water resources on
a sustainable basis.
The QWQG identifies three levels of
ecosystem condition for which different
levels of protection can be applied:
Level 1 - high conservation /
ecological value systems;
Level 2 - slightly to moderately
disturbed systems; and
Level 3 - highly disturbed systems.

March 2010

2-24

Department of Transport and Main Roads


Road Drainage Manual

Both the ANZECC guidelines and QWQG


are primarily focussed upon deriving
guideline values for Level 2 aquatic
ecosystems, as these represent a significant
proportion of Australian waters.
The sensitivity of the receiving aquatic
ecosystems and the potential impact of the
road corridor, both in construction and
operation, should have been identified in
the projects environmental assessment
documentation
and/or
Environmental
Impact Statement (EIS). Traffic volumes
and heavy vehicle content will be
documented in the planning report.
This information, in conjunction with the
water quality requirements noted within the
environmental assessment documentation
and/or EIS, will assist in identifying and
selecting appropriate key water quality
objectives which may include:
capture of pollutants upstream of a
sensitive water body; and
discharges
achieve
(ANZECC
sensitivity
ecology.

from sediment basin to


water
quality
levels
guidelines) owing to the
of downstream aquatic

In 2008, the EPA (now the Department of


Environment and Resource Management)
released the EPA Best Practice Urban
Stormwater Management - Erosion and
Sediment Control Guideline in support of
the State Coastal Management Plan (2002)
for planning schemes and development
assessment.
Although this guideline has been
specifically prepared for urban land
development within the coastal zone, it has
relevance to this manual for the
construction and operation of drainage
infrastructure located in this zone. Best

Chapter 2
General Design Requirements

practice environmental management of an


activity is defined in the guideline as:
...the management of the activity to
achieve an ongoing minimisation of the
activitys environmental harm through costeffective measures assessed against the
measures currently used nationally and
internationally for the activity.
In urban catchments, reference should also
be made to Australian Runoff Quality A
Guide to Water Sensitive Urban Design
(EA 2006) and to QUDM (NR&W 2008).
These reference documents together with
EPA Best Practice Urban Stormwater
Management guidelines provide the
standards for erosion and sediment control.
Pollution control guidelines for a particular
drainage infrastructure project are specific
to that project and are detailed in the
environmental assessment documentation
and/or EIS for that project. No specific
guidance is therefore provided in this
manual.
Guidelines for the selection of specific
pollution control options for a project are
set out in Chapter 7.
Erosion and sediment control guidelines for
a particular drainage infrastructure project
are specific to that project may be detailed
in
the
environmental
assessment
documentation or EIS for that project.
Guidelines for the selection of design
standard (average recurrence intervals) for
erosion and sediment controls are contained
within Chapters 7 and 13.
2.5.2.11 Road Closure Periods /
AATOS/AATOC
Consideration of times of closure is
important in some situations to supplement
the flood immunity assessments. The time
of closure is a measure of the disruption to

March 2010

2-25

Department of Transport and Main Roads


Road Drainage Manual

traffic and in some ways is a better measure


of the performance of the road. This
measure can be expressed at either the
average annual time of submergence or
closure, the average time each year when
the road is affected or as the duration of
submergence or closure.
More details on this topic are provided in
Chapter 10 where the methods of
calculation are included.
2.5.2.12 Inundation of Adjacent Land
Roads can provide a restriction to flow
across a flow path or floodplain and can
cause ponding upstream. This inundation
must be considered carefully (extent and
duration of) in the planning and design of
the road and any adverse impacts identified
and mitigated. These impacts are important
in urban areas, where there may be
development or infrastructure that may be
affected. However there may also be
concerns in rural areas, where there may be
impacts on agricultural land.
Generally the drainage systems for roads
are sufficiently large enough that the
duration of ponding is not increased greatly,
but this may be possible in some situations.
These cases need particular attention.
2.5.2.13 Maintenance of Flow
Patterns
The road is a linear structure across the
floodplain and therefore may divert flow
across the floodplain, especially in flat
areas. This diversion may have impacts on
both economic and environmental factors.
Any diversions should be identified and
generally there should be minimised to
maintain the existing flow patterns as well
as possible.
In some situations diversions may be worth
considering especially where there are
March 2010

2-26

Chapter 2
General Design Requirements

benefits to the cost and complexity of the


drainage system, but the potential impacts
must be carefully assessed to determine if
they are acceptable.

2.5.3

Cross Drainage Criteria

The design criteria for cross drainage for a


particular project may be set either by the
client or by departmental strategies and may
be based on any of the following
conditions:
Flood immunity - This is defined as
the average recurrence interval (ARI)
of a flood that just reaches the height
of the upstream shoulder. In other
words, the road surface remains dry
/ is immune to flood of set ARI.
Furthermore, freeboard may be
required to lower the water level
further to keep the pavement dry
and/or provide a buffer in case of
error
in
calculation.
Another definition used is the (ARI)
of a flood that just reaches the point
of overtopping the highest point of
the road. This definition is used to
calculate the flood immunities in the
Bridge Information System (BIS).
Trafficability - In some instances, it
is desirable to allow traffic to
continue to use the road while
floodwater crosses the road surface.
The design criteria therefore may be
specified in terms of the ARI of the
flood at the limit of trafficability.
This limit is based on a combination
of depth and velocity of flow over the
road or floodway and is defined as
occurring when the total head (static
plus velocity) at any point across the
carriageway is equal to 300 mm. The
road is defined as closed if the flow

Department of Transport and Main Roads


Road Drainage Manual

is greater than this limit, as used


below. This standard is in line with
recommendations
made
by
Austroads.
Time of Submergence (TOS) - This
is a measure of the expected time that
the road is submerged in any flood
but especially in a major flood such
as the ARI 50 year event.
Submergence is defined as the point
where the road is just overtopped,
even by very shallow water.
Average
Annual
Time
of
Submergence (AATOS) - This is a
measure of the expected average time
per year of submergence of the road
caused by flooding. It is expressed
as time per year.
Time of Closure (TOC) - This is a
measure of the expected time of
closure of a road (road not
trafficable) in any flood but
especially a major flood such as a
ARI 50 year event.
Average Annual Time of Closure
(AATOC) This is a measure of the
expected time of closure of the road
due to flooding, expressed as time
per year.
The times of submergence and closure
provide useful data to supplement the flood
immunity results. They give an indication
of the extent of disruption to transport that
may result from flooding on the road. In
some cases, low flood immunity may be
acceptable if the times of closure are low
and the expected disruption is relatively
minor.
The average annual times of submergence
and closure depend on the frequency of
submergence / closure as well as the
duration of each occurrence. For example,

Chapter 2
General Design Requirements

two streams may have a similar average


annual time of submergence, but a quite
different flood immunity, if one is closed
frequently for short durations, while the
other is closed more rarely for longer times.
The impacts of these different patterns can
be analysed to determine the most
appropriate design for each particular
crossing.
The time of submergence / closure is
related to catchment area and response
times as well as the flood immunity. These
times are calculated either from design
flood events or from stream flow data, as
described later in this manual.
Refer Sections 2.5.6, 2.5.7 and 2.5.8 for
specific detail regarding flood immunity
criteria for cross drainage for various road
types.

2.5.4

Longitudinal Drainage
Criteria

The requirements for longitudinal drainage


will vary from project to project. The
design considerations for the site will
dictate the choice between alternative
longitudinal drainage options such as kerb
and channel, grassed swales, and lined or
unlined table drains. It is also important
that the longitudinal drainage (drain type
and capacity) of the adjoining projects be
considered when determining the criteria
for the site being planned or designed to
ensure consistency of drainage capability
and to mitigate potential system failure.
In urban environments, kerb and channel
has historically been favoured for most
roads, though grassed channels and swales
are also common on divided roads.
Reference should be made the Road
Planning & Design Manual to determine
the cross sectional components of table

March 2010

2-27

Department of Transport and Main Roads


Road Drainage Manual

drains and other drains associated with the


formation / carriageway.

The following criteria are to be considered


in determining the standard for longitudinal
drainage. It is important to note that the
standard for longitudinal drainage should be
compatible with the standard adopted for
cross drainage as these two components of
the drainage system typically work in
combination.
Refer Sections 2.5.7 and 2.5.8 for specific
detail regarding flood immunity criteria for
longitudinal drainage for various road
types.
2.5.4.1 Shape of Table Drains

Chapter 2
General Design Requirements

vertical curve apex, the table drain invert is


gradually shifted away from and then back
closer to the shoulder edge, in order to
deepen the drain and effect sufficient grade.
However, this solution may not always
work, therefore modification / adjustment
of the road geometry may need to be made.
2.5.4.3 Flow Velocities
Flow velocities in longitudinal drainage
should be limited to prevent erosion.
Limiting flow velocities is preferred over
maintaining high flow velocities and
providing armouring. Acceptable velocities
should be based on the soil conditions and
characteristics of the site.

Flat-bottomed drains are the preferred type


or shape. Parabolic can also be used
although these are difficult to construct /
maintain. The use of V drains is to be
limited / confined to constrained sections
where cross sectional width is critical. The
flat-bottom of the drain is to be sloped away
from the carriageway and be wide enough
to allow access for maintenance machinery.

2.5.4.4 Flow Depths

2.5.4.2 Minimum Grades

Median longitudinal drainage will usually


have a concrete lined invert to assist
maintenance and reduce the risk of errant
vehicles rolling after hitting ruts caused by
tractor mowing.

The minimum grade for unlined drains,


including table drains, is 0.5% and 0.2% for
lined drains however 0.3% may be regarded
as the minimum practical slope for
construction (allowing for construction
tolerances). This is to ensure that the drain
will flow and, if applicable, minimise
ponding against formations and pavements.
This criterion also applies to both crest and
sag vertical curves where grades fall below
0.5%. Generally, to achieve the required
minimum grades, widening of the table
drains is needed over the critical length
(length where grade is less than that
required). Widening of the table drain
means that when travelling away from the

March 2010

2-28

Flow depths should be limited to prevent


erosion and inundation of the pavement.
An increase in the number of outflow points
(e.g. turnouts or level spreaders) from the
longitudinal drainage should be considered
to assist in managing depth of flow.
2.5.4.5 Median Drainage

2.5.4.6 Bridge Runoff


Road runoff from bridge scuppers should
be discharged into a sediment basin, gross
pollutant trap or other relevant first flush
containment removal device.
This is
particularly important where the scupper
would direct bridge run-off into a base flow
channel or upstream of a sensitive
environment (e.g. wetland, fish habitat
reserve).

Department of Transport and Main Roads


Road Drainage Manual

Reference should also be made to Standard


Drawing No: 1178 (DMR 2009b).

2.5.5

Road Surface Drainage


Criteria

The requirements for surface drainage


primarily relate to safety (e.g. aquaplaning
and ponding) and are dealt with in Chapter
11.
For surface drainage, the main criterion is
the allowable flow width on the road.
However flow velocity also needs to be
addressed particularly when pedestrian
movement is adjacent to or crosses the
flow.
Refer Sections 2.5.7 and 2.5.8 for specific
detail regarding flood immunity criteria for
road surface drainage for various road
types.

2.5.6

Immunity Criteria for the


National Land Transport
Network

Generally, the adopted flood immunity


criteria for cross drainage on the National
Land Transport Network (Auslink) is ARI
100 years and for other drainage
components, refer Tables 2.5.7 and 2.5.8 as
applicable.
For major projects on the Auslink network,
the Australian Governments Department of
Infrastructure,
Transport,
Regional
Development and Local Government will
set project specific drainage design criteria
which may differ from the ARI values
specified in this manual in and may also
include time of closure requirements. The
criteria will be agreed to and set in relevant
project documentation and approvals.

Chapter 2
General Design Requirements

2.5.7

Immunity Criteria for State


Controlled Roads - Rural
Catchments

For rural catchments, the generally accepted


design criteria for various drainage
components are specified in Table 2.5.7.
In some situations, it might not be possible
to design for this level of flood immunity
without causing intolerable impacts on
existing development or because of
extensive flooding that could not be
managed without unacceptable cost. In
such situations the ARI may be relaxed to a
lower level. In this instance, assessment
and use of time of closure / submergence
(TOC/TOS) for design criteria may be more
appropriate. Refer discussion in Chapter 1
regarding road infrastructure delivery.
This criterion also applies to rehabilitation
and reconstruction projects where existing
structures are assessed as hydraulically or
structurally deficient and need to be
completely replaced.
Designers should check departmental
strategies for flood immunity or
trafficability requirements for specific
routes and individual projects (refer
Chapter 3).

2.5.8

Immunity Criteria for State


Controlled Roads - Urban
Catchments

The design ARI for a project in urban areas


will often be influenced by the capacity or
capability of the existing drainage system or
network that the new work needs to connect
into.

March 2010

2-29

Department of Transport and Main Roads


Road Drainage Manual

Table 2.5.7 Design ARI for State


Controlled Rural Roads

Location

ARI

Cross drainage excl.


floodways

50 years

Diversion channels

50 years

Road surface drainageA

10 years

Bridge deck drainage

10 years

Road surface drainage


of pavements

1 year

Water quality treatment


devices

1 year

. Road surface drainage includes kerb and


channel, table drains, diversion drains,
batter drains and catch drains.

For urban catchments, the generally


accepted design criteria for various
drainage components are specified in Table
2.5.8. The Department of Transport and
Main Roads has also adopted the ARI
criteria as described in Table 7.02.1 of
Queensland Urban Drainage Manual
(QUDM) (NR&W 2008). Key values are
included in Table 2.5.8. Designers should
confirm the requirements of any existing /
connecting systems with the relevant
authority.
Urban drainage systems are generally based
on the major / minor drainage system or
dual drainage system. This type of system
or drainage concept has two distinct
components:
The minor drainage system is
designed to fully contain and convey
a design minor stormwater flow of
specified ARI with road flow limited
in accordance with the requirements
set out in Chapter 11 on this manual.
Refer to the glossary for full
definition.
March 2010

2-30

Chapter 2
General Design Requirements

The major drainage system


conveys the floodwater beyond the
capacity of the minor drainage
system and up to a specified ARI.
Refer to the glossary for full
definition.
The minor and major design storms
correspond to the rainfall events for the
ARI chosen for the design of the minor and
major systems respectively.
Designers should note that the design
discharge for the major system ARI may
require that the capacity of the gully inlets
and underground pipes be increased beyond
that required by the design discharge for the
minor system ARI, in order to meet the
major system design criteria.
Another important design consideration is
that with any proposed drainage system
adjacent to sensitive areas where flood
inundation will not be tolerated, the design
of the major drainage system should also
consider the flow conveyed in the
underground minor drainage system should
this system fail due to malfunction or
blockage.

2.5.9

Environmental Criteria

The environmental considerations and


strategies for managing aspects of a project
(refer in Section 2.3.4) that are predicted to
cause environmental harm will most likely
become environmental criteria for the
project.
Chapter 7 deals further with the
development of environmental criteria.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 2
General Design Requirements

Table 2.5.8 Design ARI for State


Controlled Urban Roads

(c) Protect water quality of surface and


ground waters.

Location

ARI

Major System
includes all above and
below ground
components

50 or 100
yearsA

(d) Integrate water into the landscape to


enhance visual, social, cultural and
ecological values.

Minor System Components


Cross drainage excl.
floodways

50 years

Diversion channels

50 years

Road surface drainage


including intersectionsB

10 years

Bridge deck drainage

10 years

Sediment basins

2 years

Road surface drainage of


pavements

1 year

Water quality treatment


devices

1 year

Conventional water management has been


compartmentalised with water supply,
wastewater and stormwater traditionally
treated as separate entities.
However
integrated water management needs to
consider the total water cycle and this
concept is increasingly being accepted
and/or adopted.
While the principal concern for the
department is related to stormwater
drainage, the department also has an
interest in a range of other issues such as
the use of water for construction and water
quality controls.

. Refer to relevant local authority for


confirmation of required Design Storm ARI
particularly where connecting / discharging
to an existing system under their control.

. Road surface drainage includes kerb and


channel, underground pits and pipe
networks, table drains, diversion drains,
batter drains and catch drains.

ARI for design of retention and detention


basins is project specific and must be
specified in design brief.

2.6

Water Sensitive Urban


Design

Water Sensitive Urban Design (WSUD) is a


particular issue for urban planning and
design, but the key principles of WSUD are
also applicable to road infrastructure in the
rural environment. These principles are:
(a) Protect existing natural features and
ecological processes.
(b) Maintain the natural
behaviour of catchments.

hydrologic

Roads may represent a relatively small


proportion of the total catchment, but they
sometimes contribute significantly to water
quality concerns. This is especially the case
on roads with high traffic volumes, where a
number of different contaminants may be
produced.
Between rainfall events,
contaminates can build up and then runoff
at a greater rate than normal into receiving
waters.
The principles that the department need to
consider include:
(a) Consider all parts of the water cycle,
natural and constructed, surface and
subsurface, recognising them as an
integrated system.
(b) Consider all requirements for water,
both anthropogenic (human activity)
and ecological.
(c) Consider the local context, accounting
for environmental, social, cultural and
economic perspectives.

March 2010

2-31

Department of Transport and Main Roads


Road Drainage Manual

(d) Include all stakeholders in the process.

(e) Strive for sustainability, balancing


environmental, social and economic
needs in the short, medium and long
term.
The department also needs to be aware of
all water related issues, not only in the road
reserve, but also both upstream and
downstream.

2.7

Extreme Rainfall
Events

While the planning and design of road


drainage systems is based on a determined
average recurrence interval or set of
average recurrence intervals, it is also a
requirement to review designs for possible
adverse outcomes that may occur during an
extreme rainfall event.
To illustrate this, most roads are designed
to an ARI 50 year standard. However
should an ARI 100 year event or larger
occur, culvert velocities may become
unacceptably high causing significant
environmental harm, afflux may increase
above the acceptable ARI 50 year limit
causing excessive flooding, the road may
overtop threatening the integrity of the road
embankment, safety of road users, and so
on.
The extent of the extreme events to be
analysed
depends
on
particular
circumstances, so the requirements cannot
be defined exactly. Furthermore, while the
risk of occurrence of these extreme events
is low, the impacts of an extreme event
must be assessed.
In the case of the likelihood of the event
occurring and the adverse outcomes / risks
being unacceptable, the design criteria may
need to be altered and the design
recalculated or appropriate mitigating
March 2010

2-32

Chapter 2
General Design Requirements

measures developed and included into the


project.
It is important to note that any outcomes
(adverse or otherwise) resulting from an
extreme rainfall event could occur within
both the road and external environments
(refer Section 2.2) therefore identification
of possible outcomes should not be limited
to the road reserve and/or chainage limits of
the project.
The following sections outline some
situations where the design of a project
should be assessed for adverse outcomes
and risks that may occur during an extreme
rainfall event. However, other situations
may also exist where assessment should be
undertaken, therefore careful engineering
consideration and judgement should be
exercised. Assistance in identifying or
confirming situations requiring assessment
and at what level (ARI) assessment should
be undertaken at can be provided by
Director (Hydraulics), Hydraulics Section,
Engineering & Technology Division.

2.7.1

Erodible Soil
Environments

Part of the road drainage design process is


the determination of acceptable or
maximum
allowable
velocities
for
stormwater flows. It should be noted that
these velocities are largely based on
research that identified the velocity when
erosion / scour started to occur in different
soil / stream types.
The maximum
allowable velocities for a project are then
used in the design of various drainage
structures / devices (for example, culverts
and channels) to ensure design discharge
through those devices is below the set
maximum allowable velocity for that
location. Some design solutions that may
be adopted, equal or are just below the set

Department of Transport and Main Roads


Road Drainage Manual

maximum allowable velocity.


In an
extreme rainfall event occurs, the maximum
allowable velocity for a given structure /
device will most likely be exceeded which
in turn could result in excessive scour,
erosion or environmental harm.
It is
therefore important that these situations are
identified and assessed.
If this situation is considered applicable on
a project, specialist advice needs to be
sought from the departments Hydraulics
Section or suitably prequalified consultant
as analysis methods are beyond the scope of
this manual.

2.7.2

Excessive Flooding

Larger floods may need to be considered in


locations where the impacts of the road on
flood levels (based on a normal design
ARI) are / will be significant / very severe.
These impacts will most likely be worse in
a large flood / extreme rainfall event. This
issue is particularly important where the
road embankment is relatively high and the
flood immunity provided by the high
embankment is much greater than the
usually adopted standard of ARI 50 years.
In this case, while larger floods may not
overtop the road, a higher peak water level
will build up on the upstream side of the
road causing excessive flooding and in
some cases may cause the overtopping of
the catchment boundary, directing or
diverting flow to an area not able to handle
the increased flow. Furthermore, the higher
peak water level may produce larger flow
velocities through the drainage structure,
which has been designed for a smaller ARI.
The higher velocity may cause scour
problems or could cause the catastrophic
failure of the structure itself.
The above issues may be further aggravated
by blockage of the drainage structure(s) (by

Chapter 2
General Design Requirements

silt and/or debris) which may lead to a


greater risk to the drainage infrastructure
and surrounding area, if the flow cannot
overtop the road.
Therefore, where flood impacts would be
significant / very severe, it is necessary (and
can be specified in design / contract
documentation) to consider floods up to the
Probable Maximum Flood (PMF). The
PMF is defined as the largest flood event
that can reasonably be expected. In some
situations, extreme events, though smaller
than the PMF, may be more appropriate.
If the situation of excessive flooding is
considered applicable on a project,
specialist advice needs to be sought from
the departments Hydraulics Section or
suitably prequalified consultant.

2.8

Self Cleaning
Sections

Self cleaning sections, for example,


culverts and channels, require a reasonably
regular flow of a specific velocity / energy,
that will pickup and transport any silt or
debris within the section to a specific
location beyond the section.
The required minimum velocity / energy for
a self cleaning flow through the section
must be determined based on the
anticipated sediment and/or debris (type /
size / weight) that may accumulate in the
section. This flow must be generated by a
design storm with a suitable Average
Recurrence Interval (ARI) such as ARI 1, 2
or 5 years depending on how often the
channel should be cleaned. Intervals of 1
or 2 years are preferred while intervals
greater than 5 years are not recommended.
The requirement for self cleaning sections
and the selected design interval (ARI) must

March 2010

2-33

Department of Transport and Main Roads


Road Drainage Manual

be specified in the design brief / contract


documents.

The location that any silt or debris can be


transported to (and deposited) must also be
considered as it must:
be accessible to allow maintenance /
clean out;
must not cause any adverse effects to
the environment (for example, water
quality and fish passage); and
must not adversely affect any future
flows (for example, cause ponding /
increase tailwater levels).
The inclusion / presence of a self cleaning
section does not remove or lessen the
requirement for regular / routine
maintenance inspections. Self cleaning
sections may reduce the requirement for
maintenance (cleaning) of the section.

March 2010

2-34

Chapter 2
General Design Requirements

Department of Transport and Main Roads


Road Drainage Manual

Chapter 3
Strategic Planning & Development Control

Chapter 3
Strategic Planning &
Development Control

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 3
Strategic Planning & Development Control

Chapter 3 Amendments Mar 2010


Revision Register

Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 3
Strategic Planning & Development Control

Table of Contents
3.1.

Introduction

3-1

3.2.

Strategic Planning

3-1

3.3.

3.2.1.

Introduction

3-1

3.2.2.

Flood Immunity

3-1

3.2.3.

Community Impacts

3-3

3.2.4.

Acceptance of a Lower Standard

3-5

Development Control

3-5

3.3.1.

Introduction

3-5

3.3.2.

Departmental Impacts

3-6

3.3.3.

Development Impacts

3-7

3.3.4.

System Augmentation

3-9

3.

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

March 2010

iv

Chapter 3
Strategic Planning & Development Control

Department of Transport and Main Roads


Road Drainage Manual

Chapter 3
Strategic Planning & Development Control

Chapter 3
Strategic Planning and Development
Control
3.1. Introduction
This chapter addresses two departmental
functions that are important in the preplanning of a road corridor or link and the
on-going stewardship of a road with respect
to drainage.
The first is concerned with strategic
planning and the major drainage
considerations required when developing
strategies and plans for state controlled
roads and roads within the Auslink
network. The second is concerned with the
possible effects on road drainage systems
due to the development of the road and
external environments (refer Section 2.2).

3.2. Strategic Planning


3.2.1.

Introduction

Chapter 2 of this manual has described


various design considerations, controls,
criteria and standards that apply when
planning a drainage system. This section
expands on these aspects with a focus on
strategic, pre-project planning.
Reference should be made to the Guidelines
for Strategic Road Network Planning
(DMR 2008) released by the departments
State-wide Planning Group for guidance in
strategic planning and the development of
documents such as Road Route Strategies
and Road Link Plans.

The design considerations for the strategic


planning of drainage include:
flood immunity;
community impacts;
safety; and
environmental impacts.
This section will discuss key aspects
affecting flood immunity and community
impacts. These aspects need to be carefully
considered and initially addressed when
allowing for road drainage at the strategic
level of planning.
Safety and
environmental considerations will not be
discussed, however it is essential that any
planned or proposed changes to the road
environment through the development of
the road should not have an adverse effect
on safety and the environment.

3.2.2.

Flood Immunity

In the planning of a road corridor, link or


project, the flood immunity expected /
required is an important consideration.
However as stated in Section 2.1.3 and
detailed later in this chapter, the assessment
/ determination of flood immunity for a
road is extremely difficult.
The accepted practice is to assess /
prescribe the flood immunity for individual
types of drainage infrastructure, such as
cross drainage and surface drainage
infrastructure, along the road.
The
departments general design criterion for
flood immunity for cross drainage on state

March 2010

3-1

Department of Transport and Main Roads


Road Drainage Manual

controlled roads is to achieve a design that


provides an average recurrence interval of
50 years (ARI 50 years). Projects on roads
within the Auslink network usual require an
average recurrence interval of 100 years.
Designers should refer to Section 2.5 for
specific requirements.

While all road planning and design projects


should aim for these objectives, there are a
number of considerations that affect this
objective in particular circumstances. This
section
discusses
some
of
these
circumstances and outlines how these
additional issues can be incorporated into
the process.
For the project concerned, additional
considerations are needed in the assessment
of flood immunity required. These include:
In some
Project economics.
circumstances, provision of a
required level of flood immunity may
come at a very high cost, which may
be difficult to justify because of the
function of the road. This situation
often occurs in large flat floodplains,
where there is an extensive length of
road across the floodplain.
Road alignment and corridor. In
some situations, the road alignment
and corridor width may make a high
level of flood immunity difficult to
achieve, and a lower level may need
to be adopted. This situation usually
occurs in areas with significant
controls or constraints, where there
are environmental issues or in urban
areas.
Community impacts. These impacts
may affect the flood immunity
standard to be adopted, especially in
flat areas. Often in these situations,
there will be a significant width of

March 2010

3-2

Chapter 3
Strategic Planning & Development Control

flow in the natural or existing


environment. If the road is to be
upgraded to a higher level of flood
immunity, a significant flow must be
directed under the road, which will
tend to concentrate the flow. In this
case, afflux is difficult to manage
without a significant amount of cross
drainage. A lower level of flood
immunity will allow extra flow
across the road and thereby result in a
better outcome related to afflux. This
benefit may be greater than the
concern with the reduced flood
immunity.
Flood immunity along a road link. In
this case, the flood immunity along a
whole road link needs to be
considered. If there is a drainage
crossing on the road link where it is
clear that it cannot be upgraded to an
ARI 50 year flood immunity, the
upgrade of other crossings on the link
may not be justified, because the
whole link may be closed anyway
whether or not the particular crossing
is upgraded. This issue can be
influenced by the availability of an
acceptable alternate route.
In all of these cases, the required level of
flood immunity would usually be
technically achievable, though at a cost that
cannot be justified for the benefit gained.
For example, the Murray River Crossing in
north Queensland, as shown in Figure 3.2.2,
has been design based on a time of closure
criteria as the cost to construct the highway
to normal levels of immunity was
prohibitive.
The consequences of adopting a lower level
of flood immunity can be analysed by
consideration of the extent of traffic
disruption and the economic impacts of this

Department of Transport and Main Roads


Road Drainage Manual

disruption. This can be considered in


conjunction with the assessment of times of
closure discussed further below.
While it may be acceptable to adopt a flood
immunity standard for a project that does
not meet the general criterion, in all cases it
is essential that the justification for the
decision should be clearly detailed.

Chapter 3
Strategic Planning & Development Control

closure. Therefore a small steep catchment


where the flood immunity of the road is low
may suffer a similar amount of traffic
disruption, over a long period of time, as a
large flat catchment with a higher flood
immunity standard.
This means that
additional data for selection of an
appropriate flood immunity standard can be
gained from consideration of time of
closure.
Because of this factor, time of closure can
be considered in conjunction with the flood
immunity, to possibly adjust the design
criteria for a particular crossing.

Figure 3.2.2 - Murray River Crossing of


Bruce Highway

3.2.3.

Community Impacts

3.2.3.1. Time of Closure


While the flood immunity of a drainage
system is important, the actual disruption to
transport depends on the frequency and
duration of the time of closure. In Chapter
2, the time of closure for a road was defined
as the periods of time when water over the
road is a hazard to vehicles. This is defined
by Austroads and the department as when
the total head of water over the road is more
than 0.3 m.
It may be the case that a road with a low
level of flood immunity is closed for short
periods of time (though frequently), in
which case there may be limited disruption
to transport, and the low flood immunity
may not be a serious concern.
Time of closure depends on the catchment
response time. Small and steep catchments
flow faster and the times of closure will be
relatively short, while large and flat
catchments will have a longer time of

Through the analysis of time of closure, the


costs of traffic disruption can be analysed
and decisions made on the level of flood
immunity that can be justified for the
investment.
This analysis is an important consideration
in the road planning process, but as with the
possible adoption of a lower level of flood
immunity, careful justification of the time
of closure is needed in the analysis.
3.2.3.2. Risk of Link Closure
When planning the desired immunity level
of a road link, an important concept, that is
often not understood or implemented well,
is the assessment of the risk of closure.
This concept or issue is particularly
important where there is no realistic
alternative route available.
The key part of the risk assessment that is
poorly understood is the link between the
design ARI for cross drainage structures
and the probability of closure of the road
link. The probability of closure for an
existing road link is not simply based on the
minimum ARI standard of all cross
drainage structures along the road link.
Equally, it is incorrect to set the acceptable
level of risk of closure for the road link and
March 2010

3-3

Department of Transport and Main Roads


Road Drainage Manual

then adopt this level (in the form of an ARI)


for the design of each drainage crossing.
This will not necessarily achieve the desired
level of immunity or risk of closure.

The department usually designs individual


cross drainage structures to a standard of
ARI 50 years. Statistically this means the
probability of the road being closed at one
of these crossings is once in 50 years.
However, due to the independence of
rainfall events over time and potentially
between catchments, any drainage crossing
along the link could close, due to a greater
than ARI 50 year event, independently from
all other crossings. This situation, when
considered across the whole road link,
could greatly increase the risk (probability)
of closure of the road link.
The assessment of probability of closure for
a road link requires the determination of
dependency between the crossings, that is,
are the crossings along a road link
independent of each other or not. The level
of dependency influences the level of
probability and therefore risk of closure.
This can be explained by use of an
example. Assume a road link with 5
drainage crossings, each designed with a
flood immunity of ARI 50 years. If all of
the crossings on the road link are totally
dependent, then a single rainfall event
greater than ARI 50 years will close all of
the crossings at the same time. This gives
the road link a probability of closure of 1 in
50. However if the crossings are fully
independent, each crossing can be subjected
to its own greater than ARI 50 year rainfall
event at different times to the other
crossings. Therefore the risk of closure of
each crossing is independent of what occurs
at the other crossings. This situation can
give the road link a probability of closure of
about 1 in 10.

March 2010

3-4

Chapter 3
Strategic Planning & Development Control

Normally the crossings are dependent to a


certain extent and the risk will be between
the two cases noted above. In this case a
single event may affect more than one
crossing at a time on some occasions, but
they will be affected independently at other
times.
Crossings will be fully dependent if the
length of the road link is small and a single
rainfall event will always affect all
crossings together.
The dependence
decreases as the length of the road link
increases and the probability of a single
rainfall event affecting all catchments
reduces. The dependence is low when there
are many small catchments. In this case,
each small catchment may be affected by a
localised short duration storm event, which
will only extend over a limited geographical
area, and these events may occur anywhere
along the road.
For example, if there is a long length of
road with 50 small catchment areas, each
with its own crossing designed for an ARI
50 year flood event, the risk of closure of
the road link is quite high in every year.
The analysis of this risk is complex and
depends on an assessment of the catchment
types, the expected rainfall mechanism and
the distance between crossings.
Calculation of the
whole link needs
immunity of each
well as the degree
crossings.

risk of closure of the


to consider the flood
individual crossing as
of independence of the

The analysis of this combined risk is a


complex statistical analysis. However it
should be considered in many projects, to
ensure that there is a good understanding of
the total risk of closure. For further
discussion and/or advice regarding this type
of analysis, contact the Director

Department of Transport and Main Roads


Road Drainage Manual

Chapter 3
Strategic Planning & Development Control

(Hydraulics),
Hydraulics
Section,
Engineering & Technology Division.

3.2.4.

3.2.3.3. Flood Impacts

While the drainage design for state


controlled roads and roads within the
Auslink network should provide for the
general required level of flood immunity,
there are situations where this standard is
impossible to meet. In this case and after
careful consideration of the above discussed
issues, a lower standard of flood immunity
may be adopted.

While the risk of traffic disruption, as


shown in Figure 3.2.3.3, is the main
criterion for selection of suitable flood
immunity, the impacts on the local
environment
and
community
also
contribute to the selection of flood
immunity.
The impacts on the community result
especially from directing flow from a wide
flow path over the road through a relatively
narrow set of drainage structures.

Acceptance of a Lower
Standard

However if this is the case, the justification


of this decision must be documented.

3.3. Development Control


3.3.1.

Introduction

There are two aspects of development


control related to drainage that the
department must consider.

Figure 3.2.3.3 - Flooding in Mackay

In order to meet the initial required


immunity level, the provision of sufficient
drainage structures may not be justified by
cost and therefore a lower flood immunity
may seem to be an appropriate option. In
this case, the additional cost of providing
the extra drainage to meet the initial
required immunity must be balanced
against the extra inconvenience (time and
cost) of more extensive road closures that
would result from selecting a lower level of
flood immunity.
However, this situation may be justified
especially in flat floodplains, where there
are extremely wide flow paths.
The
justification also depends on the traffic
volume and the nature of the traffic using
the roadway.

Firstly, the department could be regarded as


a developer because it controls and directs
the construction of road infrastructure and
this development may have an impact on
the surrounding environment, both natural
and built. Road planning therefore must
determine, assess and mitigate any impacts
to an acceptable level.
Secondly, the department needs to be aware
of development near roads and/or in a
catchment that may impact on existing
departmental drainage infrastructure. This
impact could be a change in flood levels or
flows or a diversion of runoff.
The
department is consulted on development
approvals when the proposed development
is within 100 m of an existing or future
planned state controlled road.
This
criterion provides for a number of
developments, but there are occasions
where the proposed development is in the
catchment draining to the road crossing, but

March 2010

3-5

Department of Transport and Main Roads


Road Drainage Manual

further away. Development anywhere in


the catchment may have an adverse impact
on the road drainage, even if it is remote
from the road. In this case, the department
must maintain surveillance of development
and make appropriate allowances or
provide advice to developers or the council.
Consultation with local authorities assists in
this provision.
Both of these aspects must be analysed to
ensure that the departments road
infrastructure is and/or remains acceptable
from the point of view of drainage
considerations. To enable analysis of these
aspects, the department must firstly
determine or establish the hydrologic /
hydraulic conditions of the site including
the capacities (and immunity level) of any
existing drainage infrastructure.
This is a complex area and it is difficult to
provide any clear-cut criteria, but there are
some general principles that should be
considered in assessment of development
and this section has some comments on
relevant issues.

3.3.2.

Departmental Impacts

When the department builds a new road or


rebuilds / upgrades an existing road, the
drainage impacts of this road must be
considered. There are three main aspects,
namely afflux, the concentration of flow
and flow diversions.
Permissible levels of afflux are discussed in
Chapter 2 of this manual, and should be
referenced in all cases. Afflux is a critical
consideration / criterion as it is often the
controlling factor in drainage designs. New
road embankments and changes to existing
embankments (even by small amounts) will
create or change water levels both upstream
and downstream of the crossing. These

March 2010

3-6

Chapter 3
Strategic Planning & Development Control

impacts need
mitigated.

to

be

determined

and

Generally, floodplain flow will tend to be


concentrated when directed through
culverts under the road. This concentration
of flow provides a higher risk of scour at
the culvert outlet. The design therefore
must consider this risk and make
appropriate allowances.
Roads may actually change the direction of
flow in some circumstances, and this
diversion could have serious adverse
impacts on the environment and
neighbouring property owners. The design
must carefully review the possible flow
redirection and generally minimise any
diversions.
The impact of road drainage on flooding is
usually analysed for the range of floods up
to an average recurrence interval of 100
years, the most commonly used flood
criterion for floodplain management in
local authorities.
However in some circumstances, it may be
appropriate to consider larger floods, even
up to extreme events. This situation usually
arises in urban areas, or where there are
particularly sensitive locations. It is also
only relevant where the road construction
may cause obstruction to flow for larger
floods, while allowing the smaller floods up
to ARI 100 years to pass through the
drainage structures. Occasions where this
may be important is where there is a high
embankment, where there may be safety or
noise barriers on the road or where
overpasses cause obstructions.
The road design should be reviewed in all
cases, and specific analysis should be
carried in cases where this is considered
necessary. A risk assessment should be

Department of Transport and Main Roads


Road Drainage Manual

carried out in each case and if necessary,


modifications should be made to the design.

3.3.3.

Development Impacts

The most important development issue of


concern for the department is urbanisation
(including residential and commercial
development).
Urbanisation increases
stormwater runoff from a catchment.
However while urbanisation may be the
most significant development that may
affect road drainage, there are other forms
of development that may also have an
impact. These include:
Levees and other farm works. These
may divert flows across the
floodplain and may therefore change
the point where flow must cross the
road.
Dams, detention basins and other
water storages may affect flood
levels and discharges. Refer Figure
3.3.3.

Figure 3.3.3 - Detention Basin in a


Residential Subdivision

Urbanisation
increases
stormwater
discharge by increasing the impervious area
in a catchment and by improving the
channel conditions. The combination of
these two factors increases the volume of
runoff and the peak discharge and changes
the time the peak discharge occurs, both of
which may affect the existing road

Chapter 3
Strategic Planning & Development Control

drainage.
Furthermore, urbanisation
generally provides artificial flow paths and
can reduce the floodplain storage by the
filling of depressions and so on. These
aspects also increase the flood discharge.
Any increase in discharge will most likely
affect the flood immunity of the road as it
would have been designed for less runoff.
The impacts to departmental drainage
structures
(located
downstream
of
development) can be the increased chance
of overtopping the road and/or increased
outlet velocities.
These factors inturn
increase the risk of scour, water quality
problems and safety concerns. Also, the
increased discharge will most likely
increase the peak water levels at that
location, increasing the level of flooding.
Urbanisation or development downstream
of departmental drainage structures may
change the condition of the outlet channel
(in the external environment). Change or
improvement in the channel will most
likely change the tailwater level at the
structure. If the channel can drain the
stormwater away more quickly than before
the changes were made, the tailwater at the
structure will drop. This can change the
operation of the culvert and in turn could
mean increased outlet velocities. If the
channel capacity is reduced or restricted,
the tailwater at the structure will increase
which will reduce the capability of the
culvert which will typically increase
flooding on the upstream side of the
structure.
Where development is planned that may
affect the departments drainage systems,
the development should be reviewed to
ensure that the existing operation and
conditions of departmental owned /
controlled drainage structures is not
adversely affected.

March 2010

3-7

Department of Transport and Main Roads


Road Drainage Manual

These reviews should not be limited to


cross drainage infrastructure and must
include the following departmental drainage
infrastructure:

diversion channels;

Stormwater management reports should be


received from developers, or consultants for
developers, for all proposed developments
where the runoff or flooding may affect a
state controlled road. These reports should
be reviewed to assess the potential impacts,
and if there are impacts acceptable
mitigation measures should be proposed.

energy dissipation measures;

Key requirements of these reports are:

longitudinal drainage (table drains,


kerb and channel and so on);

Chapter 3
Strategic Planning & Development Control

retention / detention basins


levees;
catch banks / drains;
underground systems (pits and pipes)
and subsoil drains;
water treatment / quality devices
(including sediment basins); and
any other environmental protection
device / measure related to drainage.
The department should check for:
worsening of flood levels (afflux)
upstream and downstream of the
road;
any increase in the risk of water
occurring on / overtopping the road;
any change in the risk of scour
because of larger flows / higher
velocities; and
any increased risk of environmental
harm or change in water quality.
As well as individual impacts, cumulative
impacts should also be considered. These
impacts are where the development
currently being proposed is one of several
(or many) that may occur. One individual
development may not have an adverse
impact, but further similar developments
may be unacceptable when they are all
combined.

March 2010

3-8

The flood report should be prepared


by a suitably qualified and
experienced consultant.
The hydrologic and hydraulic
modelling should be appropriate for
the required assessment, and should
be described fully in the report.
The analysis should calculate the
flood discharges and flood levels for
a range of ARIs.
The base case should calculate the
flood discharges and levels for the
existing conditions, and clearly show
the results where the flow crosses the
state controlled road. It is possible
that the base case shows that the road
has a flood immunity that does not
meet the departmental criterion, but
the objective of this analysis is to
show no worsening.
The developed case should include
the proposed development and
should calculate the flood discharges
and levels at the state controlled road.
If there is an adverse impact,
mitigation measures must be
provided. Adverse impacts include
an increase in flood discharge or
flood level at the road. If there is an
increase in flood level or discharge,
but the road still maintains the
required flood immunity, this may

Department of Transport and Main Roads


Road Drainage Manual

still be regarded as adverse, since


other similar developments could
make conditions worse. The flood
discharges and levels for the
mitigated case should be shown at
the state controlled road crossing and
these must be no worse than for the
base case.
Mitigation measures may include
detention basins, channel works,
diversions or other works that ensure
that the flood conditions are not
worsened.
It is important that
mitigation measures at one crossing
should not worsen conditions at other
locations.
The study should be supported by a
comprehensive report that describes
the analysis undertaken and presents
assumptions with the results.

Chapter 3
Strategic Planning & Development Control

To allow discussion and negotiation in


regard to any financial contribution from a
developer, a reasonable basis for
negotiation needs to be established. The
following cases outline different situations
that can inturn form the basis of discussion
/ negotiation with developers.
3.3.4.1. Case 1
If it is determined that the existing
departmental drainage infrastructure meets
current and planned (immunity and
environmental) requirements and currently
performs / operates satisfactorily, then any
change required to the existing drainage
infrastructure to enable it to adequately
handle the changed hydrologic, hydraulic
and/or environmental conditions caused by
the development should be met by the
developer.
3.3.4.2. Case 2

3.3.4.

System Augmentation

As stated in Section 3.3.1, the department


must firstly determine or establish the
hydrologic / hydraulic conditions at the site
including the capacities (and immunity
level)
of
any
existing
drainage
infrastructure.
When these existing
conditions are compared to the drainage
outcomes of any proposed development, the
differences and potential impacts can be
determined and understood.
In the event that the department believes
that a proposed development will have
adverse affects on its existing drainage
infrastructure a financial contribution from
developers can be requested to allow the
department to undertake appropriate work
to augment or upgrade the existing drainage
infrastructure in order to handle the
changed conditions (hydrologic and/or
hydraulic) caused by the development.

If it is determined that the existing


departmental drainage infrastructure meets
current and planned (immunity and
environmental) requirements but does not
currently perform or operate satisfactorily,
then the department would be responsible to
undertake remedial work to enable the
infrastructure to adequately perform /
operate while any change required to the
existing drainage infrastructure to enable it
to adequately handle the changed
hydrologic, hydraulic and/or environmental
conditions caused by the development
should be met by the developer.
3.3.4.3. Case 3
If it is determined that the existing
departmental drainage infrastructure does
not meet current or planned (immunity and
environmental) requirements, then the
changes required to the existing drainage
infrastructure to meet current or planned

March 2010

3-9

Department of Transport and Main Roads


Road Drainage Manual

(immunity
and
environmental)
requirements is the responsibility of the
department
while
any
additional
augmentation to the infrastructure required
to adequately handle any additional
hydrologic, hydraulic and/or environmental
conditions caused by the development
should be met by the developer.

March 2010

3-10

Chapter 3
Strategic Planning & Development Control

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Chapter 4
Data Collection

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Chapter 4 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Table of Contents
4.1

Introduction

4-1

4.2

Types of Data

4-1

4.3

4.2.1

Strategic Data

4-2

4.2.2

Project Data

4-4

4.2.3

Sources of Data

4-5

Environmental Assessment

4-6

4.3.1

Vegetation

4-6

4.3.2

Fauna

4-7

4.4

Forms and Checklists

4-7

4.5

Field Inspections

4-8

4.6

Rainfall

4-8

4.7

Flood Data

4-9

4.8

Drainage and Flow Patterns

4-10

4.9

Waterway Characteristics

4-11

4.10 Water Quality Data

4-12

4.11 Topography

4-12

4.12 Soils

4-13

4.12.1

Natural Soils

4-13

4.12.2

Acid Sulphate Soil Information

4-14

4.13 Existing Infrastructure

4-14

4.14 Survey

4-15

4.14.1

Aerial Imagery

4-15

4.

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

March 2010

iv

Chapter 4
Data Collection

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Chapter 4
Data Collection
4.1

Introduction

In this chapter, general guidance is


provided on how to source and collect data
and how to conduct site surveys /
assessments to assist in the planning and
design of road drainage infrastructure. This
data is required to quantify the design
requirements as described in Chapter 2.
The various forms of data used in the
planning and design of drainage
infrastructure are broadly categorised as
either strategic data or project data.
Designers should ensure that collected data
is appropriately stored for easy retrieval,
not only during the preconstruction
activities of the project, but also in the
future.
This chapter discusses:
(a) Types of data;
(b) The importance
strategic data;

and

sources

of

(c) Types of project data, sources and


application;
(d) The importance of site surveys and
assessments to an overall project; and
(e) Methods for collecting and recording
data.

4.2

Types of Data

Data is progressively collected, analysed


and used throughout all preconstruction
activities at a level of detail that is
appropriate for the purpose being
considered. Data that is collected and used

for network planning and the development


of Road Route Strategies (DMR 2008) is
strategic and regional in nature. The data
may become more focussed and
geographically specific as the strategies are
used to prepare Road Link Plans (DMR
2008). Data collected for these purposes is
defined as strategic data.
In the development of specific Project
Proposals, the strategic data needs to be
reviewed and expanded with the
introduction of more detailed, project
specific data.
As a project proposal
progresses through various preconstruction
activities, refinement of data occurs through
various investigations and studies and as
new design specific data is obtained.
During the Construction activities, more
data is collected, usually as as-constructed
detail. Once a project is completed and
becomes operational, further data regarding
the operations and maintenance of the road
should be recorded as part of the asset
management process.
Data collected during Preconstruction,
Construction
and
Operational
/
Maintenance activities is defined as project
data.
Both strategic and project data, with respect
to drainage, is useful not only to the
department but also to others who are
interested in information such as flood
levels and so on.
Local authorities,
developers and consultants may refer to the
department for assistance in providing
observed flood levels and so on in areas of
interest and this assistance should be given
where relevant / appropriate.
March 2010

4-1

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Data that has been obtained from various


sources for use in planning and design work
should be retained as part of the
documentation for the project.

4.2.1

Strategic Data

Strategic data is usually regional in nature


and is required for network planning and
the preparation of Road Route Strategies /
Road Link Plans (DMR 2008). It may also
be required for the planning and design of
drainage infrastructure.
It can be
considered in four types of information:
Type 1 - Planning Instruments such as:

Type 3 Drainage and Water Management


Infrastructure such as:
specific drainage infrastructure;
water catchment storages
aquaculture, fish); and

(e.g.

irrigation schemes.
Type 4 Private or Public Utility Plant
(PUP) such as:
communications systems;
municipal services;
trunk distribution systems for oil,
gas, water and effluent;

regional strategic land-use plans;

electricity transmission lines; and

statutory and advisory


management plans;

state and interstate railways and


industry narrow gauge rail systems.

land-use

land based and marine national parks;


land based and marine estuarine
environmental
protection
and
management plans;
other land and
management plans;

water

based

commonwealth government planning


instruments;
local authority
schemes; and

town

planning

urban
and
rural
drainage
management plans, initiated under
State Government Legislation.
Type 2 Naturally Occurring Events such
as:
storm event data;
flooding event data;
abnormal highest astronomical tide
event data; and
storm surge event data.

March 2010

4-2

4.2.1.1 Planning Instruments


The department may be a participant in the
planning processes that create some of these
instruments to ensure appropriate road
service delivery is provided through
Queensland.
However, planners and
designers need to work within the overall
statutory and advisory planning framework
when developing various strategic network
plans and when planning and designing
specific projects.
As these instruments may change over time,
it is not advisable to attempt to store this
type of data but rather obtain current
information at the start of each new project
and review the currency of this information
as the phases of a project progress.
Land-use planning is one form of data that
can change within the departments
planning and design time frame. In rapidly
developing urban areas, upstream and
downstream land-uses could change
through:

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

issue or the review of a Regional


Land-use Plan;

lasting benefits for many years after the


flood has receded.

amendments to the planning scheme;


or

During site visits, observations and


measurements need to be documented to
record matters such as:

the completion of a new planning


scheme.

flood levels;

All relevant Department of Transport and


Main Roads regional and district offices,
need to be part of the regional planning and
local authority planning processes to ensure
that drainage infrastructure is consistent
with land-use planning.

inundated areas;

4.2.1.2 Naturally Occurring Events

culvert embankment performance


under prolonged headwater; and

An important part of the data collection and


site assessment work for drainage
infrastructure is the collection and analysis
of data following naturally occurring events
such as seasonal storms and extreme events
such as floods, abnormally high tides and
storm surge events that can accompany
cyclones.
These events occur from time to time and
are an excellent opportunity to gather data
on the performance of the road. The data
collected as part of this programme needs to
be archived in an appropriate database to be
available as a historical record for planning
and design.
Routine inspections following seasonal
storm events provide opportunities to assess
drainage performance and document
maintenance requirements.
When a surge or flood event has occurred, a
departmental officer needs to visit that
region as soon as possible after the event or
even during the event if safe access is
possible.
While there are other
requirements for staff during and after
extreme events, collected data will provide

water flow patterns;


scour and erosion behaviour;
debris accumulation;

floodway performance;

fauna assistance measures.


Observed features (e.g. flood and debris
marks) should be photographed or videoed
and marked for later survey and
documentation. Departmental staff, land
owners and residents are sources of
anecdotal information that could be useful
in confirming or calibrating measured
information.
4.2.1.3 Drainage
and
Water
Management Infrastructure
Local authorities and various statutory
authorities manage urban and rural drainage
systems that are designed for:
existing and future land-uses in the
catchment;
specific
hydrological
environmental parameters; and

and

local drainage parameters.


Planners and designers need to obtain data
relating to the design of these facilities so
that
the
departments
drainage
infrastructure is compatible with local
authority planning and design.

March 2010

4-3

Department of Transport and Main Roads


Road Drainage Manual

Drainage infrastructure located in the


catchments of existing or planned
municipal water storages may need to
conform to requirements of the relevant
catchment authority particularly in matters
of water quality and erosion and sediment
control. Planners and designers should
consult with the catchment management
authority to ascertain requirements for
drainage in the catchment under review.

Authorities managing irrigation schemes


may have similar requirements for drainage
infrastructure to municipal catchment
management
authorities
and
these
requirements should be obtained. Details of
existing irrigation infrastructure should be
confirmed by survey and any expansion
plans obtained from the authority.
4.2.1.4 Private / Public Utility Plant
As the department generally approves the
location of service infrastructure within the
road corridor, documentation associated
with these approvals provides an initial
source of data for new drainage projects.
This data needs to be verified with the
agency involved and any information
supplied should be confirmed with site
measurements and ground survey.
Existing services in the vicinity of drainage
infrastructure needs to be located by survey
to ensure:
that the service installation does not
impair drainage performance;
maintenance
of
drainage
infrastructure can be completed
without damage to the service or the
drainage; and
the extent of any PUP relocation
requirements to enable drainage to be
correctly installed can be determined.

March 2010

4-4

Chapter 4
Data Collection

4.2.2

Project Data

Project data is more relevant to the planning


and design of specific projects and largely
relates to the physical characteristics of a
site and the surrounding catchment. It may
be collected or measured at varying times in
the different phases of the planning process
and at different levels of detail.
Project data includes:
land-use;
topographic information;
catchment information;
rainfall data;
stream flow and flooding information
stream flow patterns;
tidal information;
waterway
stability;

characteristics

and

water quality;
sedimentation issues;
soils data;
erosion history;
vegetation constraints;
acceptable time of inundation;
fauna habitats;
downstream conditions;
service installations; and
obstructions.
Specific project and routine maintenance
inspections provide opportunities to obtain
data and to review the in service
performance
of
the
infrastructure.
Inspections should have similar objectives
to those outlined in Section 4.2.1.2 for
inspections following extreme events, and

Department of Transport and Main Roads


Road Drainage Manual

the findings recorded in the district


database.

Chapter 4
Data Collection

field investigations, studies and recorded


information in various forms such as:

The quality of data collected has a direct


bearing on the successful design and
implementation of drainage infrastructure
and is strongly linked to an effective site
assessment and planning process. It is
important that adequate data is collected in
the early stages of a project and that it is
stored in a readily available format for use
in all subsequent phases.

existing field inspection records;

For example, the collection of soils data at


the planning or design phase of a project
will facilitate the selection of appropriate
erosion and drainage controls and the
preparation of an appropriate Erosion &
Sediment Control Plan (ESCP) for the
construction phase.

land resource manuals;

Site assessment is also strongly linked to


risk assessment.
A thorough site
assessment, where data is added at each
stage of the project, will lead to a reduced
risk of adverse impacts to the surrounding
environment or to the road itself. This in
turn will lead to reduced costs in the long
term.
The identification of special environmental
characteristics of a project site is a key
requirement whilst undertaking a site
assessment, though it is expected that most
such characteristics will be identified as
part of the environmental assessment
process. Knowledge of special conditions
and factors which influence sensitive
environments facilitates environmentally
responsible drainage design, as defined in
subsequent sections of this manual.

4.2.3

Sources of Data

Different phases and steps during the


preconstruction process may utilise the
same data. This data may be obtained from

topographic maps;
documentation obtained during the
environmental assessment process;
existing design drawings;
geotechnical investigations;
survey records;

aerial photographs;
published references (e.g. Australian
Rainfall & Runoff);
previously published reports and
investigations
(i.e.
Feasibility
Studies);
Concept and Link Studies;
Acid Sulphate Soils maps;
vegetation maps;
flood maps; and
various electronic data sources (eg
geospatial data).
data is available from various
departmental sources, from land
owners and organisations such as:
Bureau
of
(Commonwealth);

Meteorology

Department of Environment and


Resource Management (State);
Department
of
Employment,
Economic
Development
and
Innovation (State);
Queensland Rail;
historical societies;
local authorities;

March 2010

4-5

Department of Transport and Main Roads


Road Drainage Manual

port authorities;
industry organisations; and
environmental groups including
catchment management groups and
river trusts.
Table 4.2.3 has been prepared to indicate
the type of data available from each of these
external organisations.

4.3

Environmental
Assessment

For every infrastructure project, the


department has a responsibility to consider
the projects potential environmental effects
and / or impacts and to then develop
appropriate
mitigating
measures
as
necessary. Therefore an Environmental
Assessment (EA) for the project is required
and this assessment should be undertaken
as early as possible in a projects
development.

4.3.1

Vegetation

The vegetation surrounding a project site


reduces raindrop impact on the soil as well
as stabilising the soil. These factors are
important in erosion control.
The
vegetation also filters runoff containing
sediment. Knowledge of the vegetation
characteristics of the project site:
assists in the determination of the
existing degree of disturbance (if
any) of the site and its potential for
erosion;
enables the protection of species with
conservation significance;
guides the selection of species for
revegetation;

March 2010

4-6

Chapter 4
Data Collection

Table 4.2.3 - Data Sources


Data Type

Rainfall data (historic)


Flood levels (historic)
Tidal data
Cross sections
Topography
Soil information
Flora and fauna
Survey data
Water quality
Existing infrastructure
Aerial photography

External
Organisation
1, 2, 6, 9
1, 2, 4, 5, 6, 9
7, 9
2, 4, 6, 7
2, 6
2, 3
2, 3, 6, 8
2, 4, 6, 7,
2, 3, 6, 8
2, 4, 6, 7, 9, 10
2, 3, 6, 11

1 = Bureau of Meteorology
2 = Department of Environment and
Resource Management
3 = Department of Employment, Economic
Development and Innovation
4 = Queensland Rail
5 = Historical Societies
6 = Local Authorities
7 = Port Authorities
8 = Environment Groups
9 = Local Residents
10 = Service Providers
11 = Web based data sites

assists in determining roughness


(Mannings n);
contributes to the determination of
the coefficient of runoff; and
assists in the identification
constraints to drainage design.

of

The collection of vegetation data, both


terrestrial and aquatic, is an important
process. This data is then used in other
assessment and design processes referred to
in this manual.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

The EA for a project should gather the


following data (with mapping) where
possible:

The location of drainage structures and


discharge points may also affect fish or
bank dwelling species such as platypus.

extent and location of all vegetation


types (terrestrial, littoral, intertidal,
aquatic, trees, shrubs, vines and
grasses) in and around the road
environment;

Again, the EA for a project should gather


the following data (with mapping) where
possible:

description and location of any


vegetation corridors that traverse the
road environment;
description of the conservation
significance
of
vegetation
communities within the study area;
description of any rare or endangered
species;
extent and location of any cleared
vegetation and incidence of exotic
species and weeds; and
description and location of flora used
traditionally for food, spiritual and/or
cultural purposes.

4.3.2

Fauna

Recognition of the impacts of road corridor


development on fauna populations has led
to modifications in the way that roads are
now designed.
Fauna can influence
drainage design significantly and collecting
data on ground dwelling fauna in a project
area is essential.
Research has been undertaken on
developing practices that help facilitate
fauna movement through passages in the
road corridor via drainage structures in a
way that minimises fauna mortalities on the
road. The provision of fauna passage is
important and may influence the physical
dimensions of a drainage structure.

species diversity and abundance for


terrestrial, littoral, inter-tidal and
aquatic and avifauna;
description and location of any rare
or endangered species;
fish habitat / passage requirements;
occurrence,
distribution
and
requirements for migratory species;
species important for traditional,
recreational
and/or
commercial
fisheries; and
any local terrestrial, aquatic or
avifauna used traditionally for food,
spiritual and/or cultural purposes.
Appendix 4A provides further information
on data collection and site assessment for
fauna
passage
through
drainage
infrastructure.

4.4

Forms and Checklists

To assist in the collection and retention of


data, the department has prepared a number
of pro-formas for use in data collection.
These forms should be used as they
provide:
a checklist to ensure all relevant data
has been obtained; and
media to record data while on site.
Forms currently in use by the department
include:
Bridge Design Request (Form M685)
for bridges over streams;

March 2010

4-7

Department of Transport and Main Roads


Road Drainage Manual

Bridge Hydraulics Design Summary


(Form M2303).

speak to landowners regarding site


issues, drainage and flooding history;

Field Report - Bridge Waterways


(Form M2759); and

identify and photograph site features


that may impact on the selection of
future drainage infrastructure.

Other forms, parts of which are relevant to


the collection and documentation of data
include:
Design Development Report Small
Projects (Form M4211); or

Chapter 4
Data Collection

Design Development Report Large


Projects (Form M4212)

In particular, field inspections should focus


on obtaining an understanding of:
drainage patterns
waterway characteristics
evidence of flooding
existence of debris levels

through

An overall checklist for the collection of


data is provided as Appendix 4A.

evidence of erosion or deposition

4.5

extent and type of vegetation


including vegetative communities

Field Inspections

Field inspections of catchments, existing


drainage and possible sites within a
proposed project area are essential for the
planning of major drainage works and for
the design of all drainage systems.
Field inspections provide opportunities to
understand the site and to assist in
formulating the risk profile for the project.
Where possible, field inspections by the
designer should be organised to be
completed in conjunction with field survey,
soil and environmental investigations to
provide a more integrated data collection
process.
More specifically, site inspections allow the
designer to:
obtain an appreciation of the site and
its constraints;
validate the reliability and currency
of existing records and information
(including anecdotal information);
verify characteristics and parameters
that are to be used in the drainage
planning and design process;

March 2010

4-8

soil types

potential sources of debris


existing infrastructure
location and
buildings

level

of

adjacent

locations for future controls (e.g.


retardation or sediment basins).
The departments Field Report - Bridge
Waterways (Form M2759) can be used as a
useful prompt for many other drainage
related field inspections. In addition, the
Departments Guidelines for Bridge
Surveys (Part G of the Standards for the
Provision of Road Transport Infrastructure
Surveys) provide a thorough guide to the
collection of data.
Data contained in each of these forms has
been combined and included in the
checklist in Appendix 4A.

4.6

Rainfall

The duration and intensity of rainfall are


major components with respect to the
determination of runoff and of erosivity

Department of Transport and Main Roads


Road Drainage Manual

potential.
Both vary with geographic
position and with the time of year. Thus,
rainfall distribution, seasonality and
intensity must be considered in order to
determine flow rates and the potential for
erosion. For more detailed explanations,
reference may be made to Australian
Rainfall and Runoff, A Guide to Flood
Estimation, Vol 1 (AR&R) (IEAust 2001).
(a) Rainfall Distribution
The distribution of median annual rainfall
across Queensland is shown in Figure 4.6.
The rainfall isohyets shown in the figure are
generally parallel to the coast except where
topographic features modify the pattern. In
particular, significantly higher rainfall
occurs between Ingham and Cooktown,
Proserpine and Sarina, and north and south
of Brisbane where there are high ranges
aligned perpendicular to the main onshore
winds.
(b) Rainfall Seasonality
Rainfall is summer dominant throughout
the state but the volume of rain that falls
during the other months varies considerably
between regions. South of the Tropic of
Capricorn and east of a line between
Emerald and Mitchell there is a significant
winter peak in many years.
(c) Rainfall Intensity
Rainfall intensity varies with the type of
rainfall event (such as advective, cyclonic
or frontal) but is generally higher during the
summer months than the winter months.

Chapter 4
Data Collection

4.7

Flood Data

Whilst the majority of structures are sized


using statistically derived flows, the
collection of historic flood data can also
provide valuable information. Flood data
can consist of:
gauging station records;
recorded peak levels;
mapping of flow patterns;
debris marks;

water stains;
photographs or videos;
anecdotal evidence.
Sources of historic flood data can include
landowners, local authorities, Queensland
Rail (e.g. design drawings often highlight
peak flood levels), Department of
Environment and Resource Management,
and the Bureau of Meteorology.
All data obtained must be evaluated for
accuracy and correlated across different
sources where possible. This is particularly
true with respect to anecdotal evidence of
flood heights provided by individuals as:
the observations did not coincide
with flood peak;
there was a lack of visibility (night
time flood);
a significant time (years) has elapsed
since the observation; and
personal observations can change as
time passes.

March 2010

4-9

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Figure 4.6 - Annual Median Rainfall for Queensland

To be useful, it is essential that all flood


height information be related to a
recognised level datum.
For large scale and some urban projects,
historic flood data may be used for the
calibration of mathematical models.
Information should also be sought in
relation to flood gradients, rates of rise or
fall, velocities, and flow patterns (directions
of flow).
For smaller projects, flood data is often
scarce, and hence may only provide an
indication of historic peaks, with no means

March 2010

4-10

available to estimate the average recurrence


interval (ARI) of the flood event.
Reference should be made to the
departments Field Report - Bridge
Waterways (Form M2759).

4.8

Drainage and Flow


Patterns

An understanding of drainage and flow


patterns is required to help ensure that
adequate provisions are made for upgraded
or future drainage infrastructure. This is
particularly important at sites where there is
no existing drainage infrastructure.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Whilst flow patterns may be simple to


ascertain
in
waterways,
careful
consideration is required in relation to
overland or floodplain flow.

Is the channel clear or obstructed by


banks or islands?

Drainage and flow patterns may be


determined through the review of available
topographic maps and aerial photography,
and through field inspection. For many
sites all three techniques should be used.

Is there a clear distinction between


channel and floodplain?

Elements that need to be considered


include:

Are there sequences of pools and


riffles?

Is the stream in a pristine state or has


it been degraded?
Are the banks steep?
Are the banks stable?

direction of flow particularly in flat


areas;

Is there any evidence of current or


past bank slumping?

width of flow;
possible backwater from downstream
impacts, such as rivers or weirs;

Are there any other signs of erosion


or deposition of material? If so, what
type of material is evident?

potential for spill into or from


adjacent flow paths; and

Does the waterway appear to be


stable in location?

obstacles to flow.

Is the low flow channel likely to alter


in location?

4.9

Waterway
Characteristics

The characteristics of a waterway may be


considered in terms of geometry, hydraulics
and the environment.
Geometric characteristics are based on the
physical dimensions of the waterway, and
include:
channel width and depth

Is the waterway consistent in


appearance, or are there pool and
riffle sequences?
Hydraulic characteristics relate to the actual
flow within the waterway. It is important to
note that most field inspections occur
during times of little or no flow, and hence
the data collected is unlikely to provide a
good indication of flood characteristics.
Hydraulic characteristics include:

cross section

flow depth;

bed slope

velocity (note locations


velocities show variation);

channel form
Channel form relates to the geomorphic
characteristics of the channel and notes
should be made of the following issues:
Is the waterway
meandering?

straight

or

where

backwater effects (that is, inundation


by downstream water levels, which
may drown out or control upstream
water levels); and
nature and state of vegetation within
channel / floodplain.

March 2010

4-11

Department of Transport and Main Roads


Road Drainage Manual

The environmental characteristics of a


waterway may also be characterised by its
water quality, soils and vegetation. These
are discussed in subsequent sections.

4.10

Water Quality Data

Water quality data may be required in those


instances where:

proposed works will be discharging


runoff into a waterway defined as
sensitive;
major works are constructed across a
waterway; and
there is a need to design pollution
control measures.
Therefore, whilst the collection of water
quality data will not be required for all
projects, water quality should always be a
consideration.
The extent of this
consideration should be dependent both on
the sensitivity of the waterway and on the
scale of the project.
In some cases, reports on water quality
investigations will be provided through the
mechanisms
of
the
environmental
assessment process.
Existing water quality data within the study
area provides an indication of the health of
the aquatic ecosystem. This data is useful
for identifying potential changes which may
be brought about by the proposed project,
and in particular, how runoff and drainage
may cause adverse impacts on the existing
characteristics.
The possibility of Acid Sulphate Soils
(ASS) should be carefully considered in
coastal regions and specialist advice sought.
Disturbance of these soils can have severe
water quality impacts.

March 2010

4-12

Chapter 4
Data Collection

On the lowest level, the assessment of water


quality may be as simple as noting the
condition of the water at the time of the
inspection (e.g. stagnant, brackish, colour,
turbidity, odour etc.).
For many waterways, water quality
monitoring may be in place. Queries
should be directed to local authorities, the
Department of Environment and Resource
Management (formerly Environmental
Protection Agency), or catchment groups to
determine whether this data exists and is
available.
For major projects (or as defined in the
REF) specific water quality monitoring may
be required. Typically, monitoring should
occur both upstream and downstream of the
site, so that background levels of pollution
can be recorded, and impacts of drainage
works monitored.
The Australian Runoff Quality A guide to
Water Sensitive Urban Design (EA 2006) is
a design guideline that provides an
overview of current best practice in the
management of urban stormwater in
Australia.

4.11

Topography

The collection of topographic data is


relevant to the assessment of both flow and
the potential for erosion. Topographic
mapping is required to allow catchment
definition, and in the absence of survey, an
assessment of the longitudinal gradient of
waterways.
Topographic data is normally obtained from
published topographic maps (Sunmap /
Department of Environment and Resource
Management), or orthophotos (Department
of
Environment
and
Resource
Management) or local authorities.
In
addition, the use of digital terrain models

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

(DTM) is becoming far more common with


time. These are based on aerial laser survey
and photography, and are often held by
local authorities or created for specific
projects.
Typically, the scale of available mapping
varies significantly in accordance with
proximity to population centres. Mapping
with a scale of 1:100,000 is available for
most areas, though this is sometimes
insufficient to provide an accurate
assessment of catchment boundaries or
waterway slope. In these cases, reference
to aerial photographs or field inspection
notes can be of assistance.

4.12

Soils

Soils data is collected for drainage planning


and design purposes from catchment
locations, drainage channels, eroded areas
and the road alignment to obtain soil
properties that can be used to determine the
suitability and limitations of the soils in
service in the following situations:
flow limitations for inlets and outlets
to cross drainage;
open channel permissible velocities;
stability or the need for lining of
longitudinal drainage;
slopes to cut and fill batters.

Whilst soils data may be collected at the


same time as geotechnical data, the uses of
the two data types may be quite different.
Soil characteristics are a major determinant
of soil erosion risk, and hence should be
collected during the early stages of the
planning and design process.
Field inspections can quickly identify
existing erosion and sediment issues along
an existing or proposed road corridor. This
data provides useful and low cost
information that be incorporated into early
planning phases. Where erosion is visually
apparent on existing road corridors,
investigations can be completed to ascertain
the causes of the erosion by:
discussing the problems
maintenance
personnel,
residents and land owners;

reviewing hydraulic and drainage


calculations for the existing designs
to determine flows and velocities.

with
local

reviewing maintenance procedures;


completing
site
measurements,
sampling and testing soil properties;
and

Soils information can be obtained from:


published soil maps and reports;
Department of Environment and
Resource Management and CSIRO
soil databases;
field investigations at each site.
The eleven major soil groups of Queensland
and their distinctive characteristics are
described in the departments Soils Manual
(TMR 2010b). This manual also contains a
description of how to recognise distinctive
soil characteristics to enable field
identification of the soil groups. It is
important to note that in the majority of
cases, soils data should be collected for
both surface and sub-surface layers, as the
upper topsoil layer is usually removed and
stockpiled prior to construction.

4.12.1 Natural Soils


Information on the distribution and
description of soils within Queensland is
mainly available from the Queensland

March 2010

4-13

Department of Transport and Main Roads


Road Drainage Manual

Department of Environment and Resource


Management and CSIRO.
Soil information may be available, from
these organisations, in the following
formats:
printed hardcopy maps produced
following soil surveys of specific
study areas;

published soil survey reports that


accompany the maps and describe the
soil mapping units in more detail some of these reports also contain
chemical and physical analytical data
for samples taken from soil profiles
representing the major soils present
within the study area;
digital GIS soil maps of specific
study areas - most commonly
provided as either ArcInfo Export,
ArcView / ArcMap Shape or
MapInfo TAB files;
digital GIS databases associated with
the maps and often providing
additional information such as
landform, geology, dominant soils
within each mapping unit, associated
soils and the proportion of the
mapping unit that each covers;
digital ACCESS databases that
contain additional information for the
mapping units; and
digital ACCESS databases that
provide soil profile information for
all sites recorded as part of the soil
survey.
Specific soil maps and associated reports
have also been produced for selected
transport corridors.
Care should be taken when using published
soils information as the quality of the data
may vary due to:
March 2010

4-14

Chapter 4
Data Collection

original purpose of the survey;


mapping scale (or level of intensity)
of the survey information; and
methodology used in the survey.
Any associated chemical laboratory
methods may vary over time and should be
checked before proceeding with data
interpretation.

4.12.2 Acid Sulphate Soil


Information
The Department of Environment and
Resource Management is the lead agency
for information and advice on ASS, and is
continually developing ASS risk maps that
show areas dominated by actual (AASS)
and potential (PASS) ASS. These maps
present information on presence or depth
to AASS horizons and presence or depth
to PASS. They are available in hard copy
and electronic forms at a range of scales.
For more information and mapping:
access
the
Department
of
Environment
and
Resource
Managements fact sheets such as
Acid sulfate soils in Queensland at
<www.derm.qld.gov.au/factsheets/pd
f/land/l60.pdf>; or

contact the Queensland Acid


Sulphate Soils Investigation team
(QASSIT),
Department
of
Environment
and
Resource
Management; or

Access CSIRO and the National


Committee for Acid Sulphate Soils
(NatCASS) soil databases.

4.13

Existing Infrastructure

At all locations, it is important to identify


existing infrastructure and Private / Public

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4
Data Collection

Utility Plant (PUP), which may act as


constraints to the design and location of
future drainage measures. At all sites, it is
important to note the location and existence
of:

requirements for bed levels, bed gradient


and channel cross sections. Reference is
also made to the need to identify additional
information, as described in other sections
of this chapter.

adjacent dwellings or other buildings


with floor levels;

When specifying requirements for survey, it


is important to ensure that cross sections are
surveyed perpendicular to the direction of
flow, both within the channel, and on the
floodplain.

existing culverts and bridges;


infrastructure associated with the
supply of services such as
communications, gas, water supply,
sewerage;
industrial pipelines; and
irrigation infrastructure.
The existence of infrastructure may exert a
strong influence on the design of hydraulic
structures. Constraints can include:
maximum allowable upstream water
levels (e.g. based on potential for
flooding of existing buildings and
infrastructure);
obstructions to flow;
diversion of flow; and

The most appropriate and cost effective


method of data capture should be assessed
for each project. Options include traditional
ground survey, photogrammetry and
Airborne Laser Scanning (ALS) / LIght
Detection and Ranging (LIDAR). Also,
any existing geographical data (within
Geographic Information Systems (GIS))
should be reviewed.
For advice and/or additional information,
refer to local departmental regional survey
manager and/or Geospatial Technologies
Section within the Engineering &
Technology Division.

4.14.1 Aerial Imagery

need to maintain pedestrian safety.


For larger projects, it may also be necessary
to obtain details of major infrastructure
such as dams or weirs.

4.14

Survey

In obtaining survey for a project, reference


should be made to the current departmental
surveying standards.
In particular,
reference should be made to the relevant
geomatic survey section as well as the
general
information
section.
These
standards are available on the departments
intranet and internet sites. The geomatic
type
Bridge
Survey
provides
comprehensive
details
as
to
the

Aerial imagery is a valuable source of


information for the design and assessment
of drainage infrastructure, though it is
important to be aware of the dates and times
at which the image was captured. Aerial
imagery may be used to determine, at the
time of capture:
extent, density
vegetation;

and

patterns

of

delineation of overland flow paths;


locations of active erosion (e.g.
meanders);
waterway dimensions where access is
poor; and

March 2010

4-15

Department of Transport and Main Roads


Road Drainage Manual

land-use.
Historical photographs and imagery can
also be a valuable source of information in
relation to assessing historic flood heights,
flow patterns, and waterway characteristics.
Care must be taken in the assessment,
where the date and time of the capture is
not known. Field inspections should be
used to confirm the currency of existing
information.

March 2010

4-16

Chapter 4
Data Collection

Department of Transport and Main Roads


Road Drainage Manual

Appendix 4A
Data Collection Checklist

Appendix 4A
Data Collection

March 2010

4A

Department of Transport and Main Roads


Road Drainage Manual

Chapter 4A
Data Collection Checklist

Appendix 4A Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

4A

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 4A
Data Collection Checklist

Data Collection Checklist


4A 4A
Region/District:
Project Name:

4A

Contract/Project Number:
Each section must be considered for applicability to project. Sub-components should be
checked when data is collected or noted as not applicable to project.

4A.1

Environmental Assessment

Item

Check

Has vegetation and fauna data been obtained from an environmental


assessment?

4A.2

Vegetation Coverage

Item

Check

Have aerial photographs been sought?

Has a field assessment of vegetation coverage been undertaken?

4A.3

Other Forms and Checklists

Item

Check

M685 - Bridge Design Request (for Bridges over Streams)

M2759 - Field Report - Bridge Waterways

M2303 - Bridge Hydraulics Design Summary

M4211 or M4212 - Design Development Report

Local forms / checklists

March 2010

4A-1

Department of Transport and Main Roads


Road Drainage Manual

4A.4

Field Inspections

Item

4A

Chapter 4A
Data Collection Checklist

Check

Existing land use

Evidence of flooding indications of flood heights (debris etc)

Evidence of erosion or deposition

Potential sources of debris

Potential sites for future control measures

Extent and nature of vegetation (refer also section 4A.2)

Drainage and flow patterns (refer also section 4A.7)

Waterway characteristics (refer also section 4A.8)

Water quality (refer also section 4A.9)

Soil characteristics (refer also section 4A.11)

Building infrastructure (refer also section 4A.12)

Drainage and other infrastructure (refer also section 4A.12)

4A.5

Rainfall

Item

Check

Has the following rainfall data been obtained?

Statistical data

Historical records

4A.6

Flood Data

Item

Check

Have gauging station records been obtained?

Have recorded peak water levels been obtained?

Have debris marks, water stains etc been identified?

Has anecdotal evidence been obtained?

For large-scale projects, has historical flood data been sought?

Has flood data been assessed for accuracy? (surveyed or antidotal)

Has reference been made to departmental form M2759?

March 2010

4A-2

Department of Transport and Main Roads


Road Drainage Manual

4A.7

Appendix 4A
Data Collection Checklist

Drainage and Flow Patterns

Item

Check

Verify direction of flow (particularly important in flat terrain)

Width of flow (including low and high banks)

Potential for overflow into or from adjacent flow paths

Identify obstacles to flow

Existing and potential drainage patterns

4A.8

Waterway Characteristics

Item
Have the following geometric characteristics been obtained?

Check

Channel width and depth

Cross section

Bed slope

Channel form

Have the following geomorphic characteristics been considered?


Is the waterway straight or meandering?

Is the channel clear, or obstructed by banks or islands?

Is there a clear distinction between channel and floodplain?

Are the banks steep?

Are the banks stable?

Is there any evidence of current or past bank slumping?

Are there any other signs of erosion or deposition of material? If so, what type
of material is evident?

Does the waterway appear to be stable in location?

Is the low flow channel likely to alter in location?

Is the waterway consistent in appearance, or are there pool and riffle


sequences?

March 2010

4A-3

4A

Department of Transport and Main Roads


Road Drainage Manual

4A.8

Chapter 4A
Data Collection Checklist

Waterway Characteristics cont.

Item

Check

Have the following hydraulic characteristics been obtained?


Flow depth

Velocity (note locations where velocities show variation)

Backwater effects (that is, inundation by downstream water levels, which may
drown out or control upstream water levels)

4A.9

4A

Water Quality Data

Item

Check

Has relevant data been sought from the following sources?

Dept of Environment and Resource


Environmental Protection Agency)

Local Authorities

Management

(formerly

Will on site water quality monitoring required?

Has the presence of Acid Sulphate Soils been checked?

4A.10

Topography

Item

Check

Has relevant data been sought from the following sources?

Dept of Environment and Resource Management / Sunmap

Local Authorities

Have aerial photographs, topographic maps, orthophotos and survey / digital


terrain models been sought?

4A.11

Soils

Item

Check

Has data been sought from the following sources?

Published soil maps and reports

Dept of Environment and Resource Management or CSIRO soil


database

Field investigations at each site

March 2010

4A-4

Department of Transport and Main Roads


Road Drainage Manual

4A.12

Appendix 4A
Data Collection Checklist

Existing Infrastructure

Item

Check

Has the existence and location of the following been identified?

Adjacent dwellings / buildings - where flooding could be a sensitive


issue, include floor levels or other control points etc.
Existing drainage infrastructure (eg. bridges, culverts, subsoil drains,
pipelines, environmental / water quality devices)
Private / Public Utility Services (eg. communications, gas, power,
water, sewerage etc)

Industrial pipelines

Irrigation infrastructure

March 2010

4A-5

4A

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Chapter 5 5
Hydrology

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Chapter 5 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

March 2010

ii

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Table of Contents
5.1

Introduction.......................................................................................... 5-1

5.2

Rainfall ................................................................................................. 5-1

5.3

Rainfall - Runoff Relationship ............................................................ 5-2

5.4

Methods Available for Runoff Calculation......................................... 5-2

5.5

Rational Method................................................................................... 5-3

5.6

Catchment Area ................................................................................... 5-4

5.7

Time of Concentration ........................................................................ 5-4


5.7.1

General .................................................................................................. 5-4

5.7.2

Rural Catchments .................................................................................. 5-5

5.7.3

Urban Catchments ................................................................................. 5-6

5.8

Rainfall Intensity - Frequency - Duration......................................... 5-16

5.9

Runoff Coefficient ............................................................................. 5-16

5.10

5.9.1

Rural Catchments ................................................................................ 5-17

5.9.2

Urban Catchments ............................................................................... 5-17

5.9.3

Adjustment Factors .............................................................................. 5-19

Partial Area Effects ........................................................................ 5-21


5.10.1 Rural Catchments ................................................................................ 5-21
5.10.2 Urban Catchments ............................................................................... 5-22

5.11

Progressive Catchments ............................................................... 5-23

5.12

Previous Methodology................................................................... 5-25


5.12.1 Time of Concentration.......................................................................... 5-25
5.12.2 Runoff Coefficient................................................................................. 5-26

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

March 2010

iv

Chapter 5
Hydrology

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Chapter 5
Hydrology
5.1

Introduction

Calculation of the flood hydrograph and


peak flood discharge is an important part of
the design of departmental drainage
infrastructure. This needs to be considered
in almost all drainage designs carried out
for the department.
This aspect of the drainage assessment
process is called the hydrologic analysis.
Understanding the hydrological conditions
of the project site is vitally important.
These conditions include:

runoff from a catchment. This is normally


expressed as the peak flood discharge,
given as a volume per unit time, usually as
The
cubic metres per second (m3/s).
calculation can be based on an assessment
of rainfall or on recorded streamflow data.
For routine analysis of small catchments, it
is almost always based on analysis of
rainfall, and this is the procedure described
in this manual.
It is well understood that actual rainfall
events or storms are variable and dynamic
in nature. That is, rainfall;

rainfall (duration and intensity);

is not of uniform intensity over the


whole of the catchment; and

topography (slope and soil type);

may not cover the whole catchment.

land use;
vegetation coverage;
water flow paths;
areas of
storage;

water

inundation

and

water harvesting (farm dams);


volume of runoff generated.
Understanding these aspects ensures that
any drainage system design:
meets the drainage design criteria;
is economical;
is safe for all users;
protects investment in the road asset;
and
protects the environment from harm.
The department defines hydrology for road
drainage design as the calculation of flood

To apply this understanding in a design


context, that is, to calculate flood
hydrographs or the peak flood discharge
generated by any actual storm, is extremely
complicated and difficult.
The departments principal requirement for
routine design peak discharges of defined
average recurrence intervals (ARI). These
can be calculated by relatively simple
procedures that are introduced and
discussed in the relevant sections of this
chapter.

5.2

Rainfall

Rainfall and in particular rainfall intensity,


is needed as the basis for procedures to
calculate design floods, for a specific
geographical location, to be used for the
design of road drainage infrastructure.

March 2010

5-1

Department of Transport and Main Roads


Road Drainage Manual

Design rainfall intensities are published by


the Institution of Engineers Australia in
Australian Rainfall and Runoff, A Guide to
Flood Estimation (AR&R). This is the key
reference for almost all design rainfall
calculation studies throughout Australia,
and it would be extremely unusual for an
alternative method to be used, especially for
routine projects.

The design rainfall intensity data is


provided for all Average Recurrence
Intervals (ARI) up to 100 years and for
standard durations from 6 minutes to 3
days. The different ARIs are needed to
assess floods of different risk levels and the
range of durations allows for different
catchment response times to be considered.
The design rainfall intensities provided in
AR&R have been calculated by the Bureau
of Meteorology and have been based on
extensive rainfall data collected throughout
Australia.
AR&R provides a detailed procedure for
calculating design rainfall intensities, and
this procedure should be followed in all
cases. Tables of intensities for selected
ARIs can be developed either manually, by
using specialised software or obtained from
the Bureau of Meteorology website.
Determination and application of design
rainfall intensities is discussed further in
Section 5.7.

5.3

Rainfall - Runoff
Relationship

As rainfall hits the ground, runoff is


generated (after initial and continuing
losses) from the catchment. Therefore, a
relationship between the rainfall that hits
the ground and the runoff generated exists
for a given catchment and discharge point.

March 2010

5-2

Chapter 5
Hydrology

This relationship can be plotted as flood


discharge against time and the resulting
graph is called a hydrograph. Should the
discharge point (for example, culvert site)
or ARI change, the relationship will change.
An actual or real discharge hydrograph
plots the flows from actual rainfall events
and could have several peaks. The rainfall
pattern for particular storm events may be
complex and may vary from one event to
another so this hydrograph may also be
complex. However, for design purposes,
design flood hydrographs are usually
calculated using an idealised, theoretical
procedure and a synthetic hydrograph is
produced. This plot will usually have only
one peak for each ARI event. Refer Figure
5.3 for examples of actual and synthetic
hydrographs.
Hydrographs are useful tools particularly
when reviewing total flow volume and time
of flow for a catchment and selected ARI.
Also,
hydrographs
are
used
for
determination of time of submergence and
time of closure.
For the design of most road drainage
infrastructure, the two key points on a
hydrograph that are of most interest to
designers are the peak discharge generated
and the time it occurs.

5.4

Methods Available for


Runoff Calculation

There are several techniques available for


flood estimation in various sized
catchments and these procedures are
described in detail in Australian Rainfall
and Runoff, A Guide to Flood Estimation,
Vol 1 (IEAust. 2001).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

3000

2500

Actual or Recorded Hydrograph

Discharge - m3/s

2000

1500

1000

500

0
03 Mar

04 Mar

05 Mar

06 Mar

07 Mar

08 Mar

09 Mar

10 Mar

11 Mar

12 Mar

13 Mar

14 Mar

Time

700

600

Several ARI
events shown

Discharge - m /s

500

Synthetic Hydrograph

400

300

200

100

0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

5.0

Time - hours

Figure 5.3 - Actual and Synthetic Hydrographs

Methods and techniques for determining


flood discharges or runoff estimation in
larger catchments are usually associated
with major structures such as bridges and
floodways are not described in this manual.
While there are many methods for flood
estimation, the standard method of routine
runoff calculation used by the department
for small rural and urban catchments is the
Rational Method.
The Rational Method is a simple, statistical
method used to calculate peak discharge
from a catchment for a given ARI.

5.5

Rational Method

The Rational Method formula is:

Qy k C y I tc , y A
Where:
Qy = flow rate, Q (m3/s) for an ARI
of y years;
k = a conversion factor. k = 0.278
when A is km2 and 0.00278 when A is
hectares (ha);
Cy = runoff coefficient, C
(dimensionless) for an ARI of y
years;

March 2010

5-3

Department of Transport and Main Roads


Road Drainage Manual

Itc,y = average rainfall intensity, I


(mm/h) for design duration of tc and
ARI of y years; and

5.7

Time of Concentration

5.7.1

General

A = area of catchment (either


hectares or km2).

In the Rational Method, the time of


concentration tc for a catchment is defined
as:

The Rational Method has its limits and


these are detailed in Chapter 1 of this
manual.
Each of the key terms in the Rational
Method equation (C, I and A) will be
examined in the following sections.

5.6

Chapter 5
Hydrology

Catchment Area

The catchment or watershed area, A (in


hectares or km2), is typically determined
from topographical mapping, aerial
photographs used as stereo pairs or the
basis of photogrammetric contour plots,
aerial laser survey or by field survey.
Catchments
interpreted
from
aerial
photographs in heavily timbered or flat
country should be verified by ground
inspection or survey to ensure that major
errors do not result from misinterpretation.
In urban areas, the determination of
catchment boundaries can be difficult.
Natural boundaries are often affected by
roadworks, railway embankments or other
earthworks, building works, underground
piped stormwater networks and property
fences. In complex situations, catchment
boundaries should be verified by site
inspection and survey.
When using the Rational Method, rural
catchments are limited to 25 km2 and due to
complexity, urban catchments are limited to
1 km2.

(a) Time taken for water to flow from the


most remote point on the catchment to
the outlet or point of interest; or
(b) Time taken from the start of rainfall
until all of the catchment is
simultaneously contributing to flow at
the outlet or point of interest.
The significance of the time of
concentration is that peak outflow will
almost always result when the entire
catchment is contributing flow from rainfall
on the catchment. The most intense rainfall
that contributes to the outflow will be that
with a duration equal to the time of
concentration.
Therefore, tc is the duration used to select
the design rainfall intensity from the IFD
table generated in Section 5.8.
If the rainfall is more intense but of shorter
duration not all the catchment will
contribute to the peak runoff. If the rainfall
is of longer duration the average intensity
over that duration will be less and the peak
runoff will be less even though the entire
catchment contributes.
The time of concentration is generally made
up of three components:
1. Overland flow time across natural or
paved surfaces including retardance due
to pondage on the surface or behind
obstructions;
2. Time of flow in natural and artificial
channels; and
3. Time of flow in pipes.

March 2010

5-4

Department of Transport and Main Roads


Road Drainage Manual

The type of flow will vary throughout the


catchment, although once channelised,
overland flow conditions do not normally
recur. Overland flow to channel flow and
pipe flow back to channel flow can be
expected to occur. There may also be
overland or channel flow parallel with pipe
flow at full capacity. Several flow paths
may need to be examined to determine
which is the longest or most critical in
terms of design flows.
The procedure for calculating time of
concentration varies depending on whether
the catchment is urban or rural. Both
procedures are described later in this
section.
The minimum time of concentration to be
used in design for both rural and urban
catchments is 6 minutes. This practice is
inline with the recommendations within
AR&R.
In designing culverts for road crossings, the
time of concentration used should allow for
future development of the upstream
catchment. This development could be the
changing of land use due to farming or
urbanisation near or within a town or city.
Consider the following example illustrated
in Figure 5.7.1.

Chapter 5
Hydrology

If the time of concentration to point A is


calculated in the Existing Catchment, it
will be made up of:
a considerable length of overland and
channel flow; and
a short length of flow in pipes.
In the case with Possible Catchment
Development where the drainage system in
the catchment upstream of the road is
improved, the overland flow time will be
reduced and the time of concentration to
A also reduced. This will increase the
intensity of design rainfall and therefore
increase the amount of runoff generated.
Designers need to check the full range of
possible cases.

5.7.2

Rural Catchments

Typically, for rural catchments, the time of


concentration is made up from two
components:
1. Some overland flow at the top of the
catchment; and
2. Time of flow in the natural channel to
drainage structure site.
For all catchment sizes within the limits of
the Rational Method, the time of
concentration, is determined using the
Bransby-Williams formula. This formula
which includes overland flow and channel
flow conforms to the accepted practices in
AR&R (IEAust. 2001).
The Bransby-Williams formula is:

tc

Figure 5.7.1 - Hypothetical Catchment


Development

FL
A Se
0.1

0.2

Where:
tc = time of concentration (min);

March 2010

5-5

Department of Transport and Main Roads


Road Drainage Manual

F = a conversion factor. F = 58.5


when A is km2 and 92.7 when A is
hectares (ha);
L = length of mainstream (km) from
the outlet to the catchment divide;
A = area of catchment (either km2 or
hectares);
Se = equal area slope (m/km) as
defined in Figure 5.7.2
If the catchment has several possible flows
paths upstream of the site, each path will
have to be assessed to determine the path
with the longest time of concentration.

Equal Area
Slope

Natural Stream
Profile

Chapter 5
Hydrology

following has been extracted from QUDM


with some minor modification.
It should be noted that the time of
concentration as used in the Rational
Method is not the same as the critical storm
duration or time to peak as determined from
runoff-routing models.
It is therefore
inappropriate to adopt the critical storm
duration determined from a runoff-routing
model and apply it as the time of
concentration for a Rational Method
analysis.
In determining the time of concentration,
the designer should adopt the appropriate
catchment conditions in accordance with
the required analysis.
Ultimate flow
conditions should be based on a fully
developed catchment in accordance with
the allowable land use shown in the
relevant local authority plans, or as directed
by the local authority.
To apply the Rational Method in an
appropriate and consistent manner, four
different methodologies for determination
of the time of concentration are presented
below for different types of drainage
catchments. Those catchment types being:
(a) Predominantly piped or channelised
urban catchments less than 1 km2 with
the top of the catchment being
urbanised.

Figure 5.7.2 - Derivation of the Equal


Area Slope of Main Stream.
Derived from AR&R (IE Aust 1987a)

5.7.3

Urban Catchments

The department acknowledges that the


Queensland Urban Drainage Manual
(QUDM) published by the Department of
Natural Resources and Water in 2008,
represents the latest thinking and practice
for urban catchment hydrology.
The

March 2010

5-6

(b) Predominantly piped or channelised


urban catchments less than 1 km2 with
the top of the catchment being
bushland or a grassed park.
(c) Bushland catchments too small to
allow the formation of a creek with
defined bed and banks.
(d) Urban creeks with a catchment area
less than 1 km2.

Department of Transport and Main Roads


Road Drainage Manual

(a) Predominantly piped or channelised


urban catchments less than 100ha with the
top of the catchment being urbanised.
Components of time of concentration:
Standard Inlet Time (preferred) from
Section 5.7.3.1. In cases where use
of a standard inlet time is not
considered appropriate, calculate
travel time from roof to kerb using
Section 5.7.3.2. The standard inlet
time includes the travel time along a
typical length of kerb / channel from
near the top of the catchment to the
first pipe or channel inlet. If the
actual length of kerb / channel travel
is unusually long, then an additional
travel time must be added to the
standard inlet time (next dot point
below). If a gully / field inlet does
not exist near the top of catchment,
then use Sections 5.7.3.2 and/or
5.7.3.3 to determine the initial travel
time to the start of the kerb / channel,
then add the travel time along the
kerb / channel.
Kerb flow time from Section 5.7.3.4
only if the length of kerb exceeds that
which would normally exist at the
top of a catchment.
Pipe flow time using actual flow
velocities determined from a pipe
network analysis or Mannings
Equation, refer Section 5.7.3.5.
Alternatively, if the pipe flow time is
not critical, an average pipe flow
velocity of 2 m/s and 3 m/s may be
adopted for low gradient and medium
to
steep
gradient
pipelines
respectively.
Creek and/or channel flow time using
actual flow velocity determined from
numerical modelling or Mannings

Chapter 5
Hydrology

Equation, refer Section 5.7.3.6.


Alternatively, if the expected travel
time in the creek is not critical, an
average flow velocity of 1.5 m/s may
be adopted (not applicable to
constructed channels).
(b) Predominantly piped or channelised
urban catchments less than 100ha with the
top of the catchment being bushland or a
grassed park.
Components of time of concentration:
Estimate the length of sheet runoff
at top of catchment using Table
5.7.3.3(a) or field observations, then
estimate the sheet flow travel time as
per Section 5.7.3.3.
Determine the remaining distance of
assumed concentrated overland flow
from the end of the sheet runoff to
the nearest kerb, pipe inlet, open
channel or creek. Then determine the
travel time for this concentrated
overland flow based on the calculated
flow velocity.
Kerb flow time as per Section
5.7.3.4.
Pipe flow time using actual flow
velocities determined from a pipe
network analysis or Mannings
Equation, refer Section 5.7.3.5.
Alternatively, if the pipe flow time is
not critical, an average pipe flow
velocity of 2 m/s and 3 m/s may be
adopted for low gradient and medium
to
steep
gradient
pipelines
respectively.
Creek and/or channel flow time using
actual flow velocity determined from
numerical modelling or Mannings
Equation, refer Section 5.7.3.6.
Alternatively, if the expected travel

March 2010

5-7

Department of Transport and Main Roads


Road Drainage Manual

time in the creek is not critical, an


average flow velocity of 1.5 m/s may
be adopted (not applicable to
constructed channels).
(c) Bushland catchments too small to
allow the formation of a creek with
defined bed and banks.
Time of concentration is determined as for
(b) above.
(d) Urban creeks with a catchment area
less than 100 ha.

Time of concentration for an urban


catchment containing a watercourse with
defined bed and banks may be determined
as for rural catchments but only if the
following conditions apply:
channel
storage
along
the
watercourse, for the catchment
condition being analysed, is not
significantly reduced from the natural
(i.e. pre-urbanisation) conditions; and
less than 20% of the catchment
drains to a pipe network.

Chapter 5
Hydrology

If the above two conditions do not apply,


then the time of concentration should be
based on the procedures outlined in (a) or
(b) above as appropriate for the catchment
conditions.
5.7.3.1 Standard Inlet Time
Use of standard inlet times for developed
catchments is recommended because of the
uncertainty related to the calculation of time
of overland flow. The standard inlet time is
defined as the travel time from the top of
the catchment to a location where the first
gully or field inlet would normally be
expected as depicted in Figure 5.7.3.1.
Recommended standard inlet times are
presented in Table 5.7.3.1. These inlet
times are considered appropriate for
traditional (i.e. non Water Sensitive Urban
Design (WSUD)) low density residential
areas where the top of the catchment is low
density residential, but not a park or
bushland.

Figure 5.7.3.1 - Application of Standard Inlet Time

March 2010

5-8

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Table 5.7.3.1 - Recommended Standard Inlet Times

Inlet Time
(minutes)

Location
Road surfaces and paved areas

Urban residential areas where average slope of land at top of catchment is


greater than 15%.

Urban residential areas where average slope of land at top of catchment is


greater than 10% and up to 15%.

Urban residential areas where average slope of land at top of catchment is


greater than 6%and up to 10%.

10

Urban residential areas where average slope of land at top of catchment is


greater than 3% and up to 6%.

13

Urban residential areas where average slope of land at top of catchment is


15
up to 3%.
Note: The average slopes referred to are the slopes along the predominant flow path for the catchment in
its developed state.

If the top of the catchment consists of high


density residential, then the local
government should be consulted for inlet
times appropriate for the catchment. In
such cases it is recommended that the
standard inlet time should not exceed 10
minutes unless demonstrated otherwise by
the designer.
If the hydrologic analysis is being
performed on a development located at the
top of the catchment, then use of a standard
inlet time will usually not be appropriate
because these inlet times are likely to be
significantly greater than the actual travel
time.
If the first gully or field inlet is located
further down the catchment slope than
would normally be expected, then the
standard inlet time shall only account for
the travel time down to the location where
the first gully or field inlet would normally
have been located.
If the urban drainage system does not
incorporate pipe drainage (i.e. no gully or
field inlet exists) then the standard inlet

time shall extend down the catchment to a


location where a gully inlet would normally
be located in a traditional kerb-&-channel
drainage system.
A standard inlet time should not be adopted
in sub-catchments where detailed overland
flow and kerb / channel flow calculations
are justified.
A local government may require that the
use of standard inlet times shall not apply
within their area and may recommend
designers to use alternative methods.
In certain circumstances the use of standard
inlet times may result in times of
concentration unacceptably short for the
catchment under consideration, such as
airports, or large flat car parks. In these
cases the designer should utilise Friends
Equation to determine the time of initial
overland flow (refer to Section 5.7.3.3
below). Inlet times calculated by these
methods should only be adopted for design
if the sheet flow length criteria discussed in
Section 5.7.3.3 are met and if due
consideration is given to the type and
March 2010

5-9

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

continuity of the surface where overland


flow is occurring.
Notwithstanding the above, it is
recommended that a maximum inlet time of
20 minutes be adopted for urban and
residential catchments, including playing
fields and park areas.
5.7.3.2 Roof
to
Main
Connection

System

In cases where use of a standard inlet time


is not considered appropriate, the roof to
main system flow travel times as shown in
Table 5.7.3.2 are recommended.

5.7.3.3 Overland Flow


Overland flow at the top of a catchment will
initially travel as sheet flow, after which it
will move down the catchment as minor
concentrated flow. Travel times for the
sheet flow and concentrated flow
components need to be determined
separately.

The sheet flow travel time is defined as the


travel time from the top of a catchment to
the point where stormwater runoff begins to
concentrate against fences, walls, gardens,
or is intercepted by a minor channel, gully
or piped drainage. This concentration of
flow may also occur in the middle of
vegetated areas as the stormwater
concentrates in minor drainage depressions.
The time required for water to flow over a
homogeneous surface such as lawns and
gardens is a function of the surface
roughness and slope. There are a number
of methods available for the determination
of sheet flow travel times and a local
government may direct which of these
methods shall be applied.
Two such
methods are presented in this section.
Irrespective of which method of calculation
is adopted, it is the designers responsibility
to determine the effective length of this
sheet flow.

Table 5.7.3.2 - Recommended Roof Drainage System Travel Times

Development Category
Rural Residential, Residential Low Density
For the roof, downpipes and pipe connection system from the building to
the kerb and channel or a rear-of-allotment drainage system (Figure
5.8.3.2(a)).
Residential Medium and High Density, Commercial, Industrial and
Central Business
For the roof and downpipe collection pipe to the connection point to the
internal allotment drainage system abutting the building (Figure 5.8.3.2(b)).

Time to point
A (minutes)

Note: The flow time from point A (Figures 5.7.3.2 (a) & (b)) through the internal allotment pipe system to
the kerb and channel, street underground system or rear of allotment system for the more intense
developments noted should be calculated separately.

March 2010

5-10

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Typical Roof Drainage Systems


Note: Point A is referred to in Table 5.7.3.2
Figure 5.7.3.2(a) - Residential

In urban areas, the length of overland sheet


flow will typically be 20 to 50 metres, with
50 metres being the recommended
maximum. In rural residential areas the
length of overland sheet flow should be
limited to 200 m (Argue 1986), however
the actual length is typically between 50
and 200 m where after the flow will be
concentrated in small rills, channels, or
tracks.
Design steps:
To determine the overland flow travel time
the following steps should be applied:
Where
practical,
inspect
the
catchment to determine the length of
initial overland sheet flow, or for new
developments measure the length of
overland flow from the design plans.
Where it is not practical to inspect
the catchment, determine the likely
length of overland sheet flow based
on Table 5.7.3.3(a).

Figure 5.7.3.2(b) - Industrial

Determine the sheet flow travel


time using Friends Equation see
discussion below.
Determine or measure the remaining
distance of assumed concentrated
overland flow from the end of the
adopted sheet flow to the nearest
kerb, channel, or pipe inlet.
Determine the concentrated flow
travel time using either Mannings
Equation (refer to 5.7.3.6 or Section
8.4.3) or for preliminary design
purposes, Figure 5.7.3.4.
Friend's Equation /
Overland Sheet Flow

Nomograph

for

The formula shown below and attributed to


Friend (1954) may be used for the
determination of overland sheet flow times.
This was derived from previous work in the
form of a nomograph, as shown in Figure
5.7.3.3, for shallow sheet flow over a plane
surface.
March 2010

5-11

Department of Transport and Main Roads


Road Drainage Manual

Friends Equation is:

107 nL
S

Chapter 5
Hydrology

Surface roughness values for Hortons n are


similar but not identical to Mannings n
values. Refer Table 5.7.3.3(b) for values
for Hortons n.

Where:
t = overland sheet flow travel time
(mins);
L = overland sheet path length (m);
n = Hortons roughness value for the
surface;
S = slope of surface (%).
Table 5.7.3.3(a) - Recommended Maximum Length of Overland Sheet Flow

5
Surface Condition

Assumed Maximum
Flow Length (m)

Steep (say >10%) grassland (Hortons n = 0.045)

20

Steep (say >10%) bushland (Hortons n = 0.035)

50

Medium gradient (approx. 5%) bushland or grassland

100

Flat (01%) bushland or grassland

200

Source: AR&R (IEAust 1977)

Figure 5.7.3.3 - Overland Sheet Flow Times Shallow Sheet Flow Only

March 2010

5-12

Department of Transport and Main Roads


Road Drainage Manual

Table 5.7.3.3(b) - Hortons Roughness


Values
Surface Condition

Paved surface

0.015

Bare soil surface

0.0275

Poorly grassed
surface

0.035

Average grassed
surface

0.045

Densely grassed
surface

0.060

Source: Soil Conservation Measures Design


Manual for Queensland (NR&M 2004)

Chapter 5
Hydrology

overall time of concentration, then an


average pipe velocity of 2 m/s and 3 m/s
may be adopted for low gradient and
medium to steep gradient pipelines
respectively.
5.7.3.6 Channel Flow
The time stormwater takes to flow along an
open channel may be determined by
dividing the length of the channel by the
average velocity of the flow.
The average velocity of the flow is
calculated
using
the
hydraulic
characteristics of the open channel.
Mannings Equation is suitable for this
purpose:

5.7.3.4 Kerb Flow


Time of flow in kerb and channel should be
determined by dividing the length of kerb
and channel flow by the average velocity of
the flow.

S
n

Where:
V = velocity in pipe (m/s);

The average velocity of the flow may be


determined in either of two ways:

R = hydraulic radius of pipe flowing


full (m);

1. Izzards Equation - refer to Section


11.2.3; or

S = slope of energy line or hydraulic


gradient (m/m);

2. For preliminary design purposes, use


Figure 5.7.3.4.

n = Mannings roughness coefficient


for pipe.

5.7.3.5 Pipe Flow


Wherever practical, pipe travel times should
be based on calculated pipe velocities either
using a Pipe Flow Chart (refer Appendix
5A), uniform flow calculations using the
Mannings Equation, or results from a
calibrated numerical drainage model.
For preliminary design purposes, pipe flow
travel time can be estimated using Figure
5.7.3.4 or alternatively, if the travel time
within the pipe is small compared to the

Note: The slope of the energy line is often


difficult to determine, therefore use a
representative slope of the channel (So) in
the vicinity of the site to estimate the slope
of the energy line, that is, So S.
Where an open channel has varying
roughness or depth across its width it may
be necessary to split the channel into
sections and determine the average flow
velocity in each section, to determine the
overall flow time.

March 2010

5-13

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

5
Figure 5.7.3.4 - Kerb and Channel Flow Time using Mannings Equation

Grass swales
Flow travel times along grassed swales can
vary significantly depending on flow depth
and vegetation. Swale roughness, n should
be determined from the vegetation
retardance charts presented in Appendix
8A.
5.7.3.7 Estimate of Kerb, Pipe and
Channel Flow Time
For checking or preliminary design
purposes, an overall flow time can be
determined from Figure 5.7.3.7. The chart
may be used directly to determine
approximate travel times along a range of
rigid channel types and, with the
application of multiplier for a range of
loose-boundary channel forms.

March 2010

5-14

Flows can reach drains / channels via roof


to gutter conduits, overland flow paths or
along gutters. In many cases, flows travel
along two or three consecutive paths, and it
is necessary to calculate a total travel time.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Figure 5.7.3.7 - Flow Travel Time in Pipes and Channels


Source: Argue 1986

NOTES:
1. Flow travel time (approximate) may be obtained directly from this chart for:
kerb-and-gutter channels
stormwater pipes
allotment channels of all types (surface and underground)
drainage easement channels (surface and underground)
2. Multiplier , should be applied to values obtained from the chart as per:
grassed swales, well maintained and without driveway crossings - = 4
blade-cut earth table drains, well maintained and no driveway crossings - = 2
natural channels - = 3

March 2010

5-15

Department of Transport and Main Roads


Road Drainage Manual

5.8

Rainfall Intensity Frequency - Duration

The design rainfall intensity, Itc,y is defined


as a rainfall depth per unit time and is
defined for a frequency and duration. The
normal units for rainfall intensity, I, are
millimetres per hour (mm/h).

Rainfall intensity varies with location and


topography as well as duration and
frequency, and this should be considered
when using the Rainfall Intensity Frequency - Duration (IFD) calculations
performed using methods described in
AR&R.
For calculation of design IFD rainfall
tables, departmental staff can use an inhouse computer program called RAIN.
Other similar programs are available. IFD
tables can also be obtained from the Bureau
of Meteorology website.
The use of an IFD program requires the
input of nine parameters determined from
maps contained in Australian Rainfall and
Runoff, A Guide to Flood Estimation, Vol 2
(IEAust. 1987b). These parameters are:

50

50

50

I1 (2 year, 1 hour log-normal rainfall


intensity);

I12 (2 year, 12 hour log-normal


rainfall intensity);
I72 (2 year, 72 hour log-normal
rainfall intensity);
I1 (50 year, 1 hour log-normal
rainfall intensity);

I12 (50 year, 12 hour log-normal


rainfall intensity);
I72 (50 year, 72 hour log-normal
rainfall intensity);

G - Skewness Factor;
F2 - Geographical Factor; and
March 2010

5-16

Chapter 5
Hydrology

F50 - Geographical Factor.


A tabulation of IFD values for durations
from 5 minutes to 72 hours and average
recurrence intervals from 1 to 100 years is a
standard output from the RAIN program
but values for non standard times are also
readily obtained. Note that in accordance
with AR&R, the minimum standard
duration is 6 minutes.
An important observation of an IFD table is
that the longer the duration of rainfall, the
lower the intensity of the storm.
AR&R also details the use of temporal
patterns and areal reduction factors.
Temporal patterns are applied to large
catchments to allow for rainfall intensity
variations across the catchment over the
storm duration. Areal reduction factors are
applied to point rainfall intensities to
address the issue that the application of
point rainfall values over a large catchment
is unrealistic as such intensities are unlikely
to be maintained across the entire area.
These two aspects are not considered for
use within the Rational Method as follows:
(a) Temporal patterns are not needed since
the Rational Method only uses a
uniform rainfall intensity based on the
time of concentration; and
(b) Because the Rational Method only
applies for small catchments up to an
area of 25 km2, the areal reduction
factor is approximately equal to 1.0.

5.9

Runoff Coefficient

The runoff coefficient C is a statistical


composite of several aspects including the
effects of rainfall intensity, catchment
characteristics, infiltration (and other
losses) and channel storage. It should not
be confused with the volumetric runoff

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

coefficient which is the ratio of total runoff


to total rainfall.

5.9.1

Table 5.9.1(b) - Adjustment Factors for


Runoff Coefficients for Other Average
Recurrence Intervals

Rural Catchments

The runoff coefficient includes effects of


catchment characteristics, infiltration and
other losses as well as rainfall intensity.
The runoff coefficient Cy, as used in the
Rational Method, is a function of the design
ARI (y in years) and depends on many
features of the catchment area including:

Average
Recurrence
Interval
(years)

Rural
Coefficient

Urban
Coefficient

0.8 C50

0.80 C10

0.8 C50

0.85 C10

rainfall intensity;

0.8 C50

0.95 C10

relief or slope of catchment;

10

0.8 C50

1.00 C10

20

0.9 C50

1.05 C10

50

1.0 C50

1.15 C10

100

1.05 C50

1.20 C10

storage
or
other
characteristics; and

detention

ground characteristics such as


vegetation cover, soil type, and
impervious areas.
The runoff coefficient for the ARI 50 year
event (C50) is determined using Table
5.9.1(a). To determine the value of Cy for
other ARIs, the C50 value is modified using
the factors from Table 5.9.1(b).
It should be noted that this method can give
a C50 value greater than 1.0. This can occur
when rainfall intensity exceeds 120 mm/hr
and the remaining characteristics are at
maximum values (possible in small, steep
catchments). In this instance, C50 should be
rounded down to 1.0.

5.9.2

Urban Catchments

Again, the department acknowledges that


QUDM (NR&W 2008) represents the latest
thinking and practice for urban catchment
hydrology. The following discussion has
been extracted from QUDM with some
minor modification.

Notes:
1. C50 determined for rural catchments using
Table 5.9.1(a)
2. C10 determined using method described in
Section 5.9.2
3. Where runoff coefficients calculated using
the above table exceed 1.00, they should be
arbitrarily set to 1.00

The runoff coefficient must account for the


future development of the catchment as
depicted in the planning scheme or zoning
maps for the relevant local government, but
should not be less than the value
determined for the catchment under existing
conditions.
The runoff coefficient is calculated in
accordance with the method summarised in
the following steps:

March 2010

5-17

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Table 5.9.1(a) - Estimation of the Runoff Coefficient for Rural Catchments

Characteristic

Runoff producing values (in brackets) as % in calculation of C for a 50


year average recurrence interval event

Rainfall
Intensity
Catchment
Relief
Catchment
Storage

Ground
Characteristics

(C ) = 0.3 I50 + 4
Very steep slopes >
15%
(10)
Well defined water
courses, negligible
storage.
(10)
Grazing land and
open forest

Hilly to steep slopes 4 - Flat to rolling slopes <


15%
4%
(5)
(0)
Overland Flow is
Poorly defined water
significant, some
courses, large flood
floodplain storage
plain storage capacity
(5)
(0)
Dense
Heath and sand
Agricultural land
Vegetation and
dunes
rainforest
(30)
(20)
(10)

(40)
Notes:
Catchment storage is defined as; a catchments ability to detain or temporarily hold water within
a streams adjacent floodplain. Water will slowly drain after flood water recedes.
Example:
Determine C50 for a Rainfall Intensity of 40 mm/h over a catchment with the following
characteristics:
Catchment Relief Hilly with average slopes 4-8%;
Catchment Storage - Well defined system of small watercourses with little storage capacity;
Ground Characteristics - Open forest.

C 50

16 5 10 40
0.71
100

STEP 1 Determine the fraction impervious


fi for the catchment under study from Table
5.9.2(a).
STEP 2 Determine the 1 hour rainfall
intensity 1I10 for the ARI 10 year event at
the locality. Refer to Section 5.8.
STEP 3 Determine the 10 year C value
from Tables 5.9.2 (b) & (c).
STEP 4 Determine the Urban Coefficient
for the required ARI from Table 5.9.1(b), if
required.
STEP 5 Multiply the C10 as per Urban
Coefficient (Step 4) to determine the runoff
coefficient for the design storm Cy.

March 2010

5-18

In certain circumstances the resulting value


In
of Cy will be greater than 1.0.
accordance with the recommendations of
AR&R, a limiting value of Cy = 1.0 should
be adopted for urban areas.
There is little evidence to support an
allowance for either slope or soil type in
fully developed (non WSUD) urban areas.
If there are significant local effects, and
reliable data is available, then adjustments
for soil type may be incorporated within the
calculations at the discretion of the
designer in consultation with the relevant
local authority.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Table 5.9.2(a) - Fraction Impervious vs.


Development Category

Development Category

Fraction
Impervious
(fi)

Central Business

1.00

Commercial, Local
Business, Neighbouring
Facilities, Service
Industry, General
Industry, Home Industry

0.90

Significant Paved Areas


eg. Roads and Carparks

0.90

Urban Residential - High


Density

0.70 to 0.90

Urban Residential - Low


Density (Including
Roads)

0.45 to 0.85

Urban Residential - Low


Density (Excluding
Roads)

0.40 to 0.75

Rural Residential
Open Space & Parks etc.

The relationships shown in Book 8 of


AR&R (IEAust 2001) and adopted in this
manual apply to areas that are essentially
homogeneous, or where the pervious and
impervious portions are so intermixed that
an average is appropriate. In cases where
separable portions of a catchment are
significantly different, they should be
divided into sub-catchments and different
values of C applied.
Notwithstanding the above notes and
limitations, it is the responsibility of the
designer to ensure each sub-catchment flow
is determined using a suitable coefficient of
discharge. The local government may set
specific C values to be used within their
area.

5.9.3

0.1 to 0.2
0

Notes:
1. The designer should determine the actual
fraction impervious for each development.
Local governments may specify default
values.
2. Typically for Urban Residential High
Density developments:
- townhouse type development fi =
0.70
- multi-unit dwellings>20 dwellings per
ha
- highrise residential development fi =
0.90.
3. In Urban Residential Low Density areas fi
may vary depending upon road width,
allotment size, house size and extent of
paths, driveways etc.

Adjustment Factors

In making the decision on whether or not to


allow for future development, the disruption
to traffic when the additional waterways are
constructed in the future must be
considered. Other considerations include
the
requirements
by
some
local
governments to not allow any increase in
water discharging into drainage structures
in the road corridor from development of an
upstream catchment.
Detention basins are therefore specified in
the design of the development, particularly
in small urbanised catchments. In this case
there is no need to consider the effect of
development.
In the case where the detention basin only
moderates the runoff from the development,
then the parameters of the detention basin
design need to be considered in the
discharge calculations at the departmental
drainage structure.

March 2010

5-19

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Table 5.9.2(b) - C10 Values

Intensity
(mm/h)
1
I10

Fraction Impervious fi
0.20

0.40

0.60

0.80

0.90

1.00

39-44

0.44

0.55

0.67

0.78

0.84

0.90

45-49

0.49

0.60

0.70

0.80

0.85

0.90

0.55

0.64

0.72

0.81

0.86

0.90

0.60

0.68

0.75

0.83

0.86

0.90

0.65

0.72

0.78

0.84

0.87

0.90

65-69

0.71

0.76

0.80

0.85

0.88

0.90

70-90

0.74

0.78

0.82

0.86

0.88

0.90

Refer to Table 5.7.2(c)

0.00

50-54
55-59
60-64

5
1

I10
C10

= One hour rainfall intensity for a ARI 10 year event


= Runoff Coefficient for a ARI 10 year event

Table 5.9.2(c) - C10 Values for 0% Fraction Impervious

Land
Description

Dense bushland

Medium density bush,


or Good grass cover,
or High density
pasture, or Zero tillage
cropping

Light cover bushland, or


Poor grass cover, or
Low density pasture, or
Low cover bare fallows

Soil permeability

Soil Permeability

Soil Permeability

Intensity
(mm/h)
1
I10

High

Med

Low

High

Med

Low

High

Med

Low

39-44

0.08

0.24

0.32

0.16

0.32

0.40

0.24

0.40

0.48

45-49

0.10

0.29

0.39

0.20

0.39

0.49

0.29

0.49

0.59

50-54

0.12

0.35

0.46

0.23

0.46

0.58

0.35

0.58

0.69

55-59

0.13

0.40

0.53

0.27

0.53

0.66

0.40

0.66

0.70

60-64

0.15

0.44

0.59

0.30

0.59

0.70

0.44

0.70

0.70

65-69

0.17

0.50

0.66

0.33

0.66

0.70

0.50

0.70

0.70

70-90

0.18

0.53

0.70

0.35

0.70

0.70

0.53

0.70

0.70

Derived from Qld Department of Natural Resources & Mines (2005)

March 2010

5-20

Department of Transport and Main Roads


Road Drainage Manual

5.10

Partial Area Effects

In general, the appropriate time of


concentration (tc) for calculation of the flow
at any point is the longest time of travel to
that point. However, in some situations, the
maximum flow may occur when only part
of the upstream catchment is contributing.
Thus the product of a lesser C.A and a
higher tIy (resulting from a lower tc) may
produce a greater peak discharge than that
if the whole upstream catchment is
considered. This is known as the Partial
Area Effect.
Usually the above effect results from the
existence of a sub-catchment of relatively
small C.A but a considerably longer than
average tc. This can result from differences
within a catchment of surface slope, or from
catchment shape.
The onus is on the designer to be aware of
the possibility of the Partial Area Effect and
to check as necessary to ensure that an
appropriate peak discharge is obtained.

Chapter 5
Hydrology

areas are presented in the following


sections.
One method is for rural
catchments and the other for urban
catchments. It is recommended that the
hydrologic assessment of catchments with
unusual or widely varying surface features
should be undertaken by Hydraulics
Section, Engineering & Technology
Division
or
suitably
pre-qualified
consultant using an appropriate numerical
runoff-routing model.

5.10.1 Rural Catchments


The occurrence of the Partial Area Effect in
the rural environment is not common, but
designers should look for catchments that
display the characteristics that may allow
Partial Area Effect to occur.
These
catchments need to be checked to ensure
that the peak discharge for the catchment is
correctly determined.
Figure 5.10.1 shows two examples of rural
catchments that may experience Partial
Area Effect.

Two generally accepted procedures for use


with the Rational Method for the
calculation of peak flow rates from partial

Figure 5.10.1 - Examples of Rural Catchments that may be subject to Partial Area Effects

March 2010

5-21

Department of Transport and Main Roads


Road Drainage Manual

In catchment (I), the whole catchment is


assessed with all parameters for the time of
concentration and runoff calculations
determined normally. The catchment is
then divided, where the catchment changes
from wide to narrow, into two portions.
The lower portion is then assessed as if it
was the whole catchment, that is, A will be
smaller, Se will be based on the channel bed
from the catchment outlet to the dividing
line between the portions and so on. In this
case tc will be shorter and therefore Iy will
be higher. The higher discharge of the two
assessments is deemed to be the peak
discharge.

In catchment (II), the whole catchment is


again assessed with all parameters for the
time of concentration and runoff
calculations determined normally.
The
catchment is then divided, where the
catchment slope changes from steep to flat.
The lower portion is then assessed as if it
was the whole catchment. In this case, Se
will be much higher and will have a bigger
impact on reducing tc.
Intensity will
therefore be higher. Again, the higher
discharge of the two assessments is deemed
to be the peak discharge.

5.10.2 Urban Catchments


The following discussion has been
extracted from QUDM (NR&W 2008) with
some minor modification.
The occurrence of Partial Area Effect in the
urban environment is much more common
than in the rural environment.
Figure 5.10.2 shows various examples of
urban catchments that may experience
Partial Area Effect.

March 2010

5-22

Chapter 5
Hydrology

A simplified procedure is given in Argue


(1986) based upon a comparison between
the full area discharge and the partial area
peak discharge for the time of concentration
of the impervious areas of the critical subcatchment. Care must be exercised as this
procedure can underestimate the peak
discharge.
The method involves the use of a time of
concentration ti corresponding to the flow
travel time from the most remote, directly
connected, impervious area of the
catchment to the point under consideration.
Thus, the calculated peak discharge is that
from the impervious portion of the
catchment plus that from the pervious part
of the catchment which has begun to
contribute up to time ti since the storm
began.
Therefore:

CA C i Ai i C p A p
tc

Where
C = overall coefficient of runoff with
Ci and Cp being the coefficients for
the impervious and pervious areas
respectively.
A = overall area with Ai and Ap being
the impervious and pervious areas
respectively (ha).
ti = time of concentration from
impervious area (min).
tc = time of concentration for the
catchment (min).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

Figure 5.10.2 - Examples of Urban Catchments that may be subject to Partial Area Effects
Source: QUDM (NR&W 2008)

5.11

Progressive
Catchments

A situation that often occurs in rural


environments is where the one stream
crosses the road several times.
The Rational Method can only estimate the
runoff at a point, usually the outlet of the
catchment, the site for a culvert. Therefore,
disregarding any upstream crossing, the
variables A and C (Q=k.C.I.A) must
describe the whole upstream catchment and
the variables Se / tc must be based on the
flow path from the site to the top of the
catchment. If there are several flow paths
or streams, time of concentration
calculations will need to be undertaken on
each path to determine the critical duration.
Furthermore, this approach also relies on
components of C (slope of catchment,
storage and ground characteristics) being
similar for all catchments.

For example, with reference to Figure 5.11,


the peak discharge at Point 5 would be
estimated from assessing Catchment C
normally. To estimate the peak discharge at
Point 3, variables A and C must cover both
catchments B and C while variables Se / tc
would be based on the critical duration
determined from paths [3, 4, 10], [3, 4, 5, 6,
7] or [3, 4, 5, 6, 8]. To estimate the peak
discharge at Point 1, variables A and C need
to cover catchments A, B and C and
variables Se / tc would be based on the
critical duration determined from paths [1,
2, 9], [1, 2, 3, 4, 10], [1, 2, 3, 4, 5, 6, 7] or
[1, 2, 3, 4, 5, 6, 8].
Technically, any upstream crossing could
act as a detention device, that is, a device
that reduces the peak flow but lengthens the
time flow occurs at that point, though this
effect would usually be small. This would
have an impact on the flow, at some
downstream point, as determined by using
the Rational Method as explained in this
section. The impact would typically be a

March 2010

5-23

Department of Transport and Main Roads


Road Drainage Manual

Chapter 5
Hydrology

reduction in the peak discharge. Therefore


the approach adopted here is considered
conservative.
If the land use or components of C vary
between the catchments or if a more
accurate estimate of runoff is required, then
use of an appropriate numerical runoffrouting model is needed and assistance
from Hydraulics Section, Engineering &
Technology Division or suitably prequalified consultant is required.

Should the situation occur in an urban


environment where the one stream crosses
the road several times then specialist
assistance from Hydraulics Section,
Engineering & Technology Division or
suitably pre-qualified consultant is required.

7
9

Catchment C
8

Catchment A

Road
5
3

Catchment B

10

Figure 5.11 - Progressive Catchments

March 2010

5-24

Department of Transport and Main Roads


Road Drainage Manual

5.12

Previous Methodology

On 13 September 2007, the department


(then Main Roads) issued a Planners and
Designers Instruction which made changes
to drainage design methodology. Key
changes made applied to hydrology
calculation of time of concentration and the
determination of runoff coefficient. The
superseded methods are detailed below for
historical purposes as future project work
may require to back calculate the discharge
used in previous projects.

5.12.1 Time of Concentration


Previously, the department had adopted two
methods of calculating the Time of
Concentration depending on the area of the
catchment.
For rural catchments greater than 5 km2, the
Modified Friend Formula was used to
estimate tc.
The Modified Friend Formula is:

tc

8.5 L
Ch A 0.1 S e

0.4

The variables L, A and Se were reasonably


straight forward to determine. Chezys
coefficient (Ch) was determined using the
following formula:

Ch

R 0.166
n

The variable R (hydraulic radius) was


determined using a simple but highly
subjective formula (R=0.75Rs or R=0.65Rs,
where Rs is the hydraulic radius at an
initially
assumed
flood
level).
Interpretation and assumptions were
required and if the assumed flood level used
differed from the flood level determined
after all hydrologic and hydraulic (stream

Chapter 5
Hydrology

analysis) calculations were made, then a


new assumed flood level was required and
the whole process repeated. Also, the
variable n was an average Mannings
roughness coefficient for the entire main
stream. This was highly interpretive
(assumptions generally made) and difficult
to determine.
Due to the assumptions and interpretations
made, it was possible for a wide range of
results and it was difficult to replicate
results between designers.
For rural catchments less than 5 km2, the
Stream Velocity method was used.
The Stream Velocity formula is:

tc

L
60 V

In this formula, V was highly interpretive


and difficult to determine. Therefore a
wide range of results was possible.
The use of these two methods was
complicated
and
time
consuming.
Assumptions and interpretations were
subjective and open to debate. There also
existed a discontinuity between methods
when assessing catchments around the 5
km2 size.
Despite both methods being published
within
the Section
Main
Roads
Department Rational Method Book 4 of
AR&R (IEAust. 2001), neither method
conformed to the recommended practices as
described by AR&R. In future releases of
AR&R, these methods should be removed.
The Bransby-Williams formula was
selected to replace the above methods to
simplify the design procedure and to
improve consistency of results. Test work
showed no significant or consistent
difference in results (estimates of tc)
between the methods. Also, the basis of the

March 2010

5-25

Department of Transport and Main Roads


Road Drainage Manual

Bransby-Williams formula conformed to


the accepted practices in AR&R.

5.12.2 Runoff Coefficient


The previous method to determine the
Runoff Coefficient for the ARI 50 year
(C50) was by using Table 3.5 as shown in
Figure 5.12.2. This table was somewhat
difficult to interpret and also contained a

Chapter 5
Hydrology

discontinuity within the component for


Rainfall Intensity. The table was reviewed
and modified to remove the discontinuity
and to improve interpretation / selection of
appropriate values for the other catchment
characteristics. The revised table is Table
5.9.1(a) within this chapter.

Figure 5.12.2 - Table 3.5 from previous MR Road Drainage Design Manual

March 2010

5-26

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

Appendix 5A
Pipe Flow Charts

Charts in this appendix have been reproduced


with permission from Rocla.

March 2010

5A

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

Appendix 5A Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

5A

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

DISCHARGE IN CUBIC METRES PER SECOND

5A

DISCHARGE IN LITRES PER SECOND

Figure 5A.1. Discharge and Velocity Graph


(Mannings n = 0.010)
March 2010

5A-1

Department of Transport and Main Roads


Road Drainage Manual

5A

March 2010

5A-2

Appendix 5A
Pipe Flow Charts

Department of Transport and Main Roads


Road Drainage Manual

DISCHARGE IN CUBIC METRES PER SECOND

Appendix 5A
Pipe Flow Charts

5A

DISCHARGE IN LITRES PER SECOND

Figure 5A.2 Discharge and Velocity Graph


Mannings n = 0.012
March 2010

5A-3

Department of Transport and Main Roads


Road Drainage Manual

5A

March 2010

5A-4

Appendix 5A
Pipe Flow Charts

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

DISCHARGE IN CUBIC METRES PER SECOND

5A

DISCHARGE IN CUBIC METRES PER SECOND

Figure 5A.3 Discharge and Velocity Graph


Mannings n = 0.013
March 2010

5A-5

Department of Transport and Main Roads


Road Drainage Manual

5A

March 2010

5A-6

Appendix 5A
Pipe Flow Charts

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

DISCHARGE IN LITRES PER SECOND


DISCHARGE IN LITRES PER SECOND

5A

DISCHARGE IN LITRES PER SECOND

Figure 5A.4 Discharge and Velocity Graph


Colebrook-White formula k = 0.06mm
March 2010

5A-7

Department of Transport and Main Roads


Road Drainage Manual

5A

March 2010

5A-8

Appendix 5A
Pipe Flow Charts

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5A
Pipe Flow Charts

DISCHARGE IN LITRES PER SECOND

5A

DISCHARGE IN LITRES PER SECOND

Figure 5A.5 Discharge and Velocity Graph


Colebrook-White formula k = 0.60mm
March 2010

5A-9

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

Appendix 5B 5B
Worked Examples

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

Appendix 5B Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

5B

March 2010
ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

Rural Runoff Example


This example describes the process to determine the stormwater runoff from a simple rural
catchment.
The example commences after the catchment area has been determined (refer Section 5.6) and
catchment data has been gathered (refer Chapter 4).
The task for this example is, given the catchment data below, estimate the peak stormwater
discharge at Point A, the site for a proposed culvert, for the ARIs 50, 20 and 10 year flood
events.
Catchment Data
Area = 6 km2
Catchment is predominately flat to rolling country used for grazing cattle.
Channel is well defined and with little storage.
Chainage and heights for stream profile, as shown, have been extracted from a
topographic map.

Solution
We need to use the Rational Method formula to solve this - Q50 = k x C50 x I50 x A
We have determined that A = 6km2, therefore k = 0.278. Therefore we need to determine C50
and I50.
Because I50 is required to determine C50, we need to calculate this first.
March 2010

5B-1

5B

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

Step 1. To determine I50, we firstly need to calculate tc for the catchment. To do this, use
Bransby-Williams formula.

tc

FL
A0.1 S e0.2

We have already determined that A = 6km2 (therefore F = 58.5) & the length of the catchment
as L = 1.9km, therefore we need to calculate Se
To calculate Se, use Equal Area Slope method.
Plot the stream profile. Set Point A as datum. Mark distances and heights relative to Point A.

5B

Now calculate the area under the stream profile.


Area = [(0+1)/2] x 500 + [(1+2)/2] x 300 + [(2+3)/2] x 300 + [(3+4)/2] x 600
+ [(4+5)/2] x 200 = 4450m2

Using the above area, calculate the right ordinate of a triangle, which has the equivalent area.
Formula is [Area x 2 / length of stream]

4450 x 2 / 1900 = 4.68m


Plot this ordinate (known as the equal area ordinate) and draw a line back to Point A, and
calculate the slope of this line.

March 2010

5B-2

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

5B
Formula is [Se = Equal Area Ordinate / Length of Stream x100]

4.68 / 1900 x 100 = 0.25 %


or 4.68m / 1.9km = 2.47 m/km

Now we can use Bransby-Williams formula, to calculate tc.

tc

58.5 1.9
6 0.1 2.47 0.2

tc = 77.55min (round to 78min)

Step 2. Now we can determine the rainfall intensity for the ARI 50, 20 and 10 year storm
events, each with a duration of 78 minutes.

Firstly we need the Rainfall Intensity Table (IFD table) for the site (refer Section 5.8). Using
the departments RAIN program, the following IFD table was obtained.

March 2010

5B-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

5B
We can see from the table that 78min falls between the standard durations of 1 and 1.5 hours.
Therefore we need to interpolate the required intensities.

Duration

Intensity ARI 50

1 hr (60 min)

88.73

1.5 hr (90 min)

69.33

88.73 69.33

90 60 90 78 69.33 77.09
Therefore I50 = 77.1mm/hr

Interpolating for the ARI 20 and 10 year events, we get I20 = 64.8mm/hr & I10 = 55.8mm/hr

March 2010

5B-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 5B
Worked Examples

Step 3. The last variable to determine is Cy the runoff coefficient. Refer to section 5.9.

Using Table 5.9.1(a), we get C50 (it is important to note that this table only establishes C50 and
not C20 or any other coefficient).

Intensity

27.1

Catchment Relief

Catchment Storage

10

Ground Characteristics

40

C50 =

77.1 / 100 = 0.77

To determine Cy for the other ARIs, use the factors given in Table 5.9.1(b).

ARI

5B

Coefficient
10

0.8 C50

20

0.9 C50

Therefore:
C20 =

0.77 x 0.9 = 0.69

C10 =

0.77 x 0.8 = 0.62

Step 4. Now calculate for each ARI;

Q10 = 0.278 x 0.62 x 55.8 x 6 = 57.7 m3/s

Q20 = 0.278 x 0.69 x 64.8 x 6 = 74.6 m3/s

Q50 = 0.278 x 0.77 x 77.1 x 6 = 98.9 m3/s

End of Example

March 2010

5B-5

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Chapter 6
Approach to 6
Drainage Design

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Chapter 6 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Table of Contents
6.1

Overview .............................................................................................. 6-1

6.2

Geometric Design of Drainage Infrastructure ................................... 6-2

6.3

Computer Modelling of Drainage Solutions ...................................... 6-3

6.4

Preliminary Selection Decisions ........................................................ 6-3

6.5

Bridges ................................................................................................. 6-4


6.5.1

Overview ................................................................................................ 6-4

6.5.2

Riparian and Wildlife Corridors .............................................................. 6-4

6.5.3

Bridge Location and Waterway Alignment ............................................. 6-6

6.5.4

Bridge Geometry .................................................................................... 6-6

6.5.5

Road Grade and Hydraulic Clearance ................................................... 6-7

6.5.6

Span Lengths and Pier Location ............................................................ 6-7

6.5.7

Scour Protection..................................................................................... 6-8

6.5.8

Overtopping............................................................................................ 6-8

6.5.9

Maintenance........................................................................................... 6-8

6.6

Culvert Size and Type ......................................................................... 6-8

6.7

Open Channels .................................................................................... 6-9

6.8

Bank Protection and Linings .............................................................. 6-9

6.9

Longitudinal Drainage....................................................................... 6-10

6.10

Location .......................................................................................... 6-10


6.10.1 Location Example................................................................................. 6-11
6.10.2 Culvert Locations ................................................................................. 6-11

6.11

Water Quality .................................................................................. 6-11

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

March 2010

iv

Chapter 6
Approach to Drainage Design

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Chapter 6
Approach to Drainage Design
6.1

Overview

This chapter outlines an approach to the


selection and design of various road
drainage infrastructure in order to provide
for a specified hydraulic capacity and other
related design requirements.
Although the emphasis in this and
subsequent design chapters is on the
design of drainage infrastructure, this
approach can also be used to complete
planning or preliminary drainage design
during the Pre-project and Concept phases
within the Road System Manager
Framework. In these instances, it is the
responsibility of the civil designer to
ascertain the extent of design work
required.
For a road project, drainage design is
completed for a specific horizontal and
vertical alignment and cross sectional detail
during Detailed Design. However, during
the Concept and Development phases,
drainage design may be used to assist in the
fixing of the road geometry and to facilitate
a more economical design for the overall
road project.
Road alignments and cross sectional details
therefore may change during these phases
and inturn, drainage design needs to be
reviewed each time the road geometry
changes.
Drainage design starts with the preliminary
selection of infrastructure that might be
suitable for a particular location in a road
project utilising the results of the analyses
completed in Chapter 5.

This initial selection and sizing is made by:


(a) establishing the cross sectional
geometry at the drainage site (for both
the road and channel) to determine
maximum width / height of structure
(e.g. allowing for appropriate cover
requirements);
(b) sizing for hydraulic requirements based
on the determined / estimated design
discharge; and
(c) sizing for fauna passage (if required).
With this initial selection and sizing made,
the designer then:
(d) identifies ancillary drainage works
(headwalls, wing walls, aprons and
erosion protection works) required in
the vicinity of the selected cross
drainage structure;
(e) reviews the geometry of the cross
drainage with the ancillary works in
place;
(f) reviews the results in terms of the
design criteria including cost and
constructability; and
(g) repeats the process until an acceptable
design is completed.
Computer
modelling
facilities
can
streamline the iterative process (refer
Section 6.3).
During this iterative process, the designer
may need to revisit the analyses completed
in Chapter 5, to address topics that may not
have been considered in the completed
analyses or that may require review.

March 2010

6-1

Department of Transport and Main Roads


Road Drainage Manual

With the selection of the components of the


drainage structure or system completed, the
design of the system components can be
completed by reference to the appropriate
design chapters; 7, 8, 9, 10, 11 and 12.
Each of these chapters incorporates sections
that address the application of local
environmental design criteria that are
necessary to:
select and design the most
appropriate type of infrastructure;
and
design
related
environmental
treatments that address identified
impacts of the infrastructure.

This manual does not include the design of


bridges but guidelines on the location and
layout of a bridge are provided in Section
6.5.

6.2

Geometric Design of
Drainage
Infrastructure

Establishing the geometry of a drainage site


is an initial part of the design process.
Early sizing of drainage in the planning
phase requires some assumption of the road
alignments and cross sections to provide
dimensional inputs (height and width) to
calculations of possible drainage solutions.
Site specific geometry is also necessary for
the calculation of quantities for use in
estimates of cost for:
comparison of drainage alternatives;
and
preparation of estimates of cost as
required in the stages of the planning
and design process.
As the planning and design process
progresses with the selected infrastructure,

March 2010

6-2

Chapter 6
Approach to Drainage Design

geometric design of the proposal is


developed by:
locating and aligning the structure
with the watercourse or drainage path
and any identified fauna passage
requirements;
determining the length of the
structure and setting invert levels
using hydraulic slope requirements;
checking maximum allowable afflux
related to neighbouring properties
and design culvert appropriately;
determining and designing any
mitigation works required to address
unacceptable afflux levels;
checking appropriate distribution of
culverts to maintain existing flow
patterns;
checking maximum allowable outlet
flow velocity and sizing culvert
appropriately or designing suitable
outlet scour protection measures;
checking both culvert inlet and outlet
control conditions with proposed
design levels - maximum outlet
velocities can be determined and
compared
against
maximum
allowable outlet flow velocities;
determining tailwater level based on
hydraulic gradient of existing
channel flow for a design event,
including a check for backwater
effect;
reviewing surface water flow paths to
ensure water is quickly shed from the
road surface (within allowable depths
of flow) to reduce potential for
aquaplaning;

Department of Transport and Main Roads


Road Drainage Manual

ensuring table drains slopes are


equal to or greater than the minimum
allowable;
ensuring diversion drains (outlets to
table drains) are available, preferably
within the existing road reserve, or
beyond, by agreement with the
affected property owner;
checking cover, backfilling and
structural requirements for culverts
including laying method and class of
culvert (where applicable);
checking consistency of longitudinal
and cross drainage;
checking for any bypass flows to
adjoining culverts/catchments and
mitigate as necessary;
checking for the risk of blockage of
culverts and design the works
suitably;
locating and sizing necessary
environmental drainage works; and
selecting headwalls, wingwalls,
aprons, cut-off walls and erosion
protection works as applicable.

6.3

Computer Modelling
of Drainage Solutions

Computer modelling can assist the planning


and design process, with the detailed
hydrologic and hydraulic calculations and
with the presentation and storage of data.
The department has standardised on 12d
Model as the modelling system for the
road delivery system as a project passes
from survey to planning to design to
construction and operation. This package
has some drainage design functionality, but
does not cover all aspects of road drainage.

Chapter 6
Approach to Drainage Design

Various software packages are available


and those approved for use in the planning
and design of road drainage infrastructure
are detailed in Chapter 1.

6.4

Preliminary Selection
Decisions

Knowing where to start is often a time


consuming process and Table 6.4 has been
prepared to assist designers with
preliminary
selections
of
drainage
structures.
This selection process of
drainage infrastructure is further developed
throughout this chapter.
In urban drainage, many of the design
requirements for drainage may have been
established and the selection process is
initiated by reference to the appropriate
local authority drainage planning scheme
and the Queensland Urban Drainage
Manual (QUDM), published in 2008 by the
Department of Natural Resources and
Water and now, as of March 2009,
managed
by
the
Department
of
Environment and Resource Management.
The Design Development Report Small
Projects (Form M4211) or the Design
Development Report Large Projects
(Form M4212) also lists many other factors
which need to be considered before a
decision is made on new infrastructure.
For the majority of designs, it will be
obvious as to whether a bridge, culvert or
possibly a floodway is required at a given
location. This decision will be made on the
basis of matter such as:
catchment area and hydrology;
road alignment;
serviceability;
existing waterway bank height;

March 2010

6-3

Department of Transport and Main Roads


Road Drainage Manual

potential
for
accumulation of);

debris

Chapter 6
Approach to Drainage Design

(incl.

environmental constraints;
geotechnical considerations;
need to allow for the passage of large
fauna; and
stream flow activity.
In the case of an active or permanent
stream, building a bridge will be easier and
will have less impact on the environment.

For multiple reinforced concrete box


culverts (RCBC), slab link box culvert
(SLBC) and multiple reinforce concrete
culverts (RCC) with a total length greater
than 10m along the road centreline in
expansive soil conditions refer to Section
9.2.7.
A floodway is a low level section of road
designed to allow flood waters to cross the
road without damaging the road. They are
only provided where traffic volumes are
low
and
where
the
time
of
submergence/closure
is
considered
acceptable.
A floodway may incorporate a culvert
designed to pass low flows under the road
or to reduce ponding and may also be
incorporated into a bridge or culvert
solution where the bridge or culvert is
designed to pass a lower level flood than
the design flood.

6.5

Bridges

6.5.1

Overview

This manual provides guidelines on the


location and layout of a bridge but does not
provide a detailed process for bridge
design. Additional information on the
design of bridges can be found in:

March 2010

6-4

Waterway Design (Austroads 1994);


and
Bridges and Retaining Walls, RPDM
(DMR 2006).
This manual does however, provide
designers with an overview of the criteria to
be investigated and addressed in
establishing:
riparian and wildlife corridors under
the bridge;
bridge location
alignment;

and

waterway

bridge geometry;
road grade and hydraulic clearance;
span lengths and pier location;
scour protection;
overtopping of the road; and
maintenance requirements.

6.5.2

Riparian and Wildlife


Corridors

Riparian land is any adjoining land to


creeks and rivers or any land that influences
the waterway bank itself and which may
also be a wildlife corridor. Rather than just
narrow strips of land adjoining these
corridors, riparian land may be quite wide
and diverse. The riparian zone provides
natural buffer zones that protect the water
quality and the watercourse banks and may
extend across floodplains.
In some watercourses, management
programs have been initiated to restore
degraded land in the riparian zone. These
issues need to be considered when locating
bridge abutments and piers. They also may
need to be considered when locating
culverts.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Table 6.4 - Preliminary Selection of Drainage Infrastructure


Selection Issue

Factors

Decision

Bridge

Significant
catchment

Floodway

Wide shallow flow


Low traffic volume

Consider floodway if time of submergence is acceptable,


however check immunity option as it may be more
economical due to possible high cost of floodway
protection

Culvert Type

Waterway
geometry
Flat topography
Fauna passage

Pipe culverts are default type selected, however box


culverts are more suitable for situations where
there is minimal available waterway area;
insufficient cover to allow pipes;
fauna passage or the passage of people/vehicles or
large animals is required;
supplier location;
fish passage

Stream alignment
Road alignment
Geomorphology

Culvert Location

Desirable skew is 0 (i.e. no skew)


Skew intervals of 5 are preferable
Skew of culvert should not be more than 45
Do not locate culverts on the bends of actively moving
watercourses
In urban drainage systems, is preferably to use
bearings for precise culvert alignment.

Sensitive flora or
fauna
Fish passage
Permanent stream

Consider a bridge rather than culverts


Control pollutant runoff
Maintain a natural stream bed
Minimise disturbance of bed and banks

Soil type
Limited available
width
Steep gradient
Velocity
Channel shape

Where dispersive soils are present, a barrier must be


placed between flowing water and the soil
Velocity reductions by flattening grade or widening
the channel will often allow a less expensive or more
natural lining to be utilised. (Such as vegetation,
crushed rock, etc) - consider the incorporation of
check dams, drop structures, or a change in
alignment.

Type of Bank
Protection (where
velocities are
considered
excessive)

Soil types
Flow velocity

Use of Swales (vs


lined table drains
or kerb & channel)

Available space
Rainfall
Water quality
Low erodible soils

Swales may be used where:


there is sufficient space
where grade is sufficient to prevent permanent
ponding of water regular rainfall or watering will occur
water quality benefits are desired

Need for Outlet


Protection

Outlet velocities
Soil type
Erosion risk rating

Outlet protection may be required where outlet velocities


are sufficient to cause erosion. Management options
include:
change culvert design by reducing slope
replace or cover dispersive soils
reduce velocities through dissipation (subject to
safety considerations)
provide protection

Environmental
Sensitivity

Consider bridge if floodway is unacceptable

Need for Channel


Lining

Revegetation - planting of riparian vegetation is


favoured
Erosion mats - long term stability will need to be
addressed
Rock riprap - must be properly designed
Gabions - can be used on steep banks
Grout mats - flexible but will not provide good habitat
Concrete - usually a last resort

March 2010

6-5

Department of Transport and Main Roads


Road Drainage Manual

In locations where terrestrial corridors are


required under a bridge, and when suitable
circumstances exist, vegetation cover that
exists upstream and downstream of the
bridge, should be extended under the
bridge.
Bridge abutments should be moved away
from watercourse banks in order to increase
the opportunity for fauna passage that is
usually required on both sides of the
waterway.
This also provides benefit
related to the risk of scour.
Lighting or gaps may be required in
culverts to assist diurnal animals to move
through the culvert. This is because they
may not be willing to enter a dark entrance
where the other end is not visible.

For guidelines on the use of wildlifeexclusion or movement guide fencing, refer


to the departments Fauna Sensitive Road
Design guidelines.
Reference should be made to the projects
Environmental Assessment for any specific
requirements that need to be considered in
the location, orientation or geometry of the
bridge.

6.5.3

Bridge Location and


Waterway Alignment

Where practicable, the alignment of the


bridge should be chosen to avoid unstable
sections of a watercourse channel, such as
sharp or obviously mobile channel bends.
If piers must be located within the channel,
and if a pool is likely to form within the
channel at the bridge location, then the
foundation design must allow for future bed
erosion.
When it is considered necessary to realign a
waterway channel as part of a bridge
design, the following issues and concerns

March 2010

6-6

Chapter 6
Approach to Drainage Design

should be investigated and appropriately


addressed:
(a) Potential environmental impacts of
sediment runoff from the construction
of the new channel;
(b) Erosion potential of the downstream
channel in response to the proposed
realignment;
(c) Possible changes to existing bed
conditions,
including
pool-riffle
systems, within the channel;
(d) The need for rock protection of the
channel bed and banks given the
potential to adversely affect the
continuity and health of riparian
vegetation and consequently the quality
of the wildlife corridor;
(e) Any reduction in the length of the main
channel and a consequential increase in
the hydraulic gradient and erosion
potential;
(f) The form, condition and location of the
low-flow channel. Where practical, all
of these should be maintained;
The location of the low-flow channel can
have a significant effect on channel stability
and aquatic habitat values. It can also
meander within the bed of the main
channel, and the form, condition and
location of the low-flow channel can vary
from flood event to flood event.

6.5.4

Bridge Geometry

In establishing the geometry of a bridge, the


following aspects need to be considered:
road grade and height clearances
(above
accesses,
environmental
features and services);
hydraulic clearance (freeboard);
span lengths and location of piers;

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

scour protection;

6.5.6

overtopping; and
maintenance requirements / access.

6.5.5

Road Grade and Hydraulic


Clearance

Hydraulic clearance in most bridge designs


refers to the amount of freeboard to be
provided between the underside of the
bridge superstructure and the design flood
level.
In some instances, clearance has also to be
provided for road accesses, Public Utility
Plant,
environmental
features
and
navigation requirements.
Reference should also be made to Chapter
22 of the Road Planning and Design
Manual (DMR 2006). It is important to
note that when in tidal waters, the
Department of Environment and Resource
Management must approve the bridge spans
and vertical clearance for boats (including
yachts, if relevant). This department must
also approve the final design drawings.
This process was previously known as a
Section 86 approval under the Harbours Act
1955, and is preserved under Section 236 of
the Infrastructure Act 1994.
In summary, clearance is dependent on:
design road levels;
whether the bridge is to be designed
with freeboard or for overtopping;
the size and nature of flood debris;
environmental features;
built features; and
navigation
relevant).

requirements

(where

Span Lengths and Pier


Location

In selecting span length, the following


factors must be considered:
(a) Bridge design and construction issues;
(b) Whether any piers will be allowed
within the main watercourse channel;
(c) Size of debris that is likely to be present
within the watercourse;
(d) Required road elevation; and
(e) Required waterway area to satisfy
allowable afflux limits.
The spacing of bridge piers can have a
significant influence on the cost of a bridge.
Hence, the selection of pier spacing is an
important element, and is dependent on:
the stability of the stream bed and
banks;
the environmental sensitivity of the
waterway (i.e. should piers be
allowed within the low flow
channel);
the presence of any existing bridge
piers where duplication or upgrade
of roadways is planned, new piers
should be aligned with existing piers
where possible;
geotechnical
foundations;

conditions

navigational requirements
relevant); and

for

pier

(where

required permits.
Where practical, bridge piers should be
located away from the low flow channel.
Large-scale turbulence caused by bridge
piers located within low flow channels can
adversely affect fish passage and can cause
bed and bank erosion.

March 2010

6-7

Department of Transport and Main Roads


Road Drainage Manual

6.5.7

Chapter 6
Approach to Drainage Design

Scour Protection

Scour of bridge foundations or abutments


can arise when bed or bank surfaces are not
designed to resist likely peak velocities.
Peak velocities can also be exacerbated
when a build up of debris results in a
decrease in waterway area.
The most common form of scour induced
failure relates to scour of the river or creek
bed in the vicinity of bridge piers, as shown
in Figure 6.5.7, and/or abutments. During
flood events, river or creek beds may be
mobilised to significant depths, and hence it
is necessary to design bridge foundations to
remain stable during or following such
events.

The selection of appropriate scour


protection measures should be based on:
an understanding of peak velocities
(adjacent to both the bed and banks);
maximum depth of bed mobilisation;
erodibility of bed and bank material;
likelihood of flow deflection
occurring (i.e. as a consequence of
river geometry); and
an estimation of whether piers are
likely to cause deflection of flow (i.e.
for a skewed bridge).
For the analysis and design of bridge scour,
reference should be made to Waterway
Design (Austroads 1994).

Figure 6.5.7 - Scouring around Bridge


Piers

6.5.8

Overtopping

The maximum flood design load on a


bridge usually occurs when the flood
carrying debris is at deck level (not relevant
for high level bridges).
Hydraulic
calculations are required for either a flood
at deck level or the ARI 2000 year flood
level (whichever is the higher). Floods at
or above superstructure level will also
require the consideration of buoyancy
factors.

6.5.9

Maintenance

The bridge design must allow for safe


access for all maintenance activities.
Anticipated maintenance activities must be
listed at the time of design. Maintenance
activities are to be in accordance with the
departments Bridge Inspection Manual.

6.6

Culvert Size and Type

The choice of culvert size is heavily


influenced by the permissible afflux or
headwater level, the likely depth of flow,
watercourse shape (i.e. channel or
floodplain) and, in some cases, the need to
cater for fauna passage, pedestrian, vehicle
or bikeway access, or stock movement.

March 2010

6-8

Department of Transport and Main Roads


Road Drainage Manual

The selection of culvert type is closely


linked to the determination of size, and
hence both must be considered jointly. In
many cases, a number of iterations will be
required before culvert dimensions can be
finalised.
For the preliminary sizing in the Proposal
Development and Options Analysis, initial
size estimates can be determined on the
basis of permissible velocity and afflux /
headwater level.
In addition, ambient conditions can dictate
the suitability of certain culvert types. For
example, in corrosive environments (e.g.
coastal regions), some types of culverts
may not be appropriate.
For preliminary design more rigorous
design work is required and the complete
design process needs to be completed for
the detailed design.

6.7

Open Channels

Open channels may form part of a specific


drainage infrastructure solution where space
permits in the road area or where it is part
of a drainage planning process by a local
authority or a drainage initiative of another
government department.
As part of a specific drainage infrastructure
solution, they can be used for:
inlet or outlet channels;
longitudinal diversion drains;
stream diversions;
overflow channels; and/or
connections to other drains.
Appropriate soil and environmental
analyses will need to be used to design the
channel to deliver the required hydraulic
capacity with appropriate environmental
safeguards.

Chapter 6
Approach to Drainage Design

Local
authorities
or
government
departments may provide the designer with
specific drainage requirements through
programs such as:
(a) Soil conservation plans;
(b) Master drainage plans; and
(c) Various environmental waterway plans.
Velocity control and protection of channel
bed and banks are key elements in open
channel design (refer Sections 6.8 and 6.9).
Open channels should be designed with
consideration of water quality issues. High
flow velocities, often found in lined and
artificial channels, cause high velocities that
increase the risk of scour and transport of
pollutants. Open channels can be designed
as swales to allow the collection of
pollutants before they can enter receiving
waters.
Where open channels are part of a drainage
design, the design needs to be completed
early in the design process to assist in
determining headwater and tailwater levels
used for cross drainage design.

6.8

Bank Protection and


Linings

The choice of bank protection or channel


lining (whether natural of artificial) is
relevant to the design of open channels,
chutes, floodways and swales. The decision
as to which type of lining is the most
appropriate for the site is dependent on
factors such as soil type, potential for
vegetation growth, available space and flow
velocity. Bank protection and lining can
also influence water quality and the
conveyance of pollutants.

March 2010

6-9

Department of Transport and Main Roads


Road Drainage Manual

6.9

Longitudinal Drainage

Site constraints and design standards dictate


that it is not always possible to choose
between longitudinal drainage options such
as kerb and channel, grassed swales and
lined or unlined table drains.
In urban environments, kerb and channel
has historically been favoured for most
roads, though grassed channels are also
common on divided roads.
In rural areas, earth drains are more
common.
However, in some cases it will be possible
to choose between the above alternatives.
When a choice is available, the designer
will need to consider the following:

Gradient of the channel.


Steep
gradients will allow narrower
drainage channels to be used, but can
lead to high velocities, which in turn
may require linings with higher
resistance to erosion.
Available flow width. The flow
width for any option must be
compared to the available width for
longitudinal drainage. Flow width is
dictated by gradient, shape of the
drainage path, and roughness of the
flow path.
Need to improve water quality.
Where water quality improvement is
required, use could be made of grass
swales, rather than an impervious
lining. Channels with high flow
velocities are more prone to poor
water quality. For grass swales to be
a viable option, there must be
sufficient room to cater for shallow
flow, and the designer must consider
issues relating to the survival and

March 2010

6-10

Chapter 6
Approach to Drainage Design

maintenance
vegetation.

of

the

proposed

Survival of vegetation in the area. In


arid areas, it is often not practical to
design a flow path that is reliant on
the existence of vegetation to provide
protection to soils. Only vegetation
suited to the climate of the region
should be considered.
Maintenance
requirements
of
longitudinal drainage. As with all
types of drainage and pollution
controls, a lack of maintenance will
lead to failure of the design. In those
locations where maintenance may not
be readily available, it may be
necessary to propose a low
maintenance solution.
Water sensitive design. Throughout
urban areas, there is an increasing
trend to adopt a water sensitive
approach to drainage design.
Riparian corridors. The design of
the
channel
should
consider
maintaining or enhancing the riparian
corridors, to prevent adverse impacts
on water quality, and to maintain
environmental values, including
fauna movements.

6.10

Location

For major drainage infrastructure, such as


culverts or bridges, the location of a
structure can have a significant bearing on
both environmental impacts and waterway
stability. This is also true of several other
types of drainage infrastructure, including
open channels and drop structures.
Geometric design is an integral part of the
location process as outlined in Section 6.2
and is an essential in determining location
in those areas where:

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

the watercourse is not stable;


bed or bank erosion may result from
the presence of new infrastructure;
soils are highly erodible;
the area has high environmental
sensitivity;
bed slopes are steep; and
the face of a structure is not
perpendicular to the watercourse.
Where potential problems have
identified, it is important that:

been

(a) an alternative location or alignment is


identified; or
(b) appropriate protective measures are put
in place to prevent or mitigate the
potential impacts.
Alternative (a) should be the first
preference, but will not always be possible
in areas where the alignment is fixed.
In those instances where a river or creek is
obviously active (eroding or accreting), a
geomorphic analysis may be required.

6.10.1 Location Example


A proposed new road will cross Sandy
Creek at a point where the creek has active
bank erosion owing to the existence of a
meander. A series of box culverts are
proposed at this location. The following
courses of action could be considered:
Option 1: Propose a local
realignment of the road such that the
crossing of Sandy Creek will occur at
a stable location; or
Option 2: Stabilise the meander if
constrained for space; or

Option 3: Consider realignment of


the creek away from the proposed
crossing.
Option 1 would be favoured wherever
possible, with Option 2 the next favoured
alternative. The use of hard solutions
(e.g. riprap lining) or creek realignment
(Option 3) is not favoured, as changes to
the creek at one location will often transfer
problems to other nearby locations.

6.10.2 Culvert Locations


To minimise environmental
culverts should be located:

impacts,

where satisfactory ground conditions


and soil conditions exist;
away from reaches of highly unstable
channel;
away from bends in the watercourse;
where possible adverse effects on
other existing bridges and hydraulic
structures can be avoided;
where
ecological
acceptable; and

impact

is

where aesthetic considerations are


favourable.

6.11

Water Quality

When water quality (pollution) control


devices are required as part of the road
drainage system, additional considerations
must be taken into account.
For example, the hydraulics of the drainage
system may not be conducive to efficient
pollutant removal, or conversely, the
proposed pollution control device may
compromise the hydraulic efficiency of the
system.

March 2010

6-11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 6
Approach to Drainage Design

Questions that must therefore be addressed


when incorporating water quality controls
into a drainage system include:
(a) Does the device require a large
hydraulic head loss to operate?
(b) Will the device lead to upstream
flooding in flat areas?
(c) Is the gradient of the system too steep
(resulting in high velocities) to allow
effective pollutant removal?
(d) Can the device be
maintenance purposes?

accessed

for

Examples of design implications are offered


below:

Wetlands or sediment basins may not


operate effectively if subject to high
velocities during flood events. A
high flow bypass is often required.
When trash racks are installed,
provision must be made for high
head losses associated with blockage
of the racks.
In this case,
consideration of the potential for
flooding of upstream property must
also be assessed.
Pollution control devices placed in
areas with high tailwater levels may
not operate, particularly where there
is reliance on a floating boom to trap
litter.
Chapter 7 provides design procedures for
pollution control.

March 2010

6-12

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Chapter 7
Environmental
Consideration and
7
Design

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Chapter 7 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table of Contents
7.1

Introduction.......................................................................................... 7-1

7.2

General Environmental Duty .............................................................. 7-1

7.3

Environmental Values and Water Quality Objectives....................... 7-1

7.4

Pollution Control: Water Quality ........................................................ 7-2


7.4.1

7.5

7.6

Establishing Pollution Control Requirements ......................................... 7-2

Water Quality Treatment Devices and Design Guidelines ............. 7-14


7.5.1

Filter Strips ........................................................................................... 7-14

7.5.2

Grassed Swales ................................................................................... 7-17

7.5.3

Trash Racks ......................................................................................... 7-18

7.5.4

Proprietary Devices.............................................................................. 7-20

7.5.5

Constructed Wetlands.......................................................................... 7-21

7.5.6

Buffer Zones......................................................................................... 7-23

7.5.7

Water Quality Ponds ............................................................................ 7-24

Fauna Passage .................................................................................. 7-25


7.6.1

Identifying Fauna Passage Criteria ...................................................... 7-25

7.6.2

Identify Terrestrial and Aquatic Fauna Pathways................................. 7-26

7.6.3

Identify the Species Group................................................................... 7-26

7.6.4

Consult with the Relevant Authority ..................................................... 7-26

7.6.5

Identify Criteria affecting Drainage Design........................................... 7-26

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

March 2010

iv

Chapter 7
Environmental Consideration and Design

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Chapter 7
Environmental Consideration and
Design
7.1

Introduction

Road infrastructure environmental issues


are identified and assessed throughout the
road planning and design process. Project
specific environmental assessment provides
information about the condition of the
existing environment, the proposed project
area, associated environmental impacts of
the proposal and the identification of any
opportunities
for
environmental
management.
In this chapter, an approach to
acknowledging
and
addressing
the
relationship between site drainage and
requirements for water pollution control,
design of water pollution control measures,
and fauna passages are examined. Basic
principles and design criteria for the
mitigation of these environmental issues
will also be discussed.
This chapter links in with Chapters 12 and
13,
where
soils,
drainage
and
erosion/sedimentation risk management is
discussed. Understanding water quality in
Queensland, pollution sources, transport
mechanisms, and removal techniques in this
chapter will assist in applying the design
principles and practices of this chapter, and
that of chapters 12 and 13.

7.2

General
Environmental Duty

The Environmental Protection (EP) Act


1994 lists some key principles that govern

the way the department must approach


environmental management within its
business activities.
Under the EP Act (1994), the department
has a General Environmental Duty not to
undertake any activity that causes or is
likely to cause environmental harm unless
all reasonable and practical measures are
taken to prevent or minimise the harm.
The EP Act (1994) also instructs the
department
to
use
best
practice
environmental management. This is where
an
ongoing
minimisation
of
the
environmental harm caused by the activity
is
achieved
through
optimising
environmental procedures and controls for
the activity by assessing recognised
national and international techniques with
cost-effective measures.

7.3

Environmental Values
and Water Quality
Objectives

Prior to establishing water pollution control


methods, or even before reviewing the
impact site drainage may have on pollutant
export, relevant water quality legislation
and terminology should firstly be
acknowledged.
To protect downstream waterways and
wetlands from pollutants, identification of
appropriate Environmental Values (EVs)
and Water Quality Objectives (WQOs) for
receiving waters should influence design
criteria for the selection of appropriate

March 2010

7-1

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

drainage infrastructure. This will determine


the need for, and type of pollution control.
Environmental Values and Water Quality
Objectives are available in Schedule 1 of
the Environmental Protection (Water)
Policy (2009) or EPP Water (2009). They
are qualitative and quantitative targets and
outcomes for natural waterway health.

7.4

Pollution Control:
Water Quality

A key environmental consideration related


to drainage and road runoff (via
stormwater, site water runoff, rainfall, litter
and spills), is pollutant export and its
resulting impact on water quality.
Pollutants contained in runoff and drainage
from road corridors have the potential to
adversely affect the water quality and
aquatic biota of receiving waters with short
or long term impacts.

For any given project, the significance and


impact of pollutant export will depend
upon:
the relative sensitivity
receiving environment;

of

the

traffic type and volume;


road project infrastructure type and
form (e.g. off-ramp, traffic lights,
bend in road, steep hill, etc); and
climatic factors experienced in the
locality.
Therefore, pollution control techniques
must be established and implemented
according to many factors, including the
type, source, concentration of pollutant
export and the risk of harm it may have on
the receiving environment.

March 2010

7-2

7.4.1

Establishing Pollution
Control Requirements

When identifying the need for pollution


control, the following steps should be
completed:
(a) Ascertain
environmental
risk,
environmental issues and water quality
objectives (Section 7.4.1.1);
(b) Identify
pollutant
sources
and
estimation of pollutant loads (Section
7.4.1.2);
(c) Identify pollutant transport processes
(Section 7.4.1.3);
(d) Identify pollutant removal processes
(Section 7.4.1.4);
(e) Assess potential pollutant
devices (Section 7.4.1.5);

control

(f) Calculate potential pollutant removal


(Section 7.4.1.6);
(g) Implement treatments (Section 7.4.1.7);
and
(h) Evaluate the working efficacy of
pollutant removal processes and review
if necessary (Section 7.4.1.8).
7.4.1.1 Ascertain Environmental
Risk, Environmental Issues
and Water Quality
Objectives
The
sensitivity
of
the
receiving
environment and the potential impact of the
road corridor works will have been
identified in the environmental assessment.
Greater catchment activities and dynamics,
traffic volumes and heavy vehicle content
should be identified in the planning report
and subsequently, the Environmental
Management Plan (Construction), (or the
EMP(C)).

Department of Transport and Main Roads


Road Drainage Manual

This information, in conjunction with the


water quality requirements documented
within the environmental assessment, will
assist in identifying and selecting
appropriate pollutant removal techniques to
achieve local WQOs.
These can include:
the capture of gross pollutants
upstream of a sensitive water body;
and
the discharge from control devices
such as sediment basins that comply
with provisions and regional WQOs
(as outlined in the EPP Water (2009).
As part of the environmental assessment
process, all (but not limited to) 13
environmental elements listed in Main
Roads Standard Specifications Roads
Environmental Management (MRTS51)
(TMR 2010c) need to be addressed. The
water quality section may contain
information regarding:
any watercourses that may be
affected by construction activities;
the potential (target) contaminants
involved
in
the
construction
activities;
stormwater quality requirements (if
necessary);
specific WQOs and provisions
assigned to the region / sub-region /
catchment (if available in the EPP
Water (2009); and
water monitoring locations and
frequencies as outlined in the
Environment & Heritage Road
Processes Manual and Annexure to
MRTS51 (TMR 2010c).

Chapter 7
Environmental Consideration and Design

Water quality and protection guidelines


applicable in Queensland.
The Australian and New Zealand
Guidelines for Fresh and Marine Water
Quality (ANZECC 2000) is commonly used
by environmental authorities as a nonstatutory framework for setting WQOs for
the sustainable management of water
resources.
The ANZECC (2000) guidelines provide a
tool to determine the water quality required
to protect regionally specific EVs. The
guidelines outlay pollutant concentration
threshold levels, or trigger values for
various types of aquatic ecosystems and
locations.
If a pollutant concentration exceeds its
trigger value, it may have the potential to
cause environmental harm to the particular
receiving environment. Therefore in the
event of the threshold level being exceeded,
the guideline advises that a management
response should be triggered.
Table 7.4.1.1 provides examples derived
from ANZECC (2000) showing water
quality trigger values for typical pollutants
(nutrients, suspended solids and salts) for
various locations within Australia and
various types of slightly disturbed
ecosystems.
The Queensland Water Quality Guidelines
(QWQG) (EPA 2006) was written to
address the principles and objectives of the
ANZECC (2000) but with a focus on the
protection of aquatic ecosystems in
Queensland.

March 2010

7-3

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table 7.4.1.1 - Examples of Water Quality Trigger Values for Typical Pollutants (ANZECC
2000)
Location and Ecosystem Type

Pollutants
TN

TP

Turbidity

Salinity

(Og/l)

(Og/l)

(NTU)

(S/cm)

Upland river

250

20

2-25

30-350

Lowland river

500

50

6-50

125-2200

Freshwater lakes and reservoirs

South East Australia

350

10

1-20

20-30

Wetlands

no data

no data

no data

no data

Estuaries

300

30

0.5-10

n/a

Marine

120

25

0.5-10

n/a

Upland river

150

10

2-15

20-250

Lowland river

200-300

10

2-15

20-250

350

10

2-200

90-900

Wetlands

350-1200

10-50

2-200

90-900

Estuaries

250

20

1-20

n/a

Marine

100

10-15

1-20

n/a

Upland river

450

20

0-20

120-300

Lowland river

1200

65

10-20

120-300

Freshwater lakes and reservoirs

350

10

10-100

300-1500

Wetlands

1500

60

10-100

300-1500

Estuaries

750

30

1-2

n/a

Marine

230

20

1-2

n/a

Upland river

no data

no data

1-50

100-5000

Lowland river

1000

100

1-50

100-5001

Freshwater lakes and reservoirs

1000

25

1-100

300-1000

Wetlands

no data

no data

no data

no data

Estuaries

1000

100

0.5-10

n/a

Marine

1000

100

0.5-10

n/a

Tropical Australia

Freshwater lakes and reservoirs

South West Australia

South Central Australia

TN = Total Nitrogen
TP = Total Phosphorus
NTU = Notional Turbidity Units
S/cm = Conductivity of water in micro siemens per centimetre

March 2010

7-4

Department of Transport and Main Roads


Road Drainage Manual

Until the QWQG (EPA 2006) was released,


Queensland fresh and marine water quality
was managed according to standards and
parameters contained in ANZECC (2000).
It is, however recognised that regionally
diverse and specific aquatic ecosystems,
WQOs and EVs exist within Queensland,
which require a more specific and locally
detailed guideline to protect water quality.
Therefore, the introduction of the QWQG
(EPA 2006) allows the user to break down
and identify WQOs, EVs and water quality
parameters for specific regions and subregions of Queensland. It acts as an
extension to the national guidelines
mentioned above and should take
preference over them when used for water
quality management with road construction
activities.
Using Queensland-specific
guidelines enables a more tailored approach
to managing land use, road works, linear /
diffuse and point source runoff, monitoring
and the health of aquatic ecosystems.
The QWQG (EPA 2006) can also be used
to assist local authorities with developing
their own localised sub-regional guidelines
for specific catchments in their area.
Creating sub-regional guidelines is useful
for areas that have little or no previous
coverage from ANZECC (2000) and
QWQG (EPA 2006), or if water quality
parameters and objectives are required for
development and ecosystem protection.
The Department of Environment and
Resource Managements approval and
previously sound water quality monitoring
techniques are required for these plans.
The EPP Water (2009) is a key legislative
document that identifies specific regional
water bodies based on environmental
significance. Schedule 1 of the EPP Water
(2009) lists specific fresh / estuarine /
marine water bodies in Queensland, the

Chapter 7
Environmental Consideration and Design

relevant basin it is located within, and the


published WQOs and EVs pertaining to that
water body.
It is a requirement of MRTS51 (TMR
2010c) that water released from a site (i.e.
clean diverted and dirty site-generated
runoff) must comply with water quality
provisions of the EPP Water (2009).
The Department of Environment and
Resource Management requires the
following
hierarchy
of
guideline
documentation when assessing WQOs and
EVs for water bodies in Queensland:
1. EPP Water (2009) (see Schedule 1);
2. QWQG (EPA 2006);
3. Local / site specific documents
(Department of Environment and
Resource Management approved) if
available; then
4. ANZECC (2000).
(Source: EPP Water (2009))
7.4.1.2 Identify Pollutant Sources
and Estimation of Pollutant
Loads
Common sources of pollutants in road
runoff that may adversely effect receiving
and downstream environments have been
documented in several departmental and
external reports (BMT WBM 2007 and
SMEC 2008).
A catchment area around a road will
provide some of the pollutants that deposit
onto roads, e.g. sediment, agricultural
nutrient-rich runoff, litter, atmospheric
deposition, etc. Traffic will, as mentioned
before, also contribute various pollutants
onto the road. Therefore, pollutant loads
exiting the road reserve into waterways and
drainage systems will be the sum of direct

March 2010

7-5

Department of Transport and Main Roads


Road Drainage Manual

and indirect pollutant concentrate from


roads and their catchment areas.
Table 7.4.1.2 has been adapted from Ball et
al. (1998), Construction Industry Research
and Information Session (2007), and Chiew
et al (1997) to represent typical pollutants
(and their sources), washed from road
corridors.
A review of the project environmental
assessment should be undertaken to identify
likely and harmful road runoff pollutants
from the project corridor. Road runoff
pollutant loads can be estimated by:
analysis of data from a good stormevent monitoring program;
simple computations; and/or
an appropriate water quality model.

The method selected will depend on the


management objective and data availability.
Average long-term pollutant loads can be
estimated from historic rainfall data,
monitoring and from simple information
about the catchment and road corridor.
It is important to note other qualitative and
quantitative methods of estimating pollutant
loads in the environmental assessment. For
example:
an understanding of the relationship
between
type,
sources
and
environmental impacts of pollutants;
and
an understanding of the physical
catchment, urban design and drainage
patterns will assist in determining the
best management technique for
mitigating road-runoff pollutant
impacts.

March 2010

7-6

Chapter 7
Environmental Consideration and Design

Event Based Monitoring


Event Mean Concentration (EMC) can be
estimated by monitoring mean pollutant
concentration and discharge over a storm
event. EMC within one catchment can
however differ significantly from storm to
storm.
The EMC depends on many
catchment and climate characteristics and
can vary in magnitude between catchments.
Therefore, a good event-monitoring
program is essential where accurate
estimates of pollutant loads are required.
Significant errors in estimating long term
pollutant loads can result without
monitoring programs. Typical errors in
estimating long term pollutant loads are as
follows:
No monitoring - 100 to more than
1,000 percent;
Some periodic monitoring - 50 to
more than 500 percent; or
Detailed event monitoring - 20 to 100
percent.
Simple Computations
In absence of reliable field data, simple
computations can be used as a guide to
calculate approximate estimates of pollutant
loads. This can be useful when assessing
potential pollutant loads in rainfall
volumes.
The average long-term pollutant load of an
area can be estimated using the following
formula:
Pollutant load = runoff x EMC
Where:
EMC = Event Mean Concentration.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table 7.4.1.2 - Common Sources of Pollutants in Road Runoff.


Source

Typical Pollutants

Source Details

Atmospheric
deposition

P, N, S, heavy metals (Pb,


Cd, Cu, Ni, Zn, Hg),
hydrocarbons

Industrial activities, traffic air pollution, & agricultural


activities (all deposited as particulates). Raindrops also
absorb atmospheric pollutants.

Traffic
exhausts

Hydrocarbons (Including
PAH (polycyclic aromatic
hydrocarbons), MTBE
(Methyl Tertiary Butyl
Ether), Pb, Cd, Pt, Pd, Rh

Vehicle emissions, unburned fuel and particles from


catalytic converters.

Traffic wear
and corrosion

Sediment, heavy metals


(Pb, Cr, Cu, Ni, Zn, Cd,
Mn)

Tyre wear, vehicle corrosion.

Leaks and
spillages (e.g.
from vehicles)

Hydrocarbons,
phosphates, heavy metals,
glycol, alcohols

Engine leaks, hydraulic and de-icing fluids, refuelling and


accidental spillage and lubricating oil can contain
phosphates and metals.

Roofs

Heavy metals (Cu, Pb, Zn),


bacteria, organic matter

Atmospheric deposition, corrosion of metal roofing or


coatings such as tar can sometimes contain
concentrations of heavy metals. Animal droppings and
vegetation contribute organic matter and bacteria.

Litter / animal
faeces

Bacteria, viruses,
phosphorus, nitrogen

Drink cans, paper, food, cigarettes, animal excreta, plastic


and glass. Some of this will break down and wash off
urban surfaces. Dead animals on roads decompose and
release pollutants such as bacteria and nutrients.
Animals (including pets) leave faeces that wash into the
drainage system.

Vegetation /
landscape
maintenance

P, N, herbicides,
pesticides, fungicides,
organic matter

Leaves and grass cuttings, herbicides and pesticides


used in landscaped areas such as gardens, parks,
recreation areas and golf courses can be a source.

Soil erosion

Sediment, P, N,
herbicides, pesticides,
fungicides

Runoff from poorly-detailed landscape or other areas can


wash onto impervious surfaces and cause runoff
pollution.

Cleaning
activities

Sediment, organic matter,


P, N and detergents

Washing vehicles, windows, bins or pressure washing


hard surfaces transports pollutants into surface water
drainage.

Illegal disposal of chemicals and oil

Hydrocarbons and
various chemicals

Pavement

Pavement wear from water, traffic, and so on.

Sediment

Occurs at small (domestic)


or large (industrial) scales.

March 2010

7-7

Department of Transport and Main Roads


Road Drainage Manual

Water Quality Modelling


Computer models may be used to estimate
runoff quantity and quality from a road
corridor to provide estimates for:
characterising peak, mean and
average annual pollutant loads; and
determining seasonal and spatial
characteristics;
A number of water quality and stormwater
systems models exist, which are designed
for use by engineers, managers, planners
and other staff from private to public
organisations.

A commonly used catchment model used to


assist water quality management is the
Model for Urban Stormwater Improvement
Conceptualisation, or MUSIC (Wong et
al 2002). MUSIC modelling is a decision
support tool that can assist in the planning
of
stormwater
quality
management
strategies.
MUSIC allows the user to determine the
likely stormwater quality resulting from
specific catchments, predict specific
stormwater treatment device performance
and to create subsequent management plans
and evaluate their success.
MUSIC can operate over a range of spatial
and temporal scales. It can be used with
several other models (e.g. models used for
soils, hydrology, rainfall, operation of
culverts, etc) Wong et al (2002) and eWater
(2008).
7.4.1.3 Identify Pollutant Transport
Processes
The determination of additional design
criteria to enhance or maintain the
downstream water quality will require the
knowledge of relevant pollutant transport
mechanisms.

March 2010

7-8

Chapter 7
Environmental Consideration and Design

Pollutant runoff from a roadway will be


generally transported by the roadway
drainage infrastructure and will concentrate
in gutters, pipes and channels.
The
pollutants associated with the stormwater
runoff will be transported as coarse or
bottom sediments, suspended (fine)
particles or in solution.
The rate of
pollutant transport is dependent on pollutant
size, water velocity, depth and the degree of
turbulence.
Fine particulates and dissolved pollutants
(e.g. heavy metals) can become attached to
sediments, or flocculate to form larger
particles.
Most of the pollutants in
sediments are found attached to smaller
particles owing to their greater surface area
relative to larger particles.
Pollutants
attached to fine particles are easily
transported because small flows (and hence
low velocities) are sufficient to mobilise
and keep them in suspension.
Fine particulates are difficult to prevent
from being transported from construction
sites into waterways after erosion has been
allowed to occur. Therefore to prevent or
minimise erosion provides the best method
to reduce the transportation of fine
sediment.
Heavy metals from motor vehicles and
atmospheric fallout may deposit directly
onto road surfaces or become entrained in
air flows and deposited some distance away
depending on their particle size. Particulate
material on the road surface such as
sediment, bituminous products, rubber from
tyre wear and particles coated with oils,
actively adsorb heavy metals.
The
particulates and associated heavy metals
temporarily bind themselves to the road
surface and particulate material until they
are dislodged and transported by rainfall
events.

Department of Transport and Main Roads


Road Drainage Manual

Heavy metals contained in road runoff will


be distributed in either bound or soluble
forms. Chromium, iron, nickel, lead and
hydrocarbons are predominantly adsorbed
to sediments and particulate matter. This
provides an opportunity for heavy metal
removal by targeting the removal of
sediments from runoff.
Cadmium, copper and zinc appear at higher
percentages in the soluble phase and thus
are required to be removed by storage
and/or uptake by aquatic biota (e.g. insects,
aquatic plants, etc) (Peterson and Batley
1992).
7.4.1.4 Identify Pollutant Removal
Processes
Stormwater quality improvement measures
rely on a variety of mechanisms for
reducing pollutant levels within stormwater.
The mechanisms employed may be either or
a combination of physical (e.g. stormwater
grate), chemical (e.g. flocculation) or
biological (e.g. macrophytes) process and
their effectiveness may be dependent on the
site
conditions
and
stormwater
characteristics.
Stormwater pollution
removal devices can be grouped into three
categories based on their dominant
treatment processes.
Primary Treatment - Physical screening
or rapid sedimentation techniques (e.g.
typically retained contaminants include
gross pollutants and course sediments).
Secondary Treatment - Sedimentation of
finer particles and filtration / chemical
techniques
(e.g.
typically
retained
contaminants consist of fine particles and
attached pollutants).
Tertiary
Treatment
Enhanced
sedimentation and filtration, biological uptake, adsorption onto sediments (e.g.

Chapter 7
Environmental Consideration and Design

typically retained contaminants


nutrients and heavy metals).

are

As regional WQOs are developed or are


redefined, best practice management
techniques for stormwater runoff treatment
need to respond to these changing
standards.
Regional WQOs should be consulted when
deciding which level of treatment (i.e.
primary, secondary or tertiary) is required
for environmental compliance.
There is general industry recognition to,
where possible, incorporate a combination
of treatment mechanisms in one location, to
optimise the amount and range of pollutants
removed from stormwater runoff (also see
Chapter 13, Section 13.13.2). In other
circumstances where space limitations and
certain practicalities impose, single
treatment measures are used to achieve
prescribed regional WQOs.
Depending on size and condition of a site,
relative need and practicality, timeframes,
materials and cost, stormwater pollution
treatment measures may be applied using
either an outlet or treatment train
approach.
The outlet approach involves a single
treatment measure at the road corridor
catchment outlet that discharges directly
into the downstream environment. An
example of this is a rock-lined outlet
channel shown in Figure 7.4.1.4(a).
The treatment train approach requires a
number or sequence of different treatments
throughout the road corridor catchment
before discharge to the receiving
environment. The sequence of treatment
measures are designed to remove different
types and sizes of pollutants, thus
optimising the amount and range of

March 2010

7-9

Department of Transport and Main Roads


Road Drainage Manual

pollutants removed from discharge waters,


see Figure 7.4.1.4(b).
The selection of the treatment controls for a
road
corridor
catchment
under
consideration will depend on a wide range
of key selection criteria to enable
achievement of regional WQOs.

Chapter 7
Environmental Consideration and Design

(c) Soil Type: Differing treatment devices


may be reliant upon either infiltration
or storage of stormwater runoff. For
example, stormwater infiltration will
yield better results on highly permeable
soils, whilst the storage of stormwater
will require soils with very low
permeability.
For more detailed soils information,
see the departments Soils Manual
2010 and Chapter 13, Sections 13.3 to
13.9.
(d) Land Availability and Catchment
Area: The availability of sufficient
appropriate land within a subcatchment that can be used for a
treatment device may be restricted,
thereby reducing the size, effectiveness
or even the option of using the device;

Figure 7.4.1.4(a) - Rock-lined Outlet


Channel from a Culvert

The selection of the most appropriate


stormwater treatment methods should be
influenced by a number of environmental
and design elements (e.g. soil type, slope
etc). Each (site) environmental scenario is
dynamic and individual and will require an
assessment on a project-by-project basis.
These elements are defined as:
(a) Slope: Treatment devices that do not
store flow may require small velocities
and hence gentle slopes;
(b) Hydraulic Head: Head losses in
treatment devices can exert a minor to
large impact upon the hydraulic grade
line. As a result head losses from a
treatment device may adversely impact
upon upstream flood levels particularly
when retrofitting a device into an area;

March 2010

7-10

(e) Habitat Enhancement: Treatment


devices that are able to offer either a
wildlife
and/or
aquatic
habitat
enhancement may improve aesthetics;
(f) Water Table: A high water table depth
may reduce the effectiveness for a
treatment device relying on infiltration;
(g) Safety Hazard: Treatment devices may
introduce new safety hazards that may
have not been present before
installation
(e.g.
water-borne
pathogens, drowning risk, etc);
(h) Water Supply: Treatment devices such
as wetlands or ponds may require a
permanent water supply to ensure the
long term effectiveness of the device;
(i) Pests: Treatment devices such as
wetlands or ponds may increase the
potential for nuisance from pests such
as mosquitoes and weeds; and

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Figure 7.4.1.4(b) - Desirable Design Ranges for Treatment Measures and Pollutant Sizes
(adapted from Wong (1999)).

(j) Maintenance: Treatment devices will


vary significantly with regard to their
maintenance
cost,
accessibility,
equipment and scheduling to ensure the
desired effectiveness is consistently
maintained.
7.4.1.5 Assess Potential Pollutant
Control Devices
Each potential pollutant control device
needs to be assessed to determine if it is
suitable for the site conditions. Each
pollutant control device can be accepted or
rejected on the basis of screening criteria to
provide a short-list. Table 7.4.1.5 provides
a means of assessing common design
elements in order to determine if a
particular control device is suitable for a
specific site condition.
7.4.1.6 Calculate Potential Pollutant
Removal

and regional WQOs. The final selection


can be achieved by comparing all potential
treatments as follows:
(a) Determine the pollutant removal of
each short-listed control device based
on relevant performance data or Tables
7.4.1.6(a) to 7.4.1.6(c) (e.g. target 90%
reduction of lead through pollutant
control device).
(b) Determine the area of the catchment for
which the device(s) can treat runoff.
(c) Factor the mean removal rate of each
pollutant parameter by the ratio of area
treatable by the device to total
catchment area. For example, if a
pollution control device has a 60%
removal efficiency and will treat 50%
of the catchment area then the overall
pollutant removal efficiency will be
30%.

The final selection of potential pollutant


control devices should be made on the basis
of achieving pollutant removal objectives

March 2010

7-11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table 7.4.1.5 - Design Elements Associated with Treatment Devices


Pollutant
Control
Device
Oil Grit
Separators
Open Gross
Pollutant
Trap
Closed
Gross
Pollutant
Traps
Trash Rack
Downward
Inclined
Screen
Extended
Detention
Basin (see
Chapter 12
for design)
Sand Filter
(depth of)

Filter Strips
Buffer Zones
Grassed
Swales
Constructed
wetlands

Area
Served
(ha)

Slope

Head
Require
ment

Soil Type

Capital
Cost

Mainten
ance
Cost

<1

Note 1

Low

NA

Moderate

Moderate

>2
>40

Note 1

High

NA

High

Moderate
High

Low

NA

High

Moderate

NA

Moderate

Note 1

Low
Moderate
High

NA

Moderate
High

Low
Moderate
Low
Moderate

>5

Note 1

Low

All

High

Moderate
High

<2
can be
designed
larger
<2

Note 1

High

Generally
housed in
concrete

High

Moderate
High

Note 1

Low

All

Moderate

Low

Note 1
<5%

Low
Low

All
Sand to
sandy
loam

Moderate

Low

<2

Moderate

Low

Note 1

Low
Moderate

Loam to
clay
feasible in
sand to
sandy
loam

High

Moderate

<15

<20
40

Note 1

General Configuration

Outlet structures
include weirs or outlet
pipes. Energy
dissipater at both basin
inlet and outlet to
control velocities
Min filtration depth of
400 mm on
recommended filtration
time. Energy dissipater
at inlet.
Requires Considerable
Land. Length of strip
generally>6 m.
Recommended min
length of 30 m. Bottom
width between 0.6 m to
2.5 m recommended

Water
>5
Note 1
Low
High
Moderate
quality
Moderate
ponds
Source: Derived from: NSW EPA (1997) and Mudgeway et al (1997).
Note: 1 - From 0 - 5% slope preferred but the range can be extended beyond 5%. Buffer zones should only be extended beyond
5% with careful design.

March 2010

7-12

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table 7.4.1.6(a) - Pollutant Removal Performance of Typical Primary Treatment Devices

Pollutant Removal Efficiency (%)


Treatment
Control

Coarse
Sediment

Fine
Sediment

Phosphorous
TP

Nitrogen
TN

Oil &
Grease

Bacteria

Litter

Oil Grit
Separators

50-75

10-50

0-10

0-10

50-75

0-10

10-50

Gross
Pollutant
Trap

60-100

20-30

20

20

10-20

0-10

50-75

Trash
Rack

10-50

0-10

0-10

0-10

0-10

0-10

10-50

Note: These percentages are indicative only and appropriate design procedures should be followed.
Derived from: NSW EPA (1997) and Mudgeway et al (1997).

Table 7.4.1.6(b) - Pollutant Removal Performance of Typical Secondary Treatment


Devices
Pollutant Removal Efficiency (%)
Treatment
Control
Extended
Detention
Sand Filter
Filter strips

Phosphorous
SS
org. sol.
TP
P
P
50-75 10-66
64
18
60-90 35-80
80
57
5-95 50-79
74
66
62

Buffer strip
Grassed
Swales

Nitrogen
TN

org.N sol.N

10-35
19
40-70
55
50-73
62

Cu

Cd

Fe

BOD COD Bacteria


20-41
30

60-80

35-70
70

50-90
70

Sample Size
6 studies
6 studies

53

10 studies
1 studies

95

8-24 -4-11 -4-13


16
4
5

4-25
15

Zn

70-90 24-62
83
45
-110-0 65-90 10-80 20-60
>0
74
53
40

100
80

Pb

-5-22
9

0-91 34-90 14-60 20-50 3-67


80
70
50
29
35

7 studies
25

Legend: SS = Suspended Solids. Shading denotes median value. Source Mudgeway et al (1997).

Table 7.4.1.6(c) - Pollutant Removal Performance of Typical Tertiary Treatment Devices


Pollutant Removal Efficiency (%)
Treatment
Control
Constructed
Wetlands
Water
Quality
Ponds

SS
40-98
81

Phosphorous
org. sol.
TP
P
P
-3397
50

39-98 0-80
69
51

Nitrogen
TN

Pb

Zn

Cu

Cd

Fe

BOD COD Bacteria

13-90
50

9-94
72

-2940-99 33-99 12-62 18-34


97
66
66
37
26
50

70-80 30-85 14-20 24-60


75
60
17
42

9-95
68

0-71
51

0-75 -9-43
59
25

20

org.N sol.N
32

40

17

0-69 20-70
44
30

Sample Size

6 studies

90-95
93

6 studies

Legend: SS = Suspended Solids. Shading denotes median value. Source Mudgeway et al (1997).

March 2010

7-13

Department of Transport and Main Roads


Road Drainage Manual

7.4.1.7 Implement Treatments


Design guidelines for the installation of
selected control devices can be found in
Section 7.5, and in IECA (2008).
7.4.1.8 Evaluate the Working
Efficacy of Pollutant
Removal Processes and
Review if Necessary.
If stormwater runoff or drainage from a site
contains levels of pollutants that do not
comply with regional and sub-regional
WQOs, then a review of the working
efficacy of the pollutant removal processes
is required.

Visual inspections of catchment waters


should be undertaken to determine the
presence of litter build-up, sediment or
chemical plumes or other contamination.
As the MRTS51 (TMR 2010c) stipulates,
water quality monitoring should be carried
out at nominated intervals to determine
pollutant levels.
An on-site assessment of the physical state
(integrity) of the pollutant removal devices
should be carried out regularly and any
maintenance should be undertaken to
restore devices to the desired working
condition and standard. Devices should be
added or taken from the site in order to
achieve the relevant WQOs.
The steps outlined in Sections 7.4.1.2 7.4.1.6 should be reviewed and repeated in
conjunction with the environmental
assessment documentation. The Erosion &
Sedimentation Control Plan (ESCP) (refer
Chapter 13) may be used in conjunction
with this process.

Chapter 7
Environmental Consideration and Design

7.5

Water Quality
Treatment Devices
and Design Guidelines

The following section will guide the user


through descriptions of pollution removal
devices for runoff, advantages and
limitations, and design guidelines.
Because mobilised sediment in drainage
and runoff is considered a pollutant
(especially when deposited into receiving
waters), the following devices and
measures should also be considered when
applying erosion and sediment control
measures.
In the example of a treatment train, a
combination of water pollution removal
measures and erosion and sediment controls
may be combined.
Therefore the following section can be used
in conjunction with Chapters 12 and 13.

7.5.1

Filter Strips

7.5.1.1 Description
Filter strips are grassed or vegetated areas
used to control polluted runoff from the
pavement surface or other disturbed areas
within the road corridor as shown in Figure
7.5.1. Flow over filter strips is usually
shallow sheet flow. They are generally
located adjacent to regions where there is a
sensitive receiving environment (e.g. water
course or wetland). They can treat runoff
containing sediments, heavy metals and
other pollutants.
Advantages of filter strips include:
increase rate of infiltration, which
can reduce and delay storm run-off;
high removal rates of pollutants;

March 2010

7-14

Department of Transport and Main Roads


Road Drainage Manual

retention of pollutants close to


source;
improved aesthetic appeal of an area;
and
relatively inexpensive construction.

Chapter 7
Environmental Consideration and Design

regular inspections are required to


assess the condition of the strips.
7.5.1.2 Design Guidelines
The primary purpose of the filter strip is
removal of sediment, with some removal of
soluble pollutants by biological uptake and
by infiltration into the subsoil.
The
objective of a filter strip is to generate a
dense and diverse vegetation cover to
maximise infiltration, provide adequate
contact time between runoff and vegetation,
and to minimise erosion.
Horner et al. (1994) cited a technique for
sizing filter strips and grass swales. It was
developed in Seattle, USA, and the results
indicate that optimum pollutant retention
occurs when the hydraulic residence time is
nine minutes. The performance of pollutant
retention deteriorates when the residence
time falls below five minutes.
This design technique is summarised in the
following 10 point process:

Source: Environmental Best Management


Practices (BCC 1996)
Figure 7.5.1 - Grass Filter Strip

Disadvantages of filter strips include:


limited removal of fine sediment and
dissolved pollutants;
sizeable land areas required with
limited public access;
a sunny aspect for plant growth;
reduced
effectiveness
for
concentrated flows and high flow
depths;
strips are only suitable for gentle
slopes (<5%); and

1. Calculate the design discharge for the


nominated average recurrence interval
(ARI). Pollutant control devices are
usually sized for storm events between
the ARI 3 month and 1 year.
2. Determine the bed slope So (m/m) of
the filter strip. Filter strip performance
has been found to reduce if located on
grades exceeding 5% and particularly if
the slope exceeds 15% (Schueler et al.,
1992).
3. Set the design flow depth. A maximum
depth of flow over the filter strip of 12
mm is recommended.
4. Solve for flow width using suitable
methods of hydraulic analysis, such as
Mannings Equation:

March 2010

7-15

Department of Transport and Main Roads


Road Drainage Manual

AR 2 / 3 S o
n

1/ 2

VA

Where:
Q = design runoff rate (m/s);
R = hydraulic radius = A/P;
A = cross-sectional area (m);
P = wetted perimeter (m);
So = longitudinal bed slope (m/m);
n = Mannings roughness coefficient;
and
V = average velocity (m/s).
Suggested Mannings n values are 0.20
for regularly mown areas and 0.25 for
natural grasses or infrequently mown areas.
A minimum width of 15 m is recommended
for water quality enhancement.

Chapter 7
Environmental Consideration and Design

9. Where flow bypass is not incorporated,


peak velocities resulting from major
storm events should be determined
from hydraulic analysis.
10. If the estimated peak velocity is greater
than the determined erosive velocity of
the filter strip (refer to Chapter 8) then
the strip must be enlarged to
accommodate the flow. Once the flow
depth is established, the final
dimensions (including depth) of the
filter strip can be specified.
More specific
include:

design

considerations

The slope of the filter strip should be


uniform and the cross-section should
be a level plane to maintain sheet
flow.

5. Determine the flow area based on the


calculated flow width and established
flow depth.

If grass filter strips are located on


slopes lower than 2%, consideration
should be given to installing a subsoil
drainage system.

6. Calculate the resulting velocity.


Reduce the flow, increase the flow
width or reduce the depth of flow if the
velocity exceeds 0.3 m/s, which is the
velocity at which most grasses are
knocked over.

Flow entering the filter strip should


be evenly distributed as sheet flow
across the upstream end.
Level
spreaders should be provided to
ensure the filter strips does not
receive direct discharges.

7. Using the resulting velocity, calculate


the flow length to achieve a residence
time in the filter strip of nine minutes.
An absolute minimum residence time
should be five minutes. To maintain
sheet flow, the minimum length of a
filter strip will generally be 6 m.

Additional design information may be


obtained from the following design
references:

8. To avoid erosion of the filter strip


major storm events should preferably
bypass the filter strip. Typically the
major storm would be defined as the
ARI 50 or 100 year event.

March 2010

7-16

Healthy Waterways WSUD Technical


Design Guidelines (2006);
Camp Dresser and McKee (1993);
Horner et al (1994);
Schueler et al (1992); and
Standing Committee on Rivers and
Catchments (1991).

Department of Transport and Main Roads


Road Drainage Manual

7.5.2

Grassed Swales

7.5.2.1 Description
Grassed swales are grassed-lined flow
paths, often running adjacent to a road
pavement, which provide an alternative to
concrete kerbing and guttering. They can
also be used in road medians and verges. If
properly maintained, grassed swales can
reduce runoff volumes, attenuate storm
flows, enhance infiltration and improve
water quality.
Water quality enhancement occurs mainly
through the removal of coarse sediments
and
attached
particulates.
The
improvement in water quality is achieved
by increased settling, filtration by swale
vegetation and some removal of soluble
pollutants through infiltration into the
subsoil. Pollutants such as hydrocarbons
may be digested and processed by soil
micro-organisms within the swale.
Advantages of grassed swales include:
increase infiltration, thereby reducing
and delaying storm runoff;
retain particulate pollutants close to
the source;
enhance aesthetic appeal; and
reduce construction costs as grassed
swales are relatively inexpensive to
construct.
Disadvantages of grassed swales include:
limited removal of fine sediment and
dissolved pollutants;
require considerable land areas
compared with kerb and channel;
may interfere with driveways in
higher density development;
less effective for concentrated flows
and high flow depths;

Chapter 7
Environmental Consideration and Design

are only suitable for gentle slopes


(less than 5%);
require adequate maintenance to
avoid weed infestation, boggy base,
mosquitoes and soil erosion;
are in general most suitable for areas
with relatively highly permeable
soils; and
should not be used on sodic or
dispersive soils.
7.5.2.2 Design Guidelines
Design guidelines for grass swales follow
the same guidelines as those provided for
filter strips (Section 7.5.1) but with the
following amendments:
1. Grass swales should be located on
grades of 4% or less. However, slopes
of up to 6% can be adopted if small
check dams (or mounds) are located in
the swale every 15 m to 30 m to reduce
flow velocities. For slopes of 2% or
less, consideration should be given to
installing a subsoil drainage system to
ensure
effective
drainage
and
infiltration.
2. Recommended depth of flow is one
third of the grass height in infrequently
mowed swales; half the grass height, to
a maximum of 75 mm, in regularly
mowed swales.
3. Swales should be trapezoidal, with a
recommended bottom width between
0.6 and 2.5 m (Horner et al. 1994). If a
wider base is required, the flow could
be diverted into more than one swale.
The side slopes should not be steeper
than 1 on 3. If steeper slopes are used,
up to 1 on 2, permanent stabilisation
may be required. Triangular crosssections are not recommended, as the

March 2010

7-17

Department of Transport and Main Roads


Road Drainage Manual

flow can become channelised in the


bottom of the swale.
4. To maintain sheet flow, the minimum
length of a swale is generally 30 m.
More specific
include:

design

considerations

The base of the swale should be


level, with the longitudinal grade of
the swale either uniform or with
gradual changes only. Particular
attention should be paid to these
requirements during construction.

The integrity of a swale may be


impaired if flows greater than the
design event enter the swale.
Velocities exceeding the design
velocity can be expected to result in
reduced swale pollutant removal
efficiency until the grass has
recovered. Such flows may also
result in scouring of the swale. A
bypass for high flows could be
installed
to
prevent
large
concentrated flows eroding the
swales.
The depth to groundwater should be
considered when designing a swale.
If the water table is shallow, the grass
species will need to tolerate this
situation. Further, a shallow soil
depth for pollutant retention presents
a possible risk of pollution entering
the groundwater.

Chapter 7
Environmental Consideration and Design

Horner et al (1994);
Schueler et al (1992); and
Standing Committee on Rivers and
Catchments (1991).

7.5.3

Trash Racks

7.5.3.1 Description
Trash racks can be installed in drainage
channels or outlets to trap litter and other
gross pollutants. They generally comprise a
series of vertical or horizontal steel bars
which form a physical barrier to objects
larger than the bar spacings. An example of
a trash rack is shown in Figure 7.5.3.1(a).
Trash racks can be designed to be
perpendicular, angled or staggered to the
direction of flow. They can be located
either on-line or off-line. With an online arrangement, trash racks are placed
within an existing channel or drainage
system. With an off-line arrangement, a
flow diversion mechanism is installed,
which directs low and medium flows into
the trash rack while high flows bypass the
trash rack.

Additional design information may be


obtained from the following design
references:
Healthy Waterways WSUD Technical
Design Guidelines (2006);
NSW Department of Housing (1998);
Camp, Dresser and McKee (1993);

March 2010

7-18

Figure 7.5.3.1(a) - Trash Rack

Source: Brisbane City Council.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Advantages of trash racks include:


simple to construct;
trap all pollutants larger than the bar
spacing and also retains smaller
pollutants when the rack becomes
partially blocked;
can be retrofitted into
drainage systems; and

existing

collection of litter at a single point.


Disadvantages of trash racks include:
potential to cause upstream flooding
when material accumulates behind
the trash rack (refer Figure
7.5.3.1(b);
scouring at the base or sides of the
rack if adequate protection is not
implemented;
maintenance intensive, requiring
manual cleaning either on an as
needs basis or as part of the
programmed works for maintenance;
and
trapped material may be resuspended
during large storm events.

7.5.3.2 Design Guidelines


The most commonly used technique for
sizing trash racks is an approach described
in Department of Urban Services ACT
Government (1994).
1. The length of the trash rack should be
determined in conjunction with the
trash rack height and the available
space.
2. The height of the designed rack should
be such that the rack is not overtopped
by the design flood flow when 50%
blocked.
3. Under submerged conditions, the
required height of the trash rack is
twice the depth at critical flow through
the unblocked trash rack. For a trash
rack consisting of vertical 10 mm bars
at 40 mm spacing, this leads to:

Q
H 1.26
Lr

2/3

Where:
H= rack height (m);
Q= design flow (m/s); and
Lr= length of rack (m).

4. The presence of downstream hydraulic


controls can lead to submergence of the
trash rack. Under these conditions,
(Beecham and Sablatnig 1994):

v2

h k
2g
Figure 7.5.3.1(b) - Trash Rack and
Pollutants

and

A
k 1.45 0.45 n
A
g

An2

A2
g

March 2010

7-19

Department of Transport and Main Roads


Road Drainage Manual

Where:
h = headloss through the trash rack
(m);
k = rack coefficient;
v = average velocity through rack
(m/s) = Q/An.
g = gravitational acceleration (9.81
m/s2);
An = net area through bars (m); and
Ag = gross area of racks and supports
(m).
5. Trash racks have the potential to
exacerbate upstream flooding if
blocked. Hydraulic analysis should be
carried out to investigate impacts
arising from the rack being 50% and
100% blocked. The trash rack should
be assessed for the ARI 2 year, 5 year,
50 year and 100 year flood events.

Additional design information may be


obtained from Department of Urban
Services (1994), and Willing and Partners
(1992).

7.5.4

Proprietary Devices

Chapter 7
Environmental Consideration and Design

only. Propriety pollutant control devices


should be designed in accordance with the
manufacturers recommendations in order
to achieve the required pollutant control
objectives.
Before devices can be considered and their
suitability assessed, the pollutants to be
removed from a flow determined need to be
determined and the treated water quality
objectives set. This involves establishing
appropriate design criteria such as:
surface and underground drainage
gradients;
catchment area;
pollutant
characteristics
sediment diameter);
location of installation; and
requirements for maintenance access.
The assessment and selection of appropriate
proprietary
devices
should
include
consideration of:
ability to meet required water quality
objectives;
proven ability to achieve the desired
pollutant removal rate;

7.5.4.1 Description

capital costs;

There are a number of manufactured


devices, which are designed to remove
specific pollutants such as coarse sediment,
oil, grit and hydrocarbons from runoff.
Each proprietary device is specifically
designed to treat one or more pollutants
associated with stormwater. The majority
of these devices are designed to be located
underground, as part of the stormwater
network.

construction materials;

7.5.4.2 Design Guidelines


The design considerations outlined in this
section are for general guidance purposes

March 2010

7-20

(e.g.

installation procedures;
drainage design criteria:
o head loss
o crossfall
o hydraulic capacity
maintenance procedures such as:
o access
o frequency and cost
o spills

Department of Transport and Main Roads


Road Drainage Manual

o disposal of pollutants
o inspection of control device.

7.5.5

Constructed Wetlands

7.5.5.1 Description
Constructed wetlands are structures built
with predominantly natural materials to
reproduce the physical, chemical and
biological processes of natural wetlands.
They are used to remove a range of
pollutants, including suspended solids,
nutrients, heavy metals and other toxic or
hazardous compounds.
Their pollutant
trapping efficiency varies with the type of
pollutant, being moderate for oil and
grease, moderate to high for sediments and
nutrients.
Wetlands typically comprise an upstream
inlet zone, a shallow macrophyte zone and
a high flow bypass channel. The upstream
inlet zone consists of a relatively deep,
open water body or sediment basin with
some fringing aquatic vegetation. The
downstream macrophyte zone is a more
permanent shallow water body with
extensive vegetation. The bypass channel
is used to protect the macrophyte zone from
scour and vegetation damage (Healthy
Waterways WSUD Technical Design
Guidelines (2006)).
Advantages of wetlands include:
potentially achieve high sediment
and nutrient retention efficiencies;
be incorporated into the road
corridor, thereby providing improved
habitat and visual amenity in
disturbed areas; and
potentially be retrofitted into existing
sediment ponds / or detention basins.

Chapter 7
Environmental Consideration and Design

Disadvantages of wetlands include:


either pre-treatment or removal
mechanisms at the inlet to remove
coarse sediment and litter;
large areas for construction; and
a reliable inflow to ensure they
remain wet, unless the wetland is
designed to be ephemeral.
In addition, wetlands may:
have a treatment performance which
is highly sensitive to hydrologic and
hydraulic design;
take up to three years to achieve
optimal performance;
have a potential impact on public
health and safety; and
have adverse interactions (pollutant
exchange) with groundwater in some
situations.
Constructed wetlands generally require
large areas of land, resulting in high
construction costs. Maintenance cost of
wetlands can be kept relatively low if the
design provides for mechanised sediment
removal facilities as well as the inclusion of
upstream pre-treatment devices for the
trapping of coarse sediment and litter.
7.5.5.2 Design Guidelines
Victorian Stormwater Committee (1999)
outlines a number of design principles to
consider before construction.
These
include:
establish a uniform flow distribution
throughout the wetland.
Avoid
creating stagnant areas;
to enhance sedimentation, maximise
the amount of time macrophytes are
in contact with flow. This can be

March 2010

7-21

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

groundwater conditions; and

done by providing low flow


velocities with healthy vegetation;
provide adequate
treatment;

wetland

pre-

minimise organic matter loading; and


maintenance requirements must be
met in order to manage sediment
build up and weed occurrence.
The design procedure outlined below
outlines five general steps in the design of
constructed wetlands and the reference, The
Constructed
Wetlands
Manual

Department
of
Land
and
Water
Conservation (NSW) 1998, is relevant to
these five design steps.
1.

Wetland Location

Constructed wetlands can be located on a


watercourse or adjacent to a watercourse.
For road corridors, the preferred location is
adjacent to a watercourse. In this case, the
drainage system of the road can be designed
to direct runoff from the local catchment
and pavement surface into the wetland,
rather than directly into the watercourse. A
high flow bypass channel around the
wetland system can be designed such that
sediments and vegetation within the
wetland are not damaged during flood
events.
The location of a wetland will depend on a
number of factors, including:
aquatic
habitat
vegetation
of
environment;

and
the

riparian
receiving

wetland size and available space in


the road reserve;
topography;
reliability of low flows to maintain a
permanent wetland;

March 2010

7-22

maintenance requirements.
2.

Wetland Size

The components of the wetland that need to


be sized are the temporary and permanent
storage volumes. The permanent storage
zone encourages biofilm growth on the
macrophytes (plants such as reeds, rushes
and sedges) and maintains sedimentation.
The temporary storage zone can be used to
attenuate peak flows and to increase
hydraulic
residence
time,
thereby
maximising the rate of pollutant removal.
As a rule of thumb, wetland areas in South
East Queensland should cover 2% - 4% of
the contributing catchment area.
3.

Pre-Treatment Measures

The removal of coarse sediment upstream


of the wetland will minimise changes to
depth profile and damage to macrophytes.
The installation of a sediment trap and trash
rack upstream of the wetland are
recommended and will assist greatly in
reducing the frequency and cost of
maintenance.
4.

Macrophyte Planting Requirements

A wetland should be divided into a number


of zones to encourage macrophyte
diversity.
An open water zone will
encourage
UV
disinfection
and
oxygenation. Shallow marsh, marsh and
deep marsh zones encourage macrophyte
diversity and uniform flow across the
wetland.
5.

Outlet Structure

Outlet structures are important as they


control the water level within the wetland.
An appropriate water level optimises water
quality improvement, achieves macrophyte

Department of Transport and Main Roads


Road Drainage Manual

diversity and provides for weed and


mosquito control.

Chapter 7
Environmental Consideration and Design

Low costs
maintenance.

and

requires

low

Four outlet varieties discussed in Victorian


Stormwater Committee (1999) are risers,
weirs, culverts and siphon outlets. These
outlet devices should be assessed to
determine the appropriate water level or
hydraulic regime each would have on a
wetland. The choice of outlet type will also
be influenced by the basin morphology and
hydrology.

Disadvantages of a buffer zone include:

Additional design information may be


obtained from:

The design procedure below provides six


general steps for the design of buffer zones.

Healthy Waterways WSUD Technical


Design Guidelines (2006);
Victorian Stormwater
(1999); and

Committee

Guidelines for Stabilising Waterways


(SCRS 1991).

7.5.6

Buffer Zones

7.5.6.1 Description
Buffer zones are areas that are left
undisturbed to provide some filtering and
trapping of sediment (and therefore some
heavy metals). Buffer zones are usually
established upstream of sensitive receiving
environments, such as natural wetlands,
streams or bushland. Buffer zones preserve
the existing vegetation and landscape, may
be used as habitat corridor for wildlife and
contribute to preserving biodiversity. They
may require fencing to exclude traffic and
to prevent damage to the vegetation.
Advantages of a buffer zone include:
Applicable to all development areas;
Able to reduce runoff volume by
30% to 50% more than grass filter
strips; and

Have a limited ability to reduce


pollutant loads;
Only trap coarse sediments; and
Are suitable for slopes between 1%
and 10%.
7.5.6.2 Design Guidelines

1. Buffer Zone Dimensions


The performance of a buffer zone generally
increases with increasing width and
decreasing slope.
As a general guide, the width of a buffer
zone (expressed in m) should be the greater
of: five times the slope (expressed as
percentage (e.g. 50m buffer width on a 10%
slope) or 1.5 times the width of the
disturbed area from which the sediments
are mobilised. The latter generally applies
to grassed buffers adjacent to the road
pavement. In some locations these design
rules may not need to be applied.
Grassed Buffer Zones
For maximum efficiency, the depth of sheet
flow should not exceed the grass length.
Grassed areas should be maintained with a
minimum grass length of 50 mm (welltrimmed lawns are less effective although
they tend to be more sought after).
2. Performance Monitoring Requirements
Buffer zones adjacent to high quality
watercourses and environmentally sensitive
areas should be wide enough to trap all
visible sediment within the first quarter of
the buffer zone width.

March 2010

7-23

Department of Transport and Main Roads


Road Drainage Manual

If buffer zones are to be effective, at least


75% of the ground should be covered by
vegetation and weed growth should be
controlled.
3. Slopes

can generally be constructed at


steeper sites than constructed
wetlands.
Disadvantages of water quality ponds
include:

Buffer zones are recommended for the


control of sheet flow on slopes between 1%
and 10%;

can be prone to eutrophication and


thus have an adverse impact on
downstream water quality;

4. Fencing

have the
mosquitoes;

Fencing can be used to exclude traffic from


buffer zones, thus preventing damage to the
vegetation and surface rutting; and
5. Planning Buffer Zones
Buffer zones should be incorporated into
the final landscaping plan and should be
constructed (or retained) early in the
development program.

7.5.7

Chapter 7
Environmental Consideration and Design

Water Quality Ponds

7.5.7.1 Description
A water quality pond is a relatively deep
open body of water, possibly with littoral
macrophytes (reeds). Wet basins achieve
pollutant removal through sedimentation.
Their pollutant removal efficiency depends
on the stormwater residence time and the
amount of runoff detained in the basin.
Pollutant removal efficiency increases with
longer residence times and greater utilised
storage volumes.
Advantages of water quality ponds
include:
can be used to trap coarse sediments
and associated pollutants;
have the potential for stormwater reuse;
have a potentially high aesthetic or
recreational value;
provide habitat for wildlife; and

March 2010

7-24

potential

to

breed

may cause habitat degradation


upstream and downstream of the
basin; and
may require flocculation.
Large pond volumes may be required in
regions with high rainfall intensities.
7.5.7.2 Design Guidelines
Ideally, a continuous simulation approach
should be undertaken due to the highly
variable nature of catchment runoff and
associated pollutant concentration. From
the continuous simulation an appropriate
storage volume could be selected on the
basis of long term performance rather than
prescribed performance for a single event.
The size and capacity of water quality
ponds should be such that stormwater is
detained for as long as possible to promote
effective treatment of pollutant, but should
also guarantee that runoff generated during
subsequent storms is captured and treated.
The longer the residence time and the more
water stored in the pond, the better the
pollutant treatment.
Ideally, the pond should be protected from
flood flows larger than the design storm
flow. Provision of a high flow bypass
channel is a means of reducing the risk of
pond scouring but it is subjected to the site
topographical constraints.

Department of Transport and Main Roads


Road Drainage Manual

To enhance pollutant removal, the


following design features should be
considered:
an effective residence time can be
achieved by a pond length to width
ratio of between 3:1 and 5:1;
the inlet to the pond should be
located as far as possible from the
outlet as possible; and
to increase the length to width ratio
or overcome problems with the inlet
being too close to the outlet, berms
and baffles may be installed to
redirect flows.
The proper design of a pond outlet is
critical to its performance. One option is to
place a low level culvert at the Mean
Operating Level (MOL) of the water
quality pond. Between storms, the water
level in the pond will drop below the MOL
because of evaporation and infiltration
losses. During storms, the water level will
rise to and beyond the MOL, and water will
flow out of the pond through the outlet.
Other outlet options include using orifice
outlets, which may enhance flow detention
during smaller storms, or broad crested
weirs.
With the latter option, flood
attenuation may not be as effective as with
culvert or orifice outlets.
An access track should be provided for the
maintenance of water quality ponds.
Maintenance may include the following:
mowing of banks and harvesting of
macrophytes;
weed removal;
litter removal; and

Chapter 7
Environmental Consideration and Design

intervals not exceeding six months to assess


the performance.
Additional design information may be
obtained from Environmental Protection
Agency (NSW) (1997), Horner et al (1994),
and Schueler et al (1992).

7.6

Fauna Passage

Recognition of the impacts of road corridor


development on fauna populations has led
to modifications in the way that roads are
now designed.
A substantial amount of research has been
undertaken to develop practices that
facilitate fauna movement through the road
corridor in a way that minimises road
mortalities. Much of the research has
focussed on passages that are integral with
drainage structures.
As such, the provision of fauna passage is
one of the key environmental factors which
may influence the physical dimensions of a
drainage structure.
This section provides an overview of what
steps need to be taken when the project
environmental assessment process has
identified a need for fauna passage to be
incorporated into drainage design.
Completion of these steps may result in the
determination of additional design criteria
which may then influence the selection of
cross drainage structures.
Further reading on this topic is available in
Fauna Sensitive Road Design Manual
Volume 1 (DMR 2000).

7.6.1

Identifying Fauna Passage


Criteria

removal of accumulated sediment.


Monitoring of a water quality pond should
be undertaken after large storm events at

When a project environmental assessment


has identified fauna passage requirements,
the Regional / District Environmental
March 2010

7-25

Department of Transport and Main Roads


Road Drainage Manual

Officer should undertake the following


steps:
(a) Identify terrestrial and aquatic fauna
pathways and areas of high mortality
on the road.
(b) Identify the species groups of concern.
(c) Consult with the relevant authority.
(d) Identify criteria affecting drainage
design (for example, ensuring dry
fauna passage at all times).

7.6.2

Identify Terrestrial and


Aquatic Fauna Pathways

A review of the project environmental


assessment should be undertaken to check
for the presence of any significant
terrestrial and/or aquatic fauna movement
pathways which could be potentially
affected by the proposal.

If fauna pathways have been identified or


are likely to exist in the study area, proceed
to identifying the species groups of
concern. If not, document the outcomes of
the environmental assessment review and
continue identifying other relevant drainage
design criteria. Bridges provide a good
solution to maintaining terrestrial and/or
aquatic fauna movement pathways as
shown in Figure 7.6.2.

Chapter 7
Environmental Consideration and Design

7.6.3

Identify the Species Group

Where fauna pathways have been


identified, identify the relevant species
group from the list below:
Fishes
Amphibians (frogs)
Mammals: Macropods
Mammals: Arboreal species (e.g.
possums, gliders, etc.)
Mammals: Koalas
Mammals: Platypus
Mammals: Bats / flying foxes
Mammals: Small-size (e.g. echidnas)
Birds (flying and ground-dwelling)
Reptiles (e.g. snakes, lizards and
turtles)
Invertebrates (insects and spiders)

7.6.4

Consult with the Relevant


Authority

The Regional / District Environmental


Officer should consult with the Department
of Environment and Resource Management
and/or the local authority to discuss
potential impacts of the project on the
known fauna species and their movement
patterns. The publication manual, Fauna
Sensitive Road Design (DMR 2000), should
also be consulted.
A listing of key authorities and relevant
legislation is provided in Table 7.6.4.

7.6.5

Figure 7.6.2 - Fauna Corridor under


Bridge.

March 2010

7-26

Identify Criteria affecting


Drainage Design

In consultation with the relevant authorities


and from information provided in the

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

environmental assessment (i.e. baseline


fauna studies), determine the following:

in consultation with the Project


Manager and Designer;

Specific
characteristics
and
requirements for the species of
concern (i.e. movement patterns,
habitat range);

The need for further preliminary


design work (i.e. implications to cut
and fill balance if a different culvert
size is required, revised dimensions,
etc); and

Opportunities to facilitate safe


passage of fauna through the
drainage system (i.e. bridging
options, culvert modifications) (refer
Table 7.6.5, and the Fauna Sensitive
Road Design Manuals Volume 1 and
2) (DMR 2000);

Specific design criteria for drainage


structures (i.e. allowable flow
velocity, minimum culvert height,
dry passage requirements, etc).

The economic and engineering


feasibility of potential opportunities

Table 7.6.4 - Key Environmental Authorities for Protection of Fauna


Species
Fish and other aquatic
fauna
Species and
communities of
national significance*
Other fauna

Relevant Authority
Department of Employment, Economic
Development and Innovation (was
Primary Industries And Fisheries QLD)
Department of Environment, Water,
Heritage and the Arts
Department of Environment and
Resource Management (was QLD
Environmental Protection Agency)

Relevant legislation
Fisheries Act 1994 (QLD),
Fisheries Management Act 1991
(Commonwealth)
Environmental Protection and
Biodiversity Conservation Act 1999
(Commonwealth)
Nature Conservation Act 1992
(QLD),
Environmental Protection Act 1994
(QLD),
Nature Conservation (wildlife
management) Regulation 2006,
Nature Conservation (wildlife)
Regulation 2006

* For a complete list of species and communities that are protected under the environmental
protection
agency
can
be
obtained
from
the
EPBC
website
http://www.environment.gov.au/epbc/about/lists.html#species (as of May 2009)

March 2010

7-27

Department of Transport and Main Roads


Road Drainage Manual

Chapter 7
Environmental Consideration and Design

Table 7.6.5 - Confirmed Use of Culverts or Underpass Types by Fauna


Fauna Type

Small pipe
<0.5m dia

Large pipe
>0.5m dia

Small box
culvert <1.2m
high

Large box
culvert <1.2m
high

Bridge
underpass

Small mammal

Medium mammal

Large mammal

Semi-arboreal
mammal

Arboreal mammal

(large only)

Bats

(adapted roof
structure)

(adapted
roof
structure)

Reptile

Bird

Amphibian

Introduced predator

Source: Queensland Department of Main Roads, 2000.


Caution: This table is based on preliminary research only. Although not confirmed at the time, fauna should pass
through all the culverts larger than the minimum ones shown. Recommended minimum sizes for design are
shown in Section 4.2.2.3 of this Manual

March 2010

7-28

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Chapter 8
Open Channel
Design
8

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Chapter 8 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Table of Contents
8.1

Introduction
8.1.1

Open Channels

8-1
8-1

8.2

General Considerations

8-2

8.3

Local Authority Requirements

8-2

8.4

Fundamentals of Open Channel Flow

8-3

8.4.1

Stream Dynamics

8-3

8.4.2

Assumptions for Analysis

8-4

8.4.3

Fundamental Equations

8-4

8.4.4

Application of Fundamental Equations

8-9

8.4.5

Energy Principles

8-11

8.4.6

Hydraulic Jump

8-14

8.4.7

Hydraulic Drop

8-20

8.5

Erosive Velocities in Natural Streams

8-20

8.6

Backwater

8-20

8.7

8.8

8.9

8.6.1

Tidal Waters

8-21

8.6.2

Downstream Tributary

8-24

Tailwater Levels

8-24

8.7.1

Tailwater Effects

8-24

8.7.2

Design Tailwater Levels

8-25

Open Channel Design

8-26

8.8.1

Design Methodology

8-26

8.8.2

Channel Transitions

8-29

8.8.3

Energy Losses in Channel Bends

8-30

8.8.4

Superelevation in Channel Bends

8-30

Grassed Channels

8-31

8.9.1

Normal Grassed Channels

8-31

8.9.2

Reinforced Grassed Channels

8-32

8.10 Channels Lined with Hard Facings

8-33

8.10.1

General

8-33

8.10.2

Riprap and Rock Filled Wire Mattresses / Gabions

8-33
March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

8.10.3

Concrete Lined Channels

Chapter 8
Open Channel Design

8-36

8.11 Channel Drops

8-36

8.12 Baffle Chutes

8-36

March 2010

iv

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Chapter 8
Open Channel Design
8.1

Introduction

Open channels are economical where large


flows are to be carried and space is not
restricted.
An example on a newly
constructed, open channel is show in Figure
8.1.
They have the advantages of
continuous collection of surface runoff, and
where the system surcharges, general
shallow flow is the most likely outcome
rather than more concentrated flooding at
upstream inlets of the closed drainage
system.
This manual focuses on the analysis and
design of smaller streams and creeks.
Assessment of larger streams, creeks, rivers
and floodplains is complex and should be
referred to Hydraulics Section, Engineering
& Technology Division or a suitably prequalified consultant.

atmosphere and/or there is no additional


pressure on the flow other than atmospheric
pressure. Flow is caused by gravity and
streams tend to follow the path of least
resistance.
Open channels may be constructed to
specified criteria:
as part of the road drainage system
where space within the road reserve
is sufficient to provide for open
channels;
as diversion channels, especially
where the road is being constructed
generally along the line of a
watercourse and severs one or more
meanders in the stream. Care must
be exercised as shortening of the
stream will increase the gradient and
hence velocity, which may induce
scouring and also prevent the
upstream passage of fish; and/or
from the outlets of culverts or
drainage systems.

Figure 8.1 - Bare Earth Open Channel

8.1.1

Open Channels

An open channel is defined as a conduit or


conveyance (artificial or natural) in which
water flows with a free surface. A free
surface means that the surface is open to the

Creeks and waterways, disturbed by


construction work, need to be left in a way
that the channel will continue to behave in a
hydraulic and ecological manner similar to
that of the undisturbed natural stream. New
open channels must be appropriately
integrated into the surrounding ecological,
visual, social and physical environments.
Department of Environment and Resource
Management and relevant local authorities
should be consulted early in the planning
process as they may have special

March 2010

8-1

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

requirements for existing channels which


could be:
part of soil conservation programs;
stream rehabilitation or restoration
plans;
riparian improvement programs;
part of natural
programs; or

channel

other
ecologically
programs.

8.2

design

sustainable

General
Considerations

Considerations in open channel design


include:
(a) Establishing
the
ecological
requirements for the channel such as:
determining the specific or generic
riparian management requirements;
stream bank stability or erosion
guidelines;
water quality requirements;

natural channel design requirements;


use of the channel by fauna as a
fishway or wildlife corridor.
(b) Evaluating the most economical route
consistent with the land use and
topography;
(c) Optimising the slope of the channel.
To reduce the velocity, a flatter slope
than that existing may be adopted, by
introducing
drop
structures
or
meanders. However, shortening the
natural channel alignment will increase
velocities;
(d) Determining the width available for an
open channel in the road reserve or
drainage easement;

March 2010

8-2

(e) Assessing the need for channel lining,


particularly for erosion resistance,
including vegetation selection and
aesthetics.
Ensure
future
maintainability of lining is considered;
(f) Calculating
flow
characteristics.
Design for subcritical flow (refer
Sections 8.4.5.2 and 8.4.5.3) is
recommended.
In only a small
percentage of natural streams under
certain conditions, flow is not
subcritical. Supercritical flow (refer
Sections 8.4.5.2 and 8.4.5.3) with its
highly erosive velocities should be
avoided in a design flood except at
channel drops or energy dissipators
(refer Chapter 11);
(g) Assessing
the
safety
or
the
appropriateness of an open channel in
certain locations, particularly in
developed areas;
(h) Providing access for maintenance.

8.3

Local Authority
Requirements

The requirements of local government


should be sought for any significant open
channel design particularly in developed
areas or where the channel forms part of a
major drainage system as defined in
Chapter 2.
There is a need to check the effects of flows
in excess of the design flood and to
understand the likely impacts.
Where
flooding of property is a possibility, the
effects of a flood in excess of the design
flood should be analysed even if there is no
local government requirement.
Other local government requirements may
include the provision of access /
maintenance berms, barrier fencing and
appropriate warning signs to prevent access

Department of Transport and Main Roads


Road Drainage Manual

to the channel and a low flow channel to


take a minimum flow.
Water Sensitive Urban Design (WSUD) is
an important aspect of pollution control and
the design of open channels should
consideration these requirements. Chapter
2 outlines the principles of WSUD while
Chapter 7 provides more information and
detail.

8.4
8.4.1

Fundamentals of Open
Channel Flow
Stream Dynamics

It is important to have a basic


understanding of how streams work or
behave to ensure that the planning and
design of drainage infrastructure is
appropriate and will work not only after the
work is completed, but also in the future.
Streams are constantly changing and
responding to external factors and are
therefore dynamic. However, to simplify
analysis and design, streams are generally
considered as static, that is the stream is
divided into small, isolated parts and/or
sections and the flow within the channel is
analysed at a point in time.
8.4.1.1 Streams and Roads
Human interference can easily change /
affect streams and it is critical to understand
that changes can affect stream behaviour
both upstream and downstream.
Typical encroachments that would affect
normal stream flow as related to road
infrastructure are:

Chapter 8
Open Channel Design

bridges.
8.4.1.2 Governing Parameters
All streams are governed by the same basic
parameters:
(a) Geological Factors
soil conditions
(b) Hydrological Factors
rainfall
(c) Geometric
channel slope and crossfall
(d) Hydraulic
fluid properties
(e) Vegetation
Native vegetation in channels / along
banks
8.4.1.3 Velocity Distribution
The flow velocities within a stream are not
uniformly distributed across the cross
section. This is due to friction along the
channel bed, banks and free surface.
The velocity of water particles near the bed
and banks (planes of highest friction) is the
slowest while the highest velocity is
typically found just below the free surface
in the deepest part of the flow.
Figure 8.4.1.3 illustrates the velocity
distribution across channel flow. To map
water particles of equal velocity, isovels are
drawn (isovels trace equal velocity similar
to contour lines tracing equal ground
height).

roadway embankments;
cross drainage culverts;
floodways; and

March 2010

8-3

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Turbulent Flow is where the water


particles move in very irregular
paths. White water rapids are a good
example of turbulent flow.
Open channel flow can be in a variety or
combination of these types and states.
Figure 8.4.1.3 Velocity Distribution
Across the Flow

8.4.1.4 Types and States of Flow


Open channel flow can be described in a
number of ways. The following outlines
the types of flow:
Steady Flow is where the depth of
flow at the same point does not
change over time.
Unsteady Flow is where the depth
of flow at the same point does change
over time.

Uniform Flow is where the depth


of flow is the same at every cross
section along the channel. This is
really only possible when the channel
flow cross section, slope and
roughness remain constant.
Non Uniform or Varied Flow - is
where the depth of flow changes
along the length of channel. Varied
flow can be further described as
either rapid or gradual (i.e. in
artificial channels).
The following outlines the states of flow:
Laminar Flow is where the water
particles appear to move in smooth
paths or in very thin sheets (laminas)
over the top of each other. Laminar
flow is very smooth flow and is
rarely
seen
in
the
natural
environment.

March 2010

8-4

8.4.2

Assumptions for Analysis

Due to the dynamic nature of open channel


flow, it is difficult to model and/or analyse
the flow as it occurs in the natural
environment.
Therefore a number of
assumptions are made to enable and
simplify analysis. These assumptions are
considered reasonable and valid in the
context of the analysis and design
procedures described in this manual.
The assumptions made are:
Steady Flow;
Uniform Flow;
Velocity is averaged over the whole
cross section; and
Flow is non-turbulent.
Generally, these assumptions can be applied
to most situations encountered in design.
However, where it is obvious that the
assumptions cannot be applied, the analysis
and/or design should be referred to
Hydraulics Section, Engineering &
Technology Division or a suitably
prequalified consultant.

8.4.3

Fundamental Equations

The most basic equation in the analysis of


open channel flow defines the relationship
between flow rate, velocity and the cross
sectional area of flow and is represented by
the following formula:

Department of Transport and Main Roads


Road Drainage Manual

Q V A

P = wetted perimeter (m).

Where:
Q = flow rate (m3/s);
V = average velocity (m/s); and
A = cross-sectional area of flow (m2).
This equation also forms the basis of the
theory behind the Continuity Equation.
The theory allows simple analysis over
changes in the channel irrespective of cross
section, slope or roughness. The theory
assumes no addition to or subtraction from
the flow between the 2 sections being
considered (i.e. Q1 = Q2).
The Continuity Equation is:

Q1 V1 A1 V2 A2 Q2
The formula most commonly used for the
calculation of steady, uniform flow in open
channels is Mannings Equation. This
equation is used to determine the velocity
of flow at a specific point in the channel,
and therefore the variables in the equation
must be representative of the point being
assessed.
Mannings Equation is:
2

R S
n
3

Where:
V = average velocity (m/s);
R = hydraulic radius (m);
S = slope of energy line (m/m); and
n = Mannings roughness coefficient.
The hydraulic radius is given by:

Chapter 8
Open Channel Design

A
P

Where:
A = cross-sectional area of flow (m2);

In determining the hydraulic radius, the


wetted perimeter is defined as the length of
line (normal to the flow) where the water
touches the surface of the ground (channel)
or the perimeter of flow less the surface
(exposed to the atmosphere) length (refer
Figure 8.4.3(a)).

Figure 8.4.3(a) Wetted Perimeter

For natural channels the slope of the energy


line, S, is almost impossible to determine.
Therefore a suitable estimation of S is
required. The slope of the water (flow)
surface, Sw, could be used, however this is
also difficult to determine. The only easily
available slope that can be used to estimate
S is the channel bed, So, provided that the
channel and the bed slope about the point
being assessed is reasonably uniform. If the
channel shape and bed slope is irregular,
Mannings Equation will not give
acceptable results and therefore the analysis
/ design should be referred to Hydraulics
Section, Engineering & Technology
Division or a suitably pre-qualified
consultant.
In
applying
Mannings
Equation,
particularly to natural channels, the greatest
difficulty lies in the determination of the
roughness coefficient n.
For artificial
channels values of n can be obtained from
Table 8.4.3(a).
For natural channels, a comparison may be
made with the photographs in Figure
8.4.3(b) and values of n in Table 8.4.3(b)
for similar channels. However, for grassed
channels, and where the hydraulic radius
March 2010

8-5

Department of Transport and Main Roads


Road Drainage Manual

for the flow is less than 1 m, vegetal


retardance curves (Figures 8A.1 to 8A.5 in
Appendix 8A) should be used to determine
n. Table 8.4.3(c) is used to determine the

Chapter 8
Open Channel Design

appropriate vegetal retardance curve to be


used.

Table 8.4.3(a) - Mannings n Values for Artificial Channels / Conduits


Type of Structure

1. Closed Conduits
(a) Concrete pipe
(b) Corrugated metal pipe or pipe-arch (small corrugation)
(i) plain or unpaved
(ii) paved invert-full flow
25% circumference paved
50% circumference paved
Fully paved

0.013
0.024
0.021
0.018
0.012

(c) Structural plate pipe or pipe-arch

0.030-0.033

(d) Monolithic concrete (box culvert)

0.012

(e) Vitrified clay pipe

0.012

2. Open channels - Lined


(a) Concrete - Smooth forms or trowelled

0.012

(b) Bituminous Concrete


(i) Smooth
(ii) Rough

0.013
0.016

3.Open Channels - Excavated (Straight alignment and natural lining)

(a) Earth - uniform section


(i) clean - new to weathered
(ii) with short grass, few weeds
(iii) in gravelly soil, clean

0.016-0.020
0.022-0.027
0.022-0.025

(b) Earth - uniform section


(i) no vegetation
(ii) grass, some weeds
(iii) dense weeds or plants in deep channel
(iv) sides clean, gravel bottom
(v) sides clean, cobble bottom

0.022-0.025
0.025-0.030
0.030-0.035
0.025-0.030
0.030-0.040

c) Dragline excavated or dredged


(i) no vegetation
(ii) light bush on banks

0.028-0.033
0.035-0.050

(d) Channels not maintained, weeds and brush uncut


(i) dense weeds, high as flow depth
(ii) clean bottom, brush on sides
(iii) same, highest stage of flow
(iv) dense brush, high stage

0.08-0.12
0.05-0.08
0.07-0.11
0.10-0.14

Source: Based on Bureau of Public Roads, 1965. Item 1(b) and 1(c) have been added based on values
recommended by Maccaferri in their publications.

March 2010

8-6

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Table 8.4.3(b) - Mannings n Values for Natural Channels


Type of Channel

Main Channel
1.

2.

Fairly Regular Section


(a)
Some grass and weeds, little or no brush
(b)
Dense growth of weeds, depth of flow materially greater than weed height
(c)
Some weeds, light brush on banks
(d)
Some weeds, heavy brush on banks
(e)
Some weeds, dense willows on banks
(I)
Trees within channel with branches submerged at high stage
Irregular Section, with pools, slight channel meander
To (a) to (f) above as applicable

0.0300.035
0.0350.05
0.0350.05
0.050.07
0.060.08
Add 0.01
0.02
Add 0.01
0.02

3.
Mountain Streams, no vegetation in channel, banks usually steep, trees and brush
along banks submerged at high stage
(a)
Bottom, gravel, shingle and few boulders
0.040.05
(b)
Bottom, shingle with large boulders
0.050.07
Adjacent Flood Channels
1.

Pasture, no Brush
(a)
Short grass
(b)
High grass

0.0300.035
0.0350.05

Cultivated Areas
(a)
No crop
(b)
Mature row crops
(c)
Mature field crops

0.030.04
0.0350.045
0.040.05

3.

Heavy Weeds, Scattered Brush

0.050.07

4.

Light Brush and Trees

0.060.08

5.

Medium to Dense Brush

0.100.16

6.

Dense Willows, Summer, not bent over by Current

0.150.20

7.

Cleared Land with Tree Stumps (250 to 450 per ha)


(a)
No sprouts
(b)
With heavy growth of sprouts

0.040.05
0.060.08

2.

9.

Heavy Stand of Timber, A Few Fallen Trees, Little Undergrowth


(a)
Flood depth below branches
(b)
Flood depth reaches branches

Major Streams (Surface width at flood stage > 30 m)


Roughness coefficient is usually less than for minor streams of similar description on
account of less effect offered by irregular banks or vegetation on banks. Values of n may
be somewhat reduced. The value of n for large streams of mostly regular section may be
in the range.

0.100.12
0.120.16
0.120.16

Source: Based on Austroads (1994)

March 2010

8-7

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

n = 0.03

n = 0.04-0.045

n = 0.05-0.06

8
n = 0.07

n = 0.08

Notes:
1) Increase in n value with an increase in grass, weeds, shrubs and trees.
2) In general, growth of trees in photographs tends to look denser than when seen on a site
inspection.
3) Except for n = 0.03, roughness is for bank full flood heights and/or floods in upper branches of the
trees.
4) Use photographs with caution. Use in conjunction with Table 8.4.3(b)
Figure 8.4.3(b) - Natural Streams in Queensland

March 2010

8-8

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Table 8.4.3(c) - Guide to the Selection of Vegetal Retardance


Density of stand
(see Note 1)

Average length of
vegetations (mm)

Degree of
retardance

Examples

Thick

Longer than 750

Rhodes grass in ungrazed scrub soil waterway

Thick

280 to 610

Wheat 660mm tall in 180mm rows Rhodes grass

Kikuyu under maximum fertility conditions, long


and green

African star grass. Lucerne (see Note 2)

Thick

150 to 250

Most grasses can be hold at this retardance with


mowing or grazing e.g. Rhodes grass, couch
grass, carpet grass, native grasses

Thick

50 to 150

African star grass, kikuyu or couch grass, all


under heavy grazing

Thick

Less than 50

Mowed lawn. Any grass burned short

Fair

Longer than 750

Fair

280 to 610

Rhodes grass under low fertility conditions

Fair

150 to 250

African star grass under low fertility conditions.

Fair

50 to 150

Fair

Less than 50

Notes:
1. Thickness of vegetation has an important bearing on retardance, possibly more important than species.
2. Vegetal retardance curves have been based on tests in experimental channels.
3. Values of n = 0.15 plus have been used for fields of standing sugar cane.
4. Refer to Appendix 8A for nomographs using different vegetal retardances.

Source: Based on Soil Conservation Handbook (DPI 1978) and Rouven et al., (1981)

It is important to understand that the depth


and velocity of any given flow in a channel
operating under normal or natural
conditions is called the Normal Depth and
Normal Velocity. The depth and velocity
of flow as determined using Mannings

Equation are the Normal Depth and Normal


Velocity

8.4.4

Application of
Fundamental Equations

The primary usage of the fundamental


equations is to determine the basic flow

March 2010

8-9

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

characteristics of a given channel. These


characteristics are:
maximum flow capacity of the
channel;
velocity of flow when at maximum
capacity;
depth of flow
discharge; and

for

specific

velocity of flow for a specific


discharge.
To determine these characteristics, a
representative cross section for the channel
needs to be chosen as well as channel slope
(So) and Mannings n value.
Mannings equation gives the velocity of
flow in a channel based on a selected depth
of flow (which enables calculation of the
cross-sectional area of flow and wetted
perimeter). To determine the flow rate (Q),
for the selected depth of flow and using
velocity as determined by Mannings
equation, use the fundamental equation:

Q V . A rewritten as V Q

Figure 8.4.4(a) Stage-Discharge Curve

To use the Stage-Discharge curve to


determine the depth of flow for a specific,
discharge, plot the required flow rate (Qy)
onto the left y-axis and extend across the
graph to the flow line and then project
down to x-axis. To determine the velocity
of this flow, the cross sectional area of the
flow needs to be calculated using the
determined depth of flow.
However, the Stage-Discharge curve can be
easily modified to include velocity by
plotted the graph using depth of flow for
the x-axis, the flow rate for the left y-axis
and velocity for right y-axis (refer Figure
8.4.4(b)).

If the selected depth of flow was based on


the maximum capacity of the channel, the
first two characteristics, as listed above,
have now been determined.
To determine the last two characteristics, a
Stage-Discharge curve (also known as a
rating curve) is required.
A StageDischarge curve plots the depth of flow
against the flow rate at a particular cross
section along a stream / channel. To plot a
Stage-Discharge curve, a series of iterations
of the above calculations / process is
required, with depths selected from zero to
maximum capacity.
The curve is plotted using depth of flow for
the x-axis and flow rate for the y-axis (refer
Figure 8.4.4(a)).
March 2010

8-10

Figure 8.4.4(b) Modified StageDischarge Curve

To use the Modified Stage-Discharge curve


to determine the depth and velocity of flow
for a specific discharge, plot the required
flow rate (Qy) onto the left y-axis and
extend across the graph to the flow line and

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

then project down to x-axis, to determine


depth, and up / down to velocity line and
then across to the right y-axis to determine
velocity.
The advantages to using a Stage-Discharge
curve is that the curve can be used to
determine the depth and velocity of flows
for several annual recurrences intervals. It
should be noted that the Stage-Discharge
curve is linked to the channel
characteristics. Should the analysis shift to
a different location where channel
parameters are different (slope, shape nvalue and so on), a new Stage-Discharge
curve will be required.
8.4.4.1 Compound Channels
Simple channels are often represented as a
basic trapezoid (refer Figure 8.4.1.3),
assessed using one Mannings n-value
and/or based on one slope.
This
methodology is reasonable for simple,
small channels. In reality, the shape of
channels is not a basic trapezoid. They are
compound in shape (refer Figure 8.4.3(a))
with:
high and low banks / possible small
overflow sections;
different n-values for banks and
channel bed; and
potentially different slopes between
main channel and overflow sections.
Assessment of compound channels uses the
same principal and methodology as
described earlier in this section, however
the channel cross section is first divided
into smaller sub-sections.
Figure 8.4.4.1 shows, diagrammatically, a
compound channel divided into two subsections (labelled A & B).

Figure 8.4.4.1 Compound Channel

Each section is analysed individually using


Mannings Equation to determine flow
velocity based on the parameters (S or So, n,
A and P) specific to the sub-section. It is
important to understand that the wetted
perimeter (P) is determined by the length of
the Water / Channel Boundary for the
sub-section. The wetted perimeter is the
component of the equation where effect of
friction is applied. The Water / Water
boundary, as shown by the red dashed line
in figure 8.4.4.1, is not to be included as the
friction along this boundary is negligible.
Once all sub-sections have been calculated,
the total flow rate can be determined by:

QTotal V A AA VB AB
Review of the calculations shows that each
section has a different average velocity.
While an average velocity can be calculated
using QTotal and the cross sectional area of
the whole channel (AA + AB), the different
velocities across the channel represents
reality reasonably well.

8.4.5

Energy Principles

The energy in open channel flow can be


expressed as the total head in metres (m) of
water, which is equal to the sum of the
elevation above a datum (elevation head),
the hydrostatic head and the velocity head.
The elevation head and hydrostatic head are
components collectively known as the static
head.

March 2010

8-11

Department of Transport and Main Roads


Road Drainage Manual

In conventional terms, energy has two


components:
potential energy; and
kinetic energy.
Now, relating conventional terms to the
energy within water, potential energy is
equal to the static head and kinetic energy is
equal to the velocity head. To illustrate
this, a static body of water such as a lake,
has potential energy while a moving body
of water such as a creek, has both potential
energy and kinetic energy.
8.4.5.1 Energy Equation
The total energy for an open channel
section is given by:

V2
H (z d )
2g
Where:
H = total energy head (m);
z = height of channel bed above
some reference datum (m), known as
elevation head;

d = depth of flow normal to the


direction of flow (m), known as
hydrostatic head;
V = velocity of flow (m/s); and
g = gravitational acceleration (9.81
m/s2).
2
The term V

2g

within this equation is

known as the velocity head.


According to Chow (1959) and Henderson
(1966), the variable d describes the depth of
flow normal to the direction of flow
(typically perpendicular to channel bed).
The variable y, as used in other equations,
describes the vertical depth of flow. Where
channels are relatively flat (So < 18), d and
March 2010

8-12

Chapter 8
Open Channel Design

y are of similar magnitude. While strictly


not
correct,
the
terms
can
be
interchangeable (based on condition) as the
difference is small, therefore the practice is
acceptable.
It is important to understand that this
equation differs slightly from the normal
form of the energy equation for open
channels, which is:

H ( z d cos )

V2
2g

Where:
= slope angle of channel bed; and
= coefficient accounting
velocity distribution.

for

Channels assessed by the department for the


installation of cross drainage culverts are
relatively flat and therefore is assumed to
be zero.
Referring to Section 8.4.2, it is assumed
that velocity is averaged across cross
section of flow, therefore, is equal to 1.
The energy equation is useful when
comparing two sections within a channel.
Typically, as water flows along a channel,
energy is lost. However, the first law of
thermodynamics tells us that energy can
neither be created nor destroyed.
Therefore, when comparing the energy
between two sections, the relationship can
be expressed as:
2

V
V
( z1 d 1 ) 1 ( z 2 d 2 ) 2 h L
2g
2g
Where:
hL = head loss (m).
This relationship is best illustrated in Figure
8.4.5.1

Department of Transport and Main Roads


Road Drainage Manual

8.4.5.2 Specific Energy


Specific Energy (m) is defined as the
energy per unit mass of water at any
section, measured with respect to the
channel bed. The Specific Energy (Es)
equation (which is essentially a subcomponent of the energy equation) is:

Es y

V2
2g

Chapter 8
Open Channel Design

When the depth of flow is plotted against


the specific energy for a given channel
section and discharge (only bed slope
changes from flat to steep), a specific
energy curve is obtained, as shown in
Figure 8.4.5.2. This curve shows that for a
given specific energy (Es), there are two
possible (conjugate) depths called alternate
depths; y1 and y2.

Figure 8.4.5.1 Comparing Energy between Points

Figure 8.4.5.2 - Specific Energy Curve (Source: Franzini & Finnemore (1997))

March 2010

8-13

Department of Transport and Main Roads


Road Drainage Manual

At point C on the curve, the specific energy


is at a minimum and there is only one
depth. The depth corresponding to this
point is known as the critical depth yc. The
corresponding velocity is the critical
velocity (Vc) and the slope is the critical
slope (Sc).
When the depth of flow is greater than the
critical depth, the flow velocity is less than
the critical velocity and is referred to as
subcritical flow. When the depth of flow is
less than critical depth, the flow is
supercritical. Subcritical flow is controlled
by downstream channel conditions, while
supercritical flow results from some
upstream control condition.
8.4.5.3 Froudes Number
Another, simpler method to determine
subcritical / supercritical flow is by
determining the Froude Number (Fr) for the
flow.
Froudes Number is a dimensionless
number, the ratio of inertial forces to
gravitational forces, and defined as:

Fr

V
gL

Where:
V = velocity of flow (m/s);
g = gravitational acceleration (9.81
m/s2); and
L = characteristic length (m).
The characteristic length (L) in open
channel flow is equal to the hydraulic
depth, which is defined as the cross
sectional area of flow divided by the
surface width of flow. Using either the
hydraulic depth or flow rate for the stream,
there are two commonly used forms of the
above equation:

March 2010

8-14

Chapter 8
Open Channel Design

Fr

g A

and Fr Q B

gA
3

Where:
A = cross sectional area of flow
normal to direction of flow (m2);
B = width of flow at surface (m); and
Q = flow rate (m3/s).
These equations are valid for irregular
channel / stream sections.
Froudes Number indicates the state of
flow. Critical Flow occurs when Fr = 1.
When Fr is greater than 1, the flow is
supercritical, and when it is less than 1 the
flow is subcritical. When Fr is close to 1 (>
0.9 but < 1.1) the flow state is unstable and
waves appear on the water surface.
Design is recommended to achieve a
Froude Number less than 0.9 (subcritical
flow).

8.4.6

Hydraulic Jump

A hydraulic jump is an abrupt rise of the


water surface which occurs when flow
changes from supercritical to subcritical in
response to a downstream control
conditional causing a higher tailwater level.
The rapid (and turbulent) rise of water
dissipates energy. The turbulence persists
for some length after the rise before
settling.
The rise and trailing turbulence is called the
roller and is shown between points 1 and
2 on Figure 8.4.6.
The depth, y1 before the hydraulic jump is
called the initial depth, and the depth y2,
after the jump is called sequent depth.
With reference to Figure 8.4.6, the sequent
depth is always lower than the high stage
alternate depth, y2, due to the loss of

Department of Transport and Main Roads


Road Drainage Manual

energy, E, in the turbulence of the


hydraulic jump. Due to this loss of energy,
the solution to hydraulic jump problems
cannot be found using the Specific Energy
curve alone. The principle of momentum
must also be used to develop a Specific
Force or Force Momentum curve as shown
in Figure 8.4.6.
The detailed analysis of hydraulic jumps is
difficult and complex. Full analysis of
jumps should be undertaken by experienced
hydraulic engineers and therefore should be
referred to Hydraulics Section, Engineering
& Technology Division or a suitably prequalified consultant.
However, several models based on accepted
theory, experimentation, laboratory testing
in rectangular channels and observation,
have been developed to approximate key
parameters of hydraulic jumps in irregular
open channels.
These simple methods are considered
reasonable for analysing undular and weak
hydraulic jumps (refer Section 8.4.6.2) and
have been adopted by the department for
general use in analysing hydraulic jumps.

Chapter 8
Open Channel Design

8.4.6.1 Specific Force


As mentioned above, the solution to
hydraulic jump problems cannot be found
using the Specific Energy curve alone. The
hydraulic jump involves internal energy
losses which cannot be simply evaluated
using the energy equation therefore the use
of the momentum principle is required.
In applying the momentum principle to a
short horizontal reach of a prismatic
channel, the Specific Force can be defined
as:

Q2
y A
gA

Where:
Q = discharge (m3/s);
g = gravitational acceleration (9.81
m/s2);
A = area of cross section (m2); and
= depth to the centre of gravity to
cross section (m).

Figure 8.4.6 - Analysis Curves and the Hydraulic Jump


Source: Chow (1959)

March 2010

8-15

Department of Transport and Main Roads


Road Drainage Manual

The first term is the momentum of the flow


passing though the channel section per unit
time, per unit weight of water. The second
is the force per unit weight of water.
An examination of Figure 8.4.6 shows that
while water flows at a depth (y1) for a given
discharge (Q) in any channel, there will
always be another depth where the sum of
the force due to velocity plus the
hydrostatic pressure for both depths will be
the same.
This means that the specific forces of
sections 1 and 2 are equal, provided that the
external forces and the weight effect of
water can be ignored. This generally can be
applied to horizontal channels with small
bed slopes (So 10%).
With respect to the rule that there is always
two depths, the exception is at critical depth
(points C & C) where energy is at a
minimum and there is only one depth of
flow.
8.4.6.2 Jump Strength

Hydraulic jumps can only form if the


upstream flow is supercritical (i.e. has a
Froude Number greater than 1).
As
Froudes Number increases, the strength of
the jump also increases.
Table 8.4.6.2 lists the different types of
jump and their defining characteristics.
This table has been derived from research
studies undertaken in 1955 by the U.S.
Bureau of Reclamation (USA).
8.4.6.3 Sequent Depth Horizontal
Channels (SO 10%)
The parameters of the supercritical,
upstream flow before the hydraulic jump
are known / can been determined. These

March 2010

8-16

Chapter 8
Open Channel Design

parameters are flow velocity, flow depth


and Froude Number.
Using these parameters, the alternate or
sequent depth (y2) can be determined.
Considering;
energy and momentum principles;
using the initial flow velocity as V1;
using the initial flow depth as y1; and
experiments
channel.

using

rectangular

Rectangular Channels
The following equation was developed to
calculate the sequent depth (y2) in a
rectangular channel:

y2 1

1 8 Fr2 1
y1 2
OR

y2

y1
2

1 8F 1
2
r

Where:

Fr V1

gy1

(rectangular section)

This relationship is shown graphically in


Figure 8.4.6.3.
Trapezoidal Channels
The determination of sequent depth (y2) in a
trapezoidal channel of reasonably small bed
slope (So 10%), requires an iterative
solution. However, when applying the
Specific Energy and Specific Force
equations, it is possible to arrange these
equations into a dimensionless form,
allowing solution by means of a table.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Table 8.4.6.2 Jump Types and Strength


Jump Type

Fr

Energy Dissipated

Characteristics

Undular

1.0 1.7

< 5%

Very weak jump; standing waves

Weak

1.7 2.5

5 15%

Weak jump; small roller; smooth rise

Oscillating

2.5 4.5

15 45%

Unstable avoid

Steady

4.5 9.0

45 70%

Stable, well defined jump; best design


range

Strong

> 9.0

70 85%

Choppy; very turbulent; intermittent

Table 8.4.6.3 assists the designer with the


calculation of the sequent depth (depth after
the jump). This table has been derived
from the Handbook of Hydraulics, 7th Ed,
1996 by Brater et al.

8
y2 1

1 8 Fr2 1
y1 2

Fr V1

gy1

Fr

Source: Chow (1959)


Figure 8.4.6.3 - Relation between Froude
Number and Depths of Flow in a
Hydraulic Jump

March 2010

8-17

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Table 8.4.6.3 - Sequent Depth for Trapezoidal Channels (So 10%)

8
Source: Brater et al. (1996)

8.4.6.4 Sequent
Depth
Sloping
Channels (SO > 10%)
While the effect of the weight of water in
the jump is negligible for horizontal
channels (So 10%), it is essential to
consider it in the analysis of hydraulic
jumps in sloping channels (So > 10%).
Thus, the momentum equations for
horizontal channels cannot be applied to
jumps on channels having a bed slope

March 2010

8-18

larger than 10%. However the momentum


principle can be used to derive analogous
equations, which will contain an empirical
function that has to be determined
experimentally. This work was undertaken
by U.S. Bureau of Reclamation (1964) and
the research shows that the depth ratio y2/y1
can be shown as a function of Froude
Number (F1) and Slope (So). The graph of
this function is shown in Figure 8.4.6.4.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

y2
y1

Fr1

V1
gA

Source: Chow (1959)


Figure 8.4.6.4 - Sequent Depth for Sloping Channels (SO > 10%)

8.4.6.5 Location of Jump


The location of the jump is dependant
primarily on the momentum of flow before
and after the jump and can occur either
before or after the trigger situation that
causes the initial supercritical flow to
slow to subcritical flow. For a closer
estimate of the jump position, the length of
the jump should be considered as described
in Section 8.4.6.6 below.
Figure 8.4.6.5 shows a channel having a
break in the bed slope (steep to mild
slope) and two possible locations for the
jump. If the sequent depth y2 is greater than
the alternate depth y1, the jump will occur
in the steep region. If the sequent depth y2
is lowered to less than approximately the

alternate depth y1, the jump will start to


move into the mild channel.

Source: Chow (1959)


Figure 8.4.6.5 Location of a Hydraulic
Jump

8.4.6.6 Length of Jump


The length of a hydraulic jump is defined as
the distance measured from the front face of

March 2010

8-19

Department of Transport and Main Roads


Road Drainage Manual

the jump to a point on the surface


immediately downstream of the roller and is
difficult to model and predict.
Two
methods to determine the length of jump
are presented here.
Method 1
Observation and measurements have shown
that a good approximation of jump length is
5 times the sequent depth (y2). In most
cases, the jump length is between 4 and 6
times y2.
Method 2
The length of the jump can be determined
from a relationship based on experimental
data prepared by the U.S. Bureau of
Reclamation (1964), as shown in Figure
8.4.6.6.
This curve was developed
primarily for jumps in rectangular channels
but may also be applied to approximate
undular and weak jumps formed in
trapezoidal channels.
The length determined by either method
should be rounded up to the nearest meter.

As the length of jump indicates the extent


of turbulence, it is therefore the minimum
length of channel that should be armoured
against scour / erosion. When determining
the length of jumps in open channels for
armouring purposes, Method 1 is
recommended as it gives a more
conservative answer. Method two should
only be used when in constrained situations.
8.4.6.7 Application
Jumps

of

Hydraulic

There are many practical applications for


the hydraulic jump. The most practical of
these applications, for the purpose of road
design, is its use as an energy dissipator to
prevent scouring downstream of a hydraulic

March 2010

8-20

Chapter 8
Open Channel Design

structure such as a culvert or to prevent


scouring in an open channel.

8.4.7

Hydraulic Drop

The converse of the hydraulic jump, the


hydraulic drop, can also occur due to an
abrupt change in channel slope or cross
section. It can frequently occur near the
entrance to a culvert flowing under inlet
control.
(An explanation of culverts
flowing under inlet control is given in
Chapter 9).

8.5

Erosive Velocities in
Natural Streams

Natural streams may be part of an open


channel design or subject to outlet
velocities from drainage structure and
therefore a key design consideration is
ensuring that any design changes to the
flow in the natural stream do not create
erosive situations in the stream.
Erosive velocities in natural stream beds are
shown in Table 8.5 and have been derived
from curves in Neill (1973).

8.6

Backwater

Backwater is the accumulation of water in a


stream that can be caused by:
an obstruction limiting downstream
channel capacity (that causes afflux
in the stream);
tidal flow (refer Section 8.6.1);
a downstream body of water such as
a weir, dam or lake; or
a higher-than-normal flow stage in a
downstream connecting stream (refer
Section 8.6.2).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Source: Chow (1959)


Figure 8.4.6.6 - Length of Hydraulic Jump in Terms of Sequent (y2) in Horizontal Channels

Table 8.5 - Erosive Velocities in Natural


Streams
Stream Bed Type

less than 0.3

Silt
Sand

Gravel

Clay

Rocks

Velocity (m/s)

have a level that is increased above normal


depth. The cause of backwater can be a
significant distance away from, but still
affect, the site being assessed. Therefore
designers should check the downstream
conditions of planned sites or designed
infrastructure for possible backwater
influence.

Fine

less than 0.3

Coarse

less than 0.3

6 mm

0.6 to 0.9

25 mm

1.3 to 1.5

100 mm

2.0 to 3.0

Soft

0.3 to 0.6

Stiff

1.0 to 1.2

Hard

1.5 to 2.0

8.6.1

150 mm

2.5 to 3.0

300 mm

3.5 to 4.0

Backwater as a result of tidal flow can


affect stream flow and in turn drainage
infrastructure design. When determining
backwater effects on tailwater levels for
streams discharging into tidal waters, three
factors may influence the final design level:

The presence of backwater typically causes


an upstream section or body of water to

Calculating the effect of backwater on


normal stream flows can be difficult and is
beyond the scope of this manual. If it is
considered that backwater could be
affecting stream flow, advice and/or
assistance should be sought from the
departments Hydraulics Section or a
suitably pre-qualified consultant.

Tidal Waters

tide levels;
March 2010

8-21

Department of Transport and Main Roads


Road Drainage Manual

storm surges; and


climate change.
8.6.1.1 Sea and Tide Levels
Tide Tables are published by the
Department of Transport and Main Roads
(previously Queensland Transport) for
various coastal locations in Queensland.
This data is used to estimate backwater and
tailwater levels for drainage designs.
The Mean High Water Spring (MHWS) tide
is often applied as a tailwater level since
this is a reasonable high tide level, but not
overly conservative. Spring tides occur
approximately once a fortnight at full and
new moon. The definition of different tidal
levels is provided in Figure 8.6.1.1,
reproduced from the tide tables.
Storm tides are an increase in ocean levels
above the expected tidal levels caused by
low pressure and wind effects. Storm tides
can occur at any stage of the tidal cycle.

As with other effects, the selection of a


storm tide needs to consider the combined
risk of occurrence of the storm tide and the
flood from the waterway. The probability
of a catchment flood at the same time as a
storm tide is the product of the probabilities
of the separate events, resulting in a much
rarer combined recurrence interval than that
of the separate events.
Storm tides have been analysed for
locations around Queensland and levels are
available from the Department of
Environment and Resource Management
(previously Queensland Environmental
Protection Agency) and local authorities.

March 2010

8-22

Chapter 8
Open Channel Design

It is important to note that:


mean sea level given for the location
may be different to Australian Height
Datum (AHD) which is the average
mean sea level of 42 locations around
Australia;
tide levels are given at or very close
to the coastline; and
tide levels are often specified in
terms of low water datum.
It is very difficult to estimate the tide level
at a location in a stream some distance from
the coastline. The time for the tide to rise
along the creek and to then flow back in the
opposite direction is one factor. Another is
the existence of local sand bars or raised
areas of the creek bed closer to the mouth
of the stream, which may prevent a tide
from reaching or falling below a certain
level at a particular road crossing. The
transient nature of sand bars adds to the
difficulty in estimating the tidal reach.
Although precise tide levels are not usually
necessary at an upstream road or bridge
crossing, tide levels at the mouth of the
tidal stream are not sufficient to give tide
design parameters at the road or bridge.
The absolute minimum survey requirements
at a road or bridge site are the times and
levels for successive low - high - low tides.
Alternatively successive high - low - high
tide information may be found. The more
tidal cycles measured the better.
This information when compared with tide
levels at the mouth of the creek will allow
experienced hydraulic engineers to predict
approximate design tides such as MHWS
and MLWS (Mean Low Water Springs) at
the project site.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

MSL
(Mean Sea-Level) The average level of the sea over a long period (preferably 18.6
years) or the average level which would exist in the absence of tides.
HAT (Highest Astronomical Tide) / LAT (Lowest Astronomical Tide) These are
the highest and lowest levels which can be predicated to occur under average
meteorological conditions and any combination of astronomical conditions. These
levels will not be reached every year. HAT and LAT are not the extreme levels
which can be reached, as storm surges may cause considerably higher and lower
levels to occur.
MHWS (Mean High Water Springs) Long term average of the heights of two
successive high waters during these periods of 24 hours (approximately once a
fortnight) when the range of tide is greatest, at full and new moon.

MLWS (Mean Low Water Springs) The long term average value of two successive
low waters over the same periods are defined for MHWN
AHD (Australia Height Datum) This Datum has been adopted by the National Mapping
Council as the datum to which all vertical control for mapping is to be referred.
Figure 8.6.1.1 - Tidal Planes

8.6.1.2 Storm Surge


A storm surge is the rise (or fall) of open
coast water levels relative to the normal
water level and is due to the action of wind
stress and atmospheric pressure on the
water surface.

Storm surges occur as part of major storms


such as cyclones where there are low
atmospheric pressures and the wind blows
over reaches of the ocean.
Some predicted surge heights have been
produced
by
the
Department
of
Environment and Resource Management
and this data should be referenced.

March 2010

8-23

Department of Transport and Main Roads


Road Drainage Manual

There is no correlation with tide levels, nor


are there any predictions for wave break
setup and wave runup on the land. These
factors will need consideration for any
design with storm surge as a factor.
Storm surges would need to be considered
with respect to coastal developments, the
protection of coastal roads, route immunity
for evacuation purposes, and for major
coastal drainage designs.
Local
government may also have specific
requirements or data in relation to storm
surge.
There is a low likelihood of occurrence of
storm surge coinciding with flood events,
so calculation of design floods does not
need to consider both together, but the two
effects should be considered independently.

8.6.2

Downstream Tributary

If the crossing is located on a stream which


joins another watercourse (larger or
smaller) downstream, other issues need to
be considered.

As the two open channels have different


catchment sizes, they will peak at different
times. The combined flow at their junction
needs to be assessed.
In this case, two situations need to be
considered:
(a) Major flood on tributary with
limited flow in the main stream;
and

Chapter 8
Open Channel Design

the relative sizes of the two streams, it may


not be realistic to expect floods to occur
together in the two streams.

8.7

Tailwater Levels

Tailwater level refers to the normal water


level, for a given flow, in a channel
immediately downstream of a drainage
structure. For a proposed culvert, it is the
depth of flow at the culvert outlet, measured
from the water surface in the downstream
channel to the invert of the culvert and the
depth can be significant. Tailwater levels
are an important control on the hydraulic
performance of road drainage structures,
therefore the estimation of tailwater levels
is required for all hydraulic analyses for
bridges and culverts.
Typically, tailwater level is defined by the
downstream channel properties. If the
downstream channel and flow is uniform,
the tailwater level may be determined using
Mannings Equation (refer Section 8.4.3).
If the downstream flow is non-uniform, or
if it is being influenced by backwater, a
more sophisticated analysis by an
experienced hydraulic engineer is required.
In this situation, advice and/or assistance in
this case can be provided the departments
Hydraulics Section or a suitably prequalified consultant.

8.7.1

Tailwater Effects

(b) Major flood on the main stream and


limited flow in the tributary.

As previously mentioned, tailwater is an


important aspect required to determine the
hydraulic operation and performance of
drainage structures.

Both cases need to be analysed to provide


an understanding of the potential flood
conditions at the road.
The risk of
coincidental flooding in the two streams
needs to be considered to determine the
combined risk of flooding. Depending on

A low tailwater typically means the channel


can drain the stormwater away fairly
quickly / easily. This is common where
drainage channels are less defined with low
banks, allowing the stormwater to spread

March 2010

8-24

Department of Transport and Main Roads


Road Drainage Manual

widely.
Drainage structures with low
tailwater levels will most likely be
controlled by inlet conditions and can have
high outlet velocities, therefore erosion can
be an issue. A high tailwater level typically
means the channel does not drain the
stormwater away quickly. This is common
where drainage channels are reasonably
defined with mid-to-high banks, allowing
the channel to contain most stormwater
flows.
Drainage structures with high
tailwater will most likely be controlled by
outlet conditions and can have low outlet
velocities, therefore high headwater levels
(afflux) can be an issue.

8.7.2

Design Tailwater Levels

Typically, tailwater levels should be


determined based on an assessment of
channel flow (and flow depths), for a given
discharge, over a significant length of
undisturbed downstream channel. This
assessment is usually commenced at a point
well downstream of the proposed structure
and calculated back / upstream towards the
proposed structure. This assessment is
complex and hydraulic modelling software
is often used.
For the design of most drainage structures
within the scope of this manual, the survey
data gathered in relation to creeks / streams
/ channels is usually limited to the road
environment (constrained by property
boundaries), unless extended downstream
survey was specifically requested.
Using available survey data and assuming
the channel is regular in shape and steady,
uniform flow conditions can be expected,
the tailwater level can be determined using
Mannings formula as follows:
(a) Adopt a stream / channel cross
section just downstream of

Chapter 8
Open Channel Design

proposed structure outlet, but


representative of the downstream
channel in general.
(b) Using a series of calculations
(Mannings Equation & Q=A.V),
from no flow to maximum channel
capacity, develop a stage-discharge
curve for the channel.
(c) Using the stage-discharge curve
and design discharge for the
proposed structure, determine flow
depth. This depth is adopted as the
tailwater depth for the proposed
structure analysis.
If the design discharge exceeds the
maximum capacity of the channel, it means
the channel will overflow and the design
discharge (flow) will spread.
Review
survey data to see if there are any high
banks or if cross section can be extended to
allow for increased flow depth and if so,
extend stage-discharge curve to incorporate
additional waterway area. If the extended
cross section cannot contain the design
discharge and the spread of flow is
significant (there would be little increase in
flow depth based on design discharge), then
the tailwater level could be set at the level
of maximum capacity for the defined
channel. Also, if there is not much of an
outlet channel (flat terrain / low banks) and
the design discharge easily overtops the
banks and/or spread of flow is significant,
then a low tailwater level should be
adopted. Use survey data to produce a wide
flat cross section and then process as above
to determine this low tailwater level. The
old rule-of-thumb to adopt a tailwater
level of barrel height of proposed culvert
is not considered accurate enough and can
result in inappropriate designs.
Furthermore, anecdotal information with
respect to channel flows should be sourced
March 2010

8-25

Department of Transport and Main Roads


Road Drainage Manual

from local residents.


Theoretical
predictions (proposed tailwater level)
should be checked against the anecdotal
evidence to see if both reasonably agree. If
so, then the level of confidence in
prediction is increased. If not, then further
checks of calculations / gathered
information and so on is required. All
gathered information should be recorded.

8.8

Open Channel Design

The problem most likely to be faced by the


designer of artificial channels is the
determination of a channel cross-section
that will convey the design flow down a
slope producing a velocity which does not
create scour.

8.8.1

Design Methodology

There are two artificial channel types:

Chapter 8
Open Channel Design

bare-earth channels. If a self-cleaning


channel is required, designers should
reference Section 2.8 for requirements.
Channel side slopes should be based on the
slope stability of the material the channel is
to be constructed in (channel is shaped
before lining applied or grass grows).
Typically, this will be in the range 1on 1 to
1 on 4. Safe access for maintenance
equipment should also be considered.
Freeboard is the additional height of
channel required above the height of the
design flow. This allows for inaccuracies in
data used in calculation and possible
surcharge due to silt / debris build up and/or
grass growth in the channel because of
delayed maintenance of the channel.
The freeboard to be adopted for open
channels, similar to QUDM (NR&W 2008),
is the greater of the following calculations:

hard lined channels; and

300 mm A;

vegetated / bare-earth channels.

20% of the flow depth; or

Each of these channel types has a specific


design methodology to determine the
channel cross section and these are detailed
in the following sections.
8.8.1.1 General Considerations
Further to the discussions in Sections 8.9
and 8.10, the following are aspects that
should be considered / established before
the design of the channel commences.
The velocity of flow in open channels is a
key aspect. Channel slope is an important
factor in flow velocity. Generally, flow in
channels is intermittent, and the channel
must be constructed to allow all stormwater
to drain away (no ponding). Therefore the
minimum channel slope (So) for hard lined
channels should not be less than 0.25% and
not be less than 0.5% for vegetated and
March 2010

8-26

velocity head B of the flow.


A

: Where flooding of adjacent land


and buildings does not represent a
risk, the 300mm requirement can be
reduced to 150mm.

: Refer Section 8.4.5.1.

8.8.1.2 Hard Lined Channels


With hard lined or rigid boundary channels,
the design procedure is relatively simple as
potentially erosive velocities are less
important.
Therefore, the cost of
excavation and lining are primary factors in
determining the geometry of the channel.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

In designing hard lined channels the


following must be initially considered:
lining
material
Mannings n-value).

(determines

Based on the proposed channel lining, the


Mannings n-value can be selected from
Table 8.4.3(a).
Research and experimentation has found
that the most efficient, hydraulic section is
when the wetted perimeter (P) is minimised
for a given cross sectional area (A).
Typically, in open channels, this occurs
when the width of flow is twice the depth of
flow. Therefore the shape of channel that
gives the best hydraulic efficiency (also
known as best hydraulic section) occurs
when the hydraulic radius is R

y
2

Based on this requirement, the most


efficient shape is a semi-circle as it has the
least wetted perimeter for a given area R
is always equal to y/2 irrespective of
dimensions.
However the cost and
difficulty of building semi-circular channels
prohibits their widespread use. Trapezoidal
channels are more commonplace. The most
efficient trapezoidal shape is the halfhexagon.
For the design of hard lined channels:
(a) Determine design discharge (Q), bed
slope (So), Mannings n-value and
channel side slope (X) (refer Figure
8.8.1.2).

Figure 8.8.1.2 Trapezoidal Channel


Section

(b) For best hydraulic section:

by yX 2
2 y 2 yX b
2

y
2

and resolve for b as a function of y.


(c) Combining Mannings and Continuity
Equations:
2

AR 3 S f

Where:
A = by + Xy2 ; and
R = y/2
So is used to approximate Sf (Friction
Slope)
(d) Substitute b in equation with function
from (b) and solve for y.
(e) With y, calculate b using function from
(b).
Channel geometry can now be determined
and freeboard added.
(f) Calculate Froude Number for channel
section / flow and together with
velocity and depth of flow, check that
conditions are acceptable where the
designed channel discharges into an
existing channel etc.
8.8.1.3 Vegetated
Channels

or

Bare-earth

Vegetated / bare-earth channels are erodible


boundary channels and the design
procedure is relatively simple, based on the
method of maximum permissible velocity.
This method assumes that the channel will
remain stable provided that the average
velocity of the design flow is below a
determined threshold value. This method is

March 2010

8-27

Department of Transport and Main Roads


Road Drainage Manual

not as realistic as shear-based / tractive


force methods but is considered reasonable.

For the design of vegetated / bare-earth


channels:

A key concern for vegetated channels is


what happens when the grass cover cannot
be maintained, such as during drought or
after a bushfire. This aspect should be
considered and if there is a reasonable risk
of occurrence and channel scour is likely /
not desirable, then design should be
undertaken assuming bare-earth design
values.

(a) Determine design discharge (Q), bed


slope (So), Mannings n-value, channel
side slope (X) (refer Figure 8.8.1.2) and
maximum permissible velocity (Vmax)
from Table 8.8.1.3. For bare-earth
channel design, adopt 0% cover.

In designing vegetated / bare-earth channels


the following must be considered:
the material the channel is to be
constructed in;
a suitable grass species for the
channel (where applicable); and
an appropriate Mannings n-value.

Chapter 8
Open Channel Design

A suitable grass species for a stormwater


channel should be quick to establish, be self
repair, have a relatively short blade length
(< 50 mm), be able to survive short
durations of inundation, be able to
withstand proposed design velocities and be
native to area. The species should be
chosen in consultation with a suitably
qualified
landscape
architect
or
horticulturalist. The reason for choosing a
short blade length grass species is primarily
that longer blade length species can reduce
flow velocity by increased flow resistance.
This increases flow depth (surcharge) and
ultimately can cause the channel to break its
banks.
Table 8.8.1.3, largely derived from Soil
Conservation Measures Design Manual
for Queensland (NR&M 2004), gives
maximum permissible velocities, for use in
design, for consolidated, bare-earth
channels and vegetated channels.

March 2010

8-28

(b) Using the basic equation Q=V.A,


determine required waterway area.

AQ

Vmax

(c) Combining Mannings and Continuity


Equations:
2

AR 3 S o
Q
n

(d) rearrange and solve for wetted


perimeter (P):

A 5 3 S 12
o
P
Qn

(e) Based on channel geometry, two


expressions for A and P exist:

A b yX y ; and

P 2y X 2 1 b
Rearrange expression for A to isolate b:

A
yX
y

and substitute back into expression for P:

P 2y X 2 1

A
yX
y

Now calculate depth of flow (y). Solving


for y requires rearrangement into the form
of a quadratic equation.

ay 2 by c 0

Department of Transport and Main Roads


Road Drainage Manual

The general solution of a quadratic equation


is:

b b 2 4ac
y
2a
Solving a quadratic equation gives 2 roots.
Substitute roots (y) into:

A
yX
y

and solve for b. Dismiss any negative


solution and adopt the positive (real)
solution for y.
Channel geometry can now be determined
and freeboard added.
If there are no real roots to the equation,
this means that the velocity will never reach
the maximum permissible for the given
discharge (Q) and roughness (n), regardless
of channel dimensions. In this case, use the
best
hydraulic
section
design
methodology as detailed in Section 8.8.1.2
for the design of the vegetated channel, but
check that the average velocity of the
resultant channel section does not exceed
the maximum permissible velocity.
Calculate Froude Number for channel
section / flow and together with velocity
and depth of flow, check that conditions are
acceptable where the designed channel
discharges into an existing channel etc.

8.8.2

Channel Transitions

Changes from one channel cross-section to


another should be smoothly transitioned as
energy losses occur with the changes.
It is recommended that maximum
contraction rates of about 1 on 1, and
maximum expansion rates of about 1 on 4
be adopted as shown in Figure 8.8.2.

Chapter 8
Open Channel Design

Table 8.8.1.3 - Maximum Permissible


Velocities (Design) for Consolidated,
Bare Channels and Vegetated Channels
Channel
gradient

Permissible velocities (m/s) when


fraction of stable surface cover
is:
0.02

0.5

0.7

1.0

Erosion resistant soils (e.g. Krasnozemsm &


Red Earths)

0.5
1
2
3
4
5
6
8
10
15
20

0.8
0.7
0.6
0.5

1.8
1.6
1.4
1.3
1.3
1.2

2.8
2.8
2.5
2.4
2.3
2.2
2.1
2.0
1.9
1.8
1.7

2.4
2.1
1.8
1.7
1.6
1.6
1.5
1.5
1.4
1.3
1.3

Easily eroded soils (e.g. Black Earths & fine


surface texture contrast soils)

0.5
1
2
3
4
5
6
8
10
15
20

0.6
0.5
0.5
0.4

1.3
1.2
1.1
1.0
1.0
0.9

2.3
2.1
1.9
1.8
1.7
1.6
1.6
1.5
1.5
1.4
1.3

1.6
1.5
1.4
1.3
1.2
1.2
1.1
1.1
1.1
1.0
0.9

Notes:
1. Assume the following species under average
conditions will provide the fraction of cover indicated:
- Kikuyu, Pangola and
well maintained Couch species 1.0.
- Rhodes grass, poorly maintained
Couch species - 0.7
- Native species, tussock grasses - 0.5
2. Applies to surface consolidated, but not cultivated

March 2010

8-29

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Where:
hb = channel bend head loss (m);
B = channel width (m);
Figure 8.8.2 - Maximum Rates of
Contraction and Expansion

Transition head losses can be calculated


using the equation:

V 2 V 2
ht C u 1 2
2g 2g
Where:
ht = transition head loss (m);
Cu = transition energy loss coefficient
(dimensionless);
V1 = average flow velocity upstream
of transition (m/s);
V2 = average flow velocity
downstream of transition (m/s).
Table 8.8.2 gives contraction and expansion
coefficients (Cu) for use in this formula.

It will be noted that the more abrupt the


transition, the greater the energy loss.
The difference in velocity head is always to
be taken as positive even if mathematically
it is negative. In other words a change in
section will always give a head loss.

8.8.3

Energy Losses in Channel


Bends

QUDM (NR&W 2008) gives an equation to


calculate the loss in a bend in an open
channel based on work by Mockmore
(1944). The equation for channel bend
head loss for bends with changes in
direction between 90 to 180 degrees is:

2 B V 2

hb
2
R
g

c
March 2010

8-30

V = average flow velocity (m/s);


g = acceleration due to gravity (9.81
m/s);
Rc = centreline radius of bend (m).
Results from using this equation should be
considered as providing a conservative
estimate of head loss.
For bends between 0 and 90 degrees, linear
interpolation is recommended.
Table 8.8.2 - Channel Transition Energy
Loss Coefficients
Transition
Type

Contraction

Expansion

Coefficient
Cu

Coefficient
Cu

(i) Gradual
channel
transition

0.1

0.3

(ii) Straight
through
bridge
transition

0.3

0.5

(iii) Square
edged abrupt
transition

0.6

0.8

Source: US Army Corps of Engineers

8.8.4

Superelevation in Channel
Bends

Water flowing around a bend in an open


channel may show a superelevation with a
rise in water level on the outside of the
bend and a corresponding dip in water
levels on the inside of the bend. The

Department of Transport and Main Roads


Road Drainage Manual

maximum difference in these water levels is


termed the superelevation (not to be
confused
with
the
percentage
superelevation in geometric road design).
The superelevation should be considered
when a flood level is reported at a bend (the
exact location needing to be defined), or
where it is essential that the design flow
does not exit from a man made channel at
the outside edge of a bend.
QUDM (NR&W 2008) gives the following
equation to calculate the superelevation:

R
hsup 2 log e o
Ri

V 2

2 g

Where:
hsup = superelevation of the water
surface across the channel (level
difference) (m);
V = average flow velocity (m/s);
g = acceleration due to gravity (9.81
m/s);
Ro= outside radius of bend (m);
Ri = inner radius of bend (m).

8.9

Grassed Channels

This section provides the guidelines for


channels that will use grass as a protective
lining. Designers should refer to Chapter 7
of this manual for further guidelines when
an open channel (such as a swale) is to also
be used for environmental reasons.

8.9.1

Normal Grassed Channels

The cross section shapes generally used for


grassed channels are trapezoidal (most
common), parabolic, or triangular. Flat side
slopes of absolute maximum steepness 1 on
4 and desirable maximum 1 on 6 should be

Chapter 8
Open Channel Design

used where space permits to assist in


minimising maintenance costs.
The maximum permissible velocity of flow
in a grass channel is the velocity which, if
maintained for a reasonable length of time,
will not cause erosion.
Permissible
velocities for different vegetal covers,
channel slopes, and soil conditions are
given in Table 8.8.1.3.
The minimum non-silting velocity is
uncertain. As a general rule the channel
velocity should not fall below 0.5 m/s.
To function well, grassed channels
normally have slopes of from 0.2 to 0.6
percent.
Where natural topography is
steeper than the desired slope of the
channel, drop structures / channel drops
(refer Section 8.11) should be considered.
Sharp curves which affect the flow in the
channel should not be used. In general,
curves should be such that the minimum
radius at the centre line is equal to about
three times the width of the top of the
design flow.
Research by the United States Soil
Conservation Service (USSCS) has found
that, for a particular channel lined with a
certain species of grass of a definite length,
the value of n varies with the product of the
mean velocity of flow V (m/s) and the
hydraulic radius R (m). The relationship is
a characteristic of the vegetation and is the
same for most common sections used.
The vegetation characteristics were defined
by five degrees of retardance (A to E)
depending on the grass species and length.
A channel lined with a grass having a high
retardance (A), such as 750 mm long
Rhodes grass, has its flow severely
restricted by the grass, while any grass less
than 50 mm long, retardance (E), has little
effect on the flow.
The relationship

March 2010

8-31

Department of Transport and Main Roads


Road Drainage Manual

between grass species, length


retardance is given in Table 8.4.3(c).

Chapter 8
Open Channel Design

and

The original USSCS curves for degree of


retardance D and E have been replaced in
this Manual by curves based on work by
Rouven et al. (1981) who found that the
original curves were not acceptable for
short, stiff vegetation on slopes smaller than
5%.
The use of erosion control cut-off walls at
regular intervals in a grassed channel is
desirable. Such cut-off walls will safeguard
a channel from serious erosion prior to the
grass becoming established or immediately
after a fire. These walls are usually of
unreinforced concrete or rock masonry,
approximately 200 mm wide and from 450
mm to 600 mm deep, extending across the
entire bed of the channel conforming to its
shape. Where the banks of the channel may
erode, these walls should extend up the
banks to the design water surface.
The designer needs to consider the impact
of these walls on the conveyance of the
channel and hence flood levels.

8.9.2

Reinforced Grassed
Channels

The use of reinforced grass (or turf) in


channel design provides the benefits of
significantly increased allowable flow
velocities, however, these products also
come with disadvantages, including:
possible public safety problems
resulting from high flow velocities;
maintenance personnel may not be
aware of the existence of the grass
reinforcing and normal grading
operations can result in costly
damage to the reinforcing mats;
damage by grass fires;

March 2010

8-32

degradation of habitat of grounddwelling wildlife in a bushland


environment,
die-off of grass owing to high
sediment deposition (reinforced grass
or turf can only be installed once the
contributing catchment is stable); and
loss of topsoil in the future when the
channel
is
removed,
rebuilt,
reconditioned or relocated (the
reinforced grass and its enclosed
topsoil will likely need to be
removed from the site, thus
preventing the reuse of the topsoil).
The latter point can be a major concern in
areas where there is limited topsoil, which
includes most of Queensland.
There are several proprietary types of
reinforced grass which are termed a form
of geotextile protection in the U.S.A.
Some tests of relatively small flows
indicate resistance to erosion for flows up
to 5 m/s. Caution should be exercised
before adopting design velocities of this
magnitude; the duration and depths of flow
on the actual job site will also need to be
considered.
The reinforced grass may be considered an
intermediate type of lining between
conventional grass and hard / impermeable
linings. Values of Mannings n for these
linings can be found in Table 8.4.3(a).
Example:
Section 3 (b) (ii) n value range 0.025
0.030 say use n = 0.028 for
conventional grass
Section 2 (a) n = 0.012 for hard /
impermeable
Average for reinforced grass n =
0.020

Department of Transport and Main Roads


Road Drainage Manual

8.10

Channels Lined with


Hard Facings

8.10.1 General
These channels are typically lined with
concrete, riprap or rock mattresses / gabions
and are highly resistant to erosion. Design
velocities may be as high as 5 m/s, much
higher than those for grassed channels.
Smaller cross-sections than those for
grassed channels are possible to take the
same discharges, as side slopes of 1 on 1
are practical. Channels with hard facings
should be considered where the width is
limited within the road environment.
Other considerations for channels with hard
faced linings include the provision of:
contraction and expansion joints to
minimise the risk of cracking and
seepage and potential undermining;
step irons (or similar device), for
access / safety reasons, where the
channel side slope is steeper than 1
on 2 and the channel depth exceeds
0.9 m;
pressure relief weep holes in
impermeable linings both within the
channel invert and within the channel
side slopes. Weep holes should be
90 mm diameter at 1.8 m maximum
centres (depends on detailed design
(structural) requirements);
lateral protection against surface
flows undermining the side slopes.
A minimum hard faced strip of width
0.45 m on both sides at the top of the
channel is recommended;

Chapter 8
Open Channel Design

possible requirements for an energy


dissipator at the end of the lined
channel section to manage the
expected high flow velocities.

8.10.2 Riprap and Rock Filled


Wire Mattresses / Gabions
Riprap is a conventional protection used to
prevent channel bed and bank damage
upstream and downstream of hydraulic
structures, at bends, at bridges and in other
channel areas where erosion potential exist.
The thickness of the riprap should be not
less than 1.5 times the largest diameter of
rock.
Figure 8.10.2(b) shows the
relationship between bed velocity and rock
diameter (Rouven et al., 1981). When
using Figure 8.10.2(b), bed velocity can be
taken as approximately 0.7 times the
average channel velocity.
Rock filled wire mattresses or gabions may
also be used to line the channel bank or bed
as shown in Figure 8.10.2(a). Smaller sized
rocks can be used because the wire basket
surrounding the rock in the mattress or
gabion tends to make the mass act as a unit
while retaining flexibility.
Beside
providing protection against scour, rock
filled gabions are useful as drops with
either vertical or stepped faces. At drops,
the gabions should be keyed into both
banks, and a downstream cut-off wall
should be considered.
Design and
construction of gabion protection should be
in
accordance
with
manufacturers
recommendations / specifications.

cut-off walls at the ends of the bed


and side linings. These should have
a minimum depth of 0.5 m; and

March 2010

8-33

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Erosion: When used on the outside


of channel bends, rock filled
mattresses and gabions effectively
form a hydraulically smooth
surface.
This smooth surface
causes high flow velocities to occur
immediately adjacent the wire
baskets. These high flow velocities
can then result in increased bed
erosion at the base of the gabion wall
(causing foundation problems), or
bank
erosion
immediately
downstream of the gabion protected
bank;
Figure 8.10.2(a) Channel Lined with
Rock Filled Wire Mattresses.

8.10.2.1 Open Channel Lined with


Riprap
When using riprap or rock filled wire
mattresses over erodible underlying
materials an appropriate geotextile fabric
should be placed under and behind the
riprap, rock filled wire mattresses, or
gabions to prevent erosion and water piping
through the gabion.

Some of the in-service problems associated


with rock filled wire mattresses and gabions
are listed below:
Vegetation Problems: In some
locations, such as South East
Queensland, rock filled wire baskets
have been known to quickly vegetate
with both native and non-native
weeds and vines. These vines can
then migrate into the surrounding
bushland causing problems to native
trees;

March 2010

8-34

Vegetation
Cover:
In
some
installations, long-term stability is
best achieved by having the gabion
and mattress structures covered with
vegetation. When this form of bank
stabilisation is placed under a widespan bridge, light and rainfall
restrictions can limit or prevent
vegetation cover, thus resulting in a
reduction of the expected life span of
the structure; and
Damage by Sediment: In channels
that transport large quantities of
coarse, bed-load sediment, the
expected design life of gabion and
rock mattress structures can be
reduced due to damage to both the
plastic and galvanised wire coatings.
Site specific designs need to be completed
for every gabion or mattress application
using design recommendations from the
suppliers of proprietary products.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Figure 8.10.2(b) - Relationship between Bed Velocity and Rock Diameter

March 2010

8-35

Department of Transport and Main Roads


Road Drainage Manual

8.10.3 Concrete Lined Channels


Concrete linings must be protected from
hydrostatic uplift forces which can be
created by a high water table after the flow
in the channel has dropped. A free draining
perforated subsoil drainage pipe may be
required in large channels.
In small
channels, weepholes and porous backfill
may be adequate.
Where undermining along the top of the
lining is a risk, a cut-off lip or wall should
be provided. In long lengths of lining,
expansion joints are required.
Care should be taken in the design of
channels required to carry supercritical flow
to ensure that:
curves are avoided if possible;
channels are free of obstructions
which may cause oscillatory waves to
extend down the remaining length of
the channels and into culverts;
pipes entering a channel are cut off
flush with the lining;

expansion joints are dowelled to


prevent
differential
movement
between segments of the lining,
which could otherwise allow high
velocity flow to enter the joints and
cause uplift; and
unintended hydraulic jumps do not
form.
Channels designed for supercritical flow
require a smooth surface which is provided
by concrete lining. The thickness of lining
and type of reinforcement required should
be determined in consultation with the
departments Bridge Design Section or a
suitably pre-qualified structural consultant.

March 2010

8-36

Chapter 8
Open Channel Design

Where channels are large enough, they


should generally be accessible to vehicles
for maintenance in which case the lining of
the base of the channel should be designed
for the appropriate vehicular loading and
soil conditions.

8.11

Channel Drops

The use of channel drops, waterfalls or


scour checks may permit flatter channel
slopes to be employed. It is generally
desirable to use several low-head drops
rather than fewer higher drops for safety
reasons. However, it is often more difficult
to control the hydraulics of low drops.
While vertical drops should generally be
avoided to minimise safety concerns and
erosion control problems, they may have to
be used in some circumstances. Also,
depending
on
tailwater
conditions,
supercritical flow often occurs on the apron
of the drop and as a result, either a
hydraulic jump or submerged jet will
usually form downstream.
The use and design of channel drops are to
be referred to Hydraulics Section,
Engineering & Technology Division or a
suitably pre-qualified consultant.

8.12

Baffle Chutes

Baffle chutes (Figure 8.12) are a


satisfactory method of dissipating the
energy of flow in a channel where the slope
is steep. They require no tailwater to be
effective.
The use and design of baffle chutes are to
be referred to Hydraulics Section,
Engineering & Technology Division or a
suitably pre-qualified consultant.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 8
Open Channel Design

Figure 8.12 Diagram of a Baffle Chute

March 2010

8-37

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

Appendix 8A
Vegetal Retardance
Curves
8A

March 2010

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

Appendix 8A Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

8A

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

8A

Figure 8A.1 - Solution of Manning Equation for Vegetal Retardance A

March 2010

8A-1

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

8A

Figure 8A.2 - Solution of Manning Equation for Vegetal Retardance B

March 2010

8A-2

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

8A

Figure 8A.3 - Solution of Manning Equation for Vegetal Retardance C

March 2010

8A-3

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

8A

Figure 8A.4 - Solution of Manning Equation for Vegetal Retardance D

March 2010

8A-4

Department of Transport and main Roads


Road Drainage Manual

Appendix 8A
Vegetal Retardance Curves

8A

Figure 8A.5 - Solution of Manning Equation for Vegetal Retardance E

March 2010

8A-5

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Appendix 8B
Worked Examples

8B

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Appendix 8B Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

8B

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Mannings Formula Example 1


This example describes the process to determine the flow rate, the velocity of a flow and the
state of flow in a stream based on discussion in Section 8.4.
The example commences after the stream data (such as cross section, terrain, condition of
channel and stream profile to determine site bed slope) has been gathered (refer Chapter 4).
The task for this example is, given the stream data and height of flow (see diagram below),
determine the velocity of flow in the channel, the flow rate and state of flow (subcritical /
critical / supercritical flow).

Stream Data
Bed slope about the site is 0.8%
Ht of channel bed is 110.60 m
Channel is regular and considered a little rough with a lot of trees and weeds along the
banks.

Ht = 111.80 m
sides are
1 on 1
2.5m

Solution
We need to determine the velocity of flow using Mannings formula first, then the flow rate
using the fundamental equation and finally, determine Froudes number to describe the state of
flow. (Refer Sections 8.4.3, 8.4.4 and 8.4.5)
Mannings equation is:

R 0.667 S 0.5
V
n
Where R is the hydraulic radius, determines as cross sectional area of flow (A) divided by the
wetted perimeter (P). Also, S is the slope of the energy line which we dont have, therefore we
can use the bed slope (So) to approximate S.

Step 1. Calculate the cross sectional area of the flow.


A = 1.22 + 1.2 x 2.5 = 4.44 m2

March 2010

8B-1

8B

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Step 2. Calculate the wetted perimeter.


P = (1.22 + 1.22) x 2 + 2.5 = 5.89 m

Step 3. Calculate the hydraulic radius.


R = A / P = 4.44 / 5.89 = 0.75 m

Step 4. Now determine an appropriate Mannings roughness coefficient.


Using Table 8.4.3(b), Mannings n Values for Natural Channels, we can see that our channel,
being the main channel and regular in shape, places us in the top portion of the table and within
Section 1. The trees and weeds would suggest (e), the range 0.06 0.08. Now, with the bed
being a little rough, a value n = 0.07 (in the middle of range) is considered appropriate.

Step 5. All variables have now been determined, therefore calculate velocity.

0.750.667 0.0080.5
V
0.07
V = 1.06 m/s

8B

Step 6. Using fundamental equation Q = V.A, we can now determine the flow rate in the
channel.

Q = 1.06 x 4.44
= 4.71 m3 /s
Step 7. To determine the state of flow, we calculate Froudes number.

Fr Q B

gA
3

We have determined that Q = 4.71m3/s, A = 4.44m2 and g is acceleration due to gravity (taken
as 9.81m/s2), therefore we need to calculate B, the width of flow across the surface.

B = 1.2 + 2.5 + 1.2 = 4.9 m

March 2010

8B-2

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Therefore:

Fr 4.71 4.9

9.81 4.44
3

Fr = 0.36

Froudes number is below 1.0, therefore the flow is subcritical.

End of Example

8B

March 2010

8B-3

Department of Transport and Main Roads


Road Drainage Manual

8B

March 2010

8B-4

Appendix 8B
Worked Examples

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Mannings Formula Example 2


This example describes the process to determine the depth and velocity of flow based on a
known discharge / flow rate in a stream, based on discussion in Section 8.4.
The example commences after the stream data (such as cross section, terrain, condition of
channel and stream profile to determine site bed slope) has been gathered (refer Chapter 4) and
the flow rate (as determine using Rational Method) has been determined (refer Chapter 5).
The task for this example is, given the stream data and flow rate (see diagram below), determine
the depth and velocity of flow in the channel.

Stream Data
Discharge / flow rate = 17.86 m3/s
Bed slope about the site is 1.2%
Ht of channel bed is 65.10 m
Max depth of flow is 2.0 m
Mannings n = 0.06.

d = ?? m

sides are 1
on 1

8B

4m
Solution

To solve for d, we need to use Mannings formula and develop a Stage-Discharge curve.

R 0.667 S 0.5
V
n
A Stage-Discharge curve plots discharge against depth of flow. Therefore several iterations
using Mannings formula are required for several depths of flow.

March 2010

8B-5

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Step 1. Using the maximum channel depth of 2.0m, calculate stream velocity and flow rate.
Calculate the cross sectional area of the flow, wetted perimeter and hydraulic radius:

A = 12.00 m2, P = 9.66 m therefore R = 1.24 m


Now,

Using Q/VA,

1.24 0.667 0.012 0.5


2.11m / s
0.06

Q=25.33 m3/s

This flow is greater than the known discharge therefore we know that the channel can easily
carry the flow.

Step 2. Now, using the same method, re-calculate stream velocity and flow rate for several
lesser depths (suggest using even increments).

8B

Depth

Velocity

Discharge

1.50

8.25

8.24

1.00

1.83

15.07

1.00

5.00

6.83

0.73

1.48

7.42

0.50

2.25

5.41

0.42

1.02

2.29

Step 3. Now draw the Stage-Discharge curve for this site / channel (refer next page).

Step 4. From the curve, we can now read of the flow depth for our design flow of 17.86m3/s.

Q = 17.86 m3/s, therefore d = 1.62 m

Step 5. Now we can use the depth to calculate flow area, then Q=V.A to determine the average
flow velocity.
A = 1.622 + 1.62 x 4 = 9.10 m2

17.86 m3/s = V x 9.10 m2


V = 1.96 m/s

March 2010

8B-6

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Stage-Discharge Curve
30

Flow Rate (m3/s)

25

20

15

10

0
0

0.5

1.5

2.5

Flow Depth (m)

End of Exercise

8B

March 2010

8B-7

Department of Transport and Main Roads


Road Drainage Manual

8B

March 2010

8B-8

Appendix 8B
Worked Examples

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Mannings Formula Example 3


This example describes the process to determine the flow rate, the average velocity of a flow in
a compound stream based on discussion in Section 8.4.
The example commences after the stream data (such as cross section, terrain, Mannings n and
stream profile to determine site bed slope) has been gathered (refer Chapter 4).
The task for this example is, given the stream data and height of flow (see diagram below),
determine the average velocity of flow in the channel and the flow rate.

Stream Data
Bed slope about the site is 0.8%
d = 1.2m
A
sides are
1 on 1

d = 2.1m
B

d = 1.0m
C

4m

3m

n values
A = 0.07
B = 0.035
C = 0.06

2m

Solution

To solve for Q, we need to use Mannings formula for each sub section of stream:

8B

R 0.667 S 0.5
V
n
After calculating V for each sub-section, use Qtotal = VA x AA + VB x AB + VC x AC to determine
total flow rate.

Step 1. For sub-section A, calculate VA using Mannings equation.


Calculate the cross sectional area of the flow, wetted perimeter and hydraulic radius for subsection A are:

AA = (1.22 /2) + 1.2 x 4.0 = 5.52 m2


PA = (1.22 + 1.22) + 2.5 = 5.70 m

March 2010

8B-9

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

It is important to remember that that water - water boundary between sub-sections A & B does
not contribute any length to the wetted perimeter.

RA = A / P = 4.44 / 5.89 = 0.97 m

Now,

VA
Using

Q/VA ,

0.97 0.667 0.008 0.5


1.25m / s
0.07

QA = 6.90 m3/s

Step 2. For sub-section B, calculate VB using Mannings equation.

AB = 7.39 m2
PB = 4.83 m
RB = 1.53 m

Now,

VB

8B

Using

Q/VA,

1.5310.667 0.008 0.5


3.40m / s
0.035

QB = 25.09 m3/s

Step 3. For sub-section C, calculate VC using Mannings equation.

AC = 3.50 m2
PC = 4.41 m
RC = 0.79 m

Now,

VC
Using

March 2010

8B-10

Q/VA,

0.7930.667 0.008 0.5


1.28m / s
0.06

QC = 4.47 m3/s

Department of Transport and Main Roads


Road Drainage Manual

Appendix 8B
Worked Examples

Step 4. Now we can calculate:

Qtotal = 6.90 + 25.09 + 4.47 = 36.46 m3/s


and
Atotal = 5.52 + 7.39 + 3.50 = 16.41 m2
therefore
Vavg = 36.46 / 16.41 = 2.22 m/s

End of Exercise

8B

March 2010

8B-11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Chapter 9
Culvert Design

January 2013

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Chapter 9 Amendments Jan 2013


Revision Register
Issue/
Rev
No.

Reference
Section

9.2.1, 9.2.6,
9.7.2,
9.7.3, 9.15
& 9.16

January 2013

ii

Description of Revision

Initial Release of 2nd Ed of manual.


Minor update.
Incorporation of Planners & Designers
Instruction No:7.
New Appendix 9D & 9E.

Authorised
by

Date

Steering
Committee

Mar
2010

M
Whitehead

Jan
2013

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Table of Contents
9.1

9.2

Introduction

9-1

9.1.1

Overview

9-1

9.1.2

Constructability

9-2

9.1.3

Computer Programs

9-2

General Requirements

9-2

9.2.1

Pipe Joint Types

9-2

9.2.2

Geometric Tolerances and Cover Requirements

9-3

9.2.3

Skew Angle / Skew Number

9-3

9.2.4

Minimum Culvert Size

9-3

9.2.5

Outlet Flow Velocity

9-3

9.2.6

Structural and Configuration Requirements of Culverts

9-4

9.2.7

Culverts in Expansive Soil Areas

9-10

9.2.8

Flap Gates (Tides and Floods)

9-16

9.2.9

Multiple Barrels

9-17

9.2.10

Reduction in Culvert Size

9-17

9.2.11

Splay Pipes

9-17

9.3

General Process for Culvert Design

9-18

9.4

Location of Culverts

9-20

9.5

Allowable Headwater

9-21

9.6

Preliminary Selection of Culvert Size

9-22

9.7

Requirements for Fauna Passage

9-22

9.7.1

Considerations for Fish Passage

9-22

9.7.2

Culvert Specifications for Fish Passage

9-22

9.7.3

Considerations for Terrestrial Passage

9-24

9.7.4

Culvert Features for Fauna Passage

9-24

9.7.5

Stock Underpass

9-26

9.8

Selection of Culvert Type

9-26

9.9

Typical Culvert Operating Conditions

9-27

9.9.1

Inlet Control Conditions

9-29

9.9.2

Outlet Control Conditions

9-29

January 2013

iii

Department of Transport and Main Roads


Road Drainage Manual

9.10

9.11

Hydraulic Calculations

Chapter 9
Culvert Design

9-29

9.10.1

Control at Inlet

9-29

9.10.2

Control at Outlet

9-30

9.10.3

Determination of Tailwater

9-33

Design Procedure

9-33

9.11.1

Hydraulic Design

9-33

9.11.2

Practical Design

9-40

9.12

Minimum Energy Loss Culverts

9-40

9.13

Blockage of Culverts

9-41

9.14

Consideration of Large or Extreme Events

9-42

9.15

Culvert Outlet Protection

9-42

9.16

Special Energy Dissipation Structures at Culvert Outlet

9-45

9.17

Self Cleaning Culverts

9-45

9.18

Inlet Structures

9-45

9.19

Managing Sediment

9-45

9.20

Safety

9-46

9.20.1

Culverts Used As Walkways and Bikeways

9-46

9.20.2

Barriers to Flow

9-46

9.20.3

Additional Safety Considerations

9-46

January 2013

iv

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Chapter 9
Culvert Design
9.1

Introduction

9.1.1

Overview

Culverts are important hydraulic structures


used to convey water across a road corridor
or in one of a range of other situations.
Culverts must be designed to convey this
flow in an acceptable way, considering the
hydraulic conditions and the required
performance (level of flood immunity) of
the road.
Environmental and/or other
requirements may also need to be
considered / incorporated depending on the
specific circumstances.
In particular, the provision of fauna
passage, including fish movement, may
need to be incorporated into the design. In
certain situations (typically where specified
by an environmental assessment document),
culvert designs will have special
requirements to allow the passage of fish or
terrestrial fauna.
Other uses for culvert structures may
include pedestrian / cyclist movement,
vehicular movement (including rail / cane
rail), fauna crossing or stock underpass.
While the requirements will differ
depending on whether passage is to be
catered for, all will require a clear
distinction between wet and dry
passageways.
The size requirements for passage as well
as hydraulic requirements need to be
considered and the culvert sized
appropriately to meet both of these
requirements.

If no passage is to be provided, then the


culvert is designed solely for hydraulic
purposes.
Hydraulic design of culverts is often an
iterative process. As such, while the layout
of this chapter reflects a typical order of
design, this will not always be the case. In
many instances it may be necessary to
revisit steps in the design process. Section
9.3 sets out a general process for the design
of the culverts while Sections 9.10 and 9.11
provide more detail. The process adopted is
considered as industry practice and while a
simplified process, it is appropriate for most
culverts designs.
Changes may arise for technical or nontechnical reasons, and hence a clear record
of the decision making process should be
retained at all times.
Procedures for assessing the erosion
potential and control at the inlet and outlets
of culverts are included as part of the
culvert design process. In some cases,
hydraulic structures for permanent erosion
control may be required with the design of
culvert inlet and outlet structures and
energy dissipaters considered.
In some locations, the design will be
complex and beyond the methods presented
in this chapter. In this instance, computer
model development, simulation and testing
may assist in design and the designer
should contact the Director (Hydraulics),
Hydraulics Section, Engineering &
Technology Division for further guidance.

January 2013

9-1

Department of Transport and Main Roads


Road Drainage Manual

9.1.2

Constructability

When designing culverts, construction and


future maintenance requirements must be
considered with appropriate treatments
incorporated into the design.
Some
guidance with respect to construction
requirements and methods can be found
within various departmental standard
drawings and specifications. However, this
must not prevent all hydraulic and
environmental requirements from being
satisfied.
Relevant standard drawings for culverts are:
1131, 1132, 1145, 1148, 1174, 1179, 1284,
1303, 1304, 1305, 1306, 1316, 1317, 1318,
1319, 1320 and 1359 (QDMR 2009b).
The main departmental specifications are
MRTS03 and MRTS04 (TMR 2010c).
For any possible non-standard or
complicated culvert configurations, it is
highly recommended that designers should
involve construction personnel early in the
design process to provide site specific
construction / constructability advice.

9.1.3

Computer Programs

With reference to Section 1.2.2.5, the


department requires the use of the
CULVERT software to hydraulically
design culverts and model them within the
12d Model application. The reasoning
for this requirement is that the software:
performs the hydraulic design of
cross road drainage consistent with
practice detailed in this chapter;
provides records of the culvert
design;
produces the drawing of the designed
drainage
cross
section
to
departmental standards, i.e. showing

January 2013

9-2

Chapter 9
Culvert Design

chainage, skew number, invert


heights, culvert component details
(including joint type), quantities of
precast components and installation
materials, hydraulic design details;
and
incorporates the culvert model into
the project electronic model within
the 12d Model application.
As with all tools, use of this computer
program will require output to be checked
and verified.
As a quick check, the head loss through a
culvert is typically around 1.4 to 1.7 times
the velocity head. Alternatively, review the
design charts provided in this chapter (and
appendices) or perform a quick manual
calculation.

9.2

General Requirements

9.2.1

Pipe Joint Types

There are three types of joints for


reinforced pipes:
Flush or Butt Joint;
Rubber Ring Joint / Spigot and
Socket; and
Jacking.
Designers should choose the most
appropriate joint type for each installation.
Flush or Butt Joint pipes are best suited
where ground movement is not expected.
They are an economical option but
installations require a high level of
compaction where the resulting soil
envelope is extremely stable.
Rubber Ring Joint / Spigot and Socket type
pipes allow for ground movement more
than flush joint pipes and should generally
be used for all sizes of pipe in unstable

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

ground, when pipes are laid in sand, or


where pipe movement is possible, such as
on the side of fills or at transitions from cut
to fill. This type should also be used where
the normal groundwater level is above the
pipe obvert.

Skew
Number

Jacking pipes are used where conventional


excavation / laying / backfill methods are
not feasible. Designers should refer to
manufacturers guidelines for selection, use
and design detail for jacking pipes.

9.2.2

Geometric Tolerances and


Cover Requirements

The geometric tolerances for location /


position of culverts and minimum cover
requirements will be as specified in the
drawings or as per the departments
specification MRTS03 (TMR 2010c).

9.2.3

Skew Angle / Skew


Number

As per MRTS01 (TMR 2010c), the


determination of the Skew Angle and/or
Skew Number for a skewed culvert
installation is shown in Figure 9.2.3. These
terms are used to specify the horizontal
orientation of skewed culvert structures
relative to the road centre line.
The skew number is the number of degrees
measured in a clockwise direction from the
road centreline to the structure centreline.
Skew Number is required for ordering
metal structures.

Figure 9.2.3 Skew Number / Skew Angle

9.2.4

Minimum Culvert Size

The minimum diameter of any pipe culvert


shall be 375 mm.
The minimum waterway dimension (height)
of any box section shall be 375 mm,
however in constrained situations and
where all reasonable attempts to fit a 375
mm high box section have failed, a
minimum 300 mm high box culvert can be
used.

9.2.5

Outlet Flow Velocity

The calculation of outlet velocity is


required for every culvert, and the type and
extent of protection should be decided after
consideration of the outlet velocity, the
natural downstream ground conditions, the
natural stream velocities, and the
performance of existing culverts in the area.
High outlet velocities can cause erosion for
a significant distance downstream of an
outlet. Where high outlet velocities are
expected, appropriate dissipation and/or
protection measures will be required.
Where possible, culverts should be
designed to provide acceptable velocities
without the need for additional channel
protection. Allowable velocities to avoid

January 2013

9-3

Department of Transport and Main Roads


Road Drainage Manual

scour vary according to soil type and


topography.

unit strength (for example, pipe


class);

Table 9.2.5 provides some target culvert


outlet velocities for various channel
materials. These target velocities must be
verified / amended by data gathered from
site inspections and relevant environmental
reports.

type of bedding and backfill material;


and

Table 9.2.5 Target Culvert Outlet


Velocities
Material
Downstream of
Culvert Outlet

Target Outlet
Velocity (m/s)

Rock

4.5

Stones 150
mm
diameter or larger

3.5

Gravel 100 mm or
grass cover

2.5

Firm loam or stiff


clay

1.2 - 2.0

Sandy or silty clay

1.0 1.5

Source: derived from Austroads GRD Part 5 (2008)

9.2.6

Structural and
Configuration
Requirements of Culverts

Loads on buried culverts include:

Chapter 9
Culvert Design

(a) Fill over the structure, which is a


function of:
height of fill
type of fill material
installation conditions (e.g. trench
or embankment)
(b) Design traffic loads;
(c) Construction traffic loads; and
(d) Other or abnormal load conditions.
The load bearing capacity of a culvert is a
function of:

January 2013

9-4

pipe diameter
culverts).

(excluding

box

The following specifies the requirements


with respect to the determination of design
loadings for buried structures and selection
of correct pipe class, based on the
departments
technical
specification
MRTS25 (TMR 2010c).
(a) Determination of Design Loads
The methodology, to determine the design
loads on a culvert is specified in Clause 6 of
AS3725. With reference to Clause 6.5.3.3
of AS3725, the calculated distribution of
working loads shall comply with the
requirements as specified in AS5100.2. For
uniformity, the department has adopted
AS5100.2 to provide the design criteria for
loads on all buried structures. The load
distribution ratio is to be in accordance with
Clause 6.12 of AS5100.2.
It is important to note that (for departmental
projects) Table B2 of AS3725 is not
accepted as it is based on a load distribution
ratio which is not in accordance with
AS5100.2. At the time of issue all known
pipe load and class determination software
available from various companies and
associations, do not use the TMR accepted
load distribution ratio. This software must
not be used until a proof of change can be
verified by Structures Branch within
Engineering & Technology.
Design loadings need to be determined for
SM1600 and HLP400 loads. Table 4-A of
MRTS25 shows correct live load pressures
(as it is based on AS5100.2) for W80 wheel
load.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

(b) Determination of Pipe Class

coatings. Refer Section 9.2.6.8 for more


details.

The design load, as determined above,


becomes the proof load for using Table 4.2
of AS4058 to determine the required pipe
class.
It is important to note that civil designers
are restricted to specifying culverts up to
and including Class 10 pipes. Any situation
or requirement to use higher class pipes
must be reviewed and approved by
Structures Branch.
(c) Support Types
The department generally specifies either
H2 or HS3 support types - refer AS3725,
Standard Drawing 1359 (QDMR 2009b)
and Section 9.2.6. Other support types can
be used, however the use of these types be
fully detailed and specified in the projects
documents as they are not covered by the
departments
Standard
Specifications
and/or Drawings.

9.2.6.1 Structural Requirements


for New Structures
The design life for all culvert types is now
100 years. The following departmental
specifications (TMR 2010c) apply to
culvert components:
MRTS24 For concrete box culvert
components;
MRTS25 for precast concrete pipe
components;
MRTS26 - for fibre reinforced pipe
components; and
MRTS29 for polyethylene and
polypropylene pipe components.
Steel culvert components do not currently
have
an
applicable
departmental
specification and do not meet the 100 year
design life requirement without protective

9.2.6.2 Structural Requirements


for Existing Structures
The road network is increasingly subjected
to vehicles carrying heavy loads and
therefore planners and designers must
consider the freight task for the road link.
Bridges and culverts designed since 2004
are in accordance with AS5100 which has
SM1600 design loads and HLP 400
vehicles. Bridges and culverts that were
designed and constructed prior to 2004 do
not have a design loading capability
equivalent to current design loads. Of
particular concern are pre-2004 culverts
that are subject to critical loading, such as
structures that have less than 2 metres of
cover (including pavement layers). These
structures may influence the freight task
capability of the road link in which they are
located and should be considered for
replacement if they cannot support the
anticipated future loads.
Specifically, pre-cast and cast-in-situ
culverts purchased or placed in the road
network before 1976, were designed for
significantly lighter loads and may also be
past their design life (which was typically
50 years). This means that pre-1976 culvert
structures on the road network with less
than 2 metres of cover are likely to be
subjected to repetitive and/or peak loadings
that exceed the structural capacity of the
aged culvert structure. This reduces their
service life and greatly increases the risk of
structural failure.
While critical loading is less of an issue for
all culverts with two or more metres of
cover, dead loads can be. These structures
should also be inspected and replaced if
found deficient.

January 2013

9-5

Department of Transport and Main Roads


Road Drainage Manual

The following requirement (policy) is not


intended as a network management tool or
guide and should only be applied on a
project by project basis.

For all planning, design and construction


projects (excluding non-drainage related
routine maintenance and sprayed reseals) it
is strongly recommended that no pre-1976
culvert be retained. Retaining any pre-2004
culverts, including pre-1976 culverts, is an
action that will potentially extend the
service life of the culvert (or parts of it; e.g.
centre units of culvert extensions) beyond
its design life and placing the structure at
greater risk of failure. Therefore each pre2004 structure to be retained, must undergo
the Structural Assessment / Decision
Process as shown in Appendix 9D. While
this process applies to all culvert structures,
major culverts (as defined in the
departments Bridge Inspection Manual
(BIM)) are currently inspected on a regular
basis, with records stored in the
departments Bridge Information System
(BIS). All other culverts are deemed minor
culverts and inspections should be carried
out based on the philosophy within the
BIM. Inspectors should take an inventory
of the structure based on BIM Appendix C
and assess the Condition State of the
structure using BIM Appendix D. This will
be based on available information and
access to structure. Use of CCTV maybe of
assistance to inspect small culverts.
The process within Appendix 9D details
both the engineering and departmental
management components of assessment in
order to determine if it is appropriate and/or
justified to retain a culvert structure instead
of replacing it. Appendix 9E is to be used
in conjunction with Appendix 9D and
assists the user in determining the risk
(probability) of structural failure of a given

January 2013

9-6

Chapter 9
Culvert Design

culvert.
Further advice in relation to
structural requirements can be obtained
from the Director (Bridge Design). Advice
in relation to inspections can be obtained
from Director (Bridge Asset Management).
It is highly recommended to replace steel
culverts greater than 15 years of age.
Departmental experience has shown that
these structures do not generally last much
longer than 15 to 20 years and structural
failure of these types of culverts can occur
quickly.
The requirement for inspection may be
relaxed if a pre-2004 culvert is assessed as
hydraulically deficient when compared to
the projects performance and capability
requirements. In this instance, the culvert
should be replaced to improve hydraulic
capability and this would remove the
requirement for structural inspection.
However if it is intended to retain the
existing structure and simply modify it to
meet project performance (hydraulic)
requirements, then the inspection/decision
to retain process is required.

9.2.6.3 General Configuration


Requirements
Culverts are laid in single or multi-cell
installations. Culverts are also laid straight
and on a constant slope / grade (also refer
Sections 9.2.10 and 9.2.11).
For multi-cell installations, each cell or
barrel must be of the same size / dimension.
There are two
requirement.

exceptions

to

this

Where a different cell size is required


in a multi-cell installation to
accommodate passage (human or
fauna), similar to the configuration as
shown in Figure 9.7.4(c). or

Department of Transport and Main Roads


Road Drainage Manual

In a Slab Link Box Culvert (SLBC)


installation where the spanning
slab(s) maybe longer than the width
of the box unit (refer manufacturer
for additional details).
Where a different cell size is required, the
designer should note that design software,
standard drawings and specifications do not
fully cover these types of installations.
Therefore the designer is required to
manually design these installations and
include all details and specification in the
model / drawings / project documents.
As per the departments specification
MRTS03 (TMR 2010c), Rubber Ring Joint
/ Spigot and Socket pipes are placed with
their spigot end facing the culvert outlet.

9.2.6.4 Reinforced Concrete


Pipes
The strength classes of reinforced concrete
pipes (RCP) generally used are 2, 3 and 4.
The minimum strength class for concrete
drainage pipes should be Class 2.
Higher strength pipes, expressed as
multiples of 2 (6, 8 and 10), are also
manufactured to order.
It is important for designers and
construction contractors (essential for the
latter) to consider construction loads when
determining pipe class during design /
installation
of
culverts
to
avoid
overstressing installed pipe units and cause
cracks / fractures. While temporary load
mitigating measures during construction
exist, it may be an economical solution to
use a higher class pipe.
With reference to AS 3725 and Standard
Drawing 1359 (QDMR 2009b), the support
condition normally used is H2 for both
embankment and trench installations. The
HS3 support condition should be used

Chapter 9
Culvert Design

where the height of the cover is critical, and


where significant savings can be made by
using a lower class pipe.
To illustrate this point, the H2 support
condition is typically cheaper to construct
than HS3. However, under the same site,
load and cover conditions, it may be
possible to use a lower strength pipe
together with HS3 support than compared
to a higher strength pipe with H2 support.
These combinations of pipe strength and
support condition should be cost analysed
to determine an economical solution.
The load on a pipe installed in a trench
depends upon the width of the trench and is
particularly sensitive to any changes in this
dimension. Therefore, the trench width
specified by design is a maximum
allowable and not a minimum.
It is important that the maximum allowable
trench widths used in the design be strictly
adhered to during construction. If any
doubt exists as to whether the design trench
width can be maintained (for example, wet
weather causing erosion and slips of the
trench wall), it is recommended that the
embankment installation be used to
determine
allowable
heights
of
embankment.
For heights of cover less than 3 m, the
allowable height depends on both live and
earth loads. At these lower heights, the live
load is a major contributor to the total load
on the pipe.
Design engineers can use other support
conditions as detailed in AS 3725, however
Standard Drawings and Specifications do
not address these.

9.2.6.5 Fibre Reinforced Pipes


Fibre reinforced concrete (FRC) pipes are
similar to reinforced concrete pipes in their

January 2013

9-7

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

requirements and are available in sizes 100


mm to 750 mm in Classes 1, 2, 3 and 4.
They are generally suitable anywhere an
RCP could be used (refer Section 2.3.5.3).
The reinforcement is a cellulose (wood
fibre) product. Hydraulic design curves for
precast reinforced concrete pipes (RCP) are
appropriate, although FRC pipes have
slightly lower roughness coefficients.

9.2.6.6 Reinforced Concrete Box


Culverts
Standard precast reinforced concrete box
culverts (RCBC) are manufactured in a
range of standard sizes. Small box culverts
are available in the range 300 x 150 mm to
1200 x 900 mm and large box culverts in
the range 1500 x 600 mm to 3600 x 3600
mm.
In accordance with the departments
specification MRTS24 (TMR 2010c),
standard RCBC and link slab components
have been manufactured to withstand a
maximum height of fill (including
pavement) of 2.0 m. For higher fills
(including pavement) a special design for
proposed installation is necessary.

For multi-cell installations, slab linked


construction (SLBC) should be considered,
as shown in Figure 9.2.6.6. Details are
shown on Standard Drawings Nos. 1303,
1316, 1317, 1318, 1319 and 1320 (QDMR
2009b).

Figure 9.2.6.6 Slab Link Box Culvert


under Construction

9.2.6.7 Reinforced Concrete Slab


Deck Culvert (RCSDC)
The standard RCSDC of 2500 mm span
allows for a maximum fill of 2500 mm
above the slab deck and is a cast-in-situ
structure (refer Standard Drawings Nos.
1131 and 1132(QDMR 2009b)).

9.2.6.8 Corrugated Steel Pipes


Corrugated steel pipes, as shown in Figure
9.2.6.8(a) offer a light-weight alternative
when compared to concrete pipes. This is a
particularly important consideration when
transport of material over long distances
becomes costly.

RCBC and SLBC can be used with nominal


minimal cover above the culverts. Nominal
cover has advantages in regions where the
height of the overall formation is critical.
In expansive soil areas, RCBC and SLBC
installations require special consideration.
Refer Section 9.2.7 for details.

January 2013

9-8

Figure 9.2.6.8(a) - Corrugated Steel


Pipes

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

The following Australian Standards are


relevant to corrugated steel pipes:
AS 1761-1985 Helical lock-seam
corrugated steel pipes;
AS 1762-1984 Helical lock-seam
corrugated steel pipes - Design and
installation; and
AS 2041-1998 Buried corrugated
metal structures.

corrugated steel pipe culverts shall be lined


with concrete as follows:
Concrete Class: 32 MPA / 9.5;
Minimum depth of concrete above
corrugations: 50 mm; and
Minimum height of lining above
invert to be D/6 where D denotes
diameter as shown in Figure
9.2.6.8(b).

The departments specification MRTS03


(TMR 2010c) also contains requirements
for supply and construction of metal
culverts.
New steel culvert components must have a
protective coating applied to both internal
and external surfaces in order to meet the
required 100 year design life.
The
protective coatings must conform to:
Galvanizing to Z600 in accordance
with AS 1397; and
A polymer coating to ASTM A742.
Before specifying in design / used in
construction, any proposed steel culvert
product with the above protective coating
applied must be approved by Bridge Design
Section, Engineering & Technology
Division).
Additional design
helical culverts are:

requirements

for

The maximum diameter of helical


pipe culverts complying with
AS1761 and AS1762 shall be 3600
mm;
The maximum flexibility factor for
installation shall not exceed the limits
in AS1762.
Apart from the above protective coating
requirement, steel culverts must also have
invert protection. The invert of all

Figure 9.2.6.8(b) - Culvert Lining Detail

It is important to note that any bolts or lugs


that are connected to the culvert to allow
anchorage for the concrete invert must also
be covered by the protective galvanizing /
polymer coating.
Further details for concrete lining are
contained in the departments specification
MRTS03 (TMR 2010c).
The design of Corrugated Metal Helical
Pipe and Arch Culverts is not as simple as
for concrete culverts and therefore should
be undertaken by an experienced design
engineer. Furthermore, installation of metal
culverts is also specialised and therefore
should be undertaken by qualified /
experienced personnel.
Assistance and
advice with respect to the design and
installation of metal culverts can be
provided by Bridge Design Section,

January 2013

9-9

Department of Transport and Main Roads


Road Drainage Manual

Engineering & Technology Division, when


required.

9.2.6.9 Plastic Flexible Pipes


Flexible plastic pipes are being used in
small quantities by some road authorities.
Conditional approval has been given for the
use of these types of pipes on some
departmental projects. They have typically
been allowed to be used in non-traffic
loaded areas.
They have been used successfully to line
corroded or damaged pipes as their light
weight allows easier installation. They
should be considered as a viable alternative
in very acidic and corrosive sites.
Disadvantages include additional care
needed during installation to prevent
excessive movement and flotation of the
flexible pipe and concerns about
performance should a fire occur.
Further work is being undertaken by the
department with regard to the specification
and use of flexible plastic pipes (including
in traffic loaded areas) and outcomes of this
work will be included in updates to this
manual.

Generally, design is to comply with


AS/NZS 2566.1:1998 - Buried flexible
pipelines Part 1. Structural design and
installation is to comply with AS/NZS
2566.1:2002 - Buried flexible pipelines Part
2: Installation.

9.2.6.10

Aluminium Culverts

Aluminium culverts have been used in


small quantities by the department.
Aluminium is suitable for use in pH range 4
to 9. The advantages of aluminium culverts
relate to their ability to be used in areas
where the natural material has a pH in the
range of 4 to 6. This pH range is aggressive

January 2013

9-10

Chapter 9
Culvert Design

to many materials and aluminium culverts


represent a durable solution in these
environments.
Design criteria for aluminium culverts are
covered in AS/NZS 2041.
Aluminium is not as strong as steel;
consequently a larger wall thickness is
required for an equivalent installation as
compared to steel. Aluminium is also more
expensive than steel.

9.2.6.11 Other Structural


Aspects
All other structural aspects should be
referred to the departments Bridge Design
Section, Structures Branch located within
Engineering & Technology for advice
and/or guidance.

9.2.7

Culverts in Expansive Soil


Areas

Expansive soils pose particular problems in


civil engineering works due to shrink-swell
behaviour.
With respect to drainage
structures, this movement in the soil places
uneven stresses on RCBC, SLBC and RCC
culverts and can damage the structure as
shown in Figures 9.2.7(a) & (b). In Figure
9.2.7(a), the longitudinal crack stops at the
slab which forms the footing of the wing
wall. While the apron has risen relative to
the wingwall, the outer edge of the apron
has dropped and now slopes away from the
culvert. Figure 9.2.7(b) shows the vertical
displacement across the break in the apron
slab. Again, the apron has risen relative to
the wingwalls.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Longitudinal Crack

Figure 9.2.7(a) Longitudinal Cracking of Apron Failure

Figure 9.2.7(a) Vertical Displacement of Apron Failure

January 2013

9-11

Department of Transport and Main Roads


Road Drainage Manual

Other issues, such as culverts which appear


to rise above their original level,
approaches to culverts that deform, and
pavement distress over culverts, are often
the result of volume changes in expansive
soil foundations and embankments. With
differential ground movements taking place
at the culvert / approach interfaces, heavy
vehicles can inflict high axle impact loads
on the structure. The design load of
culverts is influenced more by high impact
loads of individual axles rather than the
behaviour of the entire vehicle. High
impacts on structures can dramatically
reduce service lives.
This section addresses:
multiple barrel, reinforced concrete
box culverts (RCBC) and slab link
box culverts (SLBC); and
multiple barrel, reinforced concrete
culverts (RCC);
with a total length greater than 10 m
along the road centreline in expansive soil
conditions.

9.2.7.1 Expansive Soil Potential

The shrink-swell behaviour of expansive


soils is caused by moisture movement in the
soil brought about by climatic changes
producing moisture variations from extreme
wet to extreme dry or vice versa. Examples
are:
soils in arid climates, usually in a
desiccated state i.e. cracked, which
are subjected to occasional unusually
high rainfall or prolonged inundation
causing the soil to saturate and
expand;
soils in semi-arid climates where the
moisture conditions of the soil reflect
the wet-dry seasonal cycle and may

January 2013

9-12

Chapter 9
Culvert Design

be subjected to occasional climate


extremes of drought and flood; and
predominantly wet soils which from
time to time are subjected to a
prolonged period of drought and
exhibit drying shrinkage.
The shrink-swell behaviour or volume
change phenomena is controlled by three
major factors:
intrinsic expansiveness of the soil
(generally characterised by shrinkswell index for the soil);
suction change (site specific and
dependent on the atmospheric
conditions); and
applied stress.
Changes in soil moisture produce suction
changes, which in turn produce a loading /
unloading effect on the soils and result in
volume changes in the soil. The two most
important site-specific issues with regard to
suction are:
the postulated suction change at the
surface;
the depth over which the suction
change manifests, called the active
depth (generally between 2 to 5 m).
Whilst some guidance is available for
expansive soil embankments in road
construction the problem with drainage
structures in expansive soils is different in
that it is a soil / structure interaction
condition.
Experience has shown that expansive soil
problems generally tend to occur in soils
which have a Linear Shrinkage greater than
8% and/or swell strains greater than 5% at
OMC (based on a multi-point soaked CBR
test).

Department of Transport and Main Roads


Road Drainage Manual

According
to
the
Unified
Soil
Classification, these soils range from
SC/CL to CH and are not necessarily
restricted to high plasticity CH clays.
Therefore
particular
design
and/or
construction considerations need to be
adopted to avert damage to culverts where
expansive soils are exposed to significant
long term moisture changes.

9.2.7.2 Postulated Mechanism of


Distress
The observed movement in some large
culverts is generally a movement of the
outer edges of the culvert relative to the
central section of the culvert, which is
generally immune to the movement. Figure
9.2.7.2 depicts this failure mechanism for
two case study sites.

Chapter 9
Culvert Design

It is considered most likely that the outer


edges, i.e. apron slabs, are subjected to
extremes of wetting / drying phenomena
which produce either high swelling
pressures or lack of base support.
Most small culverts (< 10 m along the road
centreline) generally only suffer small
movement which is satisfactory or exhibit a
uniform heave due to their inherent
geometric stiffness.

9.2.7.3 Standard Drawings


The current Standard Drawings for culvert
bases do not state the design assumptions
on which the drawings are based and, most
importantly, situations when the drawings
are inappropriate for use.

Figure 9.2.7.2 Failure Mechanism of Base Slabs in Expansive Soils

January 2013

9-13

Department of Transport and Main Roads


Road Drainage Manual

The design assumptions on which those


drawings are based include:

swell > 5%). This preliminary testing can


be undertaken at the Regional level.

the base slabs are designed as a beam


on a moisture insensitive, elastic
foundation, i.e. differential settlement
due to moisture changes are not a
design consideration in the standard
drawing; and

It is imperative that a vertical profile is


established to determine the extent of the
actual expansive zone.
A field
investigation should include:

the minimum ultimate bearing


capacity of the strata under the
culvert base is at least 150 kPa but
preferably in excess of 200 kPa.
Sites subject to large settlements or large
differential settlements, arising out of
moderate or highly expansive soils below
the culvert base, are outside the design
method of these Standard Drawings.

9.2.7.4 Amended Design


Procedure

Chapter 9
Culvert Design

For culverts with a base greater than 10 m


along
road
centreline,
specialist
geotechnical and structural advice (from
TMR E&T Division) should be obtained
where highly reactive or expansive clay
soils (linear shrinkage > 8% and/or CBR
swell > 5%) occur below the culvert bases.
This is to determine if non-standard base
slabs or other foundation treatments are
required.

9.2.7.5 Foundation Investigation


An appropriate, special investigation for
culvert bases on expansive soils should be
undertaken in a similar manner to the
proven need for special bridge site
investigations.
This work should be
undertaken under the direction of specialist
geotechnical engineers and/or geologists as
appropriate.
This is required only if
preliminary testing indicates the subgrade
to be expansive (i.e. LS > 8% and/or CBR

January 2013

9-14

Trenching or drilling to 2 m depth


under or in the vicinity of the
proposed culvert location;
In-situ moisture content (Q102A) and
density testing (Q111A) at every 300
mm in depth or at change of soil
horizon, whichever is earlier, to
determine the active zone. (Below
the active zone, no significant
moisture content changes occur over
time). Due to lack of data, AS 2870 1996 gives little guidance on active
depths for Queensland conditions.
For most other States, e.g. Victoria,
where the reactivity of clay profiles
has been the subject of extensive
research,
useful
guidance
is
available;
50 mm undisturbed tubes taken from
each soil horizon, for shrink-swell
index testing and filter paper suction
measurements.
Adequate materials to be sourced
from each location for the following
laboratory tests.
A laboratory investigation is required for
each soil horizon. The required tests are
detailed in Table 9.2.7.5.
If instrumented sites are established in
different soil / climatic regions, enabling a
rational classification of soil / climatic
behaviour response patterns, the level of
testing can be reduced in the future.

Department of Transport and Main Roads


Road Drainage Manual

Table 9.2.7.5 Required Tests


Parameter
Test Method
Particle size
distribution (Sieve
analysis)
Liquid Limit
Plastic Limit
Linear Shrinkage
Shrink-swell index*
Filter paper suction
measurement*

Q103A
Q104A
Q105
Q106
AS2870-1996
BRE-IP 4/93

Chapter 9
Culvert Design

ripping,
scarifying
and
then
compacting the soil with moisture
and/or density control.
These processes are carried out to a depth
beyond the level of seasonal moisture
variation within the soil. The areas to be
treated would be under the aprons and 1 m
beyond the cut-off wall of the apron.
Control
of
Foundation
Moisture
Fluctuations - Moisture Stability

9.2.7.6 Options for the Control of


Distress in Culverts
For drainage structures using culvert bases,
special measures need to be undertaken to
avert distress. Options for control of
distress of culvert bases may be categorised
into either geotechnical alternatives or
structural alternatives. In many cases,
geotechnical methods may be used
successfully in conjunction with structural
methods.
(a) Geotechnical Methods
There are broadly two geotechnical
methods for limiting damage to light
structures such as culverts constructed on
expansive soil foundations. These either
reduce the expansive potential of the soil or
minimise the seasonal fluctuations of the
subgrade moisture.
Reducing Expansive Potential of the
Foundation - Volume Stability
Methods for reducing the expansive
potential of the foundation may include one
of the following:
excavation of the foundation and
replacement with a low permeability
granular or non-swelling material;
chemically treating the natural
material (e.g. lime stabilisation); and

The aim of these methods is to control the


moisture fluctuations in the foundation
within acceptable limits.
Methods of
control may include one of the following:
pre-wetting or ponding a foundation
prior to construction;
stabilisation of foundation moisture
conditions by a physical limit, e.g.
vertical moisture barriers.
This
involves the placement of a
geomembrane
(generally
a
waterproof fabric) in a trench along
the perimeter of the slab to the limit
of the estimated active zone depth.
These vertical barriers minimise
seasonal lateral migration of moisture
to and from the foundation soils
beneath the foundation slab. Details
of backfilling and other technical
issues will need to be covered by
supplementary specifications; or
extending the concrete apron with a
flexible apron, e.g. grout-filled
erosion mattress ( 3 m wide)
underlain by an impermeable
membrane
(horizontal
moisture
barrier). This is aimed at shifting the
moisture fluctuation zone to be under
the extended apron, thus shielding
the concrete apron slab from the edge
effects.

January 2013

9-15

Department of Transport and Main Roads


Road Drainage Manual

Prediction of moisture infiltration under


sealed areas by numerical methods may be
used in estimating the required lateral
extent that needs to be provided by the
flexible apron.
(b) Structural Methods
The structural options to control distress of
culvert bases are as follows:
Improved Layout of Culverts
The risk of damage to culvert bases may in
some circumstances be reduced by limiting
the size of banks of culverts. In wide
floodplains, it is considered that a number
of banks of culverts distributed across the
watercourse will result in a better hydraulic
and structural solution.
Other Structural Solutions

The use of stiffened raft foundations (AS


2870) are technically proven solutions
widely used in the building industry. As
the culvert distress is commonly observed
within the apron area of the slab, any
stiffening needs only be confined to the
apron slab. Swell pressures can be as much
as 200 kPa, i.e. much greater than the
applied pressure at the base of the slab
(typically up to 50 kPa). Each case has to
be considered on its own merit.
Other Options - Bridges
Consideration should be given to using
short span bridges founded on free standing
piles extending to the stable material below
the active zone. Due to the limited nature
of contact between the volumetrically
active soil and the foundation elements, i.e.
the free standing piles, limited upward
thrusts are transmitted to the deck.
Therefore, these foundation systems are less
influenced by the movement of the ground
and allow such designs to be optimised.
However, expansive soil issues with bridge

January 2013

9-16

Chapter 9
Culvert Design

abutments and general bridge maintenance


requirements would still need to be
resolved.
Bridges would not be a practical option for
low height structures, but the actual height
limit has not currently been determined, and
it may vary for different sites.

9.2.7.7 Improved Construction


Practice
Consideration should be given to restricting
construction practices which adversely
affect the moisture content of the soil. The
following practices should be excluded:
placement of permeable fill behind
the culvert (either granular or cement
stabilised sand);
opening a culvert base up for a
prolonged period when the moisture
content is low; and
not allowing adequate time for the
culvert base to reduce moisture
content after a prolonged wet period.
Supplementary specifications may be
required in such situations, and may require
specialist input from geotechnical and
structural
specialists
within
the
departments E&T Division.

9.2.8

Flap Gates (Tides and


Floods)

Where the outlet to a culvert may be


submerged by tide or a flood from
downstream sources, and where it is
necessary to prevent the flow of such
waters into the culvert, it may be necessary
to install flap gates.
However, flap gates will generally cause a
higher head loss to occur. Therefore,
reference should be made to loss

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

the

particularly when slab link box culverts


(SLBC) are used.

Where flap gates are required, it will be


necessary to ensure that only those culvert
types for which gates are available are
selected. Obvious impacts on fish passage
must also be considered.

For large culverts, it is often possible to


utilise link slabs to reduce costs. In these
situations, selection of an odd number of
box culvert cells is preferred, though by no
means essential. An odd number of cells
should only be sought if no other design
criteria are compromised.

coefficients
as
manufacturers.

provided

by

Regular maintenance of flap gates is


required to ensure their efficient operation.
This is especially important for locations
where there is significant debris or sediment
transport.
If regular maintenance is
unlikely, then flap gates may not be
appropriate.
In addition, if the culvert is expected to pass
large quantities of sand, then the outlet
should be raised above the downstream
invert to avoid sediment blockage of the
gate(s).

9.2.9

Multiple Barrels

When two or more barrels of pipes are laid


parallel, they should be separated by the
dimensions as shown on the relevant
standard drawing and within MRTS03
(TMR 2010c).
The spacing allows for thorough
compaction of the back-fill material which
is essential for haunch support and the
prevention of settlements.
Multiple pipe culverts should always be
treated as an embankment installation when
determining the class of pipe required from
the allowable height of cover. Even for
culverts installed in trench conditions, the
height of cover should be calculated for
embankment installation.
For multiple cell culverts in a restricted
natural waterway, box culverts can make
better use of the width available,

9.2.10 Reduction in Culvert Size


New culvert installations must maintain the
same size diameter pipe / box dimensions
for its whole length.
For culverts being extended:
a reduction in culvert size is not
permitted on the downstream side as
the discontinuity between the
different pipe sections can catch
debris causing blockage which inturn
reduces the capacity of the culvert
and/or can cause failure of the
culvert; and/or
a reduction in culvert size is
permitted on the upstream side
provided that the hydraulic capability
is not compromised.
This also applies for projects where the
inflow to an existing culvert has been
reduced and the culvert requires extension.
Where culverts require strengthening by
insertion of a sleeve or similar, the internal
dimensions / diameter must be maintained
for the length of the culvert.

9.2.11 Splay Pipes


The use of splay pipe components to
construct bends in culverts is not
permitted.

January 2013

9-17

Department of Transport and Main Roads


Road Drainage Manual

However, a relaxation of this requirement


may be approved for individual cases
within a project subject to the following
requirements:

site including detailed survey of the site,


site inspection data, and other site specific
information including environmental and
geotechnical reports.

Specific locations where it is


proposed to use splay pipes to
construct bends in culverts, are to be
reviewed and approved by the
departments design representative;

A generalised approach to the design of the


culvert follows.

Culverts must conform to relevant


departmental specifications;
Bends to be constructed in large
culverts only - for pipe diameter of
1200 mm or greater using propriety
splay pipe units and for box culverts
with widths and heights of 1200 mm
or greater using a cast in-situ
chamber without access;
Bend angles are restricted to a total
or maximum 22.5 degree culvert
deflection in the horizontal plane
only (grade of culvert must not
change);
Only one bend allowed in a single
culvert installation; and

Chapter 9
Culvert Design

The detailed design of any culvert


that includes a bend will require
formal approval by departments
Bridge Design Section / Structures
Branch
located
within
the
Engineering & Technology Division.

9.3

General Process for


Culvert Design

This section describes an approach or


general process for the design of culverts.
Some of the work described in the
following sections may not be required for
each design of a culvert.
The design of a culvert commences with
assembling the data related to the drainage

January 2013

9-18

(a) Collate Site Data


Review survey, topographic information,
locality
map,
photographs,
aerial
photographs and details from field visit to
determine / understand:
catchment / waterway details
including
natural
constrictions,
bends, low / high flow channel,
vegetation, potential overflow to
other crossings etc;
upstream and downstream conditions
and details;
location of any geotechnical issues;
location of environmental constraints
or identification of environmental
issues;
location of Private / Public Utility
Plant (PUP) or other physical
constraints;
any stream or channel diversion
issues;
any culvert skew requirements;
soils data;
existing and allowable stream flow
details (depth, velocity, energy and
so on) (refer Chapter 8);
any possible / identified inlet and
outlet erosion issues; and
possible sediment / debris issues.

Department of Transport and Main Roads


Road Drainage Manual

It is also essential to review:


recorded and observed flood data,
and local anecdotal data from
residents or local Council.
design data for nearby structures; and
studies by other authorities near the
site, including small dams, canals,
weirs, floodplains, storm drains.
From this information, the designer can
then commence the design to determine the
size and alignment of the culvert.
(b) Determine Road / Channel Geometry
Where the road alignment crosses the
stream / channel, the designer needs to
determine / understand:
the grade height of the road over the
crossing;
location of adjacent low points in the
road alignment;
height and width (location) of
shoulders (both sides) considering
crossfall / superelevation application;
embankment or batter slopes (and
possibility of steepening the slopes if
height of embankment over culvert is
too high);
pavement thickness;
shape of channel (width of bed and
bank slopes) including uniformity of
shape over the reach where the
culvert will be located; and
channel bed slope.
These parameters are required to determine
the maximum height, width and minimum
length of the culvert. This allows the
designer to fit the culvert under road and
within the channel. The parameters also
allow the determination of the height of

Chapter 9
Culvert Design

headwater that can be developed upstream


of the culvert.
(c) Determine Culvert Type and Location
Select a suitable culvert type (RCP, RCBC
etc) for site and locate the proposed culvert
along the best initial alignment and
determine the following:
an initial trial culvert size and
configuration that fits (refer (b)
above) and check cover requirements
for selected culvert type;
incorporate
any
environmental
requirement such as fauna passage;
set an outlet invert level;
based on the channel bed slope and
an initial slope for the culvert,
determine the inlet invert level;
check that the inlet invert level is at
or just below the natural bed level;
if the inlet level is well below the
natural bed level, assess the extent of
inlet works needed to avoid /
minimize siltation over time;
check / determine available space for
possible
ancillary
erosion
or
environmental protection devices;
and
identify and document any possible
issues / limitations that may
necessitate a review of the culvert
design or its location.
(d) Determine Tailwater Level
Calculate the tailwater level within
the existing channel immediately
downstream of culvert outlet.
(e) Undertake Hydraulic Design
Determine / set the maximum
allowable
headwater,
including
freeboard, for the design;
January 2013

9-19

Department of Transport and Main Roads


Road Drainage Manual

Complete the hydraulic design of


proposed culvert to determine
headwater, control, outlet velocity
Froude Number; and

As the designer works through this step by


step process, there will be some outcomes
that indicate that revisions to the proposed
culvert need to be made.

Undertake several trials based around


the initial culvert size / configuration,
including cost comparison, to select
optimum design.

The designer needs make amendments to


the proposal and restart the geometric
design at an appropriate point and continue
the design with the amended culvert
proposal.

(f) Review Hydraulic Design Output.


Check if hydraulic design is
reasonable / realistic and if the flood
immunity requirements is being
achieved;
Check generated headwater against
maximum allowable headwater;
Check outlet velocity against
permissible channel velocities;
Check extent of any additional
inundation due to afflux and review
impacts to adjacent property; and
Assess the likelihood of road
overtopping or excessive / erosive
outlet velocities in an extreme
rainfall event (how close is generated
headwater to shoulder point / outlet
velocity to maximum permissible).
(g) Check Connections

Chapter 9
Culvert Design

Can surface drains such as catch


drains and diversion drains be
drained to the culvert inlet and/or
outlet (Refer Chapter 11); and
Can underground drainage be drained
to the culvert inlet and/or outlet
(Refer Chapter 11);
(h) Assess Mitigation Treatments
Determine any inlet or outlet
protection devices to address
pollution or erosion control concerns.

January 2013

9-20

The culvert design will not be able to be


finalised until all related components of the
drainage infrastructure are defined by the
relevant sections of in Chapters 8, 9, 10, 11
and 12 are addressed.

9.4

Location of Culverts

In general, a culvert should be located to fit


the natural channel in line and grade as far
as is practical.
The culvert should be designed to suit the
outlet conditions even if inlet conditions
have to be modified, e.g. a drop inlet to
reduce potential scouring velocities through
the culvert.
High outlet velocities can cause erosion for
a significant distance downstream of an
outlet. Where high outlet velocities are
expected, appropriate dissipation measures
will be required.
The natural skew in a channel should be
adopted wherever possible, as a culvert on a
different skew could cause progressive
bank erosion, possibly entering private
property. Protection of a bank at risk could
be costly.
In most cases, culvert locations are largely
predetermined by the intersection of a
watercourse and an existing roadway.
However, where circumstances allow,
culverts should be located away from:

Department of Transport and Main Roads


Road Drainage Manual

Erodible / unstable banks


Meandering channel bends or banks;
Critical or isolated aquatic habitat
areas; and
Isolated sections of remnant, valued,
or protected riparian vegetation.
Realigning short sections of an existing
channel to fit the culvert alignment should
be avoided.

9.5

Allowable Headwater

Culverts that are designed for hydraulic


purposes only, are designed to pass the
design discharge from one side of the road
embankment to the other in a cost effective
manner in accordance with individual
project requirements.
The velocity of the water through the
culvert is usually greater than the approach
velocity in the stream because the culvert
presents a smaller cross sectional area of
flow than the stream.
As detailed in the previous chapter, the
energy in water is measured as Head (m).
When water flows through a culvert,
several losses of energy occur:
Firstly as the flow accelerates into
the culvert (known as entrance loss);
then there is friction loss along the
length of the barrel as water flows
against the culvert; and
finally as the flow decelerates out of
the culvert (known as exit loss).
In order to pass the design discharge
through the culvert, extra energy within the
flow is required on the upstream side to
overcome these losses. This extra energy is
generated by damming the flow on the
upstream side of culvert / road formation
which raises the water level (increases the

Chapter 9
Culvert Design

height of water or head), above that of


normal flow levels. This increase in level is
known as afflux and it is at its highest just
upstream of the culvert entrance.
The overall increase in water depth (afflux)
generated by the culvert cannot exceed the
allowable headwater for the site.
The allowable headwater or Maximum HW
level for a culvert will usually be
determined by one or more of the following
conditions:
Increase in Upstream Water Levels: The
increase in upstream water levels should not
cause unacceptable damage or adverse
effects to adjacent properties. Existing
flooding of adjacent properties should be
ascertained in field investigations or by
design calculations.
Upstream Freeboard: Where a road is
designed not to be overtopped during a set
ARI flood event, it is desirable to provide
freeboard of at least 100 mm between the
design upstream floodwater surface and the
upstream road shoulder edge. Where this is
not economically acceptable, pavement
design should make allowance for higher
water levels, and the likely duration of
inundation.
Outlet Velocity: In the event that both of
the above conditions permit a high
headwater, the associated outlet velocity
may be unacceptably high. In this case, the
headwater may have to be reduced to limit
the outlet velocity to a level that does not
cause unacceptable scouring.
This is
normally achieved by increasing the
waterway area of the culvert, however this
is not always possible and therefore outlet
protection, as shown in Figure 9.5 may
have to be designed.

January 2013

9-21

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

For culverts used as walkways and


bikeways
the
minimum
recommended invert level is equal to
the level / height of ARI 2 year flood
event.
For
the
minimum
recommended cell height and widths,
refer to the departments Road
Planning and Design Manual (TMR
2010a).
Figure 9.5 - Box Culverts with Riprap

9.6

Preliminary Selection
of Culvert Size

Not withstanding the minimum culvert size


requirements as specified in Section 9.2.4,
the considerations for the preliminary
selection of culvert size are:
A culvert flowing full makes most
use of the waterway opening
available. Where outlet control is
likely, the culvert opening height
should desirably be not greater than
1.25 times the depth of the tailwater
in the unrestricted channel at the job
site. This will help to ensure that the
culvert flows full. Culverts will
generally not flow full if operating
under inlet control.

The larger the waterway opening is,


the less likely it will be blocked by
debris.
An assessment of the
minimum culvert size for debris
should take precedence over any
requirement for the culvert to flow
full.
The cover requirements for various
culvert types are discussed in Section
9.2.2.
Environmental considerations such as
those nominated in Chapter 7 should
be taken into account.

January 2013

9-22

9.7

Requirements for
Fauna Passage

9.7.1

Considerations for Fish


Passage

Culverts are considered as a waterway


barrier and can limit fish movement along
the waterway. Defined waterways require a
culvert design that supports fish passage.
Definitions and key requirements are
provided in Section 2.5.2.6.
Chapter 7 provides an understanding of and
reasons for fish movement.

9.7.2

Culvert Specifications for


Fish Passage

The department is currently working with


Queensland
Fisheries
to
develop
requirements and self-assessable codes for
the design of culverts that support fish
passage, however the guidance provided in
this section should assist until updated
information can be provided.
To allow fish to swim upstream through a
culvert, the culvert invert level, slope, and
velocity may have to satisfy the following
specifications.
Preferred Culvert Type: The minimum,
culvert for fish passage is a 1200 x 900 mm
box culvert (preferred) or an 1800mm pipe.
Natural Watercourse Bed: Where
practicable, the bed of the culvert should

Department of Transport and Main Roads


Road Drainage Manual

reproduce the natural conditions of the


watercourse bed, and ideally be recessed
below natural bed levels.
Culvert Floor: The floor of the culvert
should be below natural bed levels to avoid
the need for fish to jump over an
obstruction. As depressed culvert floors
may silt up over time, especially with
undue lowering of the floor, an invert level
about 100 mm below natural surface levels
is considered appropriate with tapers to
natural surface level.
Recessed Floors: Some fishway guidelines
recommend that culverts should be recessed
into the natural bed by at least 20% of their
cell height (increases the height of culvert
over hydraulic requirements). The culvert
is then backfilled with bed material won
from the site, to the same gradient / levels
of the natural bed prior to construction.
This should be considered on a site by site
basis based on current best practice.
Floor Slope: Generally, the slope of the
floor of the culvert should not exceed 1% in
order to limit velocity increases to flows
through culvert.
Flow Velocities: As fish need a minimum
water depth of 0.2 to 0.5 m to ensure their
passage through the culvert, and can only
swim against relatively low flow velocities,
the velocity through the culvert during
periods of migration should not exceed 0.3
m/s at a water depth of 0.5 m.
To achieve this, it may be necessary to
locate an additional culvert at the shallower
edge of a stream where natural velocities
are lower, or to install a low flow channel.
Floor Surfacing: If necessary, roughening
of the floor of the culvert to achieve this
velocity may be necessary by lowering the
floor even further than 100 mm and placing

Chapter 9
Culvert Design

cobbles or boulders over the floor up to the


design level.
However, the roughness coefficient of
dumped rock is much higher than that of a
concrete floor and should be taken into
account during the design process. For
Mannings
roughness
coefficients,
reference may be made to Tables 9.4.3(a),
9.4.3(b) and Figure 9.4.3.
Low Flow Channel: Where possible, the
low flow channel should satisfy the
following conditions:
Maximum flow velocity 1 m/s;
Maximum flow velocity of 0.3 m/s at
a depth of 0.2 to 0.5 m;
Where practicable, a minimum flow
depth of 0.2 to 0.5 m; and
Absence of areas of large scale
turbulence (relative to flow depth),
i.e. whirlpools / eddies.
However, the above flow conditions are
usually difficult to achieve, especially if the
culvert cannot be recessed below natural
bed elevations. In such cases, the next
preferred option is to size the culvert flow
area such that there is minimal change in
channel flow area at the culvert.
To assist in the development of suitable low
flow conditions and to assist in the control
of sediment flow, an inlet weir can be
formed. Inlet weirs are designed to direct
all base flows to one wet cell.
The typical height of the weir that is placed
in front of the dry cells is 0.3 to 0.5 m.
For hydraulic reasons, the weir needs to be
located at least four times the height of the
weir away from the culvert entrance.
Light Conditions: When fish passage is an
issue, light should be encouraged to enter
the culvert, or in the case of a multi-cell

January 2013

9-23

Department of Transport and Main Roads


Road Drainage Manual

culvert, into the expected migration cell(s).


In multi-lane roads, stormwater drop inlets
can be installed into the median strip to
allow the entry of light.

Chapter 9
Culvert Design

Table 9.7.3 - Minimum Desirable Culvert


Cell Height for Terrestrial Passage
Size and/or Type of

Minimum

Current best practice guidelines suggest that


additional lighting is recommended when
the culvert is longer than 14 m (depending
on culvert height).

Small mammals

0.5m

Medium mammals

0.5m

Large mammals

1.2m

Further guidance and information regrading


requirements for culverts and fish passage
can be obtained from Queensland Fisheries
website.

Semi-arboreal (e.g. Koala,

>1.2m

Microchiropteran bats

>1.2m

Reptiles

0.5m

9.7.3

Considerations for
Terrestrial Passage

A guide to the appropriate culvert size to


accommodate terrestrial passage is provided
in Table 9.7.3.
To allow for terrestrial passage through a
single cell culvert, a raised ledge may need
to be formed as shown in Figure 9.7.3. The
ledge may consist of a 0.3 to 0.5 m wide,
stone pitched path elevated at least 100 mm
above the base flow water level.
The ledge must extend from upstream bank
to downstream bank, including the area
along the wing walls, thus providing a
continuous dry path.

If the ledge is placed on just one side of a


waterway, then it is important for this side
to be consistent with the side of the
waterway that is the dominant movement
corridor. In some cases it may be necessary
for a ledge to be placed on both sides of the
low flow channel.
The ledge must extend from upstream bank
to downstream bank, including the area
along the wing walls, thus providing a
continuous dry path.

January 2013

9-24

Figure 9.7.3 - Dry Passageway in Box


Culvert

9.7.4

Culvert Features for


Fauna Passage

Other features associated with terrestrial


passage are discussed below. The value
and hence need for each of these features
must be assessed on a site by site basis.

Department of Transport and Main Roads


Road Drainage Manual

Separate wet and dry cells: Most culvert


designs will have requirements for some
degree of terrestrial or aquatic movement.
Thus most multi-cell culverts will require
both wet and dry cells.
Separate wet and dry cell can be formed by
placing individual cells at different invert
levels. This may be achieved by having the
two outside cells raised at least 100 mm
above normal water level, or lowering just
one central cell below normal water level.
Dry cells can also have a duel use as
pedestrian tunnels and bikeways.
Low flow channel: In some cases, in order
to satisfy both aquatic and terrestrial
passage requirements, it may be desirable to
construct a low flow channel into the base
slab rather than constructing a raised ledge.
Fencing: Fencing, such as koala-proof
fencing, as shown in Figure 9.7.4, or
wildlife exclusion fencing, can be used to
direct animals to the culvert to avoid
animals moving over the road.
For guidelines on the use and design of
fauna control fencing, refer to the
departments Fauna Sensitive Road Design
(2000).

Chapter 9
Culvert Design

Habitat: As a general rule, the cells of a


culvert should not be modified to provide or
promote habitat of birds and terrestrial
animals. The reason for this is that these
habitats are not natural and are likely to be
destroyed during flood events.
Lizard Run: If for any reason a dry ledge
or dry cell cannot be constructed, then a
lizard run may be considered.
Lizard runs are approximately 100 mm
wide strips of timber bolted to the waterway
embankment side of a culvert cell. They
are introduced to a culvert to enable smaller
reptiles to move through the culvert at an
elevated height.
Lizard
runs
should
be
located
approximately 300 mm below the obvert of
the cell and must extend from ground level
at the upstream wing wall, through the
culvert to ground level at the downstream
wing wall.
Side Wall Roughness: As an alternative to
constructing a lizard run, the culvert cell
wall adjacent to the watercourse bank could
be roughened with texture paint, grout or
other suitable material to allow for the
movement of fauna such as lizards.
Any increase in culvert wall roughness
must be taken into account in the hydraulic
analysis when designing the culvert.
However, it is noted that in a typical road
culvert, sidewall friction only represents
around 12% of the total head loss.
Street Lighting: Many animals move only
at night. To assist in the passage of such
animals, street lighting adjacent to culverts
should be fitted with metal shields to
prevent the lighting of the culvert entry and
exit.

Figure 9.7.4 - Koala Fencing

In some circumstances it may be desirable


to paint the concrete wing walls and apron

January 2013

9-25

Department of Transport and Main Roads


Road Drainage Manual

in a dark colour (dark green) to minimise


the reflection of light.
Lighting: Many terrestrial animals will not
enter a dark culvert, so some means of
lighting inside the culvert is important. For
example, this lighting could be in the form
of a break in the median for a major road.
Vegetation: The provision of vegetation at
the entrance and exit of a culvert is a key
determining factor as to whether native
fauna will use the culvert.
Bank vegetation should be extended up to
the edge of the culvert. This is especially
important if a lizard run or fauna path
has been installed.
In critical flood control regions, this bank
vegetation may need to consist entirely of
flexible (non-woody) species that provide
minimal hydraulic resistance (i.e. no
shrubs).

9.7.5

Stock Underpass

Stock underpasses (also known as cattle


creeps) are primarily designed for the
purpose of allowing cattle to be driven
under road formations and therefore should
remain dry most of the time.

If the culvert is to also be used to convey


water under the road during storm events,
channels leading to and away from the
culvert must have sufficiently flattened side
slopes to allow easy passage for cattle. The
outlet channel must allow the culvert to
completely drain after a storm event.
The culvert must be at least 2.4 m in height
to allow a person on horseback to be able to
ride through the culvert (they can bend low
in the saddle).
For cattle to enter a stock underpass,
sufficient daylight from the other side needs
to be seen by the cattle, otherwise they will

January 2013

9-26

Chapter 9
Culvert Design

baulk and not enter. Suggested number and


width of cells is 3 x 2.4 m minimum for
culvert up to 15 m in length.

9.8

Selection of Culvert
Type

Further to Section 2.3.5, the selection of the


most appropriate type of culvert is
dependent on a range of factors including
economics,
site
conditions,
and
environmental considerations.
Box culverts are generally used where:
Insufficient embankment depth or
cover for pipes exists;
Where channel is narrow and it
would be difficult to fit a pipe
culvert; or
Fauna passage is required.
In multi-cell construction, slab linked box
culverts (SLBC) are often an economically
choice.
Historically, whilst the majority of culverts
installed consist of concrete pipes or box
culverts, the installation of corrugated metal
pipes, pipe-arch or arches, may be
appropriate and economic in some
situations.
Metal culverts have some advantages in
lower cost and ease of transport and
installation. However, disadvantages such
as corrosion due to construction damage,
high compaction standards and higher cover
requirements mean that unless there are
large
financial
savings,
or
other
construction restraints, other more robust
and more durable materials should be used.
Table 9.8 provides guidance in the selection
of the most appropriate culvert type for
different exposure conditions.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Table 9.8 - Culvert Types for Different Conditions


Exposure
condition

Concrete
Box
Culverts
(Normal
cover)

Concrete
Box
Culverts
(Saltwater
Cover)

Concrete
Pipes

Steel
Corrugated
Arch1&2

Steel
Helical
Pipe1&2

Aluminium
Helical
Pipe2

Saltwater

Aggressive soil
(e.g. Low pH, high
chloride high
sulphate)

Invert in fresh
water for
prolonged periods

N/A

Not
economic

Typical condition
(i.e. none of
above)

N/A

Not
economic

1 Refer Appendix C, AS/NZS 2041 1998 Buried Corrugated Steel Pipes


2
Refer Section 9.2.6.1 for structural requirements

Reference should also be made to Standard


Drawing No 1359 (QDMR 2009b) for
installation requirements, cover, spacing
and details for H2 and HS3 support
conditions
(refer
Section
9.2.6.4).
Additional details are discussed below.

9.9

Typical Culvert
Operating Conditions

Many parameters influence the flow pattern


of culverts. Laboratory tests and field
observations show culvert performance is

governed by whether
operating under:

the

culvert

is

Inlet control; or
Outlet control.
The hydraulic capacity of the culvert may
vary considerably depending on the type of
control.
Figure 9.9 shows eight typical conditions
under which culverts may operate.

January 2013

9-27

Department of Transport and Main Roads


Road Drainage Manual

9
Source: Bureau of Public Roads (1965)

Figure 9.9 - Typical Conditions under which Standard Culverts Operate

January 2013

9-28

Chapter 9
Culvert Design

Department of Transport and Main Roads


Road Drainage Manual

However, a study of this figure will reveal


that the prediction of inlet or outlet control
is not precise and the standard procedure
for calculations is to assume each in turn
with the most conservative answer adopted
for design.
Typically, the more conservative answer is
that which causes the highest headwater
level for a given flow and this in turn
implies a submerged inlet type (HW >
1.2D). Submerged inlets usually have more
driving head which helps improve culvert
capacity / performance, which is preferable.

9.9.1

Inlet Control Conditions

For inlet control, the capacity of the culvert


barrel is greater than that of the inlet.
Hence, culvert capacity is dictated by
conditions at the inlet.
The parameters which determine the
discharge under inlet control are the cross
sectional area of the culvert barrel, the inlet
geometry and the depth of headwater or
ponding at the culvert entrance.
In inlet control the losses from the
roughness and length of the culvert barrel
and outlet conditions (including depth of
tailwater) are not factors in determining
culvert capacity.
In Figure 9.9, Types 3, 4, 5 and 8 are
operating under inlet control.

9.9.2

Outlet Control Conditions

With outlet control the culvert flow is


restricted to the discharge which can pass
through the conduit for a given level of
water in the outlet channel (tailwater level).
The slope, cross-sectional area, roughness
and length of the culvert barrel have to be
considered as these losses exceed the inlet

Chapter 9
Culvert Design

losses. However, inlet edge geometry can


still affect the capacity.
In Figure 9.9, Types 1, 2, 6 and 7 have
outlet control. A tailwater depth equal to
80% or more of the height of the culvert
barrel / cell will usually indicate outlet
control, except in rolling or mountainous
country with the culvert on natural surface
slopes.

9.10

Hydraulic Calculations

The hydraulic design of culverts must be


undertaken using the CULVERT program
as discussed in Section 1.2.2.5.
The following design methodology (this
section) and design procedure (Section
9.11) are based on the use of nomographs
and form the basis of the CULVERT
program.
The methodology and procedure not only
provides an understanding of the hydraulic
design of culverts but also an alternate,
manual method for the design or checking
of culverts.
Better methods of culvert design, including
the use of formula, are available however
further investigation by the department is
required before adoption of these methods.
The content of this section and Section 9.11
largely aligns with Austroads Guide to
Road Design; Part 5: Drainage Design
(2008).

9.10.1 Control at Inlet


When the capacity of the culvert barrel is
greater than that of the inlet, the culvert is
subject to inlet control. Then the important
factors are the cross-sectional area of the
culvert barrel, the depth of headwater or
ponding at the entrance and the entrance
conditions, including the entrance type,

January 2013

9-29

Department of Transport and Main Roads


Road Drainage Manual

presence and angle of headwalls and


wingwalls, and the projection of the culvert
into the headwater.
For one dimensional flow, the relationship
between the discharge and the upstream
energy can be computed by an iterative
process or by the use of nomographs.
Inlet control can occur with the inlet
submerged and the outlet not submerged.
Under these conditions, the flow contracts
to a supercritical jet immediately
downstream from the inlet. When the
tailwater depth exceeds critical depth, dc
and the culvert is laid on a steep grade, flow
remains supercritical in the barrel and a
hydraulic jump will form near the outlet. If
the culvert is laid on a slope less than
critical, then a hydraulic jump will form in
the barrel.
When the culvert flows under inlet control,
the roughness and length of the culvert
barrel and the outlet conditions (including
the depth of tailwater) are not factors in
determining culvert capacity. An increase
in the slope of the culvert reduces
headwater only to a small degree, and can
normally be neglected for conventional
culverts flowing under inlet control.

9.10.2 Control at Outlet


With outlet control the culvert flow is
restricted to the discharge which can pass
through the conduit for a given level of
water in the outlet channel (tailwater level).
The slope, cross-sectional area, roughness
and length of the culvert barrel have to be
considered as these losses exceed the inlet
losses. However, inlet edge geometry can
still affect the capacity.
In general the control will be at the outlet if
the culvert slope is less than critical. A
tailwater depth equal to 80% or more of the

January 2013

9-30

Chapter 9
Culvert Design

height of the culvert barrel / cell will


usually indicate outlet control, except in
rolling or mountainous country with the
culvert on natural surface slopes. However,
a check of the design assuming inlet control
is such an easy process that it forms part of
standard design procedure.
Culverts flowing with outlet control can
flow with the culvert barrel full or with the
barrel part-full for all of the culvert length.
With outlet control, and both the inlet and
the outlet submerged the culvert flows full
and under pressure. The culvert, also, can
flow full over part of its length with partfull flow at the outlet. The point at which
the water surface breaks away from the
barrel obvert depends on the tailwater depth
and culvert grade, and can be determined by
using flow profile calculations.
If the culvert is laid at a flat grade, outlet
control can occur with both inlet and outlet
not submerged, and part-full flow
throughout the culvert length will be
flowing under subcritical conditions.
Variations of these main types can occur,
depending on the relative value of critical
slope, normal depth, culvert height and
tailwater depth. While the potential flow
conditions shown in Figure 9.10.2(a) are
the most common for simple culverts,
different flow conditions are possible where
complex culvert structures are required and
advice should be obtained from Director
(Hydraulics),
Hydraulics
Section,
Engineering & Technology Division for
further guidance.
Flow under outlet control can be calculated
from the formulae below, the parameters
for which are illustrated in Figure 9.10.2(a).
The total head (H) required to convey water
through a culvert flowing under outlet
control is determined by:

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

H Hv He Hf

and
V = mean velocity of flow in the
culvert barrel (m/s);

Where:
Hv = velocity head (m)

g = acceleration due to gravity


(assume 9.81 m/s);

V2

2g

ke = entrance loss coefficient (refer


Table 9.10.2;

He = entrance loss (m)

ke

V2
2g

n = Mannings roughness coefficient;


L = length of culvert barrel (m); and

Hf = friction loss (m)

R = hydraulic radius (m).

19.6n 2 L V 2

2g
R 1.33

V1
2g

V
2g

WS = water surface
SSo = slope of culvert

W.S
.
V1

He
Energy Lin

e
Hydraulic G
rade Line

HW
h1

Hf
Hv

LS
LSo

dhcc

W.S
.

ho2
DATUM

Source : Austroads (1994)

Figure 9.10.2(a) Hydraulics of Culvert Flowing Full under Outlet Control

9
H

HW
S

dhcc

LS
LSo

TW

+D
ho = Greater of hdc c+
D and TW
2
Source : Austroads (1994)

Figure 9.10.2(b) - Determination of Adopted Tailwater (ho)

January 2013

9-31

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Table 9.10.2 - Culvert Entry Loss Coefficient


Type of Structure and Design of Entrance

Coefficient ke

Concrete / Fibre reinforced / Poly Pipe


Projecting from fill, socket end
0.2
Projecting from fill, square cut end
0.5
Headwall, with or without wingwalls:
Socket end
0.2
Rounded edge (radius = D/12)
0.2
Square edge
0.5
End section conforming to fill slope (precast end unit)
0.5
Mitred / cut to conform to fill slope (field cut)
0.7
Reinforced Concrete Box
Headwall parallel to embankment (no wingwalls):
Rounded on 3 edges (radius of 1/12 cell dimension)
0.2
Square on 3 edges
0.5
Wingwalls at 30 to 75 to cell:
Crown edge rounded (radius of 1/12 cell dimension)
0.2
Crown edge square
0.4
Wingwalls at 10 to 30 to cell: square edged at crown
0.5
Wingwalls parallel (extension of sides) square edged at crown
0.7
Corrugated Metal Pipe
Headwall rounded edge
0.2
Headwall or headwall and wingwalls: square edged
0.5
End section conforming to fill slope (manufacturer end unit)
0.5
Mitred / cut to conform to fill slope
0.7
Projecting from fill (no headwall)
0.9
Notes:
1. The effect of wing walls reduces with multi-cell culverts.
2. For 3 - 6 cell culverts, assume entrance loss for wing walls 10 to 25 to barrel.
3. For culverts with more than 6 cells, assume wing walls parallel (extension of sides),
regardless of actual wing walls.
Sources: Bureau of Public Roads (1965); Hydraulic Design Manual. CPAA (1991); Handbook of Steel Drainage
& Highway Construction Products. American Iron and Steel Institute (1994).

Substituting in the first equation above and


simplifying:

(1 k e

19.6n 2 L V 2
)
2g
R 1.33

This equation can be solved for H by the


use of the full flow nomographs in
Appendix 9A which covers the common
culvert types.
From the development of this energy
equation and Figure 9.10.2(b), H is the
difference between the elevation of the

January 2013

9-32

hydraulic grade line at the outlet and the


energy line at the inlet. Since the velocity
head in the entrance pool usually is small
when ponded conditions occur (V2/2g 0),
the water surface of headwater pool
elevation can be assumed to equal the
elevation of the energy line.
Headwater depth under outlet control is
calculated according to the equation:

HW H h0 LS 0

Department of Transport and Main Roads


Road Drainage Manual

Where:
ho = adopted tailwater depth (m);
L = length of culvert (m); and
So = slope of culvert barrel (m/m).
The various components of this equation
are illustrated in Figure 9.10.2(b).
The tailwater level (ho) to be adopted is TW
if TW > D, otherwise the greater of:
TW, or

dc D
2

Where:
dc = critical depth (m); and
D = diameter or height of culvert
(m).

9.10.3 Determination of Tailwater

Chapter 9
Culvert Design

Special culverts are those that do not fit


the normal design procedure and these must
be referred to Director (Hydraulics),
Hydraulics Section, Engineering &
Technology Division for further guidance.
The designer should be familiar with all the
equations in the previous sections before
using these procedures. Following the
design method without an understanding of
culvert hydraulics can result in an
inadequate, unsafe or costly structure.
Because of the difficulty in determining
whether a culvert will operate under inlet or
outlet control, the standard procedure for
calculating the headwater level is to check
both inlet and outlet control cases and adopt
the scenario that results in the higher
headwater level.

9.11.1 Hydraulic Design


The tailwater level is an important input to
the culvert design process. Tailwater level
(TW) can be calculated using the method
described in Section 8.7.2.

9.11

Design Procedure

The following procedure for standard


culverts will be acceptable for all but a very
small percentage of culverts designed for
departmental projects.

The standard culvert design procedure is


illustrated in the flow chart in Figure 9.11
and the following sub-sections align with
this figure.
For this manual design method, calculations
should be recorded on a form similar to the
one shown in Appendix 9B. The following
procedure provides direction in completing
this form.

The procedure does not address the


unsteady flow hydraulic effect of changing
storage / ponding upstream caused by the
culvert / road embankment, which can
modify the discharge through the culvert
when flows are changing significantly.

January 2013

9-33

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

HYDRAULIC DESIGN PROCESS FOR CULVERTS

Collect Design Data


- Required flood immunity (ARI)
- Design discharge (Q) (also consider
extreme events).
- Tailwater height (TW).
- Shoulder height and freeboard
requirements.
- Maximum / allowable headwater (AHW).
- Proposed culvert slope (So).
- Proposed culvert length (L).
- Inlet / outlet invert heights.
- Maximum allowable velocity (Vmax).

Determine Controlling Headwater

Select a Trial Culvert


Determine an initial trial culvert waterway
area using A=Q/Vmax
- If Vmax is unknown, use Vmax = 2.0 - 2.5
m/s.

Is HWi < HWo?


If yes - HWo controls - go to step 7.
If no - HWi controls - go to step 8.

Outlet Velocity Outlet Control

Use area (A) to determine initial trial


culvert(s).
- Height of culvert (D) should be approx.
TW.
- Size and fit culvert to channel / road.

Calculate outlet velocity using Vo=Q/A.


Where A is depth of flow based on D, dc or
TW as appropriate.

Outlet Velocity Inlet Control

Design Discharge for Trials


For culverts, divide design discharge Q by
number of cells.

For box culverts, a ratio (Q/B) is also


required. Divide Q per cell by box width
(B).

Calculate outlet velocity using Vo=Q/A.


Use appropriate nomograph to determine
part full area for pipes and for boxes, use
Mannings to determine normal depth.

Outlet Flow Energy (Froudes Number)

YES

Determine Inlet Control Headwater Depth


Use appropriate nomograph to determine
HW/D.
Now determine HWi.

Determine Fr for flow at outlet.


Compare to Fr for channel is there a
hydraulic jump?

Check if HWi, > AHW?


- If yes, return to step 2 and select a larger
culvert.
- If no go to step 5.
Design Check

10

Determine Outlet Control Headwater


Depth
Select entrance loss coefficient ke.

If Vo < Vmax then OK, if not, check that


outlet protection can be provided.
If an alternative culvert configuration is to
be elevated return to step 2.

Use appropriate nomograph to determine


H.
Use appropriate nomograph for pipes /
formula for boxes to determine dc.
Determine (dc + D).
YES

Is TW > D at the outlet?


- If yes, use h0 = TW.
- If no, h0 is the larger of TW and (dc +
D).
Determine length by slope, LSo.
Determine HW0 = H + h0 LS0
Check if HW0 > AHW?
- If yes, return to step 2 and select a larger
culvert.
- If no go to step 6.

A
Figure 9.11 - Culvert Design Flow Chart

January 2013

9-34

A
AHW
B
D
dc
Fr
HWi
HWo
h0
ke
L
Q
S0
TW
Vmax
V0

waterway area of culvert


allowable headwater level at inlet (m)
width of culvert (m)
diameter (pipe) / height (box) of culvert (m)
critical depth of flow in culvert (m)
Froudes Number
generated headwater depth at inlet by Inlet control (m)
generated headwater depth at inlet by Outlet control (m)
calculated head at the outlet of the culvert over the
invert (m)
entrance loss coefficient
length of culvert (m)
design discharge through culvert (m3/s)
slope of culvert
tailwater depth at outlet (m)
maximum average velocity in natural channel (m/s)
calculated outlet velocity (m/s)

No

Department of Transport and Main Roads


Road Drainage Manual

9.11.1.1

Chapter 9
Culvert Design

Collect Design Data

Further to Section 9.3, the following data is


to be collected / determined, and then
recorded in the top sections of the design
form:
Required flood immunity (ARI);
Design discharge (Q) and extreme
event discharge (if reqd) (refer
Chapter 5);
Tailwater height (TW) for each
discharge flow (refer Chapter 8);
Road shoulder height
freeboard requirements;

and

any

Maximum / allowable headwater


height (AHW) (refer Section 9.5);
Proposed culvert slope (So);
Proposed culvert length (L);
Inlet / outlet invert heights; and
Maximum allowable stream velocity
(Vmax) for outlet channel (refer
Section 9.2.5).

9.11.1.2

Select a Trial Culvert

To select an initial trial culvert or culverts,


first determine an initial trial culvert
waterway area (A) using:

AQ

Vmax

Where:
Q = design discharge (m3/s); and

allowable afflux to limit flooding level


impacts on upstream development), then the
initial trial culvert area may be estimated
using the following formula:

Q
3.6 H 0.5

Now, choose culvert material, shape (pipe


or box) and entrance type, allowing for
minimum and maximum allowable cover
heights over the culvert.
Select a culvert trial size / configuration
using the waterway area determined above.
For example, the initial trial culvert
waterway area may have been determined
as 1.25 m2. Review of product guides show
that a single barrel 900RCP has a nominal
diameter of 0.9 m and a waterway area of
0.64 m2. Two barrels of 900RCPs would
have a waterway area of 1.27 m2, which is
slightly larger than required waterway area
and therefore suitable as an initial trial.
If possible, also select culvert size such that
the obvert of the outlet is at or just below
the tailwater level.
This ensures the
probability of the culvert running full,
which is usually desirable. This is not
always practicable in wide shallow flood
plains, or in steep country. If the trial size
is too large because of limited embankment
height or availability of size, multiple
culverts may be used by dividing the
discharge equally between the number of
cells used.

Vmax = maximum allowable outlet


velocity (m/s).

In locations where fauna passage is an


issue, compare trial culvert size with the
minimum fauna requirements.

Maximum allowable outlet velocity should


be based on Section 9.2.5 or if no data is
available, assume an outlet velocity of 2.0
to 2.5 m/s depending on channel conditions.

Several initial trial culverts of different size


/ configuration may be selected to start
design process.

Alternatively, if the allowable head loss (H)


through the culvert is known (from the

January 2013

9-35

Department of Transport and Main Roads


Road Drainage Manual

9.11.1.3 Design Discharge for


Trials
Where a single barrel culvert is not
sufficient, multiple barrels or cells will be
required and these configurations constitute
a parallel system.
Provided that all barrels of a multi-cell
culvert are the same type, size and
roughness (equal conveyance), and also
have the same invert levels / bed slope,
flow will distribute evenly.
Design nomographs as shown in Appendix
9A, are based on a single barrel installation,
therefore divide design discharge (Q) by
number of barrels for trial culvert.
For boxes, a ratio (Q/B) is also required.
This ratio is determined by dividing Q per
cell by the nominal box width (B).
Any proposed, multi-cell culvert that does
not have all cells the same will require
specialist advice / assistance as the
methodology presented in this manual does
not allow for these type of configurations.
These culverts must be referred to Director
(Hydraulics),
Hydraulics
Section,
Engineering & Technology Division or
suitably pre-qualified design consultant.

9.11.1.4 Determine Inlet Control


Headwater Depth
Using the trial culvert(s) from previous
section, find the HW/D value by use of the
appropriate inlet control nomograph (Figure
9A.1 to Figure 9A.4). In this case, tailwater
conditions are to be neglected. Three lines
are presented in the nomographs for HW/D
and the designer needs to select the
appropriate line based on:
entrance type for pipe culverts; or
wingwall flare angle for box culverts.

January 2013

9-36

Chapter 9
Culvert Design

Headwater under inlet control conditions


(HWi) is found by multiplying HW/D by the
height of culvert, D.
Now check HWi against AHW and if HWi is
greater (or much less) than AHW, try
another size until HWi is acceptable for inlet
control before computing headwater for
outlet control.

9.11.1.5 Determine Outlet


Control Headwater Depth
Several steps are required to determine the
headwater under outlet control conditions
(HWo).
Firstly, determine the entrance
coefficient, ke from Table 9.10.2.

loss

Now calculate the losses through the


culvert, H using the outlet control
nomographs, Figure 9A.5 to Figure 9A.10.
In using these nomographs, some
interpolation can be used for ke and L.
The next step is to determine the critical
depth (dc) for the culvert. For pipe culverts,
use nomographs as shown in Figures 9A-13
to 9A-15 and for box culverts, use the
formula:

d c 0.467 Q
B

Q in this formula is Q per cell.


exceeds D then take dc as D.

If dc

Now calculate (dc + D)/2.


Tailwater (TW), as gathered in Section
9.11.1.1, is required for determination of
the next variable.
The next step is to establish the adopted
tailwater (ho) for design. Determine if TW
> D:
If yes, ho = TW;
If no, ho = the larger of TW and (dc +
D)/2.

Department of Transport and Main Roads


Road Drainage Manual

Now, multiply the proposed culvert length


(L) and slope (So), as gathered in Section
9.11.1.1, to calculate LSo.
Headwater under outlet control conditions
(HWo) is calculate using:

HW0 H h0 LS 0
Now check HWo against AHW and if HWo
is greater than AHW, try another size until
both HWi and HWo are acceptable (< AHW).

9.11.1.6 Determine the


Controlling Headwater
Compare the values of HWi and HWo. The
higher headwater governs and indicates the
flow control existing under the given
conditions for the trial size selected.
If HWi > HWo the culvert is under
inlet control and Controlling HW =
HWi.
If HWi < HWo the culvert is under
outlet control and Controlling HW =
HWo.

Chapter 9
Culvert Design

a different culvert size / configuration,


where D TW.
Sometimes, calculations show HWi and
HWo to be equal or nearly equal. In this
instance, it is not clear whether or not the
culvert will perform under inlet control or
outlet condition. In reality, the culvert
could also operate under both conditions
(swap from one to the other) during the
same rainfall event.
Therefore, outlet
velocity calculations should be done for
both Inlet Control and Outlet Control
conditions with the higher velocity (and
associated control condition) being adopted.

9.11.1.7 Outlet Velocity Outlet


Control
The average outlet velocity for all culvert
types can be calculated using:

Vo Q

Where:
Q = design discharge per culvert
barrel / cell (m3/s); and

The controlling headway for trial should be


clearly shown in design workings (design
form).

A = cross sectional area of flow from


culvert barrel / cell (m2).

There are now two checks that should be


undertaken before completing trial design:

The cross-sectional area of flow (A)


depends on the flow depth at the outlet.

(a) If controlling HW is less than 1.2D,


then the culvert is most likely not
operating with a submerged inlet and
therefore may not be operating
efficiently. Design should be revised
using a different (slightly smaller)
culvert size / configuration.

Flow depth will be one of the following:

(b) If controlling HW is less than 0.75D


and the culvert is under outlet control,
then the culvert may be flowing only
part-full and using (dc+D)/2 to
calculate ho may not be giving accurate
results. Design should be revised using

critical depth (dc) if the tailwater is


below critical depth;
tailwater depth (TW) if the tailwater
is between critical depth and the top
of the barrel; or
the height of the barrel (D) if the
tailwater is above the top of the
barrel.
Refer Figure 9.11.1.7(a) for guidance in
determining flow area for box culverts.

January 2013

9-37

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

9.11.1.8 Outlet Velocity Inlet


Control
The average outlet velocity for all culvert
types can be calculated using:

Vo Q

Where:
Figure 9.11.1.7(a) Flow Area for Box
Culverts

Determination of flow area for pipes is a


little more difficult. The area can be
determined using a CADD package such as
AutoCAD or calculated using:

Area R 2 when y = D
or

Area R 2 1 R 2 sin
2
Where:
R = internal radius of pipe (m); and
= angle in radians
Figure 9.11.1.7(b) provides guidance in
determining flow area for pipes.

Q = design discharge per culvert


barrel / cell (m3/s); and
A = cross sectional area of flow from
culvert barrel / cell (m2).
The cross-sectional area of flow (A)
depends on the flow depth at the outlet
which can be approximated by the normal
depth of open-channel flow in the barrel,
computed by Mannings equation for the
discharge flow, barrel size, roughness and
slope of culvert selected.
For Pipe Culverts
For a pipe culvert, the culvert will not be
flowing full at the outlet when under inlet
control, meaning only a part of the full
waterway area will be used. To determine
this Part Area, the designer needs to firstly
establish the relationship between Full
Flow (Qf) and Part Flow (Qp) for the
culvert trial where:
Part Flow (Qp) is the design
discharge per cell; and
Full Flow (Qf) capacity of the trial
culvert, and its velocity, can be
determined using the Figure 9A-11.
Now determine the ratio Qp / Qf.

Figure 9.11.1.7(b) Flow Area for Pipe


Culverts

January 2013

9-38

Using this ratio and Figure 9A-12,


determine the percentage factors for
velocity (Vo), depth of flow (y), area of flow
(A) and hydraulic radius (R).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Figure 9A-12 is used as follows:


Step 1 - Plot Qp / Qf. ratio on x-axis.
Step 2 - Draw a line up to Discharge
line.
Step 3 - Draw a line left and right to
both edges (the left edge being the yaxis).
Step 4 Read off % Depth of Flow
from y-axis.
Step 5 Drop lines to x-axis from
each intersect between horizontal line
drawn in Step 3 and the Velocity,
Area and Hydraulic Radius curves.
Step 6 Read of % Values from xaxis for each of these hydraulic
elements.
Now draw a table as shown in Figure
9.11.1.7(c) and enter values as follows:
A is Qf
B is Qp
C is outlet velocity determined using
Figure 9A-11
D is nominal diameter of pipe
E is waterway area of pipe
F is hydraulic radius of pipe when
flowing full
G are the hydraulic element values
determined using Figure 9A-12.
The remaining spaces of table, including
the determination of Vo, are calculated by
multiplying the Full values by relevant
Factor.

Figure 9.11.1.7(c) Full Flow / Part Flow


Table

For Box Culverts


As for a pipe culvert, a box culvert will not
be flowing full at the outlet when under
inlet control. To determine the normal flow
conditions, depth (y) and velocity (Vo), use
Mannings Equation to develop a Modified
Stage-Discharge Curve for the culvert cell
(Refer Chapter 8).
Using the Modified Stage-Discharge Curve,
the flow depth (y) and outlet velocity (Vo)
can be read directly.

9.11.1.9

Outlet Flow Energy

To complete the hydraulic calculations for a


trial culvert, the designer must determine
Froudes Number (Fr) for the flow at the
outlet (refer Chapter 8). This is important
as the designer can check this against
Froudes Number for the channel flow and
determine if a hydraulic jump will occur.
Froudes Number can be determined using:

Fr Q B

gA
3

Where:
B = 2 y D y
refer Figure 9.11.1.9

January 2013

9-39

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Add another barrel or cell if channel


width permits;
Increase barrel or cell height if
vertical clearance permits;
Alter culvert slope (Note: desirable
minimum is 0.25%).
Figure 9.11.1.9 Flow Width in Pipes

When a pipe or box culvert is flowing full,


technically there is no water surface and
therefore no B. In this situation, it is
considered suitable to calculate Fr at the
moment just before the water surface
touches the obvert of the culvert and adopt
this Fr value for the trial culvert. Therefore:
For box culverts, use full waterway
area and B = nominal width of box
section; and
For pipe culverts, use full waterway
area, but assume B as 10% of D (
99.7% of A).

9.11.1.10 Design Check

At this stage, the outlet velocity (Vo) should


be checked against the maximum allowable
stream velocity (Vmax) for outlet channel.
Preferably, Vo should be less than Vmax. If
Vo exceeds Vmax then the designer must
include suitable outlet protection for the
culvert into the design (refer Sections 9.15
and 9.16). Where outlet protection is not
suitable (size / cost) then a larger / wider
culvert trial is required.

9.11.2 Practical Design


If a culvert trial is considered unacceptable,
the designer is required to redesign the
culvert by trialling another culvert size /
configuration. The choices the designer has
in determining a new trial culvert are:

January 2013

9-40

A combination of the above.


Where a designer decides to alter invert
levels, it is preferable to lower inlets and
leave outlets as close to natural surface as
possible. Drop inlets (and structures) are
better than hanging or buried outlets.

9.12

Minimum Energy Loss


Culverts

The early designs based on this principle


were known as minimum energy / constant
energy and no afflux culverts, although the
last title is no longer used.
Bridge
waterways have also been designed as
minimum energy loss structures.
To minimise energy loss through a culvert
and the resulting afflux, the design requires
carefully shaped inlets and outlets (refer
Figure 9.12) and usually a dropped culvert
barrel such that critical flow passes through
the system in the design flood. In a
minority of these structures flow is not at
critical depth in the barrel in the design
flood.
Furthermore, due to the size, material
requirements and increased difficultly in
construction, minimum energy culverts are
generally very expensive options and the
benefits of these types of culverts is
arguable.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

The cost of shaped inlets and outlets


will be more costly than those for
conventional culverts;
An anti-ponding pipe may be
necessary to prevent ponding of
water after a flood;

Figure 9.12 Minimum Energy Loss


Culvert on Gateway Motorway

Advantages promoted by others over


conventional structures are that:
by constricting the natural flow to a
greater extent, the body or cell has a
minimum
width,
reducing
construction
costs
for
this
component;
the flow through the culvert is
streamlined and therefore, has
reduced turbulence which, in turn,
reduces the erosion potential of the
flow and minimises the need for
protection; and
the minimisation of energy losses
results in little or no adverse effect on
upstream flood levels.
Problems have been observed in some
existing structures and some designers see
them as indications for caution in the future
design of minimum energy loss structures.
Stated reservations include:
Critical flow is inherently unstable
and therefore, sensitive to small
changes in energy and depth of flow.
Discharges both higher and lower
than the design discharge have the
potential to give higher affluxes than
the design flood. The range of flows
for which streamlined flow may
occur in any structure is questioned;

Sediment may be deposited after a


flood and may not be removed in
future flows with the possibility of
growth of vegetation requiring
removal
i.e.
Provision
for
maintenance required; and
The possibility of debris blocking the
culvert barrel is increased because of
the smaller cell.
The significance of some key points above
is that too much debris or siltation will
change the geometry such that streamlined
flow would not occur.
Minimum Energy Loss Culverts are no
longer used or recommended by the
department. If an existing minimum energy
loss structure requires extension or
alteration, then the design must be referred
to Director (Hydraulics), Hydraulics
Section, Engineering & Technology
Division.

9.13

Blockage of Culverts

The likelihood of blockage should be


considered for all culverts. Blockage can
occur through siltation or by debris (such as
vegetation).
Blockage reduces the waterway area of the
culvert and therefore adversely affects the
capacity / performance of the culvert. The
result of blockage is typically:
an increase in upstream peak water
levels / flooding;

January 2013

9-41

Department of Transport and Main Roads


Road Drainage Manual

an increased potential for water to


overtop the road; and
an increased risk of failure to road
embankment / culvert.
Silt deposits and some debris can be
detected and removed via maintenance
processes outside of a rainfall event
therefore would not impede the any flows,
however this is highly dependant on the
efficiency of maintenance in the area.
Where debris blockage during an event is
considered likely (typical in catchments that
contain
significant
woody
riparian
vegetation), larger culvert sizes may be
required, in accordance with the extent of
adverse impacts that could occur to the
roadway or to surrounding properties.
Where large or long branches and/or tree
trunks are a possibility, sloped extensions to
piers, as shown in Figure 9.13, can be used
to turn long objects into the culvert barrel.

Chapter 9
Culvert Design

9.14

Consideration of
Large or Extreme
Events

Irrespective of the design ARI of a culvert,


consideration should be given to the
impacts of flood events which are larger
than the design event (refer Section 2.7). In
determining the appropriate degree of
consideration, issues such as the frequency
and consequence of overtopping should be
addressed.

9.15

Culvert Outlet
Protection

Outlet protection is required in situations


where:
outlet velocity exceeds the scour
velocity of the bed or bank material;
an unprotected channel bend exists
within a short distance of the culvert
outlet;
the outlet channel and banks are
actively eroding; and

9
Figure 9.13 - Flood Water Flowing into
Box Culverts

Designers must consider the potential for


and impacts of blockage for each catchment
/ culvert installation and where impacts are
considered unacceptable, design mitigating
treatments accordingly.

January 2013

9-42

if an erodible channel bank exists


less than 10 to 13 times the pipe
diameter downstream of the outlet,
and this bank is in-line with the
outlet jet (i.e. likely to be eroded by
the outlet jet) the bank should
adequately protected to control any
undesirable damage as a result of the
outlet jetting.
The most appropriate outlet protection is
determined by considering the hydraulic
performance of the outlet in the prevailing
stream environment.
At outlet structures, the best hydraulic
performance is obtained when the confining
sidewalls are parallel and the distribution of
flow across the channel is uniform.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Culverts, however, are generally narrower


than the natural waterway and a transition
section is required to return the flow to the
natural channel.
When culvert outlet
velocities are high, additional measures at
the outlet may prove to be necessary for
energy dissipation. Section 9.16 below
provides additional information on more
specialised energy dissipators.
In all types of culvert outlets, protection of
the stream bed would normally be provided
by TMRs standard apron treatment, as
shown on Standard Drawings 1131, 1317
and 1318 (QDMR 2009b), for a minimum
distance of 1.5D downstream where D is
the diameter of a pipe or the height of a box
culvert..
To check whether standard inlet and outlet
structures with headwalls, wingwalls,
aprons and cut-off walls are adequate, the
outlet velocity for the culvert requires
examination with respect to:
natural environment
vegetation cover);

(soil

and

size of peak flow; and


duration of large flows.
If outlet velocities exceed the acceptable
limits, it may be necessary to check for
potential bed scour problems.

velocities less than 5.0 m/s, an extended


concrete apron or rock pad (commonly
used) protection is recommended.
The recommended minimum rock size (d50)
and length (L) of rock protection
downstream of culvert outlets may be
determined from Figure 9.15(a).
The minimum recommended width of the
rock pad is defined as:
Immediately downstream of the
outlet: the width of the outlet apron;
At the downstream end of the rock
pad: the above width plus 0.4 times
the length of the rock pad (L) as
shown in Figure 9.15(b).
When the width of the channel is less than
the recommended width of the rock
protection, then the rock protection shall
extend up the banks of the channel to a
height equal to the obvert of the outlet.
The recommended depth or thickness of the
rock pad is 1.5 2.0 times d50.
For Froude Number (Fr) values greater than
1.7 and outlet velocities equal or greater
than 5.0 m/s, an alternative energy
dissipator structure should be considered as
discussed in Section 9.16.

Where the outlet flows have a Froude


Number (Fr) less or equal to 1.7 and outlet

January 2013

9-43

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Culvert Outlet Velocity (m/s)

Seek specialist advice for alternative solution

Culvert Diameter or Cell Height (mm)

Figure 9.15(a) Minimum Rock Size and Length of Apron

Figure 9.15(b) Rock Apron Detail

January 2013

9-44

Department of Transport and Main Roads


Road Drainage Manual

9.16

Special Energy
Dissipation Structures
at Culvert Outlet

When rock pad protection is not appropriate


due to high velocity and/or high energy
flows, non standard and more specialised
energy dissipators are required. These
options can be expensive to design and
construct and generally have a potential for
high maintenance costs (i.e. debris
collection). These devices can also be a
potential public and wildlife safety hazard.
Use of such energy dissipators should be
considered the exception, not the norm, and
should only be used when no other
reasonable options are available.
The
design of these devices / structures should
be referred to Hydraulics Section,
Engineering & Technology Division or a
suitably prequalified consultant.

9.17

Self Cleaning
Culverts

If a self cleaning culvert is required,


designers are referred to Section 2.8 for
requirements / design parameters.

9.18

Inlet Structures

For culvert inlet structures, headwalls and


wingwalls of concrete or rock masonry are
usually needed to provide embankment
stability and protection against erosion.
Culverts with wingwalls should be designed
with an apron extending between the walls.
Aprons may be grouted rock pitching,
reinforced or unreinforced concrete or rock
mattresses. The actual configuration of the
wingwalls will vary according to the
direction of flow and so protection against
scour is maximised through inclusion of the
apron. Refer to Standard Drawings 1131,

Chapter 9
Culvert Design

1148, 1179, 1303, 1304, 1305 and 1306


(QDMR 2009b).
The use of a cutoff wall along and under the
edge of the inlet apron should also be
investigated / checked.
An important aspect that designers must
check for is the occurrence of a hydraulic
jump in the inlet of the culvert.
Supercritical flows in the existing channel,
at the site of a proposed culvert, will be
forced back to subcritical flow by the
presence of the culvert. This will force a
jump at the culvert entrance which can:
cause erosion about the inlet if not
adequately protected; and/or
affect the hydraulic design of the
culvert due to the turbulence.
If this is the case, it is recommended to
force the jump to occur some distance
upstream of the culvert (by placing a hump
across the stream bed) so that the energy
and turbulence can dissipate before arriving
at the culvert.

9.19

Managing Sediment

Sediment
deposits
within
culverts,
especially multi-cell culverts, can cause
significant operational and maintenance
problems.
Occasionally sediment traps (basins) are
constructed upstream of culverts. In these
cases, an access ramp for maintenance must
be provided to allow de-silting of the trap.
In critical areas, or for long culverts where
maintenance is extremely difficult, a small
sediment trap / weir can be constructed at
the inlet to divert low flows to just one or
two culvert cells. This will allow the flow
to enter the remaining cells only during
high flows.

January 2013

9-45

Department of Transport and Main Roads


Road Drainage Manual

These sediment weirs should be designed to


be fully drowned during major flood events
so that no adverse backwater effects occur.

9.20

Safety

Chapter 9
Culvert Design

If these barriers are likely to interfere with


flood waters or terrestrial movement, then
the road layout should be designed to avoid
the need for such barriers, subject to
satisfying safety requirements for road
users.

9.20.1 Culverts Used As


Walkways and Bikeways

9.20.2.3

Provision for pedestrians and cyclists can


be made in sizing a waterway culvert.
However, in order to be effective, the
approach ramps must allow clear vision
through the culvert cell.

Noise control fencing must allow for the


free passage of overtopping flood flows in
most cases. Even where culverts have been
designed with a large capacity (e.g. 50 yr
ARI event) consideration of the impacts of
the fence on overtopping flow is also
required.

It should also be kept in mind that


pedestrians will often prefer to cross over a
road than under it for reasons of security.

Noise Fencing

9.20.3 Additional Safety


Considerations

9.20.2 Barriers to Flow


9.20.2.1 Trafficability
Considerations
Consideration needs to be given to the
desired trafficability of the road during
overtopping flood flows.

Raised median strips can result in traffic


movement only on one side of the road
during overtopping flows. In critical flood
control areas, it may be necessary to use a
painted median.
Raised kerbs or pedestrian pathways on the
downstream side of a road can cause
ponding to occur across the full width of
the road, thus reducing trafficability in both
directions during flood events.

9.20.2.2

Traffic Safety Barriers

Median traffic barriers such as the GM


Barrier can significantly interfere with the
passage of overtopping flood flows and the
migration of terrestrial wildlife across the
road.

In the design and location of pipe


installations, designers must consider the
hazards / risks which may exist at particular
sites.
Workplace Health and Safety
requirements should be considered in
culvert design, installation and subsequence
maintenance. Some of the issues which
should be addressed may include, but are
not restricted to:
Excavation / Trenching Stage
Geotechnical analysis and the need
for shoring;
Placement of excavated material
close to trench walls;
Location of underground services;
Placement of materials;
Proximity
of
excavations;

9-46

to

Crane capacity and reach;


Proximity of machinery to overhead
powerlines and obstructions;
Backfilling;

January 2013

machinery

Department of Transport and Main Roads


Road Drainage Manual

Chapter 9
Culvert Design

Probability of wall collapse arising


from vibration or traffic movement;
Amount of time excavation will be
open;
Working within confined space; and
Anticipated weather conditions.
The cost of control measures to be used to
ensure that the risks associated with the
above issues are minimised should be
included in the installation costs of the
drainage structure.
Also, where culvert inlets are located in
areas likely to be accessed by people
(particularly children), consideration must
be given to appropriately fence the inlet to
ensure someone is not swept into the
culvert during a flood event. An example
of this type of fencing is shown in Figure
9.20.3.

9
Figure 9.20.3 - Fencing Around Pipe Inlet

A serious problem with fencing a culvert


inlet is that the potential for blockage is
greatly increased and designers need to
check for and mitigate any possible adverse
affect due to the blockage.

January 2013

9-47

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

Appendix 9A
Design Nomographs

9A

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

Appendix 9A Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

9A

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Steering
Committee

Date

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: Q is Discharge per culvert cell

9A-1 Inlet Control - Headwater depth for box culverts

March 2010

9A-1

Department of Transport and Main Roads


Road Drainage Manual

9A

Note: Q is Discharge per culvert cell

9A-2 Inlet Control - Headwater depth for concrete pipe culverts

March 2010

9A-2

Appendix 9A
Design Nomographs

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: N = no. of pipes and Q/N = Discharge per pipe

9A-3 Inlet Control Headwater depth for Corrugated Metal Pipe (CMP) and Structural
Plate Corrugated Steel Pipe (SPCSP) culverts

March 2010

9A-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: N = no. of pipes and Q/N = Discharge per pipe


9A-4 Inlet Control - Headwater depth for Corrugated Steel Pipe Arch (CSPA) and
Structural Plate Corrugated Steel Pipe Arch (SPCSPA) culverts

March 2010

9A-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A
Note: Q is Discharge per culvert cell

9A-5 Outlet Control - Head for concrete box culverts flowing full (n=0.012)

March 2010

9A-5

Department of Transport and Main Roads


Road Drainage Manual

9A
Note: Q is Discharge per culvert cell

9A-6 Outlet Control - Head for concrete pipes flowing full (n=0.012)

March 2010

9A-6

Appendix 9A
Design Nomographs

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: N = no. of pipes and Q/N = Discharge per pipe


9A7 Outlet Control Head for Corrugated Metal Pipe (CMP) flowing full (n=0.024)

March 2010

9A-7

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: N = no. of pipes and Q/N = Discharge per pipe

9A8 Outlet Control Head for Structural Plate Corrugated Steel Pipe (SPCSP) flowing
full (n=0.0328 to 0.0302)

March 2010

9A-8

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

Note: N = no. of pipes and Q/N = Discharge per pipe

9A-9 Outlet Control Head for Corrugated Steel Pipe Arch (CSPA) flowing full (n=0.024)

March 2010

9A-9

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A
Note: N = no. of pipes and Q/N = Discharge per pipe

9A-10 Outlet Control Head for Structural Plate Corrugated Steel Pipe Arch (SPCSPA)
flowing full (n=0.0327 to 0.0306)

March 2010

9A-10

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

9A-11 Discharge and Velocity in Round Pipes Flowing Full

March 2010

9A-11

Department of Transport and Main Roads


Road Drainage Manual

9A

9A-12 Velocity and Discharge for Part Full Pipes

March 2010

9A-12

Appendix 9A
Design Nomographs

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

9A-13 Critical Depth in a Circular Pipe

March 2010

9A-13

Department of Transport and Main Roads


Road Drainage Manual

9A

9A-14 Critical Depth in a Corrugated Steel Pipe Arch (CSPA)

March 2010

9A-14

Appendix 9A
Design Nomographs

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9A
Design Nomographs

9A

9A-15 Critical Depth in a Structural Plate Corrugated Steel Pipe Arch (SPCSPA)

March 2010

9A-15

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9B
Design Form

Appendix 9B
Design Form

9B

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9B
Design Form

Appendix 9B Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

9B

March 2010

ii

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9B
Design Form

9B

March 2010

9B-1

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Appendix 9C
Worked Examples

9C

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Appendix 9C Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

9C

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Culvert Design Example


Refer to Chapter 9 - Culvert Design
This example describes a culvert design procedure for a road crossing of a small channel based
on the following requirements and data:
Design Discharge (Q) is 6.25m3/s for a 50 year ARI event.
Road is 10m wide with two traffic lanes and a road crossfall of 3% towards shoulders from
centreline. The centreline in the vicinity of the proposed culvert is located at 125.65m height
(Ht).
The road pavement thickness is 460mm and it is to be the minimum cover over the pipe.
The Allowable Headwater (AHW) is to be 100mm below shoulder height to avoid any
potential damage to adjacent properties.
Invert Inlet and Invert outlet Heights (Hts) are 123.85m and 123.75m respectively.
The channel is of a trapezoidal section with bed width 5.0m, bed slope 0.65%, side slopes
1:1 and Mannings value n = 0.035. (Stage-Discharge curve has been provided). Maximum
stream velocity is 2.5m/s.
The culvert is 15.6m long and must be Reinforce Concrete Pipe (RCP).
Culvert Entrance type to be square edge with headwall.
This example can be solved following design procedure as per Section 9.11. (Refer also to
Figure 9.11 Culvert Design Flow Chart)

Step 1.
Collect Design Data. Use the hydraulic calculation form given (Appendix 9B Design Form)
and fill out the known information. From the above information, this include: design flow, road
and culvert heights, length, slope and maximum allowable velocity.
The shoulder height can be calculated based on the Centreline height, road crossfall and traffic
lane width.
10.0m Road Width

100mm
Freeboard

5.0m

5.0m

Traffic Lane

Traffic Lane

Shoulder Height
3%

470mm
Pavement
Thickness

125.65m
3%

March 2010

9C-1

9C

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Shoulder height = Centreline Height Road crossfall


Shoulder Ht = 125.65m - 5m x 3%
Shoulder Ht = 125.50m
A 100mm freeboard below shoulder height is required, therefore the Allowable Headwater
depth (AHW) will be:
AHW = Shoulder Height Invert Height at Outlet Freeboard
AHW = 125.50 123.85 0.100 = 1.55m
The Tailwater depth can be obtained from the Stage-Discharge curve provided as shown in the
graph below (For details on how to determine a Stage-Discharge curve refer to Section 8.4.4).

Complete top section of Design Form as follows:


Culvert Design Example

9C
6.25

March 2010

9C-2

50

0.695

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

125.50 m

0.0065
15.6

1.55

123.85m

123.75m
2.5

Step 2.
Select a Trial Culvert (Refer to Section 9.11.1.2)
Firstly, determine an indicative waterway area of culvert(s) based on Q and Maximum
allowable velocity
A = Q/V
A = 6.25 / 2.5
A = 2.50m2
Use area to select one initial culvert configuration that meets area required. Size and fit culvert
to channel / road.
It is required to use a RCP Culvert.
At this stage and based on preliminary considerations given above, a RCP Culvert 1200 will
not be adequate as height of barrel will not fit between bottom of pavement thickness and invert
level of culvert at inlet. That is:

9C

125.5 0.460 123.85 = 1.19m


Initial trial size selection area = 2.50m2
Try 3/1050 RCP, Area = 2.60m2 (Check Pipe OD 1.19 - OK)
Note that for ease of Nomographs use, Nominal Pipe Diameter will be used.

Step 3.
Design Discharge for Trials (Refer to Section 9.11.1.3)
For multiple cells culvert determine Q per cell:
Q per cell = 6.25 / 3 = 2.08 m3/s (3/1050 RCP)

March 2010

9C-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Step 4.
Determine Inlet Control Headwater Depth (Refer to Section 9.11.1.4)
Analyse Culvert assuming inlet control.
Use nomograph 9A-2 to determine HW/D and then calculate HWi
Q = 2.08 m3/s, D = 1050, reading from nomograph HW/D = 1.33
HWi = HW/D x D = 1.33 x 1.05 = HWi = 1.40m < AHW (1.55m)

Step 5.
Determine Outlet Control Headwater Depth (Refer to Section 9.11.1.5)
Analyse Culvert assuming outlet control.
Select Entrance Loss Coefficient. Per job description above, Culvert Entrance Type to be square
edge with headwall, therefore ke = 0.5 (Refer to Table 9.10.2 Culvert Entry Loss Coefficient).
Determine Head (H) for RCP culvert flowing full from nomograph 9A-6
Plot L = 15.6m on ke = 0.5, plot D = 1050 and draw a line
Locate pivot on turning line
Plot Q = 2.08 m3/s and draw line crossing pivot on turning line to Head Line
Read off H = 0.50m
Determine Critical depth in a pipe, dc from nomograph 9AB.13, read off dc = 0.81m
Calculate (dc + D) / 2 = (0.81 + 1.05) / 2 = 0.93m
Determine Tailwater. TW was earlier determined as 0.695m from Stage Discharge curve.
Determine ho , which is the greater of TW=0.695m and (dc + D) / 2 = 0.93, therefore ho = 0.93m
Calculate Length x Slope = 15.6m x 0.0065m/m = 0.101m

9C

Determine HWO = H + ho - LSo = 0.5 + 0.93 0.101 = HWO = 1.33m < AHW (1.55m)

Step 6.
Determine the controlling Headwater (HW) (Refer to Section 9.11.1.6).
That is, assess larger between HWi = 1.40m and HWO = 1.33m
Therefore 3/1050 RCP culverts are operating under Inlet control.

March 2010

9C-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Step 7.
Calculate outlet velocity Inlet control pipe (Refer to Section 9.11.1.8)
Determine relationship of Part Full pipe (Qp) and Full flow pipe (Qf) = (Qp / Qf)
Using nomograph 9A-11, determine Qf and Vf read Qf = 2.4 m3/s and Vf = 2.5 m/s
Qp / Qf = 2.08 / 2.4 = 87%
Using nomograph 9A-12 read 87% on the axis x to discharge and obtain the following factors:
71.5% depth of flow, therefore y = 1.05 x 71.5% = 0.75m = y
112% velocity part full, therefore Vp = 2.75 x 112% = 3.08m/s = Vp

Step 8.
Calculate Outlet Flow Energy (Refer Section
Calculate Froude Number

Fr

Voutlet

g A

In order to calculate Froude Number, we need to calculate Area and Top width of flow.
Q per cell = 2.08 m3/s
A o = Q / Vo
Ao = 2.08 / 3.08
Ao = 0.67 m2

B 2 y D y

9C

y = 71.5% depth full flow = 0.715 x 1.05 = 0.75m

B 2 0.75 1.05 0.75


B = 0.95m

Fr

3.08

9.81 0.67

0.95

Fr = 1.17 (Supercritical flow)


Since Fr < 1.7, a rock pad will be suitable.

March 2010

9C-5

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Step 9.
Design Check (Refer to Section 9.11.1.10)
Outlet velocity = 3.08 m/s greater than Allowable stream velocity = 2.5 m/s. Should this culvert
be chosen, it is recommended to determine a suitable type of Erosion control measure.

Step 10.
Design a Culvert outlet protection as detailed in Section 9.15. (Also refer to Culvert Outlet
protection example shown further in this Appendix).

Discussion.
This example required a reinforce culvert pipe (RCP) to be designed.
Two different combination of RCPs size have been sought in this example. This included 4/900
RCP and 3/1050 RCP respectively. (Refer to attached design form)
Hydraulic performance of both culvert trials for the 50 year ARI design event indicates that
culvert will be operating under inlet control condition. Additionally, outlet velocity for both
trials, will be exceeding Maximum Allowable Stream Velocity at culvert outlet.
During the detailed design, apron and headwall structures need be determined in order to
establish their impact on the channel geometry and any potential impact on the tailwater by any
increase on the channel width in the vicinity of the culvert.
A further analysis needs to be undertaken to consider the likelihood of blockage (Section 9.13)
and determine the impacts on the culvert performance. That is, as the headwater increases and
the design rate is exceeded, what the consequences and impacts are to the culvert, roadway and
adjacent land

9C

March 2010

9C-6

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

9C

March 2010

9C-7

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Culvert Outlet Protection Example


Refer to Section 9.15 - Culvert Outlet Protection
The following procedure calculates the distance; a standard culvert outlet protection has to be
provided when design velocities are higher than acceptable.
For velocities in excess of 5 m/s, the use of energy dissipators as in Section 9.16 should be
considered.
Design an erosion control measure for 2/1200 RCPs (Inlet Control) that have been designed to
convey a total discharge of 4.5 m3/s. The Maximum Allowable Stream Velocity for the channel
is 2.0 m/s. The depth and velocity at the exit are 64% depth full flow and 2.97 m/s (per cell)
respectively. (Apron width has been calculated as 4.9m)

Step 1.
First determine a suitable type of control measure by calculating Froude Number:

Fr

Voutlet

g A

In order to calculate Froude Number, it is needed to calculate Area and Top width of flow.
Q per cell = 4.5 m3/s / 2 = 2.25 m3/s
Ao = Q / Vo
Ao = 2.25 / 2.97
Ao = 0.76 m2

B 2 y D y

9C

y = 64% depth full flow = 0.64 x 1.2 = 0.77m

B 2 0.77 1.2 0.77


B = 1.15m

Fr

2.97

9.81 0.76

1.15

Fr = 1.16
Since Fr < 1.7, a rock pad will be suitable.

March 2010

9C-8

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9C
Worked Examples

Step 2.
Determine stone size and length of pad from Figure 9.15 (a) as shown below:

2.97

1200

From figure above:


Length = 8D = 8 x 1.2 = 9.6m, however it is recommended that for multiple pipes Length be
extended 25%, hence Length = 9.6m + 25% = 12m
d50 = 300mm

Step 3.
Determine upstream and downstream width based on 5:1 expansion ratio. (Refer to Figure 9.15
(b)):
L = 12.0m

9C

L/5

W1 = 4.9m

W2

L/5

1
5

March 2010

9C-9

Department of Transport and Main Roads


Road Drainage Manual

Apron width = 4.9 m = W1


W2 = W1 + 2L / 5 = 4.9 + 0.4 x 12.0 = 9.7 m
Adopt d50 = 300mm, L = 12.0m, W1 = 4.9m and W2 = 9.7 m

9C

March 2010

9C-10

Appendix 9C
Worked Examples

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9D
Structural Assessment / Decision Process

Appendix 9D
Structural
Assessment /
Decision Process
9D

January 2013

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9D
Structural Assessment / Decision Process

Appendix 9D Amendments Jan 2013


Revision Register
Issue/
Rev
No.
1

9D

January 2013

ii

Reference
Section
-

Description of Revision

Initial Release.

Authorised
by

Date

M
Whitehead

Jan
2013

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9D
Structural Assessment / Decision Process

Pre-2004 Culvert identified


on a project and planned to
be retained

Can culvert be visually inspected?


(Is culvert clean enough to allow
visual inspection of whole length
of barrel(s))
NOTE: It may be prudent to clean
out structures to enable inspection

Pre-2004 Culvert Structural Assessment / Decision Process

NO

This appendix must be used in


conjunction with Appendix B
Structural Failure Risk Assessment
Consequences of Culvert Failure

YES

Key considerations when assessing / understanding the consequences of failure

Conduct visual inspection for


deformation (cross sectional and/or
longitudinal), separation of units,
cracking, spalling, rusting

Road user safety (damage / injury)


Traffic management (time, cost, availability)
Availability of suitable resources / plant / materials to repair / replace structure
Delays to other projects where resources and plant may be pulled from
Delays to traffic / disruption to freight movement and effect on regional economy
Community expectations to use road (traffic volumes and road importance)
Use by overdimensional vehicles (particularly over weight)

Inspection by any civil qualified


(or experienced) person
including a Level 1 inspector

Can a Level 2 Inspection be undertaken?


(Resources, size, debris, access to CCTV etc)

Pre-2004 culverts do not meet


current design specifications and
therefore should be considered as
sub-standard.

NO

NO

Does culvert show any


signs of failure / distress?

While major culverts should be within


the departments BIS, they should still
be visually inspected. However they
may not require a Level 2 inspection if
last inspection has been conducted
within last 2 years. Responsible RPEQ
should review BIS and determine if
new a Level 2 inspection is required.

YES

YES

A Level 2
inspection may
escalate to Level 3

Culvert has failed and should be


replaced as part of project / work

Pre-1976 culverts assessed as Condition


State 3 or 4 should be replaced. If past their
design life, they must be replaced.
Culverts between 1976 and 2004 assessed
as Condition State 3 or 4 should be
replaced, however Condition State 3
culverts could be retained based of advice
of a Level 3 Inspector.

Conduct a Level 2 Inspection,


based on philosophy within
TMRs Bridge Inspection Manual
to determine a Condition State
for structure

Responsible RPEQ uses BIM


inspection report and Appendix
B to determine if culvert can be
retained or should be replaced

Also need to consider expected


increase in service life of structure
and future traffic loadings (i.e.
freight task) of the road

Combined engineering /
management process /
decision

Document data /
process / decision

Is culvert suitable to be
retained?

YES

With information available, review


or determine risks and costs
associated with culvert failure in
near future.

NO

YES

Use Appendix B to determine risk rating


for structural failure of culvert.
Review consequences of failure of the
structure.
Consider cost to replace now within
project (economy of scale) against one-off
cost to replace sometime in near future.

9D

Document data /
process / decision

Document data /
process / decision

Is risk(s) acceptable and can TMR


justify / defend decision to retain?
NO
Proceed with project on basis
that culvert is to be retained

Replace culvert as part of


project / work

Document data /
process / decision

Culvert is deemed structurally


deficient and replaced as part
of project / work.
YES

Does Management agree to


replace culvert?

Proceed with project on basis


that culvert is to be retained

NO

It is a requirement that documented processes / decisions be


RPEQ certified for engineering component and signed by
appropriate / delegated TMR Management for acceptance of any
associated risks etc.

January 2013

9D-1

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9E
Structural Failure Risk Assessment

Appendix 9E
Structural Failure
Risk Assessment

9E

January 2013

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9E
Structural Failure Risk Assessment

Appendix 9E Amendments Jan 2013


Revision Register
Issue/
Rev
No.
1

9E

January 2013

ii

Reference
Section
-

Description of Revision

Initial Release.

Authorised
by

Date

M
Whitehead

Jan
2013

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9E
Structural Failure Risk Assessment

Structural Failure Risk Assessment


General
This appendix assesses the risk (probability) of structural failure in simple terms.
Culverts are categorised as either a major culvert or a minor culvert. Major culverts are defined
within the departments Bridge Inspection Manual (BIM). These structures are currently
inspected on a regular basis with records stored in the departments Bridge Information System
(BIS). All other culverts are deemed minor culverts.
Link slab box culverts, cast insitu culverts and arch culverts are not covered by this Appendix
and assessing risk of structural failure requires specialist advice.

Risk Assessment - Major Culverts


For major culverts, a risk assessment can be undertaken using the Whichbridge component
within the BIS.
Risk Rating Major
Score
1500+
751 - 1500
0 - 750

Risk
High
Medium
Low

Recommended Action
Replace structure
Monitor structure
No action required at this stage

Risk Assessment - Minor Culverts


Risk of structural failure for minor culverts (under critical loading / cover is 2m) is a function
of:
W - maximum horizontal dimension of a culvert (e.g. nominal diameter for circular
culverts, nominal span width for box culverts); and
C depth of cover which is defined as the minimum depth of material (under outer wheel
path) from the road surface to top surface of culvert.
Note: When cover exceeds 2 metres, inspection is still required, but specialist advice is required.
Assessment method is same for both single cell and multi cell structures.

Risk Rating W = 375 to 600mm


Assessment Criteria
When C is 600mm

Risk
High

Recommended Action
Replace structure

All other situations

Low

No action required

January 2013

9E-1

9E

Department of Transport and Main Roads


Road Drainage Manual

Appendix 9E
Structural Failure Risk Assessment

Risk Rating W = 675 to 1200mm


Assessment Criteria
When C is 600mm
When C is between 600mm and W
When C is W

Risk
High
Medium
Low

Recommended Action
Replace structure
Monitor structure
No action required

Risk
High
Medium
Low

Recommended Action
Replace structure
Monitor structure
No action required

Risk Rating W = 1350 to 1800mm*


Assessment Criteria
When C is W/2
When C is between W/2 and W
When C is W

* - RCBC culverts greater than 1800mm (in width) but less than 3.0m2 in waterway area
are to be included.

The above Risk Rating tables are combined and shown diagrammatically in the chart below.

9E

January 2013

9E-2

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Chapter 10
Floodway Design

10

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Chapter 10 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

10

March 2010

ii

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Table of Contents
10.1

Introduction

10-1

10.2

Additional Considerations

10-1

10.3

Geometric and Safety Issues

10-1

10.4

Environmental Factors

10-2

10.5

Hydraulic Design

10-2

10.6

10.7

10.5.1

Floodway Terminology

10-2

10.5.2

Flow Over the Road

10-4

10.5.3

Full Floodway Design Calculations

10-4

Time of Submergence / Closure

10-8

10.6.1

Time of Submergence

10-9

10.6.2

Time of Closure

10-9

10.6.3

Issues Related to Times

10-9

10.6.4

Calculation of Time of Submergence or Closure

10-11

10.6.5

Procedure for Estimating AATOC / AATOS

10-12

Floodway Protection

10-13

10.7.1

Types of Protection

10-13

10.7.2

Floodways with Grassed Batters

10-13

10.7.3

Floodways with Other than Grassed Batters

10-14

10

10

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

10

March 2010

iv

Chapter 10
Floodway Design

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Chapter 10
Floodway Design
10.1

Introduction

Floodways are sections of roads which have


been designed to be overtopped by
floodwater during relatively low average
recurrence interval (ARI) floods. An
example of a flood is shown in Figure 10.1.

requirements in the selection of final road


level.
Floodways may offer environmental
advantages over culverts or bridges, since
they will tend to spread flows more widely.
This means that the risk of scour to
waterway and surrounding land is generally
reduced because flow is less concentrated.
It is also important that a floodway be
designed so that it is not covered by water
from ponding or backwater for any
significant period of time after a flood
event.

10.3

Figure 10.1 - Little Annan River Floodway

In Queensland, these overtopping floods are


usually in the ARI 10 to 20 year range
although some roads may be overtopped by
smaller flood events.
Chapter 2 outlines many factors which
should be considered before deciding on the
design flood immunity for new road works.

10.2

Additional
Considerations

Further to the requirements discussed in


Chapter 2, floodways may require costly
batter protection and therefore a higher
level road together with a larger culvert or
bridge option may be more cost effective.
Floodways also have smaller waterway
(under road) requirements and may be more
prone to blockage by debris. These cost
related performance factors should be
considered as well as trafficability and other

Geometric and Safety


Issues

It is important that adequate approach sight


distance be provided to allow drivers time
to recognise water over the road and to
stop. It is also important that the length of a
floodway be limited at about 300 m so that
drivers do not become disorientated when
confronted with wide open stretches of
water. Where a proposed floodway would
be longer than 300 m, it is recommended
that the proposed floodway be broken into
shorter lengths by providing sections of
road that are raised above the maximum
flood level.
As a general principle, floodways should be
designed so that the depth of water over the
road should be as uniform as possible over
the flooded section. Building a floodway
on a level grade avoids the possibility of a
driver unexpectedly encountering deeper
water and possibly stalling or being swept
downstream.

March 2010

10-1

10

Department of Transport and Main Roads


Road Drainage Manual

Exceptions to the level grading in


Queensland occur where bridges have been
built significantly higher than the flooded
approaches on both sides. The bridges have
been built on the basis that the approaches
will be raised sometime in the future.
Floodways should not be placed on
horizontal curves as:

there are problems in defining the


edge of the pavement for motorists;
any superelevation may change the
normal flow distribution i.e. push
more
water
to
the
nonsuperelevated sections of road; and
the water depth will be deeper on
one side of the road than the other
in a superelevated section of road
and there is the possibility of the
high side being trafficable but not
the other, thus creating a safety
problem.

Floodways should also not be located on


vertical curves to avoid variations in depths
of flows.

10

For further geometric requirements of


width, crossfall, vertical and horizontal
alignment, refer to the relevant chapters of
the departments Road Planning and
Design Manual. Signage of the floodway is
also important and designers are referred to
the latest release of the Manual of Uniform
Traffic Control Devices for warrants /
guidance.

10.4

Environmental Factors

For floodways that contain floodplain


culverts (i.e. culverts located away from a
watercourse channel) only terrestrial
movement generally needs to be
considered. However, if fish migration is
expected to occur across the floodplain

March 2010

10-2

Chapter 10
Floodway Design

during times of flood, then a check should


be done on allowable flow velocities.
The procedures outlined in Section 9.7
should be followed for fauna passage
through culverts.
As noted above, floodways reduce the
concentration of flow, compared to culverts
or bridges, so the risk of scour damage to
waterway and surrounding land is reduced.

10.5

Hydraulic Design

10.5.1 Floodway Terminology


A floodway consists not only of the
roadway embankment to accommodate
flow over the road but also waterway
openings to provide for flow under the road.
These openings may be required for one or
more of the following functions:

reduce the afflux or rise in water


level upstream due to the
obstruction (embankment);

raise the tailwater level so that less


batter protection is required on the
downstream side e.g. grass instead
of concrete; and/or

act as anti-ponding structures for


low flow stream conditions.

Flow over roadways may be:

free flow; or

submerged flow.

In the initial stages of overtopping a low


tailwater usually exists and free flow
occurs. Under these circumstances flow
passes through critical depth over the road
and the discharge is determined by flood
levels upstream.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Free flow may be either:

plunging flow which flows over the


shoulder and down the downstream
face of the embankment. The flow
then penetrates the tailwater surface
producing a submerged hydraulic
jump on the downstream slope.
Velocities are likely to be high and
erosive; or
surface flow which separates from
the surface of the road embankment
and rides over the surface of the

tailwater. This flow will have less


erosion potential downstream.
Submerged flow occurs when the
discharge is controlled by the tailwater level
as well as the headwater levels. This occurs
when the depth of flow over the road is
everywhere greater than the critical depth.
Typical velocities of flow over a floodway
are shown on Figure 10.5.1 as sourced from
Waterway Design (Austroads 1994) after
Cameron and McNamara (1966).

Q = 1.16m3/s (HEAD DROP = 1.77m)

Q = 1.16m3/s (HEAD DROP = 0.70m)

10

Q = 1.16m3/s (HEAD DROP = 0.36m)

Figure 10.5.1 - Indicative Velocities of Flow over a Typical Floodway

March 2010

10-3

Department of Transport and Main Roads


Road Drainage Manual

10.5.2 Flow Over the Road

Chapter 10
Floodway Design

Where:

H h

The broad crested weir formula used for


flow over a road is:

C
Q C f LH s
C
f
1.5

V2
= Specific Head
2g

On Figure 10.5.2;

for H/l > 0.15 (usual case) Use


curve B to obtain the free flow
coefficient of discharge, Cf.

Where:

Q = discharge over floodway


(m/s);

for H/l < 0.15 Use curve A to


obtain value of Cf.

Cf = coefficient of discharge free


flow;

Cs = coefficient of discharge flow


with submergence;

Calculate D/H x 100 and use curve


C to obtain the submergence factor
Cs / Cf.

Calculate the discharge over the


road using the broad crested weir
formula.

L = length of floodway (m); and

H = specific head or specific


energy (m).

With reference to Figure 10.5.2:

10.5.3 Full Floodway Design


Calculations

h = level difference between the


floodway crown and the upstream
water surface (m);

Floodways incorporating culverts and


bridges will require calculations in addition
to those above.

V = approach velocity of the stream


(m/s); and

l = top width of road formation (m).

The basic principle is that the total flow


over the road and through the waterway
structures equals the flow downstream in
the unrestricted channel.

The flow over the floodway may be


calculated by means of the following design
procedure:

10

Calculate the stage-discharge curve


(height versus discharge) for the
unrestricted section, from open
channel hydraulics (refer Chapter
8);

Select a tailwater level and a


headwater level (giving h and V)
from the stage-discharge curve.

Calculate

March 2010

10-4

H
l

Given the many combinations of headwater


and tailwater possible, it is necessary to fix
at least one of these parameters for each
design calculation.
With flow over the road, the issue of
whether or not road batter protection is
needed becomes important as the
calculations require the tailwater to be fixed
when the flood is at the point of
overtopping the road.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

10
Figure 10.5.2 - Discharge Coefficients for Flow over Floodways

March 2010

10-5

Department of Transport and Main Roads


Road Drainage Manual

Guidelines relating to the need for


protection are given in Section 10.7 with
the tailwater level generally in the range
0.4m to 0.6m below the crown of the road
when the road is about to be overtopped.
Full floodway design must satisfy Case A
and Case B conditions.
Case A: When the flood is at the point of
overtopping the road:

Tailwater level is not below the


level specified for the type of
protection to be adopted; and

Velocities through the bridges


and/or culverts are acceptable.
(Scouring velocities may be
acceptable, if additional outlet
protection or the formation of scour
holes is acceptable).

The step by step design procedure for this


case follows:
Figure 10.5.3 illustrates results of
calculations performed using the steps (a) to
(g) below for an actual floodway.
(a) Calculate the rating curve (height
versus discharge) for the unrestricted
channel.

10

(b) Fix the level of the road as a first trial.


The initial level of the road may be
based on trafficability eg for required
trafficability in a ARI 20 year flood, the
initial road level may be the level of the
ARI 20 year flood level in the
unrestricted channel.
(This allows
culverts to be designed for a maximum
head of 300 mm with the tailwater at
the crown level of the road. It will be
found that a lower road level will give
less fill in the embankment, but require
more culverts under the road in the ARI
20 year flood.)

March 2010

10-6

Chapter 10
Floodway Design

(c) Fix the headwater level at the crown of


the road (or the highest edge if
superelevated). Avoid the design of
superelevated floodways, if possible, as
these can result in depth variations
laterally and surface debris on the
floodway surface can compromise
safety. Residual, silt and gravel on
floodways after isolated rainfalls events
can provide hazards on their own,
causing serious injury. Effective
maintenance programs must be in place
by relevant authorities and new designs
of new floodways need to minimise
risk.
(d) Find the velocity Vx through a suitable
major culvert / bridge with the
headwater in (c) above and the tailwater
adopted for the type of floodway
protection to be adopted. Say this
tailwater level is RLx on Figure 10.5.3
Case A, corresponding to a total
discharge, Qx in the unrestricted
channel.
(e) With this velocity Vx find the total area
of waterway required from Total Area
Required = Qx/Vx and select other
culverts to give this total.
(f) For the fixed headwater and the
tailwater level, RLx, calculate the actual
flow through each culvert and bridge
and total the discharges to give Qy
which may or may not be equal to Qx
because of the different size openings.
(g) Adjust the culvert and bridge areas so
that with the fixed headwater and
tailwater level, RLx, the total flow
equals Qx.
(With progressive
adjustments the total discharge may be
Qy or Qz so on.)

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

CASE A

CASE B

Figure 10.5.3 Example Floodway Calculations

Note that the initial assumption of all the


waterways having the same velocity is
approximate only as different size and
shape openings have different hydraulic
efficiency. This is why the same head will
give different velocities.
The structures giving Qx are to be
considered minimum requirements for the
floodway regardless of the requirements
calculated for a peak flood above the
roadway, as all overtopping floods pass
through this stage.

Note: The design velocities through


culverts are normally in the range 1.8 m/s
to 3.0 m/s with 2.4 m/s commonly adopted,
depending on the scour resistance of the
surface material downstream. Higher and
lower velocities have been adopted.
Chapter 9 discusses the hydraulic design
of culverts and Chapters 7 & 13 methods
of erosion control.

March 2010

10-7

10

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

through the waterways. Similarly,


Q3 etc may be obtained.

Case B: At the peak of the flood above


the floodway:

Velocities through the waterway


structures are acceptable; and

Afflux caused by the floodway is


acceptable.

The step by step design procedure for this


case follows:
At the peak of the overtopping flood,
allowance for flow over the road must be
made and afflux calculated for the design
flood (ARI 20 year in the example in Case
A above).
The flow over the road is added to that
through the culvert(s) and bridge(s) and the
calculations initially include the structures
from Case A. With reference to Figures
10.5.3 and Case A and Case B, the
procedure is:

10

Find
the
tailwater
level
corresponding to the design flood
discharge from the rating curve for
the unrestricted channel.
Select a headwater level slightly
above the tailwater level in (a), say
0.1 m above, and calculate the flow
over the road and through each
culvert and/or bridge with this
headwater - tailwater combination.
Add these flows together to obtain
a total flow, say Q1.

Compare this flow with the design


flood discharge. If Q1 is smaller, a
greater head / afflux is required to
ensure that the check flood is
passed over the road and through
the waterways.

Adopt a higher headwater and the


same tailwater level and obtain the
total Q2 for flow over the road and

March 2010

10-8

Plot Q1, Q2, Q3 etc. against their


respective headwater levels. Join
these points to make the curve
shown on Figure 10.5.3 Case B.

From the curve, find the headwater


which gives the required design
flood discharge.

If the afflux is acceptable,


calculations are complete. If the
afflux is too high, additional bridge
or culvert waterways are required.
Waterways from Case A are not to
be reduced.

Calculate the afflux for the check


flood (usually of ARI 50 years).

Because of flow over the road, maximum


velocities are obtained from Case B.
Both velocity and afflux requirements are to
be met.

10.6

Time of Submergence
/ Closure

Some of the following text and diagrams


have been taken from Bridge Waterways
Hydrology and Design (NAASRA 1989)
and Waterway Design (Austroads 1994).
Where a crossing is designed for
overtopping, it is usually important that an
estimate be made of the frequency and
duration of the periods during which the
crossing will be submerged as well as the
times it is closed to traffic due to flooding.
The time of submergence is of importance
with respect to stability of embankments
and pavements and consequent maintenance
costs. The time of closure is of importance
in consideration of acceptable delays to
traffic.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

10.6.1 Time of Submergence


Time Of Submergence (TOS) is defined as
the period of time that the road is inundated
by flood water, no matter the depth.
The time of submergence is expressed in
one of two main ways. The first is the time
of submergence during a major flood - for
example, the number of hours of
submergence during the flood with an
average recurrence interval of 50 years.
The second is as the Average Annual Time
Of Submergence, abbreviated as AATOS.
This is the average time per year that the
road is submerged, expressed as hours per
year.
Both expressions are an indication of the
frequency and duration of submergence.
For example a crossing that is submerged
frequently for short periods of time may
have a similar AATOS as one that is
submerged less frequently for longer times.
To understand the concept and to compare
options, three parameters are needed,
namely:

Flood immunity;

TOS for the ARI 50 year flood (or


other large flood);

AATOS.

The absolute value of the


submergence is not particularly
itself for a crossing, but it is
compare several crossings or
options.

time of
useful by
useful to
upgrading

10.6.2 Time of Closure


Time Of Closure (TOC) is similar to time
of submergence. However this parameter
takes account of the fact that some
inundation of the road may not necessarily

close the road, though there may be some


hazard in travelling on the inundated road.
Different types of vehicles can travel on
roads with different amounts of inundation
with large and heavy vehicles capable of
travelling in water that is deeper and
flowing at a higher velocity than a light car.
Safety of vehicles in flood water is poorly
understood and there has been considerable
research on the topic.
There are different definitions of the flow
conditions when the road is closed by flood
waters. The normally adopted limit in the
department is when the total head across the
road exceeds 0.3 m. That is:
Total head (H) = d + v2/2g
Where:

d = depth of inundation

v = flow velocity

g = gravity (9.81m/s2)

Similar to AATOS there is an Average


Annual Time Of Closure, abbreviated as
AATOC

10.6.3 Issues Related to Times


The concept of time of closure, in
association with the time of submergence,
adds some additional information to the
question of flood immunity.
Large and flat catchments will respond
more slowly than small and steep
catchments so the time of submergence for
these catchments will generally be longer
for an equivalent flood immunity.
In small steep catchments, where the
response time is short, the time of
submergence will be low, even for a
crossing with low flood immunity.

March 2010

10-9

10

Department of Transport and Main Roads


Road Drainage Manual

In small steep catchments, and also for


urban catchments where the response time
is short, floods may occur and then recede
very quickly. In this case the disruption to
traffic may be short. Similarly in small
catchments, the depth of inundation may be
low even if the flood immunity is low. In
this situation, the time of submergence may
be short so the cost of disruption may be
very low and the additional cost of
providing for a higher level of flood
immunity may not be justified.
On the other hand, large catchments that
have a long response time may inundate the
road for extended periods of time. In this
case the cost of disruption may be very high
and a higher level of flood immunity may
be more easily justified.
There are occasions where the road crosses
a tributary close to the junction with a
major stream. If the road is inundated by
both the local catchment runoff and
backwater from the major stream, the time
of submergence may be quite different for
the two flood mechanisms. The local
catchment runoff, from a small catchment,
may have a short time of submergence,
while the larger catchment may inundate
the road for longer periods of time. Local
residents can often identify these two
sources of closure and they may be quite
distinct.

10

When calculating the time of submergence


or closure in these situations, the
calculation must consider both sources of
inundation as well as the risk that the
inundation is independent, in which case
the times must be added together.
When the flood immunity is very low, the
time of closure may not be represented
accurately by the flood event hydrographs.
In this case, long term records may need to
be analysed, but this situation would be
March 2010

10-10

Chapter 10
Floodway Design

unusual for state controlled roads in


Queensland though it does occur on
interstate main roads and on more minor
roads.
These situations can occur where the road is
inundated for months at a time during the
wet season or even sometimes the road may
be inundated by tides. The calculation in
this case is very difficult and should use
either a stream gauge or calculated
continuous discharge records.
Average time of submergence or closure
may be assessed for a range of selected
grade levels and a plot of average time of
submergence against level may be produced
as in Figure 10.6.3.

Figure 10.6.3 - Typical Deck Level / Time


of Submergence Relationship

In many cases the plot will reveal a


particular grade level above which a
relatively large increase in level will result
in only a small decrease in time of
submergence, and a small reduction in level
results in a large increase in average time of
submergence. Such a level may be selected
as a starting point for economic analysis.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

submergence) or close the road (for


time of closure).

10.6.4 Calculation of Time of


Submergence or Closure
There are two approaches to calculating
TOS / TOC of a road by flooding. The first
approach is for the situation where there is a
stream gauge located at the bridge site,
which is not a common occurrence. The
second approach is based on analysis of
flood hydrographs, and is the method that is
most commonly adopted.
Analysis with stream flow records is the
most reliable approach and is recommended
if there is a nearby station. However since
this is often not the case, the theoretical
approach is more commonly necessary.
If there is a stream gauge located on the
stream at or very near the bridge, this gauge
can be used to analyse the TOS / TOC.
Whilst this does not occur commonly, it is
the most accurate method for calculating
the TOS / TOC. Use of a stream gauge is
also useful since calculated hydrographs are
prepared considering the maximum
discharge and the length of the hydrograph
may often vary and longer duration floods
with a lower peak discharge may be more
critical in consideration of time of
submergence or closure than the actual peak
discharge.
If there is no stream gauge located close to
the crossing, it is possible to calculate a
continuous record of discharge from a
hydrologic model and to use this sequence
of flows in the calculation of TOS / TOC in
exactly the same way as a stream flow
record would be used.
The procedure for applying the stream
gauge records is as follows:

Calculate the discharge that will


just inundate the road (for time of

Calculate the total period of time


where the road is submerged (or
closed) from the complete record of
the stream gauge.

Calculate the average duration per


year that the road is submerged (or
closed). This is the AATOS (or
AATOC).

This procedure does not provide the


duration of submergence (or
closure) for a particular flood event.

As with all hydrologic analysis, it is


important that the stream flow record is
sufficiently long to provide a representative
sample of flow at the bridge. If the period
of record is too short, or the period of
record is not representative, the result from
the calculation may not be reliable.
It is hard to define the period of record that
would be adequate for this analysis, since it
depends on the variability at the site as well
as other conditions. However it is likely
that if there is less than 20 years of record at
the site, the result may not be reliable.
If there is only a short record, careful
analysis can be used to assess the best
means of extracting data from the record,
and useful information can be extracted.
It is much more likely that there will not no
stream gauge located near the bridge, and
the theoretical method discussed here must
be used.
The first step in this process is to calculate
design flood hydrographs for the crossing
site. This can be done with a catchment
hydrologic model, such as RORB or
RAFTS. It is noted that the Rational
Method calculates only the flood peak
discharge so is not suitable for application

March 2010

10-11

10

Department of Transport and Main Roads


Road Drainage Manual

of time of submergence or time of closure


calculations.
It is also noted that the actual time of
inundation of the road depends on the flood
levels and not necessarily the discharge so
in some cases, especially on large flat
floodplains, the hydrographs of water levels
may not be exactly the same as the
hydrographs of flood discharges so this
issue may need to be considered at times.
However the normal procedure is to
calculate flood hydrographs of discharge
with a catchment model and use these
directly for the calculation.
To calculate the time of submergence or
time of closure the following is required:

10

A hydrograph of the flood, which


may be obtained from actual
measurements at the stream
crossing, or for ungauged streams,
by the use of a runoff routing
method or synthetic hydrograph
method. (An experienced hydraulic
engineer is required to undertake
these analyses.)

Flow capacity of the crossing at the


point of submergence or closure.

The time of submergence or closure


may then be calculated by drawing
a horizontal line on the hydrograph
at the flow capacity level and
measuring the time for which the
flow is above this level.

10.6.5 Procedure for Estimating


AATOC / AATOS
The procedure for calculating the average
annual time of closure (AATOC) is as
follows:

Chapter 10
Floodway Design

Step 1.
Determine the ARI of the flood for which
the stream crossing is trafficable, i.e.
floodway with or without culvert or bridge.
Step 2.
Determine the times of closure ty for a
series of floods greater than the trafficable
capacity flood, and the ARIs of each of
these floods.
Estimate tmax by extrapolating a graph of
time of closure versus ARI or by estimating
the probable maximum flood for the
catchment. It should be noted that the
probable maximum time of closure cannot
be disregarded in the calculation of the
AATOC.
Step 3.
Calculate the probability FT(t) of the road
being closed for each ARI y year flood:
FT(t) = 1 1/y
Step 4.
Using the times of closure, ty and the
probability, FT(t) of the road being closed
for each ARI y year flood, draw the
cumulative probability distribution.
This distribution gives the probability of the
road being closed in any year for less than t
hours.
Step 5.
Determine the probability density function:
fT (t) = FT(t) / t
This is the slope of the line connecting each
point on the cumulative probability
distribution.
Step 6.
Determine AATOC as:

AATOC p t

March 2010

10-12

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Where:
p = area of each rectangle in the
probability density function; and
t = centroidal distance of each rectangle
from the fT(t) axis of the probability
density function.
The same procedure is used for AATOS
except a lower flow is used to derive critical
flood level.

10.7

Floodway Protection

10.7.1 Types of Protection


Selection of the form of protection of
floodways against scour is governed by:

Whether flow across the floodway


is free or submerged; and
Under free flow conditions,
whether plunging or surface flow
occurs downstream from the
floodway.

The tailwater level when the flood is at the


point of overtopping the road usually
controls the degree of protection required
for a particular floodway. Therefore, the
cost of providing adequate bridge and/or
culvert waterways to raise the tailwater to a
high enough level to require minimum
protection becomes a prime consideration
as well as the cost of the protection itself.

The tailwater level for floodways with other


than grassed batters is usually more than
300 mm and up to 600 mm or even 700 mm
below the downstream edge of the road
formation when overtopping first occurs.
Overtopping flows of long duration and at
frequent intervals may cause pavement
failures, and softening of the embankment,
aggravating any tendency to scour. Even
with a high tailwater when the flood is at
the point of overtopping the road, the time
of submergence may indicate the need for
more elaborate protection other than natural
grass.

10.7.2 Floodways with Grassed


Batters
Grass in this type of floodway is defined as
turf or seeded grass. Reinforced grass is
discussed in the Chapter 8.
Because the physical properties of grass
such as species, stiffness, cover density and
rooting pattern varies with soil type and
climate, only general guidelines based on
constructed floodways in Queensland are
possible.
Floodways with grassed batters should have
the following features:

Bitumen seal or asphalt pavements


with concrete or other rigid margins
/ shoulders (stone pitching, cement
stabilised gravel etc.) containing
the bitumen in place;

As an alternative to (a) above,


concrete blocks / nib walls along
the top edges of the formation with
bitumen seal or asphalt pavement
between them may be constructed.
These containing blocks may be as
simple as 10% by volume cement
stabilised gravel strips 600 mm
wide at the top by 180 mm deep;

Floodway protection is considered in two


categories:

Minimum protection
grassed batters; and

Other than grassed batters.

such

as

The tailwater level for floodways with


grassed batters is usually not more than 300
mm below the downstream edge of the road
formation when overtopping first occurs.

March 2010

10-13

10

Department of Transport and Main Roads


Road Drainage Manual

Culverts under the floodway


section should raise the tailwater to
not more than 300 mm below the
downstream edge of the road
formation when overtopping first
occurs;
Overtopping should occur for a
period of less than 12 hours in a 50
year average recurrence interval
flood.
However, the type of
material in the embankment and its
saturated strength may require
reduction of this allowable time of
submergence.

Chapter 10
Floodway Design

and savings in these areas should not be


considered.
Five types of floodway which have
performed satisfactorily are described in
this section with their associated limitations
specified. All these types of floodways
should have the following general features:

Culverts to raise the tailwater to not


more than 600 mm below the
downstream edge of the road
formation when overtopping first
occurs (actual range 300 - 700
mm). It is important to note that
this afflux may not be acceptable in
some areas and non-standard
additional outlet protection at
culverts may also be necessary;

Full protection of the top surface of


the road formation, as for
floodways with grassed batters;

Protection of at least the


downstream batter. Although not
clearly defined, it appears that
protection of the upstream batters
may only be required in floodways
of low flood immunity in major
streams. As a precaution where
only downstream protection is
adopted, protection for a distance of
about 3 metres on each side of
major culverts on the upstream side
may be placed to offset possible
scour due to turbulence from the
mixing of longitudinal and direct
flows at the culvert inlets;

Adequate downstream aprons. For


height of road embankment, H,
equal or more than 2.0 m, the
downstream apron should extend at
least 1.5H metres away from the
toe of the embankment. For H less
than 2.0 m, the downstream apron

Conversely there are some low floodways


that can withstand submergence for much
longer.
For this type of protection, it is desirable to
have good grass cover when the
overtopping flood occurs and this in turn
requires an ability to maintain grass cover
during the dry season.
Concrete pavements if used instead of
bituminous types will, of course, cover the
full width of the formation and not require
grass batter protection.

10.7.3 Floodways with Other


than Grassed Batters

10

There have been many types of floodways


successfully constructed in Queensland
with
other
than
grassed
batters.
Recommended types of protection are
described and illustrated in this section and
the selection of the type of protection would
be based on a cost comparison of
recommended types considered suitable to
that locality.
Most failure of floodways with downstream
batter protection in the past commenced by
scouring at the downstream aprons and/or
the downstream edge of the road formation
March 2010

10-14

Department of Transport and Main Roads


Road Drainage Manual

should extend at least H away from


the toe of the embankment unless
other specified; and

Weepholes in the downstream rigid


protection to relieve hydrostatic
pressure.
These weepholes are
normally about 90 mm diameter at
1.8 m (maximum) centres with 300
x 300 x 150 no-fines concrete
blocks behind the weepholes. The
weepholes should be placed about
300 mm above the apron level or
just above long standing water level
if higher. The more porous types of
protection such as rock on filter
cloths or layers and some cement
stabilised gravels, depending on the
grading, do not require weepholes.

A brief description of the five successful


types of floodway protection follows and
sectional details of these types are shown in
Figures 10.7.3(a) and 10.7.3(b). These
types are not in any order of preference and
comparative cost comparisons should be
made where more than one suitable
protection is considered.
A brief mention of reinforced grass, or a
type of geotextile protection as it is also
known in the USA, is given in Other Types.
Although it appears to have been successful
in some stream channels, little is known
about its performance in floods overtopping
embankments.

Chapter 10
Floodway Design

Type 1 Floodway
This type of reinforced concrete floodway
has been constructed in many areas of the
State, particularly in the west.
The
reinforcement selected should not only
satisfy strength requirements, but also
prevent temperature and shrinkage cracks.
Where cut-off walls have been used without
the downstream apron, failures have
occurred.
With a suitable width of downstream apron
and weepholes, the Type 1 Floodway is
recommended as suitable for all crossings
where other than grass protection is
required, cost permitting.
Type 2 Floodway
This is an example of a reinforced concrete
floodway where the tailwater depth is
uncertain but probably quite low (perhaps
700 mm or more below the downstream
edge of the formation when the flood
begins to overtop the road). No adverse
reports are known. Costs are higher than
Type 1 floodways.
Type 3 Floodway
This previously used type is no longer
recommended.
Type 4 Floodway
This is considered by some to be an
improvement on Type 3. The increased use
of stone mattresses and gabions has given
confidence to this type of construction.

March 2010

10-15

10

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Most common type.


Widely Used.

TYPE 1 Concrete Protection

Type performs well, but


need to justify cost.
Requires specialist
design.

TYPE 2 Concrete Protection


An improvement over
previous Type 3.
Mattresses must be
pinned / anchored.
Consider a cutoff wall.
Cut off walls may not be
necessary as mattresses,
usually achieve their
optimum position with a
little scour by dropping
into a scour proof position

TYPE 4 Rock Mattress Protection


Figure 10.7.3(a) - Downstream Floodway Protection (Types 1,2 & 4)

Type 5 Floodway

10

This type of protection incorporating a


bituminous seal is probably the lowest cost
of the types shown, but its use is limited.
It should only be used only where:

Fill height is not higher than 900


mm;
Tailwater at overtopping is not
more than 300 mm below the
crown of the road; and

March 2010

10-16

Time of submergence is small


(hours).

Rock fill with size 70 - 100 mm and median


diameter 85 mm would be adequate for
most uses.
Type 6 Floodway
This previously used type is no longer
recommended.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 10
Floodway Design

Type 7 Floodway
This floodway was designed for use in
Western Australia and is described in
Waterways Design (Aust roads 1994).
As the riprap consists of rock with grading
requirements, it may have limited
application in Queensland where supplies
of such rock are scarce where floodways
are constructed.
Further details of the required grading and
riprap thickness may be obtained from the
reference.

Common, low cost type.


Suitable for low velocities
over floodway situations
only.

TYPE 5 Bitumen Seal Protection

Variation to Type 4 where


mattresses are not
readily available.
Consider a cutoff wall.

10

TYPE 7 Dumped Rock (RipRap) Protection


Figure 10.7.3(b) - Downstream Floodway Protection (Types 5 & 7)

March 2010

10-17

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

Appendix 10A
Worked Examples

10
A

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

Appendix 10A Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

10
A

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

Floodway Example
Refer to Chapter 9 - Culverts Design and Chapter 10 Floodway Design.
This example illustrates the principles of floodway design, as described in Chapter 10.
The task for this example is to design a floodway with 20 year trafficability, over a natural open
channel approximately trapezoidal in shape. The floodway would be approximately 90 to 100
m long and for cost reasons, road batters will be grass covered only for protection against scour.

Preliminary Considerations:

Because the batters will be protected by grass only, culverts will be required to build up the
tailwater to not more than 300 mm below the edge of the downstream shoulder when
overtopping of the road first occurs. Allowing for crossfall, there will be a head of 450 mm
and a velocity of about 2.30-2.45 m/s through the culverts if this minimum tailwater is
adopted. Is this acceptable on this job? (This will be answered in the example).

For 20 year ARI trafficability, the floodway level should be at the 20 year unrestricted flood
level to allow the maximum amount of water over the road and save on culvert and overall
costs.

In general it is only in very long floodways with very little velocity in the open channel and/or
where costly protection is unavoidable, that increasing culvert requirements by raising the road,
thus decreasing the flow over the road (to the extreme of a flood free road) may reduce the
overall cost of the job.

Step 1.
List all relevant criteria:

Required standard: Trafficable in a 20 year ARI flood.

Time of closure: Maximum of 1 day in a 50 year ARI flood. (Calculated hydrograph shows
this maximum only a matter of hours - not included here).

Batter protection: Grass.

Width of floodway: 10 m.

Road crossfall: 3 %.

10
A

Step 2.
Calculate the rating curve for the unre-stricted channel.
Using open channel hydraulic calculations (as discussed in Chapter 8):
Q50 = 162 m/s @ Ht 322.76 m,
March 2010

10A-1

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

Q20 = 130.4 m/s @ Ht 322.58 m (V = 0.68 m/s),


Q10 = 108.1 m/s @ Ht 322.44 m, and
Q5 = 70 m/s @ Ht 322.13 m.

Step 3.
Adopt a road level and calculate the maxi mum allowable depth of water over the road.
Adopt the road level at the unrestricted 20 year ARI flood level and show the cross-sectional
details in figure below.

From Section 10.3, closure to traffic occurs when:

H d

V2
300mm
2g

The maximum allowable depth of water over the road, y, occurs when H = 0.300 m.
Therefore:

dH

V2
0.68 2
0.3
2g
2 9.8

d = 0.28 m

Step 4.

10
A

Calculate the discharge over the road and through the culverts in a 20 year ARI flood.
The total discharge over the road and through the culverts must equal the discharge in the
unrestricted channel downstream with flow at height 322.58 m.
i.e.

Qtot = QR + QC = 130.4 m/s

Flow over the road:

C
Q C f LH 1.5 s
C
f
road)

March 2010

10A-2

(Section 10.5.2 for free flow, tailwater not above crown level of

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

where Cf = coefficient of discharge.


From Figure 10.5.2:
H/1 = 0.30/10 = 0.03, which is < 0.15
Therefore:
Cf = 1.674

L = length of floodway = 94 m

H = 0.30 m

Therefore:

Q 1.674 94 0.31.5
Q = 25.9 m/s
Discharge through culverts:
QC = Qtot - Q = 130.4 - 25.9 = 104.5 m/s
Therefore, culverts are required to take a discharge of 104.5 m/s operating under a head of 0.30
m and outlet control for this design condition.

Step 5.
Detailed culvert design.
Proceed with the design of culverts to take 104.5 m/s.
Height of culvert opening, D, from:
D = Crown level of road - crossfall - minimum fill above culvert - thickness of deck slab invert level
D = 322.58 - 5 x 0.03 - 0.100 - 0.180 - 320.55 = 1.60 m.
For culvert design:
Outlet control with H = 0.30 m

10
A

Tailwater at Ht 322.58 m, HW at Ht 322.88 m


Invert at Ht 320.50 m, say
From culvert design procedure in Chapter 9:
13 / 2700 x 1500 RCBC have a capacity of 104m/s with Outlet Velocity = 1.98 m/s.

Adopt 13 / 2700 x 1500 RCBC and Floodway Ht 322.58 m and Length 94 m.

March 2010

10A-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

Step 6.
Check for culvert requirements when the flood is at the point of overtopping the road.
By trial and error, calculate flow through culvert at point where headwater height is 322.58 m.
Confirm that flow velocity through culvert is less than 2.5 m/s.

Step 7.
Confirm that velocity criteria is met.

Step 8.
Calculate afflux for ARI 50 year flow and confirm that criteria is met.

End of Example

10
A

March 2010

10A-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

AATOC Example
The design floods have been calculated for the bridge for a range of probabilities. In this case
the bridge is overtopped for a flood with a discharge of 400 m3/s, so has a flood immunity of
approximately ARI 3 years.

The design floods are plotted as follows.


1200

1000

Discharge - m3/s

800
ARI 100 years
ARI 50 years
ARI 20 years
ARI 10 years
ARI 5 years
ARI 2 years

600

400

200

0
0

10

15

20

25

30

35

40

Time - h

The times of submergence are listed in the following table.

ARI - years
2

TOS - hours
0.0

4.0

10

5.0

20

6.0

50

7.0

100

7.5

10
A

March 2010

10A-5

Department of Transport and Main Roads


Road Drainage Manual

Appendix 10A
Worked Examples

The average annual times of submergence are calculated as shown in the following, assuming
that the time of submergence for the PMF is 12 hours.

ARI (yrs)

ToS
(hrs)

FT(t)

fT(T)

Area

Area*ToS

0.500

0.667

0.800

0.033

0.133

0.267

10

0.900

0.100

0.100

0.450

20

0.950

0.050

0.050

0.275

50

0.980

0.030

0.030

0.195

100

7.5

0.990

0.020

0.010

0.073

1E+99

12

1.000

0.002

0.010

0.098

AATOS

1.357

End of Example

10
A

March 2010

10A-6

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Chapter 11
Road Surface
and Subsurface
Drainage Design

11

January 2013

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Chapter 11 Amendments Jan 2013


Revision Register
Issue/
Rev
No.

Reference
Section

11.2.10.7 &
11.2.11

11

January 2013

ii

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Minor update to figures and removal of Section


11.2.10.7 (c).

M
Whitehead

Jan
2013

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Table of Contents
11.1

11.2

11.3

Introduction .................................................................................... 11-1


11.1.1

Road Surface Flows ......................................................................... 11-1

11.1.2

Subsurface Flows ............................................................................. 11-2

Road Surface Drainage.................................................................. 11-2


11.2.1

Pavement Runoff.............................................................................. 11-2

11.2.2

Roadway Flow Width Criteria ........................................................... 11-2

11.2.3

Kerb and Channel Flow .................................................................... 11-7

11.2.4

Edge and Median Drainage.............................................................. 11-8

11.2.5

Table Drains and Table Drain Blocks ............................................. 11-11

11.2.6

Diversion Drains and Diversion Blocks........................................... 11-13

11.2.7

Batter Drains................................................................................... 11-14

11.2.8

Catch Drains and Catch Banks ...................................................... 11-15

11.2.9

Road Batter Stabilisation ................................................................ 11-16

11.2.10

Drainage Pits .............................................................................. 11-17

11.2.11

Access Chambers....................................................................... 11-26

11.2.12

Access Chamber Tops................................................................ 11-27

11.2.13

Reduction in Pipe Size................................................................ 11-27

11.2.14

Surcharge Chambers.................................................................. 11-28

11.2.15

Pipeline Requirements................................................................ 11-28

11.2.16

Splay Pipes ................................................................................. 11-29

11.2.17

Structural Requirements of Pipelines.......................................... 11-29

11.2.18

Flow Velocity Limits .................................................................... 11-29

11.2.19

Pipe Grade Limits ....................................................................... 11-29

11.2.20

Discharge Calculations ............................................................... 11-29

11.2.21

General Design Procedure ......................................................... 11-31

11.2.22

Hydraulic Calculations ................................................................ 11-31

Aquaplaning ................................................................................. 11-31


11.3.1

What is Aquaplaning?..................................................................... 11-32

11.3.2

Causal Factors ............................................................................... 11-33

January 2013

iii

11

Department of Transport and Main Roads


Road Drainage Manual

11.4

11

11.3.3

Road Surfacing............................................................................... 11-34

11.3.4

Tyres............................................................................................... 11-37

11.3.5

The Road-Tyre Interface ................................................................ 11-37

11.3.6

Skid Resistance.............................................................................. 11-39

11.3.7

Assessment Water Film Depth .................................................... 11-41

11.3.8

Assessment Aquaplaning Potential ............................................. 11-47

11.3.9

Quick Assessment.......................................................................... 11-48

11.3.10

Time of Flow ............................................................................... 11-49

11.3.11

Puddles / Wheel Ruts ................................................................. 11-49

11.3.12

Guidance to Reduce Aquaplaning Potential ............................... 11-50

Subsurface Drainage ................................................................... 11-51


11.4.1

Moisture in Roads........................................................................... 11-51

11.4.2

Control of Road Moisture................................................................ 11-53

11.4.3

Types of Subsurface Drainage ....................................................... 11-53

11.4.4

Requirements of Filter Materials..................................................... 11-57

11.4.5

Design Procedure........................................................................... 11-58

11.4.6

Location of Subsoil Drains.............................................................. 11-59

11.4.7

Transverse Subsurface Drains ....................................................... 11-62

11.4.8

Cut-off Drains ................................................................................. 11-62

11.4.9

Design of Cut-off Drains ................................................................. 11-62

11.4.10

Size of Drain ............................................................................... 11-63

11.4.11

Materials ..................................................................................... 11-63

11.4.12

Access to Subsurface Drains...................................................... 11-64

11.4.13

Lowering of Ground Watertable .................................................. 11-64

11.4.14

Schilfgaardes Method ................................................................ 11-64

11.4.15

Draining an Inclined Aquifer........................................................ 11-67

11.4.16

Design of a Filter Blanket to Lower a Water Table ..................... 11-68

11.4.17

Capillary Rise in Soils ................................................................. 11-68

11

January 2013

iv

Chapter 11
Road Surface and Subsurface Drainage Design

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Chapter 11
Road Surface and Subsurface
Drainage Design
11.1

Introduction

Road surface drainage deals with the


drainage of stormwater runoff from the road
surface and the surfaces adjacent to the road
formation. Several elements can be used to
intercept or capture this runoff and facilitate
its safe discharge to an appropriate
receiving location. These elements include:
kerb and channel;
edge and median drainage;
table drains and blocks;
diversion drains and blocks;
batter drains;
catch drains and banks;
drainage pits; and
pipe networks.
Subsurface drainage deals with the
interception and disposal of subterranean
(groundwater) flows with predominate
drainage element being sub-soil drainage.

11.1.1 Road Surface Flows


After falling onto road surfaces, rainfall
runoff drains to the lowest point and in
moving across the road surface forms a
layer of water of varying thickness. This
water can be a hazard to the motorist.
Splash and heavy spray are thrown up by
moving vehicles reducing visibility, whilst
the water on the pavement reduces friction
between the tyres and road surface.

Excessive water on the pavement, whether


ponded or flowing, can represent a real risk
of aquaplaning or the build-up of a layer of
water between the vehicle tyre and the road
surface which leads to a total loss of grip.
While part of road surface drainage,
aquaplaning is a critically important aspect
of road surface drainage and is discussed
within its own section, Section 11.3.
On reaching the lowest point, runoff is
channelled along the pavement edge via
kerbing / kerb and channelling or
discharged over the shoulders to a suitable
collection system such as a natural
watercourse, table drain, or piped drainage
system (pipe network).
Some degree of water quality treatment
may be needed between the road and the
receiving water to remove litter, heavy
metals, nutrients and oils. In this regard,
there is a growing trend to place some form
of grass filter between the road surface and
any concrete-lined drain. This form of
drainage is known as indirectly connected
impervious surface area and is a form of
Water Sensitive Urban Design (see Chapter
7).
In all cases, design of the elements for this
runoff must adequately cater for the safety
and convenience of road users, including
pedestrians and protect adjacent properties
and the road pavement from damage.
Where erosion of the batters is not
considered likely, pavement runoff
discharged over the shoulders and batters,

January 2013

11-1

11

Department of Transport and Main Roads


Road Drainage Manual

directly to the natural surface, may be


acceptable in some rural situations such as a
level stretch of road in flat country.
Where batter erosion is likely / possible, the
use of a concrete or asphalt kerb / dyke
should be investigated.

11.1.2 Subsurface Flows


Subsurface or subsoil drains are required to
intercept and drain excessive moisture or
groundwater flow in order to avoid
premature pavement failures. This moisture
can come from:
seepage or capillary rise from the
watertable (when in close proximity
to subgrade / pavement);
seepage from ponded stormwater into
embankment /pavement; and/or
seepage from an aquifer or other
groundwater flow.
The design and installation of subsurface or
subsoil drains beneath / adjacent to road
pavements is essential where groundwater
or seepage is known or considered to be
present.

11

It is important to note that the construction


of an underground stormwater drainage
system with associated granular pipe
bedding can result in the interception of
seepage and the concentration of this
intercepted water at drainage structures.
The installation of subsoil drains should be
considered in conjunction with the drainage
pipes to allow seepage water to be collected
and discharged into the drainage system.
Detailed recommendations in respect of
design and installation of subsoil drains
have been prepared by the Australian Road
Research Board, (ARRB 1987).

January 2013

11-2

Chapter 11
Road Surface and Subsurface Drainage Design

11.2

Road Surface
Drainage

11.2.1 Pavement Runoff


Pavement runoff is calculated by the
Rational Formula:

Q y k .C y .I tc , y . A (refer Chapter 5)
The contribution to the flow at the kerb or
median channel is given by a modification
of the Rational Formula and is expressed
as:

qy

C y .I tc , y .W

3.6 X 10 6

Where:
qy = contribution per longitudinal
metre of pavement (m/s) for an ARI
of y years;
=
runoff
coefficient
Cy
(dimensionless) for an ARI of y years
(refer Chapter 5);
Itc,y = average rainfall intensity
(mm/h) for design duration of tc and
ARI of y years (refer Chapter 5); and
W = width of contributing cross
section (m).
A runoff coefficient C50 of 0.95 (or higher)
is typical for most road surfaces.
Where the pavement width varies or the
runoff coefficient is different, then total
runoffs or lengths for given runoffs have to
be calculated algebraically.

11.2.2 Roadway Flow Width


Criteria
For the safety of vehicular traffic other than
requirements against aquaplaning, flow
width criteria apply. Water depths and
velocities are also limited by the width
restrictions.

Department of Transport and Main Roads


Road Drainage Manual

Flow widths in both the minor and major


storms need to be considered.
11.2.2.1 Minor Storm Flow Limits
Adopting the ARI 10 year flood as that
arising from the design minor storm of the
same average recurrence interval, flow
width criteria are shown on Figures
11.2.2.1(a) and 11.2.2.1(b).
These diagrams represent the following:
For two lanes (or more) in the same
direction plus parking lane, the
maximum allowable width of spread
leaves the inside and any lane-locked
lanes clear plus 2.5 m clear width in
the remaining lane i.e. water is kept
out of the wheelpaths of lanes. The
term lane includes auxiliary lanes and
any parking lane that has the
potential to become used as a full or
part time through lane.
For two lanes (or more) in the same
direction, the maximum allowable
width of spread leaves the inside and
any lane-locked lanes clear plus 2.5
m clear width in the remaining lane.
The term lane includes auxiliary
lanes.
For one lane plus parking lane, water
is not allowed to spread past the edge
of the through lane.
For one lane, a minimum clear width
of 3.5 m is to remain in the lane.
At medians, the allowable spread of
water leaves 2.5 m clear width in the
traffic lane next to the median. The
term lane includes auxiliary lanes.
At intersections without left slip
lanes, the allowable width of spread
adjacent to the kerb is 1.0 m.

Chapter 11
Road Surface and Subsurface Drainage Design

At intersections with single left slip


lanes, the allowable width of spread
leaves 3.5 m clear width in the slip
lane.
At intersections with dual left slip
lanes, the allowable width of spread
leaves 2.5 m clear in outer turning
lane.
In situations where it is difficult to achieve
the required clear width of 2.5 m in cases
(b), (e) and (h) above, the clear width may
be reduced to 1.0 m for roads of lesser
importance.
This practice is not
recommended for reasons of consistency
and the use of a reduced clear width must
be specified in design brief and/or contract
documents or approved by the department.
Where pedestrians will cross the road,
allow no more than 0.45 m width of spread
in an ARI 1 year flood. The 0.45 m
requirement is based on the typical over
step / short jump of most people. Checks
should also be undertaken on the flow
velocity. Where the risk of injury is
reasonably foreseeable, velocities should be
limited by:
dg.Vavg 0.4 m2/s.
Where:
dg = flow depth in the channel
adjacent to the kerb (m); and
Vavg = average velocity of the flow
(m/s).
There is also a water depth-velocity
relationship which is applicable for both
minor and major floods in the channel next
to a kerb. This is for pedestrian safety in
longitudinal flows along the kerb and is
shown on Table 11.2.2.2.

January 2013

11-3

11

Department of Transport and Main Roads


Road Drainage Manual

11

Chapter 11
Road Surface and Subsurface Drainage Design

Notes:
1. Lane includes auxiliary lanes and any parking lane that has the potential in the future to become used as a through
lane for full or part time.
2. In situations where it is difficult to achieve the required clear width of 2.5 m, the clear width may be reduced to 1.0
m for roads of lesser importance (refer text in Section 11.2.2.1).

Figure 11.2.2.1(a) - Allowable Flow Widths on Roadways ARI 10 year Flood (crosssection views).

January 2013

11-4

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Notes:
1. Refer to Figure 11.2.2.1(a).
2. In situations where it is difficult to achieve the required clear width of 2.5 m, the clear width may be reduced to 1.0
m for roads of lesser importance (refer text in Section 11.2.2.1).
3. At pedestrian crossings check both width and velocity (refer text in Section 11.2.2.1).
4. See Section 11.2.2.2 for allowable widths in Major Storms.

Figure 11.2.2.1(b) - Allowable Flow Widths on Roadways ARI 10 year Flood (plan views)

January 2013

11-5

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

11.2.2.2 Major Storm Flow Limits


The major storm is usually of an ARI of 50
to 100 years, depending on the local
authority. The ARI 100 year flood should
be used as a check flood at least, to allow
consideration of any detrimental effects.
Table 11.2.2.2 gives roadway flow limits
for a major storm, with particular reference
to floor levels of adjacent buildings,
pedestrian and vehicle safety.

At sags in State-controlled roads, additional


inlets and underground drainage should be
provided, if necessary, to limit ponded
water in an ARI 50 year storm so that there
is:
one lane in each direction of travel,
free of water, in a multi-lane road; or
a width of 3.5 m clear of water down
the centre of a two-lane road.

Table 11.2.2.2 - Roadway Flow Limitations - Major Storm


Situation
Where floor levels of adjacent
buildings are above road
level.
Where floor levels of adjacent
buildings are less than 350
mm above top of kerb.
- where fall on footpath
towards kerb is greater than
100 mm;
- where fall on footpath
towards kerb is less than 100
mm;
Where no kerb is provided.

Pedestrian Safety #
(a) no obvious danger
(b) obvious danger

Roadway Flow Width and Depth Limitation


Total flow contained within road reserve.
Peak water levels at least 300 mm below floor level of
adjacent buildings (i.e. freeboard of at least 300 mm)

Water depth to be limited to 50 mm above top of kerb.

Water depth to be limited to top of kerb in conjunction with a


footpath profile that prevents flow from the roadway entering
onto the adjacent property.
Above depths shall be measured from the theoretical top of
kerb.

dgVavg 0.6 m2/s


dgVavg 0.4 m2/s

Vehicle Safety

11

Maximum energy level of 300 mm above roadway surface for


areas subject to transverse flow.
Notes: dg = flow depth in the channel adjacent to the kerb ie. at the invert (m)
Vavg = average velocity of the flow (m/s)
#
Obvious danger is interpreted as areas where pedestrians are directed to or most
likely to cross water paths (such as marked crossings and corners of intersections).
Note: Table aligns with requirements within QUDM (2008).

January 2013

11-6

Department of Transport and Main Roads


Road Drainage Manual

11.2.3 Kerb and Channel Flow


The cross section of flow in channels is
basically triangular and the actual width of
flow is governed by how far the water is
allowed to encroach on to the pavement.
The adopted channel cross section is shown
in Figure 11.2.3.

Chapter 11
Road Surface and Subsurface Drainage Design

ng = Mannings roughness of kerb


(refer Table 11.2.3);
np = Mannings roughness
pavement (refer Table 11.2.3);

of

dg = depth of flow at kerb invert (m);


dp = depth of flow at edge of
pavement (m);
dc = depth of flow at crown (m); and
S = longitudinal grade (m/m).
The normal form of the Izzard formula can
also be used for shallow v-shaped channels
where:

Source: QUDM (2008)


Figure 11.2.3 - Channel Profile for Flow
Analysis

The flow in a triangular channel is given by


the Izzard formula:

2.667
Q 0.375 F g S 0.5 d g
n

For composite flow as shown in Figure


11.2.3 where the pavement and channel
have different roughness and crossfall, the
formula becomes:
Q = 0.375.F.[(Zg/ng).(dg2.667- dp2.667) +
(Zp/np).(dp2.667- dc2.667)].S 0.5
Where:
Q = longitudinal flow in channel
(m/s);
F = flow correction factor (assume
0.9 triangular shaped flows);

Zg

Surface _ width
Depth _ at _ invert

Table 11.2.3 - Mannings Roughness


Coefficient - Flow in Triangular Channels
Channel lining
Concrete gutter (trowelled
finish)

n
0.012

Asphalt pavement:
smooth texture
rough texture

0.013
0.016

Concrete pavement:
float finish
broom finish

0.014
0.016

Brick and Pavement Blocks

0.016

Gutter with vegetation and


cracks

0.020

Sprayed Seal

0.018

11

Zg = cross slope gradient of kerb (part


of ratio);
Zp = cross slope gradient of pavement
(part of ratio);

January 2013

11-7

Department of Transport and Main Roads


Road Drainage Manual

11.2.3.1 Kerb Types and Uses

11.2.4 Edge and Median


Drainage

Approved kerb types with channels are


shown in Standard Drawing 1033 (DMR
2009b). General kerb profiles or shapes are
shown in Figure 11.2.3.1(a).

In choosing the type of channel to be


adopted, consideration should be given to
the following factors:

The use of a concrete or asphaltic concrete


kerb at the edge of embankments, as shown
in Figure 11.2.3.1(b), is justified if, the
material forming the embankment will be
eroded by flow off the pavement or if for
property protection, it is necessary to
restrict runoff to particular locations.
Temporary protection must be provided
until the slope is completely stabilised by
grass.
The criteria for flow in the roadway due to
shoulder kerb treatment are as per Section
11.2.2. Flow calculations are by the same
method as Section 11.2.3.
The actual dimensions of a shoulder kerb
will be job specific. Installation of a
shoulder kerb / dyke could be a danger to
vehicles running off the side of the
carriageway. If safety is compromised,
other forms of erosion protection control
will have to be considered.

11

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.2.3.1(b) - Edge Treatment at


Erodible Slopes

January 2013

11-8

Capacity: The channel should have


adequate capacity for the design
flow.
Erosion: Erosion control is a
necessary part of good drainage
design. Scour may occur unless the
channel is protected where velocities
exceed those likely to cause erosion
to the material forming the channel.
Erosion control involves the selection
of a suitable and economical channel
lining (including vegetative cover)
which will give the desiredprotection.
For further information on erosion control
see Chapters 7, 8 and 13. The type of
lining should be consistent with the degree
of protection required, overall cost
including maintenance, safety requirements
and aesthetic considerations.
Erosion
control in the form of grass growth may be
used in combination with other types of
lining. A channel may be grass lined on the
flatter slopes and lined with more resistant
material on the steeper slopes.
Maintenance:
Without
proper
maintenance, a well designed channel
becomes unsightly and will perform
unsatisfactorily at the design flow.
Maintenance methods should be
considered in the design of drainage
channels so that the type of channel
section adopted will be suitable for
the methods and equipment that will
be used-for-maintenance.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.2.3.1(a) - General Kerb Profiles or Shapes

For example, in the majority of cases a


concrete invert should be considered for V
shaped channels because of the difficulty in
maintaining the section with the
maintenance machinery available.
On
grades less than 1.5% and for most cross
sections, a concrete invert is essential in
assisting the discharge of low flow.
Maintenance operations are extremely
difficult with saturated conditions in the
vicinity of an unlined invert.
11.2.4.1 Design of Edge and Median
Drainage
It is preferable to allow the pavement runoff
to discharge across the shoulder and into
the channel but where the batter slope is
steep due to restricted width and erosion of
the slope is likely, a kerb should be
provided.
However, on high speed
carriageways, kerbs or steep batter slopes
should be avoided for safety reasons.
The pavement runoff is calculated using the
Rational Method formula (refer Section
11.2.1), but consideration must be given to
the different runoff coefficients for
pavement and the median surfaces.
The maximum velocities allowed in grass
channels so as to prevent erosion are given
in Table 8.8.1.3 and the minimum velocity
to prevent silting as a general rule should
not fall below 0.5 m/s.

Where a grassed channel is the preferred


treatment, the channel must have sufficient
hydraulic capacity to prevent spread of the
water onto the through pavement and also
the bottom of the channel should desirably
be at a lower elevation than the pavement
sub-base in cuts, otherwise protection may
be necessary by subsoil drains. Grassed
channels are further discussed in Chapter 8.
In very narrow and steep medians, concrete
drains may be required to provide sufficient
capacity. The hydraulic design of concrete
drains is outlined in Chapter 8.
In very restricted areas, special drainage
inlets and conduits may have to be
provided.
The inlets used to intercept flow in the
median should desirably be set flush with
the ground so as not to be hazardous to
mowing equipment and vehicles out of
control, and the bars set parallel to the flow.
A concrete apron should be used around the
inlet to accelerate the flow into it and to
prevent the grate from being overgrown by
vegetation.
If a slightly raised grated inlet with the top
above natural surface is adopted to
minimise debris collection, concrete aprons
should still be provided to allow mowing.

January 2013

11-9

11

Department of Transport and Main Roads


Road Drainage Manual

11
Figure 11.2.4.1 - Median Inlets with Levees.

January 2013

11-10

Chapter 11
Road Surface and Subsurface Drainage Design

Department of Transport and Main Roads


Road Drainage Manual

Complete interception of all median flow at


each inlet gives the most efficient design.
On grades, a levee across the median just
downstream of the inlet creates a sag
condition and ensures complete interception
(see Figure 11.2.4.1).
The levee should be of just sufficient height
to maintain the design head over the grate,
and have slopes 1 on 6 or flatter (1 in 20 is
desirable) for safety of vehicles out of
control. The height of the levee is usually
that of the ARI 10 year flood. Furthermore,
designers need to also ensure that any
overtopping of the levee will not force
stormwater onto the road surface. Where
this cannot be achieved, the road
geometrics may need to be reviewed /
modified.
The design procedure for inlets is given in
Section 11.2.8. The allowable depth of
ponding is determined as the depth to
contain the flow within the median.

11.2.5 Table Drains and Table


Drain Blocks
Table drains are located on the outside of
shoulders in cuttings or alongside shallow
raised carriageways in flat country. For the
location and cross sectional requirements of
table drains see relevant chapter within the
departments Road Planning and Design
Manual - A guide to QLD practice and
Austroads Guide to Road Design series..
Also, refer to Section 2.5.7 for the general
requirements of table drains and Standard
Drawing 1178 (DMR 2009b) for general
location and arrangement.
11.2.5.1 Table Drains
Flat-bottomed table drains are the preferred
type or shape. Refer Figure 11.2.5.1 for
general location and shape details. The base

Chapter 11
Road Surface and Subsurface Drainage Design

of the drain is not flat as the name suggests,


but is to be sloped away from the
carriageway (at least 3%). These drains:
spread the flow, reducing depth and
velocity (reduced scour potential);
keep flow and any ponding away
from the embankment;
allow easier access for maintenance;
and
allow safer passage through the drain
for errant vehicles.
As discussed in Section 2.5.7.1, flatbottomed drains are the preferred type or
shape. The use of V drains is to be
limited / confined to constrained sections.
For some projects, flat-bottomed table
drains can be used as a source of borrow
material.
Determination of depth and velocity of flow
within the table drain can be undertaken
using Mannings Equation.
The invert of the table drain should be at
least 150 mm below the bottom of the
pavement (i.e. subbase), and deeper where
subsurface drains discharge into the table
drain. Furthermore, the design water level
in the table drain should be below the
subgrade level of the road pavement. The
depth of the drain will depend upon the
design capacity required to safely convey
the stormwater.
Where the material between the table drain
and the road pavement is impermeable, the
table drain may flow up to a level which
provides 150 mm freeboard against
overtopping, or 150 mm below the level of
the outer shoulder, provided that this does
not flood any subsurface drain outlets.

January 2013

11-11

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Slope

Machine
Width

Figure 11.2.5.1 Table Drain Location and Shape

Extreme care is required on very erodible


soils or soils with highly dispersive
subsoils, as exposure of such material can
have disastrous consequences. Where scour
is likely because of the nature of the
material or because of the longitudinal
grading, some type of protection of the
drain invert is required. This protection
could take the form of rock lining or
concrete. In areas known to have dispersive
subsoil, soil chemistry should be analysed
to determine whether soil properties can be
economically improved to aid soil stability.
The terminal treatment at the bottom of a
steep table drain is also important so as to
dissipate the energy of the stormwater.
11.2.5.2 Table Drain Blocks

11

On vertical grades (excluding sags), cross


drainage culverts can be used to take table
drain flow from the upstream side of the
road formation to the downstream side. In
order for these culverts (generally of small
diameter) to reasonably operate, sufficient
head must be generated.
Table drain blocks are small earth
embankments (typically 1 m in length)
located within the table drain just
downstream of a culvert inlet. These blocks

January 2013

11-12

dam and divert the flow into the culvert


inlet. The design height of the block should
theoretically coincide with the depth of
headwater required for efficient operation
of the culvert, however the design height
should be no higher than where the
subgrade meets the side slope. Designers
need to also ensure that any overtopping of
the block will not force stormwater onto the
road surface.
The use of a table drain block for a
particular culvert must be specified on the
drainage plans and include its design
height.
A marker post (refer Standard Drawing
1358 (DMR 2009b)) should be placed on /
adjacent to table drain blocks to alert
maintenance personnel of their existence.
Designers of widening and overlay /
rehabilitation projects should check
previous project documents for use of table
drain blocks in order to check / preserve
existing culvert operation.
11.2.5.3 Adjoining Projects
Another important design aspect regarding
table drains is where new and previous
projects join. The impacts of discharge

Department of Transport and Main Roads


Road Drainage Manual

from one project to the next must be


considered / incorporated. The shape and
grade of adjoining table drains should also
match. Where the shape and grade do not
match, a suitable transition or other
mitigating treatment must be designed to
ensure that scour does not occur and/or that
stormwater does not flow out onto the road
surface.

11.2.6 Diversion Drains and


Diversion Blocks
On vertical grades, the depth and velocity
of flow within a table drain can exceed
acceptable limits if no relief mechanism is
included in the design. The steeper the
grade, the quicker the limits are reached.
Diversion drains are required, at intervals
based on the grade of the road, to turn the
stormwater out of the table drain and away
from the road. On the upstream side of the
road, the diversion drain could either
discharge flow into a catch drain located
roughly parallel to the road and generally
near the boundary (drain discharges into a
culvert located further down the slope) or
onto the natural surface.
On the
downstream side of the road, the diversion
drain would normally discharge onto the
natural surface, allowing the stormwater to
naturally flow away from the road.
Normally, the shape of the diversion drains
should match that of the table drain,
however where V shaped table drains are
used, flat-bottomed diversion drains should
be used unless conflicting requirements
exist that limit the width of the formed road
such as:
minimising disturbance or loss of
existing roadside trees, and/or

Chapter 11
Road Surface and Subsurface Drainage Design

The channel of the drain should have an


excavated cross-sectional area at least equal
to that of the table drain discharging water
into it. Outlets should discharge onto
grassed areas rather than onto exposed soil.
They should be constructed either before, or
during the early stages of any road
construction.
Diversion blocks are earth embankments
that are located along the downstream side
of diversion drains and divert stormwater
flow from the table drain into the diversion
drain. The block extends from the road
formation for several metres along the side
of the diversion drain.
The height of the block typically starts
(against the road formation) at subgrade
level and should taper down to closely
match the natural surface at the end of the
block.
Refer Standard Drawing 1178 (DMR
2009b) for the general location and
arrangement of diversion drains and blocks.
Diversion drains and block are typically
constructed at a 45 angle to the road
however this can be adjusted in order to
reduce the slope within the diversion drain.
The initial grade (turnout grade) in the
diversion drain should approximate the
grade of the table drain to avoid energy
loss, and hence siltation and possible bank
failure. As the drain increases in length, the
grade in the drain should progressively
decrease.
Where the diversion drain
discharges onto the natural surface, the last
half of diversion drains length should
ideally have a surveyed grade of 0.2% to
slow the flow velocity.

reducing earthworks costs.

January 2013

11-13

11

Department of Transport and Main Roads


Road Drainage Manual

Recommended spacing of diversion drains


along the grade are:
120 m for slopes up to 2%;
60 m for slopes from 2% to 4%;
30 m for slopes from 4% to 8%; and
15 m for slopes greater than 8%.
If the diversion channel is built through /
under a fence (with landholder written
agreement / approval), it is essential that
landholders ensure that the outlet is kept
clear to allow free draining of the drain and
that departmental maintenance crews are
allowed easy access via an adjacent gate.
Depending on the fence type and
configuration, the impact of the fence on
channel flows will need to be considered.
Discharge from a diversion channel may
also be spread over a pasture to assist grass
growth.
In areas with known dispersive
subsoils, diversion banks (that extend
from the diversion block) are
preferred over excavated diversion
drains.
Where new and previous projects
join, designers are required to
appropriately consider the use and
effects of diversion drains and blocks
similar to the requirements of Section
11.2.5.3.

11.2.7 Batter Drains

11

Batter drains or chutes are structures that


are designed to convey runoff down the
face of a cut or fill batter and discharge at
either non-erosive velocities or onto a nonerodible surface.
Batter drains can be both permanent and
temporary. Temporary batter drains are
used during the construction period to

January 2013

11-14

Chapter 11
Road Surface and Subsurface Drainage Design

control water flow and to protect mulched


or newly seeded batters from the erosive
forces of concentrated flow.
11.2.7.1 Design Procedure
A four step process is provided for the
design of batter drains.
Step 1 - Dimensions
Batter drains should have a minimum depth
of 300 mm.
Hydraulic capacity of a batter drain is
normally defined by the allowable head
water level upstream of the drains inlet.
Step 2 - Foundations
The lining of the batter drain should be
adequately anchored to the foundations to
avoid slippage or separation, with a
maximum distance of 3 m between
anchorage points.
In some cases, prefabricated units may need
to be bolted together. It is important that all
bolt holes are sealed with a flexible sealant
to allow for flexural movement.
Step 3 - Inlet Design
The inlet area should be protected against
possible scour resulting from accelerating
flow velocities (usually more important on
temporary batter drains). This is to prevent
water from either undermining the top of
the batter drain, or being diverted along the
edge of the lining (the most common cause
of failure).
For temporary batter drains (i.e. drainage
chutes) and during the early revegetation
stage of permanent batter drains, sand /
gravel bags can be used to direct inflow
towards the centre of the chute.

Department of Transport and Main Roads


Road Drainage Manual

Step 4 - Outlet Design


An energy dissipator will be required.
The outlet may consist of a bed of nominal
150 mm rock (minimum) placed with a
minimum bed thickness of 250 mm or at
least 1.5 times the maximum rock size.

Chapter 11
Road Surface and Subsurface Drainage Design

Catch drains and/or catch banks can also


protect embankments, disturbed areas and
stockpile sites from surface water.
These devices are generally located no
closer than 2.0 m from the edge of the
cuttings in order to minimize possible
undercutting of the top of the batter.

Typical dimensions of the rock bed are:


L = 6De metres long (minimum); and
W = T + 0.6 metres wide at the batter
drain outlet, expanding to T + 0.5L + 0.3
metres at the end of the dissipator.

Figure 11.2.8(a) Catch Drain

Where:
De = equivalent pipe diameter (m) of
the batter drain flow area, and
T = top width (m) of flow in the
batter drain.
Figure 11.2.8(b) Catch Bank

11.2.7.2 Design Notes


Batter drains should be lined each side with
a minimum 300 mm wide (turf) grassed
filter strip or rock to control side erosion
caused by splash. In areas where the total
disturbance is to be minimised, or where
introducing turf is undesirable, other forms
of erosion control such as geotextiles or
concrete may be preferred.

11.2.8 Catch Drains and Catch


Banks
Catch drains and catch banks are separate
devices, but can be used together. They are
generally located on the high side of
cuttings clear of the top of batters to
intercept the flow of surface water and
upper soil seepage water (refer Figures
11.2.8(a) & (b)). Their purpose is to
prevent overloading of the table drain and
scour of the batter face.

Catch banks are sometimes used instead of


drains to reduce effects of seepage on
stability of the batter slopes and also to
minimize disturbed ground (increasing
scour potential). However use of catch
banks alone means importation of suitable
embankment material.
11.2.8.1 Design Procedure for Catch
Drains
The design of catch drains and banks is
based on the same methodology as for open
channels (refer design process outlined in
Chapter 8). A freeboard of at least 150 mm
should be used.
Where site conditions or some other
constraint restricts the construction of a
channel
with
suitable
dimensions,
supplementary channel treatments such as
synthetic channel linings, rip-rap or
concrete inverts should be utilised to
withstand the higher velocities. Figure

January 2013

11-15

11

Department of Transport and Main Roads


Road Drainage Manual

11.2.8.1 shows a typical concrete lined


catch drain.
Flow discharging from the catch drain
should not be allowed to cause or aggravate
erosion. Flow from a catch drain may be
discharged
to
an
existing
drain,
watercourse, or to a chute.

Figure 11.2.8.1 - Concrete Lined Catch


Drain

11.2.9 Road Batter Stabilisation


An integral part of surface drainage is the
design of erosion control measures to the
finished surfaces.
The downstream face of an elevated road
embankment across a floodplain will need
to be protected from the erosive forces that
occur during overtopping. The extent of
this protection will depend on the expected
tailwater level at the commencement of
overtopping. (refer Chapter 10).

11

Some roads are designed to be overtopped


and the erosive potential is not as severe as
those for floodways.
Rather than
overtopping, rainfall, wind and runoff from
the top of the road formation (unless kerbs
are in place) are the key agents for erosion.
Designers need to develop and assess
options and determine the most suitable
solution for the situation.
The types of protection measures available
for consideration are:

January 2013

11-16

Chapter 11
Road Surface and Subsurface Drainage Design

Revegetation;
Biodegradable blankets in association
with permanent revegetation (the
relevant specification is MRTS03
(TMR 2010c));
Chemical surface stabilisers and soil /
cement treatment (the relevant
specifications are MRTS03 and
MRTS16 (TMR 2010c));
Bank
protection
techniques
commonly associated with bridge
abutments (relevant standards are
MRTS03 and MRTS16 (TMR
2010c) and Standard Drawing 1117
(DMR 2009b) which also shows
other types of protection for bridge
abutments);
Benching to create permanent
drainage lines to reduce surface
drainage (relevant specifications are
MRTS03, MRTS04 and MRTS16
(TMR 2010c));
Kerbs at the top edges of the road
formations diverting runoff from
rainfall on the pavement and
shoulders to batter chutes (relevant
standards are MRTS03 (TMR 2010c)
and Standard Drawing 1033 (DMR
2009b));
Catch drains and catch banks used to
divert water to batter chutes or
completely away from the batter
slope (relevant standards are
specifications
MRTS03
and
MRTS16 (TMR 2010c) and Standard
Drawing 1178 (DMR 2009b)); and
Proprietary batter chutes.

Department of Transport and Main Roads


Road Drainage Manual

11.2.10 Drainage Pits


Drainage pits are field inlets and gullies
collecting surface flows to the underground
drainage system and access chambers at
pipe junctions and for maintenance.
Inlet locations should be optimised to
collect the design surface flows with the
minimum number of installations and, of
course, to reduce surface water to an
acceptable width.
This requires computations for each area
contributing flow to the inlets. Areas may
comprise both road pavement and adjacent
urban, suburban or rural land.
Proprietary pre-cast pit segments and grates
or covers are available. The department has
standardised some of the more common
field inlets, gullies and access chambers as
Standard Drawings Nos.1307, 1308, 1309,
1310, 1311, 1312, 1313, 1321, 1322, 1442,
1443, 1444, 1445 and 1449 (DMR 2009b).
Four types of kerb inlets are in common
use, they are:
grate only e.g. field inlets and antiponding gullies on kerb returns;
side inlet these inlets rely on the
ability of the opening under the
backstone or lintel to capture flow.
They are usually depressed at the
invert of the channel to improve
capture capacity;
combination grate and side inlet
these inlets utilise the backstone
arrangement of the side inlet with the
added capacity of a grate in the
channel; and
special site specific designs for high
inflow.
The capacity of the various categories of
drainage inlet may be varied by the amount

Chapter 11
Road Surface and Subsurface Drainage Design

of depression allowed in the gutter adjacent


to the kerb opening.
A flush inlet is one in which the normal
channel section is continued to and past the
inlet without any alteration to its cross
section.
A depressed inlet is one in which the
crossfall of the channel is increased, so that
the grade of the channel line against the
kerb is depressed for the length of the inlet.
Depressed inlets provide greater efficiency
than flush ones and are shown on the
standard drawings with suitable transitions.
All pits should be as shallow as practical.
As indicated on standard drawings, pits
deeper than 3 metres will require a special
design.
This Manual does not include inflow
capacity charts for drainage pit / kerb inlets.
Charts approved for use on departmental
projects are:
Brisbane City Council charts,
available from Councils website;
and
Max Q charts for Drainway and
Stormway products, available from
Max Q and as printed in Volume 2
of QUDM (2008).
It is understood that there are pit / kerb inlet
configurations currently available that do
not exactly match any of the configurations
as presented in the above approved charts.
In these situations, the designer can use
engineering judgement and first principles
to match, as close as possible, an approved
inflow capacity chart to the proposed pit /
kerb inlet (citing: opening area, grade,
crossfall and approach flow), however the
selected chart must be accepted / approved
by the departments design representative
before use.

January 2013

11-17

11

Department of Transport and Main Roads


Road Drainage Manual

Charts for other configurations / types of


pits may become available in the future.
Such charts should reflect the theoretical or
measured capacity of the inlet. Before use
on departmental projects, the supplier of the
charts must have them independently tested
/ verified and then submit them (incl
verification) for approval to Principal
Engineer (Road Design Standards), Road
Planning & Design Section, Engineering &
Technology Division.
11.2.10.1 Provision for Blockage
Blockage of drainage pits (either at the
grate or within the system) needs to be
considered as it reduces the capacity of the
underground pipe network (part of the
minor system) and therefore increases the
flow within the major system (refer Section
2.5.4).
This generally increases flood
levels.
Table 11.2.10.1 indicates the
percentage of blockage that is to be applied
to the theoretical inflow capacity of inlets.
11.2.10.2 Kerb Inlets in Roads
Kerb or gully inlets are used where
vehicular traffic is expected, to reduce the
flow width on roadways as well as to drain
low lying areas.
The standard departmental Concrete Gullies
shown on Standard Drawings Nos 1311 and
1312 (DMR 2009b) have a combined inlet
with a precast side entry (lintel) and grated
pit. Precast lintel details are shown on
Standard Drawing No. 1313 (DMR 2009b).

11

In general, kerb inlets should be provided /


considered at the following locations in
kerb and channel:

January 2013

11-18

Chapter 11
Road Surface and Subsurface Drainage Design

(a) In the low points of all sags in kerb and


channel.
(b) On grades, to ensure compliance with
the flow width limitations discussed in
Section 11.2.2.
(c) At the tangent point of kerb returns or
small radius convex curves (kerb radius
less than 15 m) such that the flow
width around the kerb return (i.e.
beyond the kerb inlet) during the Minor
Design Storm does not exceed 1.0 m
measured from the invert of kerb and
channel. This limitation will also be
applicable at important vehicular
turnouts or footpath crossovers, where
high traffic volumes are anticipated,
such as at entrances to shopping
centres.
(d) Immediately upstream of a potential
pedestrian crossing, set-down point
and/or bus stop such that the flow
width does not exceed 450 mm from
invert of kerb and channel during the
Minor Design Storm.
(e) Immediately upstream of any reversal
of crossfall (for example, application of
superelevation) to prevent flow across
the road during the Minor Design
Storm. The extent to which such flow
onto the pavement is permissible
depends upon the catchment area
involved and the risk of vehicle
aquaplaning.
The question of
aquaplaning is addressed in Section
11.3.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Table 11.2.10.1 Provision for Pit Blockage


Percentage of Theoretical
Capacity Allowed

Condition

Inlet Type

Sag

Kerb inlet
Grated
Combination

80%
50%
[See Note 2]

Continuous Grade
(On-Grade)

Kerb inlet
Longitudinal bar grated
Transverse bar grate or longitudinal
bar grate incorporating transverse
bars
Combination

80%
60%
50%
90% [See Note 3]

Notes:
1. This table does not prevent the setting of project specific, alternative blockage factors for site specific inlet
designs. Alternative factors must be specified in design / contract documentation.
2. In a sag, the capacity of a combination inlet should be taken to be the theoretical capacity of the kerb opening,
the grate being assumed to be blocked.
3. On a continuous grade the capacity of a combination inlet should be taken to be 90% of the combined theoretical
capacity of the grate plus kerb opening.

(f) Where superelevation or reverse


crossfall results in flow against traffic
islands and medians. Kerb inlets shall
be provided along the length of the
island or median as necessary to meet
the flow width limitations as stated in
Section 11.2.2 and at the downstream
end of the island or median to minimise
the flow continuing along the road.
Where sufficient width of island or
median is available, grated kerb inlets
should be recessed so that the grate
does not project onto the road
pavement. Alternatively side entry
inlets with no grate should be installed.
(g) Where it is anticipated that a parking
lane may become an acceleration,
deceleration or turn lane in accordance
with Section 11.2.2.
(h) Consideration should be given to the
positioning of kerb inlets relative to the
side property boundaries. In residential
and industrial locations, a kerb inlet
located near the side property boundary
may cause difficulties with driveway

access. In commercial areas and those


where there is likely to be a high
volume of pedestrian traffic, kerb inlets
should be located to avoid set down
points or locations where pedestrian
movements are likely to be highest.
(i) On any higher abutment end of bridge
approaches on a grade to minimise
flow on to the deck.
For kerb inlets on grade;
Kerb inlet capacity is controlled by
the crossfall of the road pavement
and the longitudinal grade.
Bypass flow from a kerb inlet must
be accounted for in the design of the
downstream kerb inlet which
receives the bypass flow. There is no
limitation to the amount of flow
which may be bypassed from a kerb
inlet provided that the flow width
criteria discussed in Section 11.2.2
are satisfied. Note that a number of
road flow capacity calculations may
be required, using actual crossfalls at

January 2013

11-19

11

Department of Transport and Main Roads


Road Drainage Manual

the intersection, to check that all


bypass flows are contained within the
1.0 m flow width limitation at kerb
returns, under Minor Storm Flow
conditions.
Where bypass flow from a kerb inlet
is required to follow a kerb return at
an intersection it may be necessary,
where the longitudinal grade is steep,
to check for the effect of

11

Chapter 11
Road Surface and Subsurface Drainage Design

superelevation upon flow spread. A


procedure for the calculation of
superelevation is given in Chapter 8.
The procedure detailed in Figure
11.2.10.2(a) is recommended for
determining the location of kerb
inlets on grade.

Figure 11.2.10.2(a) Procedure to Determine Kerb Inlet Positions on Grade


Notes:
1. Changes in catchment area may result in changes in time of concentration for a catchment.
2. The above procedure is iterative.
3. Selection of the initial trial kerb inlet location may be based on changes in road grade (e.g. steep to flat),
physical restrictions in road (e.g. median or Residential Street Management devices), or by driveways, entrances or
intersections etc.
Source: QUDM (2008).

January 2013

11-20

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Kerb inlets in sags must have sufficient


inflow capacity to accept the total flow
(including bypass flows from upstream)
reaching the inlet. Ponding of water at sag
inlets should be limited to the widths
discussed in Section 11.2.2 particularly at
intersections where turning traffic is likely
to encounter ponding water.
Where the longitudinal grades on either side
of the sag are different, or where the flow
from one direction is dominant, the location
of the effective sag may move from the true
sag and a hydraulic jump may form beyond
the sag. Care should be taken, by the
provision of extended or additional inlets,
to ensure that capture capacity is
maintained and that the water level does not

cause flow over the footpath into the


adjacent property.
A procedure for
checking whether this effect is occurring
has been proposed by Black (1987) and is
detailed in Figures 11.2.10.2(b) and
11.2.10.2(c).
At intersections;
Consideration shall be given to the
steepness of grade of the road and the
possibility of momentum carrying water
past the stormwater inlet(s), across the road
and
into
properties
opposite
the
intersection. Solutions to such problems
may require extra inlets to be installed.

HJ = Hydraulic Jump

Figure 11.2.10.2(b) Sag in a Road with Supercritical Approach Flows


Source: QUDM (2008).

11

January 2013

11-21

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.2.10.2(c) Limiting Condition for a Sag Inlet to Act as an On-grade Inlet
(n=0.013)
Source: QUDM (2008).

Where two falling grades meet at an


intersection, every endeavour should be
made to locate the low point of the kerb and
channel at one of the tangent points of the
kerb return.
Where both grades are steep it may not be
practical to locate the low point at a tangent
point. In this case, kerb inlets should be
provided at both tangent points, with
additional inlets provided upstream of the
tangent points, if necessary, designed to
limit the flow width beyond the kerb return.
An anti-ponding kerb inlet (grate only)
installed within the width of the channel nominally 450 mm long by 300 mm wide
with no kerb inlet should be provided at the
low point.

11

The location of a kerb inlet, or a grated inlet


that protrudes onto the pavement within a
kerb return is considered unsatisfactory
because of the risk of damage by and to
vehicles.
11.2.10.3 Opening Size for Kerb
Inlets
Drainage pit inlets can present an important
safety issue that designers must consider.

January 2013

11-22

Considerable debate exists regarding the


recommended maximum clear opening for
kerb inlets to provide safety for small
children. Even though past history has
shown the likelihood to be low, the
consequences of a child being swept down
a flooded kerb and into a stormwater inlet
can be extreme.
After consideration of the various
arguments presented to the QUDM
Reference Group during the development of
the 2008 release of QUDM, the
recommendation for 125 mm maximum
clear opening was accepted.
The
Department of Transport and Main Roads
also accepts this recommendation.
It
should be noted the 125 mm opening still
presents a risk of a small child partially
entering (i.e. feet first) the inlet.
A maximum clear opening of 88 mm is
required where it is necessary to exclude
the entry of the torso of child (based on test
procedures in AS 4685.1-2004). Such
consideration applies in parks, schools and
childcare centres.

Department of Transport and Main Roads


Road Drainage Manual

11.2.10.4 Design Loads


Australian Standard, AS 3996-2006
specifies design loads for access covers,
road grates and frames. They are to be
designed to support, without structural
failure, the specified minimum ultimate
limit state design loads.
The Class D loading is used where normal
vehicular traffic (includes heavy duty
commercial vehicles) may be expected.
The standard departmental access chamber
tops and roadway gullies are designed for
this loading.
Class B loading is used for units designed
for a footway loading.
Class C loading is used for units in
locations where slow moving (light duty
commercial) vehicles are expected such as
light maintenance vehicles (light trucks and
driven
grass
cutters
/
mowers).
Departmental field inlets, Types 1 and 2 are
in this category.
11.2.10.5 Bicycle Safe Covers and
Grates
AS 3996-2006 also specifies two test
wheels with pneumatic tyres to ensure the
covers and grates are bicycle safe.
At all times where there is a possibility of
bicycles, the designer should ensure that
bicycle safe covers and grates are specified.
11.2.10.6 High Efficiency Grates
The term high efficiency hydraulic grates
refers to non-bicycle safe grates with wider
bar spacing (the bar may be a flat
section).
As indicated on the drawings, such grates
should only be used in locations where
bicycles are prohibited or not likely to have

Chapter 11
Road Surface and Subsurface Drainage Design

access. The opening between the flat


sections should not exceed 52.5 mm.
Of a different design, the high efficiency
vaned grate with water deflectors shown on
Standard Drawing Nos 1321 and 1322
(DMR 2009b) is bicycle safe.
11.2.10.7 Field Inlets
Field inlets are used to drain low lying areas
and located in areas where vehicular traffic
would not be expected (except for
maintenance). Such locations are medians,
drainage easements, table drains and catch
drains.
Entry of water is from the top only in Field
Inlets Type 1 (Standard Drawing No. 1309
(DMR 2009b)).
The frame and grate of Field Inlet Type 2
(Standard Drawing No. 1310 (DMR
2009b)) is raised to allow side entry of
water as well to a depth of 175 mm before
water reaches the top of the grate. It is,
therefore, more efficient than the Type 1
Inlet and less prone to obstruction.
However, it should only be used where the
possibility of pedestrians, bicycles and
vehicular traffic is remote.
The following discussion has been
extracted from QUDM (2008) with some
minor modification.
Field inlets (also known as drop inlets)
should be provided in footpaths and
medians etc. as necessary, to drain all low
points.
Where there is considerable pedestrian
traffic adjacent to a field inlet e.g. in a
footpath, a grate with close bar spacing
should be used - recommended bar spacing
is provided in section (d) below. Elsewhere
a grate with wide bar spacing is preferable,

January 2013

11-23

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

because of the reduced risk of blockage by


debris.
In all situations an allowance for blockage
of 50% of the clear opening area of the
grate should be made.
(a)

Inflow Capacity

The inflow capacity of a field inlet depends


upon the depth of water over the inlet. For
shallow depths the flow will behave as for a
sharp crested weir. For greater depths the
inlet will become submerged and inflow
will behave as for an orifice.
It is
recommended that the capacity of the inlet
be checked using both procedures and the
lesser inlet capacity adopted.
Under weir flow conditions
11.2.10.7(a)) the equation is:

(Figure

Q g BF 1.66.L.h 3 / 2
Where:
Qg = flow into field inlet (m3/s);
BF = blockage factor = 0.5;
1.66 = weir coefficient;
L = weir length (m) (see note below);
and
h = depth of water upstream of inlet
(relative to weir crest) where flow
velocity is low (i.e. velocity head is
insignificant) otherwise use the
height of energy level above the weir
crest (m).

11

Note: The length referred to in this case is


the effective weir length. Thus for a grated
inlet adjacent to a kerb, the side along the
kerb should be ignored. For a side inlet the
length referred to is the length of the inlet.

Figure 11.2.10.7(a) Field Inlet Under


Weir Flow

Under orifice flow conditions (Figure


11.2.10.7(b)):
The orifice flow equation depends on the
pressure gradient across the orifice. The
standard orifice flow equation applies when
atmospheric pressure conditions exist
downstream of the grate, such as would
exist if the design Water Surface Elevation
(WSE) is 150 mm below the grate (as per
Table 7.16.1 of QUDM (2008) and Figure
11.2.10.7(b)).
The following equation is based upon a
pressure change coefficient of Kg = 2.75:

Q g BF 0.60. Ag 2 g .h

1/ 2

Where:
Qg = flow into field inlet (m3/s);
BF = blockage factor = 0.5;
0.60 is a constant = (1/Kg)1/2 =
(1/2.75)1/2;
Ag = clear opening area of grate (m2);
g = acceleration due to gravity (9.81
m/s2);
h = depth of approaching water
relative to the orifice (m); and
Kg = pressure change coefficient for
the grate.

January 2013

11-24

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

(c) Safety Issues


While inlet screens / grates shall comply
with the requirements of AS 3996-2006,
safety risks should be reviewed in
circumstances where a field inlet is located
within areas accessible to the public. The
maximum spacing of bars must be in
accordance with the following:
Figure 11.2.10.7(b) Field Inlet Under
Free Orifice Flow

The pressure change coefficient (Kg) can


vary significantly for unusual grate designs.
The coefficient used in the equation for
orifice flow is based on a typical open mesh
grate. It is noted that the pressure change
coefficient for the old cast iron City Grate
has been adopted as 2.23. Designers of
unusual hydraulic structures should seek
expert advice or review reference
documents on orifice flow.
If the field inlet is fully drowned (i.e. no air
gap exists below the grate and thus the
hydraulic pressure below the grate is not
atmospheric) then an estimate must be
made of the head loss through the structure
as per a normal Hydraulic Grade Line
(HGL) analysis. Such calculations require
considerable experience and hydraulic
judgement.
Guidance on head losses
through screens is provided in Section
7.16.14(c) of QUDM (2008).
(b)

Freeboard considerations

Where the inlet is contained within a pond


formed by earth mounds or similar,
freeboard should be 20% of the depth of the
pond with a minimum of 50 mm under
minor storm conditions. However where
overflow must be avoided the design storm
shall be the major storm event.

Horizontal inlet screens maximum


spacing is 125 mm, however if there
is a risk of a child being swept by
stormwater towards the screen, then a
maximum clear opening of 88 mm is
required between bars. This spacing
excludes the entry of the torso of
child and is based on test procedures
specified in AS 4685.1-2004.
Vertical or inclined inlet screens maximum spacing is 125 mm.
Other safety considerations include the
following:
Possible tripping hazard of a
horizontal grate / screen (e.g.
particularly if not set flush with the
ground);
Flow velocity through the screen /
grate should be sufficiently low
enough to prevent a child from being
held against the screen / grate by
hydraulic pressure.
Raised, horizontal screens are generally not
acceptable adjacent to footpaths, bikeways
or public areas where significant numbers
of people gather as these inlets may
represent an unacceptable safety risk. In
such circumstances, flush screens should be
used, or possibly large dome screens if such
screens are likely to be clearly visible and
not represent a safety risk. Alternatively,
marker posts or fencing may be used.

January 2013

11-25

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

11.2.11 Access Chambers

The geometry of pipes at access chambers


is critical in respect of hydraulic head loss.
The main principles to be followed to
minimise head loss are:

Access chambers, or manholes in older


literature, are placed in pipelines or
drainlines:
to provide access for maintenance;

Minimise changes in flow velocity


through the chamber.

at changes of direction, grade or


level; and

Minimise changes in flow direction.

at junctions.
Consideration should be given to the
placement of an access chamber at an
obstruction or penetration by a conduit or
service, to facilitate the removal of debris.
The maximum spacing for access chambers
is 100 m for pipes less than 1200 mm
diameter and 150 m for pipes 1200 mm or
larger. They should also be located a
maximum of 100 m upstream of the outlet
of all pipes discharging into tidal waters
Standard departmental access chambers
1050 to 2100 mm in diameter are shown on
Standard Drawings Nos 1307 and 1308
(DMR 2009b).
The tops of access chambers in roadways or
paved surfaces should be flush with the
finished surface. Although designed for
wheel loads, the tops of access chambers
should be located away from wheel paths, if
possible, to minimise damage to pavements.
The tops of manholes elsewhere should be
25 mm above natural surface and tapered
down to its surrounds.

11

In cases where precast chambers are used,


the connecting stormwater pipes should not
protrude into the chamber and should be
sealed and finished in accordance with an
approved construction detail.
The following discussion has been
extracted from QUDM (2008) with some
minor modification.

January 2013

11-26

Avoid opposed lateral inflows, i.e.


all incoming pipes should ideally be
contained within a 90 degree arc, but
certainly less than 180 degrees.
Limit the deflection from inflow to
outflow for pipes smaller than 600
mm diameter to 90 degrees, or 67.5
degrees for pipes 600 mm and greater
in diameter.
Avoid vertical misalignment i.e.
drop pits, unless deliberately
intending to induce high head loss.
Additionally, where change in pipe
size occurs, it is preferable to align
inlet and outlet pipes obvert to
obvert.
Where practical, direct inlet pipes wholly
into the barrel of the outlet pipe (Figure
11.2.11(a)). It is noted that for various
reasons, inflow pipes often need to be
directed towards the centre of the pit
(Figure 11.2.11(b)), however, this will
increase losses.
Rounding the entrance to the outlet pipe at a
radius of one-twelfth of the outlet diameter
will help to reduce losses (Figure
11.2.11(c)).
Where practical, the change of direction of
flow should occur at or near the
downstream face of the chamber.
Head losses resulting from surface inflows
(Figure 11.2.11(d)) are reduced if the
design water level in the chamber is well
above the outlet pipe obvert.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.2.11(d) Inlet Chamber


showing Water Level well above Outlet
Obvert
Figure 11.2.11(a) Inflow Pipe Directed
at Centre of Outflow Pipe

11.2.12 Access Chamber Tops


Where an access chamber is located within
a carriageway, the chamber top, or access
point, should be positioned to avoid wheel
paths and should be finished with the top
flush with the finished surface.

Figure 11.2.11(b) Streamlines


Resulting from Inflow Pipe Directed at
Pit Centre

Elsewhere, access chambers should be


finished 25 mm above natural surface with
the topsoil or grassed surface around the
chamber graded gently away. On playing
fields they may be finished 200 mm below
the finished level, but only when located in
a straight line between two permanently
accessible chambers.

11.2.13 Reduction in Pipe Size


A reduction in pipe size along a pipeline /
drainline is not permitted as the
discontinuity between the different pipe
sections can catch debris causing
blockage which inturn reduces the capacity
of the system and/or can cause failure of the
system.

Figure 11.2.11(c) Bellmouth Entrance


to Outlet Pipe

For some projects where the inflow to an


existing pipe network has been reduced and
the network requires extension, a reduction
in pipe size (for single barrel pipelines
only) may be permitted provided that an
access chamber is placed between the
different pipe sections in order to allow
access to remove any debris. Each case

January 2013

11-27

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

must be assessed individually and written


approval to reduce the pipe size in a
pipeline must be given by the departments
design representative. Table 11.2.13 lists
the recommended maximum reduction in
pipe size.

avoid displacement of the chamber


lid / screen.

Table 11.2.13 Maximum Allowable


Reduction in Pipe Size
Upstream Pipe
Diameter (mm)
Less than 600
675 to 1200
Greater than 1200

Allowable Change
in Diameter
No change
ONE pipe size
TWO pipe sizes

Notes:
Any reduction must be approved, see Section 11.2.13.
The above recommendations are based upon the
nominal sizes of pipes as manufactured in
accordance with AS 4058-2007.

11.2.14 Surcharge Chambers


This section has been extracted from
QUDM (2008).
Prior to incorporating a surcharge chamber
into a drainage line, the following should be
considered:
The potential for a person (that has
been swept into the upstream
drainage system) being trapped
inside the surcharge chamber unable
to exit the chamber or the outlet pipe.
Potential surcharge of the upstream
system and flooding problems caused
by debris blockage of the outlet
screen.

11

Structural integrity of the chamber,


outlet screen, top slab and concrete
coping, and its ability to withstand
high outflow velocities and high
pressure forces caused by debris
blockages. There is a need in many
cases to ensure the surcharge screen
is securely anchored to the top slab,
and the slab to the chamber walls, to

January 2013

11-28

Safe maintenance access to allow


removal of debris trapped within the
surcharge chamber.
The hydraulic analysis of surcharge
chambers is presented in Section 7.16.14 of
QUDM (2008).

11.2.15 Pipeline Requirements


11.2.15.1 Pipe Joint Types
The description and requirements for types
of pipe joints are detailed in Section 9.2.1.
11.2.15.2 Geometric Tolerances and
Cover Requirements
The geometric tolerances for cover
requirements are detailed in Section 9.2.2.
11.2.15.3 Minimum Pipe Size
The minimum diameter of any pipe in a
drainage system should be 375 mm.
11.2.15.4 Box Sections
The requirements for use of box sections
are detailed in Section 9.2.4.
Furthermore, where box culverts are
constructed
on
a
skew,
special
consideration is required where units join
pits and access chambers.
11.2.15.5 Location in Urban Areas
This section has been extracted from
QUDM (2008) with some modification and
provides general guidance for the location
of pipe networks in urban areas (local
authorities would generally follow this
guidance).

Department of Transport and Main Roads


Road Drainage Manual

Minor pipes connecting one kerb inlet or pit


directly to another is acceptable at the top
of the drainage system and these pipes may
be located under the kerb and channel.
For pipelines greater than 600 mm it is
recommended that the location for pipelines
in the road pavement - other than a kerb
inlet to kerb inlet connection - be 2.0 metres
measured towards the road centreline from
the invert of the kerb and channel, however
access chamber tops or access points should
be located to avoid wheel paths.
Where sufficient verge width is available
stormwater pipes may be located in the
verge to suit the services allocations of the
relevant authorities / owners.
In divided roads, drainage pipelines may be
located within the median, normally offset
1.5 metres from the centreline (as street
lighting poles are normally on the
centreline).

11.2.16 Splay Pipes


The use of splay pipe components to
construct bends in pipelines is not
permitted, particularly between pits / access
chambers.
Refer Section 9.2.11 for possible relaxation
of this requirement when not between pits /
access chambers.

11.2.17 Structural Requirements


of Pipelines
The structural requirements for pipelines
are detailed in Section 9.2.6. However, for
pipe networks in urban areas, Trench
Installation is typical.
All other structural aspects should be
referred to the departments Bridge Design
Section / Structures Branch located within
the Engineering & Technology Division or

Chapter 11
Road Surface and Subsurface Drainage Design

suitably prequalified structural engineering


consultant.

11.2.18 Flow Velocity Limits


The velocity of stormwater in pipes and box
sections should be maintained within
acceptable limits (refer Table 11.2.18) to
ensure that:
some self cleaning of the pipe or
box section is maintained (See
Section 2.8); and
scouring and erosion of the conduit,
(particularly the invert) does not
occur.
In steep terrain, the velocity of flow should
not be greater than the absolute maximum
velocity of 6.0 m/s under pipe full
conditions. To achieve this requirement, it
may be necessary to construct access
chambers with drops to dissipate some of
the kinetic energy of the flow, or to limit
the pipe diameter.
Notwithstanding the above suggested
velocity limits, hydraulic considerations
may require the velocity be controlled to
well below the Desirable Maximum
and/or the pipe size increased to minimise
structure losses and the slope of the
hydraulic grade line.

11.2.19 Pipe Grade Limits


To conform with the requirements of
Section 11.2.18 and construction limitations
the maximum and minimum grades as
detailed in Table 11.2.19 are recommended
for design purposes.

11.2.20 Discharge Calculations


The following section has been extracted
from QUDM (2008) with some minor
modification.

January 2013

11-29

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

If the major (surface) system does not have


the capacity to carry the difference between
the design peak flow and the calculated
pipe flow (based on normal ARIs) then
additional inlets and hence larger pipes are
required to ensure that the surface system
operates within the specified limits.

The system objectives must seek to limit


flooding and to ensure a reasonable level of
pedestrian and vehicular traffic safety and
accessibility. These objectives are met by
ensuring that major and minor storm flows
are managed within specified limits and by
designing both major and minor system
components in conjunction.

Table 11.2.18 Acceptable Flow Velocities for Pipes and Box Sections
Absolute
Minimum [1]
(m/s)

Desirable
Minimum [1]
(m/s)

Desirable
Maximum [2]
(m/s)

Absolute
Maximum [2]
(m/s)

Partially full

0.7

1.2

4.7

7.0

Full

0.6

1.0

4.0

6.0

Flow Condition

Notes:
1. Minimum flow velocities apply to ARI design storm based on selected maintenance period (see Section 2.8), and
apply to all pipe materials.
2. Maximum flow velocities apply to concrete pipes. For other pipe materials, refer to manufacturers advice.

Source: Based on QUDM (2008) with modification.


Table 11.2.19 Acceptable Grades for Pipes Flowing Full

11

Pipe Diameter (mm)

Maximum Grade (%)

Minimum Grade (%)

300
375
450

20.0
15.0
11.0

0.50
0.40
0.30

525
600
675

9.0
7.5
6.5

0.25
0.20
0.18

750
900
1050

5.5
4.5
3.5

0.15
0.12
0.10

1200
1350
1500

3.0
2.5
2.2

0.10
0.10
0.10

1650
1800
1950

2.0
1.7
1.5

0.10
0.10
0.10

1.4
1.3
1.2

0.10
0.10
0.10

2100
2250
2400
Source: QUDM (2008)

January 2013

11-30

Department of Transport and Main Roads


Road Drainage Manual

The general principles or objectives are:


The drainage system as a whole is
provided to mitigate flooding and to
ensure the safety and convenience of
pedestrians and vehicles.
The
minor
drainage
system
comprising underground pipes and/or
surface flowpaths is designed to
provide for the safety and
convenience of pedestrians and
vehicles.
Where flood immunity cannot be
provided
under
major
storm
conditions via overland flowpaths,
the capacity of the underground pipe
system and the inlets leading to it
need to be increased in order to
reduce surface flows to acceptable
levels.
Under normal conditions the capacity of the
underground pipe system should not be less
than its minor storm flow conditions while
the system is operating under major storm
conditions. The exceptions would be when
tailwater levels downstream have a
significant effect on the systems hydraulic
grade line, or the surface gradient is
considerably flatter than the pipe gradient,
thus causing the HGL. to rise above the
ground surface.
The underground system should be
designed with a suitable allowance for
blockage at kerb inlets as described in
Section 11.2.10.1. In this way the full
design capacity of the underground system
can be taken into account under both major
and minor storm conditions.

Chapter 11
Road Surface and Subsurface Drainage Design

11.2.21 General Design Procedure


A general design procedure is detailed in
Section 7.15.3 of QUDM (2008) and
provides a sound basis and understanding
for the design of pipe networks required for
departmental projects. It is recommended
that designers should refer to this procedure
for the initial design assessment.

11.2.22 Hydraulic Calculations


The detailed hydraulic grade line (HGL)
method is recommended for the analysis of
underground stormwater pipe systems. It is
further recommended that this be based on
an analysis proceeding from downstream to
upstream through the system.
Section 7.16 of QUDM (2008) provides
guidance, understanding and the hydraulic
calculations required to undertake the
design of a pipe network for departmental
projects. All subsections of Section 7.16
QUDM (2008) apply except for Section
7.16.10 where Section 11.4.17 of this
manual applies.

11.3

Aquaplaning

For road users to be able to stop or steer


their vehicles, the tyres must grip the road
surface. Grip is achieved by means of
friction generated in the areas where the
tyres make contact with the uppermost
particles of the road surface. If the friction
available is insufficient to resist the forces
generated by acceleration, braking or
cornering manoeuvres, the tyre may slip
over the road surface. In dry conditions,
surface friction is generally at a level that
supports most normal manoeuvres,
however the level of friction available
decreases when a road surface is wet or
flooded (adapted from Austroads (2005)).

January 2013

11-31

11

Department of Transport and Main Roads


Road Drainage Manual

Many factors exist, including tyre tread and


pressure, road surface and geometry,
vehicle speed and depth of water over the
surface that influences effective, lateral and
longitudinal friction between the tyre and
road surface. When a loss of grip or
traction occurs, the effect on a moving
vehicle is either skidding or aquaplaning,
both of which represent a major driving
hazard.
Aquaplaning is the least
understood, yet most dangerous condition.
The information, analysis method and
criteria presented in this section is primarily
intended for geometric road design
purposes, to identify and minimise
aquaplaning potential. This section can
also be used for incident investigation
purposes.

11.3.1 What is Aquaplaning?


Aquaplaning or hydroplaning occurs when
the vehicles tyres are separated (partially
or fully) from the road surface by a film of
water and which results in loss of control of
the vehicle. The effect of aquaplaning on
vehicle handling is directional instability,
with the worst case being a complete loss of
directional control. Furthermore, as contact
with the road surface is lost, braking
becomes impossible.

11

Notwithstanding the above definition, it is


technically more accurate to define
aquaplaning
as
the
build-up
of
hydropressure beneath a tyre to the extent
that it partially or fully exceeds the capacity
of the tyre to absorb it, thereby reducing the
contact area between the tyre and the
pavement.
There are three types of aquaplaning as
identified by Horne (1968). The two main
types of aquaplaning are Viscous and
Dynamic:

January 2013

11-32

Chapter 11
Road Surface and Subsurface Drainage Design

Viscous Aquaplaning
Viscous aquaplaning can occur at
low speeds where the texture of the
road surface is low (i.e. it is tending
towards smooth). It requires minimal
water depth to occur as the
mechanism is dependent on the
viscosity of water, which prevents it
from escaping from under the tyre
footprint. This type of aquaplaning is
most likely to occur during braking,
such as at an intersection or at traffic
signals, and its occurrence can be
exacerbated if the vehicle is running
on tyres with limited or no tread.
Viscous aquaplaning can occur at
water film depths as low as 0.01 mm.
Almost any condition of pavement
wetness, and even the lowest vehicle
speeds, may trigger the onset of a
viscous aquaplaning condition. The
single most important variable which
controls
and
regulates
the
phenomenon is the microtexture at
the tips of the pavement asperities
(Moore, 1975).
Dynamic Aquaplaning
Dynamic aquaplaning is the partial or
full separation of tyre and pavement
which
occurs
under
flooded
conditions. Flooding is said to occur
when the thickness of the water film
on the pavement surface is such that,
at a given speed, the combination of
tyre
tread
and
pavement
macrotexture
is
incapable
of
discharging the bulk water from the
contact patch. It may equally occur
if the tyre is free rolling or locked.
Dynamic aquaplaning will occur
once the vehicle has exceeded a
critical speed (a function of the tyre
pressure), at which time surface

Department of Transport and Main Roads


Road Drainage Manual

water in front of the tyre, acting as a


wedge, can penetrate the tyre
footprint and reduce the surface
contact area. At high speeds only a
small portion of the tyre footprint has
dry contact. Under full dynamic
aquaplaning, virtually no part of the
tyre will have any contact with the
surface and, even though the vehicle
is travelling at great speed, the tyre
can be fully locked.
Once
aquaplaning has occurred, vehicle
speed must be reduced significantly,
and to well below the critical speed,
for dynamic aquaplaning to cease
completely.
Full dynamic aquaplaning is likely to
be rare based on current vehicles
operating within existing speed limits
and with tyres in good condition.
Partial dynamic aquaplaning is more
likely to occur. Loss of control may
then occur with change in speed and
direction as high demands are placed
on either longitudinal or lateral
friction.
The third type of aquaplaning is called
tyre-tread rubber reversion, however this
only occurs when heavy vehicles such as
trucks or aircraft lock their wheels when
moving at high speeds on wet pavements
with macro-texture but little micro-texture
(refer Section 11.3.5). Rubber reversion
aquaplaning is not discussed further.
It is important to understand that
aquaplaning and skidding are not the same
condition. In a skid, the tyre still has
contact with the road surface; there is no
separation between the tyre and road. On
straight sections of road, a skid generally
results from hard braking or acceleration.
On curves, the vehicle may skid or slip over
the surface and track wider than the

Chapter 11
Road Surface and Subsurface Drainage Design

steered path or a yaw condition may occur


where the vehicle will skid / slide sideways,
usually with the vehicle ending up on the
inside of the curve. A partial aquaplaning
situation is essentially part aquaplaning,
part skidding.
Aquaplaning potential is assessed via a two
part process:
1. Determine water film depth (particularly
in the anticipated wheel paths) for a flow
path across the pavement; and
2. Based on operating speed of the road
section, check estimated water film depth
against acceptable depths limits.
These steps are discussed in detail in
Sections 11.3.7 & 11.3.8, however the
following sections (11.3.2 to 11.3.6)
provide relevant information to help the
designer understand aquaplaning and
reduce aquaplaning potential.

11.3.2 Causal Factors


Key factors which influence (or cause) the
occurrence of aquaplaning are:
road geometry;
road surface texture, porosity and
rutting;
operating speed;
rainfall intensity;
water film depth;
tyre tread depth, vertical load, width
of tyres and tyre pressure; and
driver behaviour.
Some of the above causal factors influence
others, but the listing gives an indication of
the problem of defining strict rules for
design.
Road geometry and road surfacing are the
two key factors that road designers have the
January 2013

11-33

11

Department of Transport and Main Roads


Road Drainage Manual

most control over. Road surfacing is an


important design consideration and the
selection of the most appropriate surfacing
is crucial.
Road surfacing is further
discussed in Section 11.3.3. With reference
to Section 2.3.2.2, aquaplaning is
considered more of a geometrics issue than
a drainage issue. The reason for this is that
the shape of the road surface has a direct
influence on the flow and build up of
stormwater runoff across the surface, which
in turn directly influences aquaplaning
potential. Therefore when water film depth
exceeds accepted limits, the problem is best
solved through adjustment of the road
shape, or more correctly geometrics
(alignment & crossfall).
There has been much research carried out
on the subject but more is needed.

11.3.3 Road Surfacing


Road or pavement surface properties are not
dependent solely on engineering factors,
and cannot be determined in isolation. The
required properties have to be developed in
the context of and within the constraints of:
driver behaviour;
environment;
vehicle characteristics;
regulatory environments (delineation,
signs, speed, constraint of driver
behaviour);
road alignment and layout;
maintenance regime; and

11

available funding.
It should be taken into account that
pavement surface properties cannot be
developed to compensate for extreme
weather conditions, excess speed and/or
deficiencies in areas such as poor alignment

January 2013

11-34

Chapter 11
Road Surface and Subsurface Drainage Design

(geometrics). Also, there is no one model


or formula that can be used to compare
varying combinations of road safety
components (vehicle characteristics, driver
behaviour, environment, road alignment,
pavement surface properties and so on).
In dry conditions the maximum friction for
a particular surface is provided by a tyre
with no tread running on a smooth
pavement surface. This is demonstrated by
racing cars operating in dry conditions and
the significance of when to decide to
change to wet weather tyres (grooved
treads) when rain threatens.
The required tyre and pavement surface
characteristics are very different when the
pavement is wet. To make provision for
water on the pavement surface, the
following has to apply:
(a) A design rainfall intensity has to be
adopted as policy because it is not
possible to design the road to
accommodate all rainfall intensities.
(b) The road alignment and drainage has
to be designed to shed the water from
the design rainfall intensity to meet a
maximum permissible water film
thickness.
(c) Pavement surface
required to:

macrotexture

is

(i) Reduce
the
potential
for
aquaplaning by providing drainage
paths for the water to escape from
beneath the tyre, in addition to the
drainage paths provided by the
grooves in the tyre;
(ii) Contribute to friction between the
tyre and road through hysteric
deformation of the tyre (i.e.
providing resistance to the tyre
through deformation of the tyre

Department of Transport and Main Roads


Road Drainage Manual

when it is passing across the


macrotexture and the recovery of
the initial shape of the tyre);
(d) Pavement
surface
microtexture
(asperity) has to be provided to enable
the tip of the stone to penetrate any
water film and provide adhesion with
the tyre.
Tyres must have a minimum groove depth
to provide a drainage path for the water to
escape. Sufficient tyre pressure is also
required such that the stress between the
tyre and the pavement surface is sufficient
to enable the tyre to displace the water and
achieve direct contact.
With reference to point c) above and Figure
11.3.3, Macrotexture can be affected by the
size, shape and spacing of coarse aggregate
particles in the surfacing material; the
presence and pattern of grooves purposely
manufactured in concrete surfaces; or the
connection between surface and internal
pores in the materials.
Macrotexture
influences water drainage capacity, which is
important in removing water from the road
- tyre contact patch, and allows the vehicle
tyre tread to deform, so creating the
hysteresis forces.

Chapter 11
Road Surface and Subsurface Drainage Design

of the aggregate particles in the upper layer


of the surfacing material, or by fine
particles such as sand in asphalt and
concrete
surfacings.
Microtexture
influences wet and dry skid resistance at all
speeds, interacting with the vehicles tyres
to generate the adhesive friction forces. It
is important to note that very high
microtextures can significantly increase the
rate of tyre wear experienced at a location.
When a road surface is wet, the water film
on the road surface reduces the intimate
contact needed between the tyre and the
microtexture, leading to a reduction in
available friction, i.e. it is dependent on the
extent to which the microtexture is able to
penetrate the water film. The surface
tension properties of water also result in a
water film forming more readily on smooth
aggregate,
further
highlighting
the
importance of the microtexture.
Microtexture is gradually polished away by
the action of traffic (particularly where high
stresses are imparted to the road surface by
acceleration, braking and cornering
manoeuvres and/or heavy vehicles).

Macrotexture also has an effect on water


spray generated from the road surface. In
general terms, the greater the macrotexture
provided, the greater the reduction in water
spray.
The terms surface texture and texture
depth are often used by practitioners and
these are simply descriptors of the
macrotexture of the road surface.

Figure 11.3.3 - Key Texture Elements of


Pavement Surfaces (Source: Austroads
(2005))

With reference to point d) above and Figure


11.3.3, Microtexture is the very fine
surface texture of the aggregate used in the
surfacing material. The fine texture is
provided either by the crystalline structure

11.3.3.1 Pavement Surface Types


The department uses several different
surface types for roads. The following
discussion presents general information

January 2013

11-35

11

Department of Transport and Main Roads


Road Drainage Manual

regarding the various surfaces, in relation to


aquaplaning. To select an appropriate road
surfacing for any project, designers must
refer to other guides as issued by the
departments Pavements and Materials
Branch, Engineering & Technology
Division.
Sprayed Seal
Sprayed seals (aggregate size > 10 mm)
generally provide the best texture depth of
all currently used surface types. The stone
aggregate also provides good microtexture
unless it has a low Polished Aggregate
Friction Value (PAFV).
Open Graded (OG) Asphalt
Open Graded Asphalt will easily achieve
good texture depth and could be considered
as the second best surface type with respect
to texture. Open graded asphalt also allows
some water flow within the pavement due
to the voids within the mix (refer Figure
11.3.3.1).
Some specialists consider that for vehicle
speeds up to 110 km/h it would be unlikely
that there will be much of a film of water
under vehicle tyres as the water will be
forced into the pores of the OG asphalt and
come out elsewhere.
Furthermore,
anecdotal evidence exists that shows OG
asphalt surfaces produce significantly less

11

Chapter 11
Road Surface and Subsurface Drainage Design

spray than other asphalt surfaces. This is


thought to be as a result of the surface being
able to better absorb and dissipate surface
flows from under the tyre. More research is
required to quantify the use of the asphalt.
Dense Graded (DG) Asphalt
It is extremely difficult to achieve good
texture depths for dense graded asphalt
pavements for traffic speeds over 80 km/h.
Unlike OG asphalt, DG asphalt does not
allow flow within the mix (refer Figure
11.3.3.1).
Stone Mastic Asphalt (SMA)
Stone mastic asphalt has relatively high
binder content and gap grading with the
finer sizes. This tended to give texture
depths similar to DG asphalt, however the
latest departmental specification for SMA
can give better results (intention is to
achieve 1.1 mm). Unlike OG asphalt, SMA
does not allow flow within the mix (refer
Figure 11.3.3.1).
Concrete
Depending on the surface treatment of the
finished
concrete
surface,
concrete
pavements can have poor to good texture
depth.

Figure 11.3.3.1 - Schematics of Aggregate Skeletons for Asphalt Mix Types (Source:
Austroads Part 4b)

January 2013

11-36

Department of Transport and Main Roads


Road Drainage Manual

11.3.4 Tyres
Tyres are beyond the control of the road
designer and therefore not discussed in
detail. However it is important to note that
the minimum legal tread depth in
Queensland is 1.5 mm which represents
approximately 80% wear of the tyre.
Section 11.3.5 shows that tyres at the legal
limit of tread offer little grip at speed and
with minimal water on the surface.
Tyre manufacturers are developing tyres to
improve their wet weather handling. One
such tyre, designed specifically to reduce
aquaplaning potential, is shown in Figure
11.3.4. This tyre has a large central void
and clear channels in the tread to improve
flow to the side of the contact patch.

Figure 11.3.4 - Tyre Designed to Reduce


Aquaplaning Potential (Source:
Goodyear website)

11.3.5 The Road-Tyre Interface


To enable a vehicle to be driven on a
roadway, adequate surface friction must be
available to accommodate the forces
required for it to safely complete its
manoeuvres. This friction is generated at
the road-tyre interface. This interface is
more commonly known as the contact patch

Chapter 11
Road Surface and Subsurface Drainage Design

and represents the area where the rubber of


the tyre is in contact with the road surface
material.
It is important to note that rubber does not
conform to the classic laws of friction. This
is because the material is viscoelastic and
its frictional properties depend on
temperature and rate of deformation.
Therefore, the frictional behaviour of
rubber needs to be considered differently.
Gillespie (1992) and others describe the two
primary mechanisms responsible for the
friction coupling at the road-tyre interface
as surface adhesion and hysteresis (refer
Figure 11.3.5(a)).
Surface adhesion arises from the shearing
of the intermolecular bonds between the
tyre rubber and the aggregate in the road
surface. The adhesion component is the
larger of the two mechanisms on dry roads,
but is reduced substantially when the road
surface is wet. Adhesion is provided by the
microtexture of the pavement surface.
The hysteresis (or deformation of the tyre)
mechanism represents energy loss in the
tyre rubber as it deforms when sliding over
the aggregate in the road. Hysteresis is not
affected so much by water on the road
surface. The deformation of the tyre is
largely provided by the macrotexture of the
pavement surface.
On dry roads, both peak and slide friction
decrease with velocity.
Under wet
conditions, even greater speed sensitivity
prevails because of the difficulty of
displacing water in the contact patch at high
speeds (Gillespie 1992).

January 2013

11-37

11

Department of Transport and Main Roads


Road Drainage Manual

In wet conditions, as the tyre rolls over the


pavement surface, the water on the surface
is displaced from the contact patch via three
paths:
the area in front and to the side of the
tyre;

Chapter 11
Road Surface and Subsurface Drainage Design

the voids created by the tread of the


tyre (refer Figure 11.3.5(b)); and
the voids under the tyre created by
the macrotexture of the surface (refer
Figure 11.3.5(c)).

Figure 11.3.5(a) - Mechanisms of Road Tyre Friction (Source Gillespie (1992))

11

Figure 11.3.5(b) - Tyre Tread Voids


(Source: M Gothi (2005))
Figure 11.3.5(c) - Pavement Surface
Voids below Top of Aggregate (Source:
M Gothi (2005))

January 2013

11-38

Department of Transport and Main Roads


Road Drainage Manual

The degree of contact that can be


established between the vehicle tyre and the
road surface is largely determined by the
following factors:
depth of the water film present on the
surface;
surface texture depth (and the
capacity of the surface to shed water,
which is also a function of the
efficacy of any positive drainage
provided at the location);
tread depth, width and pressure of the
tyre; and
vehicle speed.
The depth on the water present on the
surface is also important as the less water
on the surface, the less water there is to be
displaced. Therefore good drainage of the
surface is required.
Good surface texture depth and tread depth
provide capacity for water on the surface to
be pushed or squeezed away from the
contact patch. The time for the surface
water to flow into the voids created by the
tread and texture, as the tyre passes over, is
considerably less than the time for the
surface water to be pushed in front and
then to the side of the tyre. Reduced tread
depth and reduced texture depth of the
surface reduce available void capacity
which restricts the flow of water away from
the contact patch.
Vehicle speed is critical as the higher the
speed, the faster the rotation of the tyre, and
the less time available for the water on the
surface to be displaced from the contact
patch.
At low speeds, surface water
displaces via the paths discussed earlier. As
speed increases, the small wedge of water
continues building in front of the tyre. A
point is reached where the surface water

Chapter 11
Road Surface and Subsurface Drainage Design

does not have the time to displace. The


wedge of water develops enough
hydropressure to lift the tyre and allow the
water to pass under the tyre separating the
tyre from the road (refer Figure 11.2.5(d)).
Gothi (2005) showed a graphic
relationship (refer Figure 11.2.5(e))
highlighting the effect of speed and tread
depth on the reduction of the contact patch.
Figure 11.2.5(e) illustrates the reduced
voids of the tyre tread (@ 50 km/h)
between a new tyre and a tyre approaching
the legal minimum depth. Also, the figure
shows that a worn tyre (> 50% wear) at 90
km/h is partially aquaplaning as the contact
patch is substantially reduced / almost non
existent.

11.3.6 Skid Resistance


Skid resistance is a very complex topic.
The available and required skid resistance is
influenced by the interaction of many
variables, including:
driver behaviour
driver expectations
vehicle characteristics
tyre characteristics
climatic conditions
traffic mix and volume
pavement surface characteristics
(microtexture and macrotexture)
water film thickness (drainage)
surface contaminants
road alignment
vertical)

(horizontal

and

lane and carriageway width


cross-fall, and
signage.

January 2013

11-39

11

Department of Transport and Main Roads


Road Drainage Manual

The pavement surface is just one of a large


number of contributory factors. The three
main areas of control which relate to the
pavement surface, and which many road
authorities use internationally to address the
issue of skid resistance are:
(a) in-service skid resistance (friction)
(b) aggregate polishing resistance, and
(c) surface texture depth.
An interesting point worth noting is that
research has found that the level of skid
resistance provided by some surfacing types
in the early period after placement, e.g.
some asphalts and materials incorporating

Chapter 11
Road Surface and Subsurface Drainage Design

polymer-modified binders (PMB), can be


less than the level that would normally be
anticipated from the individual or combined
properties of the constituents.
It is
suggested that the reason for this early life
phenomenon is the binder coating on the
uppermost surfacing aggregate is taking
time to be worn off, and is masking the
potential microtexture of the aggregate.
While the binder coating on the uppermost
surfacing aggregate remains, it could have
the potential to act as a lubricant to any
skidding mechanism in dry weather
conditions.

Figure 11.3.5(d) - Water Wedge at Separation (Source: Foucard (2005)

11

January 2013

11-40

Department of Transport and Main Roads


Road Drainage Manual

@ 50km/h

Chapter 11
Road Surface and Subsurface Drainage Design

@ 70km/h

@90km/h

New Tyre
Tread Depth =
8mm

Worn Tyre
50% Wear

Worn Tyre
Tread Depth =
1.6mm

Figure 11.3.5(d) - Tyre Tread Depth Influence: 1 mm Water Depth


Source: M Gothi (2005)

While there is an implied relationship


between aquaplaning potential and skid
resistance of the pavement, there is
currently no established direct relationship
between them. Therefore, skid resistance
test results can only provide background
information when assessing aquaplaning
potential at any given location.

11.3.7 Assessment Water Film


Depth
Several theoretical and empirical methods
and formula exist to predict the depth or
thickness of the water film over the surface.

11.3.7.1 Adopted Method


The method adopted by the department is
the one developed by Gallaway et al (1979)
for the Federal Highway Administration.
The metric version of the equation adopted
is given below:

0.103 T 0.11 L0.43 I 0.59


T
S 0.42

Where;
D = water film depth above top of
pavement texture (mm) ;
T = average pavement texture depth
(mm) (refer Section 11.3.7.3);

January 2013

11-41

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

L = length of drainage path (m) (refer


Section 11.3.7.4);
I = rainfall intensity, adopt 50 mm/h
(refer Section 11.3.7.5); and
S = slope of drainage path (%) (refer
Section 11.3.7.4).
Several versions of this formula have been
published however the key difference is
generally the units used for the slope
variable.
In 1986, Austroads (then
NAASRA) published the formula with an
incorrect constant (0.135 instead of 0.103)
which is an incorrect conversion of units.
Figure 11.3.7.1 provides an example of
several drainage paths based on road
surface contours, about the superelevation
change when exiting a right hand curve.
Path A crosses the road from one side to the
other where the 3% superelevation is
applied. The contour spacing is uniform
(constant slope) and slope would be slightly
flatter than 3%. Where the superelevation
starts to roll-down, for example, when
exiting a curve, drainage paths start to
change and slopes can flatten off (contour
spacing increases), as shown by Path B.
Paths D & E indicate drainage paths over
normal crossfall situations. Path C is of
interest and possible concern. The path

starts to cross the road and then, due to the


superelevation rotation, turns back. This
situation can create long drainage paths
with a flat section in the middle, as
indicated by the large spacing of the
contours. Assessment of this path must
check water film depths at both the end of
the flat section and end of the drainage
path, in the vicinity of the anticipated wheel
paths.
11.3.7.2 Basis / Limits
The Gallaway formula is an empirical
formula based on the following experiment
parameters:
drainage lengths up to 48 ft (14.6 m);
rainfall intensities up to 2 in/h (50.8
mm/h);
slopes up to 8%; and
several surfaces tested (incl. sprayed
seals, asphalt and concrete).
The flow path tested was over a simple,
planar surface and the formula does not
contain a term for hydraulic resistance of
the pavement.

11
D
Figure 11.3.7.1 Aquaplaning Example (Road Surface Contours)

January 2013

11-42

Department of Transport and Main Roads


Road Drainage Manual

The research found that:


increasing surface texture resulted in
a decrease in water depth for a given
rainfall intensity, cross slope and
drainage length. This effect was
more pronounced at the flatter cross
slopes and lower rainfall intensities;
and
greater drainage lengths increased
water depths however, the rate of
increase in water depth became
smaller as drainage length increased.
Gallaway concluded that water depth as a
function of cross slope, texture depth and
rainfall intensity can be reliably predicted
for drainage lengths up to 48 ft (15 m) and
probably considerably beyond. While it is
suggested that the formula can be used for
flow path lengths greater than 15 m, no
evidence proving or disproving the use of
the formula over longer paths has been
found. Therefore the use of Gallaway
formula is still considered appropriate.
It should also be noted that the method is
one dimensional and only assesses depth of
flow along a single (zero width) flow path.
Flow velocity and width or spread of the
flow over the pavement surface is not
assessed. Some situations can occur where
water runoff from off the road surface can
flow onto the road and/or where runoff
from one flow path crosses a boundary and
joins another flow path. The Gallaway
formula is unable to assess these situations
properly and cases such as these should be
referred to Principal Engineer (Road Design
Standards), Road Planning & Design
Section, Engineering & Technology
Division.

Chapter 11
Road Surface and Subsurface Drainage Design

11.3.7.3 Texture Depth


The term texture depth refers to the
average depth of the macrotexture of the
road surface (refer Figure 11.3.7.3).

Figure 11.3.7.3 - Texture Depth

Suitable values of Texture Depth (T) for


various surface types can be determined
using Table 11.3.7.3 which provides a
range of texture depths from values
researched by the RTA (1994) and Dash
(1977) and measurements taken on typical
Queensland pavements.
Lower values
within the table can be assumed to represent
worn surfaces.
To consider future surface wear or possible
change of surface type, design should check
flow paths using T = 0.40 mm and any
difference with the proposed surface
assessment must be risk assessed and
mitigated if appropriate.
For incident investigation purposes, actual
test measurements are usually obtained and
should be used. Where testing is not
available, use Table 11.3.7.3.
11.3.7.4 Drainage Path
The length of the drainage path is relatively
simple to determine using contours of the
road surface (as shown in Figure 11.3.7.1).
It is important to:
include the contours of intersections,
turnouts and entry / exit ramps;
assess areas where superelevation is
applied; and
assess hilly and/or winding roads.
January 2013

11-43

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Table 11.3.7.3 - Road Texture Depth


Wearing Course Surface

Texture
Depth*
(mm)

Dense graded asphalt


10 mm or larger

0.40.8

Dense graded asphalt, 7 mm

0.3-0.5

Open graded asphalt**

>0.9

Stone mastic asphalt

>0.7

Fine gap graded asphalt

0.20.4

Slurry surfacing

0.40.8

Spray seals, 10 mm or larger

>1.5

Spray seals, 7 mm

0.6-1.0

Grooved Concrete

1.2

Exposed aggregate concrete

>0.9

Tyned concrete

0.40.6

Hessian dragged concrete

0.30.5

Broomed concrete

0.20.4

* Texture depth is usually measured by


the sand patch test using either sand or
glass beads (Department Test, Q7052008).
** As high as 2 mm when new, but clogs
up and needs cleaning.

However, incorrect (or less accurate)


answers are generally given as the slope
doesnt represent the full flow path slope
that well.
The best single slope representation of a
flow path which contains two or more
slopes is the Equal Area Slope (also
known as a Weighted Average Slope) and
this slope is to be used for the variable S in
the Gallaway formula. To help explain this,
if the flow path is predominately flat with
some steep sub-paths, then the Equal Area
Slope will be relatively flat. Alternatively,
if the flow path is predominately steep
with some flat sub-paths, then the Equal
Area Slope will be relatively steep.
Figure11.3.7.4(a) illustrates the difference
between the Point-to-Point Slope and Equal
Area Slope for a predominately flat flow
path and how the Equal Area Slope better
represents the slope of the flow path.
If the water film depth is required at several
points along the flow path, then the Equal
Area Slope needs to be determined from
each point of analysis back to the start of
the flow path.
The procedure to determine the Equal Area
Slope from a point of analysis back to the
start of the flow path is as follows:
1.

Plot the profile of the flow path (long


section).

Source : (RTA 1994) and (Dash 1977).

11

The slope of the drainage path used to


develop the Gallaway formula was over a
simple, planar surface. Unfortunately, in
reality, the full drainage path often contains
several sections or sub-paths, each with a
different slope. The Point-to-Point Slope
or Average Slope, as determined by
calculating the slope from the point of
analysis straight back to the start of the flow
path, has been used for some time.
January 2013

11-44

2.

Working in metres, calculate total area


under profile.
Divide area by length of profile then
multiply by 2.

3.

4.

Start at 0 (ie. point of analysis) and


move upstream to start of flow path.

This calculates the vertical ordinate of


the equal area triangle = Equal Area
Ordinate.

Plot this new ordinate (at highest point


of flow path) and join back to point of
analysis.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.3.7.4(a) - Point-to-Point vs. Equal Area Slope

Point of Analysis

Equal Area Ordinate

Figure 11.2.7.4(b) - Equal Area Slope

11

January 2013

11-45

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.3.7.4(c) - Equivalent Length Method

5.

Now calculate slope of this line the


Equal Area Slope.

Express as a percentage (%).

Figure 11.2.7.4(b) illustrates the Equal Area


Slope over a flow path containing several
sub-sections.
Another method, known as the Equivalent
Length Method has been developed / used
recently to assess the flow path. This
method attempts to assess the flow path,
sub-path by sub-path, determining the water
film depth at each point the slope changes.
The basis of this method is as follows, with
reference to Figure 11.2.7.4(c):
The water film depth over the first
section or sub-path (A-B) is
determined based on the slope of that
sub-path (S1).
An equivalent length sub-path,
based of the slope (S2) of the next
sub-path (B-C) is then back
calculated using the water film depth
determined in the preceding step (at
Pt. B).

11

The equivalent length is then added


to the length of the next sub-path (BC) and this total length is used to
determine the water film depth at the
end of the next sub-path (at Pt. C),
based on the slope of that path-path
(S2).

January 2013

11-46

Review of this method by the department


shows that it tends to over-estimate the
water film depth as it does not properly
account for the change in flow due to the
change in slope (a fundamental flaw) - it is
very erroneous when slopes steepen. This
method is not to be used.
11.3.7.5 Rainfall Intensity
Several research papers and Austroads
Guide to Road Design recommend the use
of rainfall IFD (refer Chapter 5) charts for a
particular site and an Average Recurrence
Interval (ARI) of 1 year (typically) to
determine the rainfall intensity to be used to
determine the water film depth over a road
surface. Across Queensland, this approach
would recommend the use of rainfall
intensities up to 100 mm/h (and even
higher).
However, research and other
design documentation (Yeager 1974,
NAASRA 1986, Ibrahim & Hall 1994,
HCM 2000, Dash 2006) suggest that drivers
tend to slow as rainfall intensity increases
and visibility decreases. This slowing
typically occurs at about 50 mm/h however
some drivers start to reduce speed at rainfall
intensities as low as 25 mm/h. As speed
decreases, the potential for aquaplaning also
decreases.
The department has adopted and used the
rainfall intensity of 50 mm/h to determine
water film depths for some time now and it

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

is considered appropriate to continue to use


this intensity until further research / review
supports change.

11.3.8 Assessment
Aquaplaning Potential
Efforts to quantify the probability of an
accident occurring due to aquaplaning for a
given rainfall intensity and pavement
conditions are made difficult due to the
following factors:
Groups of vehicles following each
other in lanes may pump the lane
dry to varying degrees;
Tyre tread thickness, vertical load
and width vary considerably because
of the many types of vehicles; and
The number of drivers slowing down
or not at the rainfall intensity needs
to be quantified.
Furthermore, there are some non-design
issues that can affect aquaplaning potential,
such as:
Signing for
conditions;

changed

weather

Driver education programmes to


raise awareness of aquaplaning
issues; and
Cleaning of open graded asphalt.
Despite the above statements, guidelines are
needed to assist designers and others to
assess the potential for aquaplaning. The
following brief discussion provides
information to assist understanding of the
assessment criteria / limits provided in
Section 11.3.8.1.
In 1970, Staughton & Williams (NAASRA
1986) stated that it is possible to infer that
for vehicles travelling below about 80
100km/h with tyres in good condition, full

aquaplaning is not likely to occur (partial


aquaplaning however may occur). While in
1978,
Welleman
(NAASRA
1986)
produced very similar results, he also found
that increasing water depth reduces friction
coefficient, with the greatest reduction
occurring up to a depth of 4mm. Beyond
4mm, full aquaplaning may result,
depending on tyre condition and vehicle
speed.
In 1986, the NAASRA publication Guide to
the Design of Road Surface Drainage stated
that it is not completely possible to define
recommended design limits for water
depths, however:
Critical depth to cause hydroplaning
occurs at about 4mm and above; and
Partial hydroplaning may commence
at depths of about 2.5mm.
Designers should also consider:
Likely speed conditions under chosen
rainfall; and
Duration of water flow and depth on
pavement after cessation of rain.
What happens immediately after rainfall
ceases is important to consider as Yu &
McNown in 1963 (Ross & Russam 1968)
reported that runoff increases immediately
after cessation of rainfall and Andrey &
Yagar in 1991 (Ibrahim & Hall 1994) found
that collision risk returns to normal
immediately after rain stops.
11.3.8.1 Assessment Criteria
The following criteria are to be adopted for
geometric road design and incident
investigation purposes. Where situations
do not comply, the aquaplaning potential is
considered too high and review of design is
required.

January 2013

11-47

11

Department of Transport and Main Roads


Road Drainage Manual

Road surface geometry should be


such that drainage path lengths are
limited to about 60 m.
Time of flow for any drainage path
(refer Section 11.3.10) should be
limited to no more than 10 minutes.
A maximum water film depth of 2.5
mm (desirable) to 4.0 mm (absolute)
applies to:
o sections where the operating or
design speed is 80 km/h or higher;
o the approaches to and exits from
intersections and roundabouts;
o intersections and roundabouts;
o steep downhill sections;
o the merge section for entry ramps
/ overtaking lanes / climbing
lanes;
o the diverge section for exit ramps
/ overtaking lanes / climbing
lanes; and
o superelevated curves (particularly
those approaching limiting curve
speed).
A maximum water film depth of 5.0
mm (desirable and absolute) to all
other situations.

11

Every effort to comply with desirable limits


and to achieve both length and time of flow
requirements should be made. To assist,
superelevation may need to be applied
using more than one point or axis of
rotation (i.e. one or more lanes
independently rotated). On high speed,
wide flat pavements, it can be difficult (near
impossible) to achieve the 2.5 mm desirable
limit, however experience has shown that
depths of about 3.25 mm are achievable.

January 2013

11-48

Chapter 11
Road Surface and Subsurface Drainage Design

It should be noted that the approach to an


intersection or roundabout is defined as the
stopping sight distance (on a wet surface) to
the end of the anticipated queue.
11.3.8.2 Basis / Limits
It may first appear that by adopting both a
texture requirement and a water film depth
requirement, an authority would be
doubling up on its control of aquaplaning.
That is, one may ask why a texture
requirement is necessary if water film depth
requirements are satisfied. However, a
texture requirement is still necessary for the
following reasons:
It is possible for a pavement to meet
a water film depth requirement, but
still have very low texture. In this
case, although aquaplaning may not
be a likely occurrence (assuming the
tyres are in good condition), skidding
may still be possible in the wet as
both the adhesion and hysteresis
components of skid resistance may
be low.
It would be difficult and impractical
to regularly survey the entire road
network and determine water film
depth. Data required would include
gradient, cross-fall, transverse profile
(i.e. rutting), carriageway width,
longitudinal slope length and texture
depth.

11.3.9 Quick Assessment


Review of a flow path, such as those shown
in Figure 11.3.7.1 may show one or two
flat sub-paths (refer Drainage Path C).
Provided that these sub-paths are about 15
to 20 m in length, designers can use the
Gallaway formula over the section, in
isolation of the whole flow path, to

Department of Transport and Main Roads


Road Drainage Manual

determine the water film depth. If the water


film depth is close to or exceeds acceptable
limits, then further calculations are not
required and review of the geometrics can
commence. If the water film depth is less
than acceptable limits, then full analysis of
the flow path, as per Sections 11.3.7 and
11.3.8, is required to confirm acceptable
depths.

11.3.10 Time of Flow


As stated in Section 11.3.8, it can be
important for designers to consider the time
of flow on the road surface after rainfall
ceases. For road surface flows, the time of
flow (time of concentration) for any given
flow path can be estimated using the
Kinematic Wave equation.
The equation adopted (Book 8, AR&R
(IEAust 2001)) is given below:

6.94 L n *
I 0.4 S 0.3

Chapter 11
Road Surface and Subsurface Drainage Design

Since the equation involves rainfall


intensity, it must be solved together with
the relationship between duration t and
intensity I. This can be done iteratively or
directly by preparing a table of values of
t.I0.4. An example of the use of this
equation is provided in Book 8 of AR&R
(IEAust 2001).
Where evidence exists that suggests flow
over the road surface continues well beyond
the estimated time (as determine using
Kinematic Wave Equation), further
investigation is required as water flow from
adjacent surface catchment or from off the
road may be contributing. In these cases,
advice should be sought from Principal
Engineer (Road Design Standards), Road
Planning & Design Section, Engineering &
Technology Division.

11.3.11 Puddles / Wheel Ruts

0.6

Where;
t = overland travel time (min);
L = length of drainage path (m);
n* = surface roughness (similar but
not identical to Mannings n);
I = rainfall intensity (mm/h); and
S = slope of drainage path (m/m).
This equation only applies to planes of flow
which are homogeneous in slope and
roughness.
For any drainage path
consisting of several sub-paths, each with a
different slope, it is recommended to use
the equal area slope for S in the equation.
This gives a reasonable estimation of the
time of flow, however if a more accurate
time is required, designers are referred to
Book 8 of AR&R (IEAust 2001).

Wheel ruts may produce long puddles


(parallel to vehicle travel) or alter the flow
path of road surface drainage. The road
crossfall and grade and the wheelpath width
may influence the potential for ponding in
wheelpath ruts.
It is important when assessing existing
roads for aquaplaning potential, particularly
for incident investigations, to determine the
depth of flow in wheel ruts. This depth can
be determined by using the method as
presented in Figure 11.3.11, which is based
on the average wheel rut width of 760 mm
(RTA 1994).
Crossfall in this equation is measured at
right angles to the road centreline and no
allowance is made for grade.

January 2013

11-49

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.3.11 - Puddle Depth Formula

11.3.12 Guidance to Reduce


Aquaplaning Potential
Aquaplaning may be a greater problem
where braking, hard acceleration or change
of vehicle direction occurs as these
manoeuvres place additional demands on
friction over that of normal driving.
Particular braking situations are:

If an unacceptable water depths result from


the procedure detailed in Sections 11.3.7
and 11.3.8, the designer must consider
methods of shortening the flow path such
as:
altering the crossfall;
adjusting the rate of superelevation
development;

the approaches to intersections and


roundabouts;

altering either the horizontal or


vertical alignments (or both); and

the diverge for exit ramps;

introducing more crown lines.

on steep downhill sections; and


Particular hard acceleration situations are:
at the merge section of entry ramps /
overtaking lanes / climbing lanes.
at the diverge section of overtaking
lanes / climbing lanes.

11

when entering / exiting roundabouts.

Particular change in direction situations


are:
intersections and roundabouts;
on superelevated curves (particularly
those approaching limiting curve
speed); and

January 2013

11-50

It is important to consider the interplay of


longitudinal grading and crossfall to
prevent water film depth developing to or
beyond set limits (refer Section 11.3.8.1).
Any identified problems should be solved
and mitigated through amended geometric
road design.
A drainage solution to
aquaplaning should be only considered as a
last resort option. If a drainage solution is
required, specialist advice is highly
recommended in the development /
assessment of design options.
Aquaplaning has occurred occasionally on
existing roads. One treatment of existing

Department of Transport and Main Roads


Road Drainage Manual

surfaces involves reconstructing the road


surface crossfall to allow surface water to
drain more quickly / reduce length of flow
path. However, the redesigned road surface
crossfall needs to comply with the
requirements in the departments Road
Planning and Design Manual - A guide to
QLD practice and Austroads Guide to
Road Design series.
Where modifying the crossfall of existing
roads to reduce the risk of aquaplaning is a
problem, the use of open graded asphalt as
an overlay should be considered. This
should be an overlay of OG14 open graded
asphalt, generally 40 mm thick (depending
on compaction requirements) on top of a
tack coat over the existing pavement. Use
of open graded asphalt at other locations
where the calculated water film depth is
marginally above the allowable maximum
may also be appropriate.
For concrete roads, problem locations
which have been found to need rectification
are usually but not always confined to
transitional areas of superelevation.
Surface texture is applied to hardened
concrete by carbide grinding, diamond
grinding and/or grooving, sand blasting or
water blasting.

11.4

Subsurface Drainage

The purpose of subsurface (or subsoil)


drainage is to control the moisture content
of the pavement and the surrounding
material in order to maintain pavement
strength and serviceability throughout the
design life. This section has largely been
taken from Austroads Guide to Road
Design - Part 5 (2008).
Subsurface or subsoil drains are provided in
order to avoid the following types of
premature failures:

Chapter 11
Road Surface and Subsurface Drainage Design

loss of subgrade strength and shape


due to an increase in moisture
content in moisture susceptible
materials;
overload of the subgrade due to
hydrostatic transmission of live load
through a saturated pavement; and/or
layer
separation
and
formation in the pavement.

pothole

11.4.1 Moisture in Roads


Some moisture is always present in the
subgrade and unbound paving materials due
to capillary moisture movement controlled
by the environment.
If this becomes
excessive, the subgrade and pavement can
be weakened appreciably. Consequently, it
is important to minimise ingress of water
into the pavement and subgrade.
Control of subsurface water is the
key to longevity of the pavement.
The main mechanisms by which moisture
can enter a road subgrade and/or pavement
are shown diagrammatically in Figure
11.4.1 and include:
Longitudinal seepage from higher
ground, particularly in cuttings and in
sag vertical curves;
Rise and fall of water table level
under a road;
Rainfall infiltration through the road
surfacing;
Capillary moisture from the verges;
Capillary water from a water table;
Vapour movements from a water
table;
Lateral movement of moisture from
pavement materials comprising the
road shoulder;

January 2013

11-51

11

Department of Transport and Main Roads


Road Drainage Manual

Water flowing or standing in table


drains, in catch drains, in median
areas, within raised traffic islands, or
adjacent to the road (not illustrated);
Leakage of water supply and
drainage lines (not illustrated); and
Passage
of
water
through
construction joints in pavements, and
back and front of kerb and channel,
between old and new pavements and

Chapter 11
Road Surface and Subsurface Drainage Design

behind bridge
illustrated).

Figure 11.4.1 - Sources of Moisture (Adapted from ARRB (1987))

January 2013

(not

It is important to note that in some flood


plains and low lying areas, a permanent,
high-level water table may exist. Subsoil
drains may be ineffective in such areas,
particularly where it is difficult to provide
an outlet. In some cases, such drains could
act in reverse and provide a means of access
for water to the pavement.

11

11-52

abutments

Department of Transport and Main Roads


Road Drainage Manual

In these circumstances, the most effective


measure which can be taken to control
subgrade moisture conditions is to raise the
subgrade above the surface of the ground.
A height of 1.2 m above the water table is
suggested (Earley 1979). This is usually
impossible in urban street construction, in
which case the pavement design should take
into consideration the soaked conditions. In
some situations, a cement or bituminous
stabilised subbase and/or base may be used.
Reference should be made to the
departments Pavement Design Manual.
Roads on a thick layer of soft, compressible
clay also need special consideration and
geotechnical advice should be sought for
requirements such as preloading and other
possible drainage mechanisms.

11.4.2 Control of Road Moisture


The three basic techniques for controlling
moisture are:
Layer protection: For example, seal
coats, plastic sheeting and other
impermeable barriers placed at
various levels in the pavement
structure. The durability of this type
of moisture control is suspect, and if
the barrier lets some moisture in,
pavement failure is likely.
Rendering Subgrade Insensitive:
Lime or cement stabilisation are
examples of this technique. The
advantage is that the load capacity of
the stabilised material does not
significantly decrease with increasing
moisture content. The disadvantages
are additional expense and a
significant reduction in permeability
(provided that cracks do not
develop).

Chapter 11
Road Surface and Subsurface Drainage Design

Subsurface drainage: Removal of


moisture from the pavement structure
via a subsurface drainage system. A
correctly designed and maintained
subsurface drainage system is the
only way of ensuring a stable
moisture condition.
This section is concerned only with the
subsurface drainage technique.

11.4.3 Types of Subsurface


Drainage
Subsurface drainage systems are generally
installed in a road either to remove water
from the subgrade and pavement materials
or to intercept water before it reaches the
road structure.
The former type is known as a pavement
drain, and the latter is called a cut-off drain
or a formation drain.
Consideration of the permeability and
capillary moisture characteristic of the
material surrounding the pavement is a
major factor in assessing the need and type
of subsurface drainage required.
Suitable drainage systems for various
conditions are presented below.
(a) Drainage for Surface Infiltration
Figure 11.4.3(a) illustrates the type of
subsoil drainage suitable for a permeable
base and surface contained in relatively
impermeable material.
Figure 11.4.3(b) illustrates an embankment
with the permeable base and surface on a
relatively impermeable subgrade. A free
draining layer is provided in the shoulders
below a low permeability material.
A variation is to carry the full permeable
base course over the full width of the
shoulders.

January 2013

11-53

11

Department of Transport and Main Roads


Road Drainage Manual

(b) Groundwater
A static water table may be lowered by
using either drainage trenches shown in
Figure 11.4.3(c) or a horizontal filter
blanket shown in Figure 11.4.3(d).
The horizontal filter blanket will also act as
an intercepting barrier for capillary
moisture in some situations.

Chapter 11
Road Surface and Subsurface Drainage Design

subsurface flow into a drainage pipe before


it can enter the road structure. The trench
should be excavated to at least the depth of
the permeable strata.
Upward flow from a pervious aquifer is
usually controlled by constructing a
horizontal filter blanket in the base of the
excavation as shown on Figure 11.4.3(f).

If water flows along an inclined permeable


layer, as shown in Figure 11.4.3(e), a trench
should be constructed to divert the

Figure 11.4.3(a) - Drainage for Surface Infiltration with Subsoil Drains

11

Figure 11.4.3(b) - Drainage for Surface Infiltration with Free Draining Layer

January 2013

11-54

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.4.3(c) - Drainage Trenches to Lower Water Table

Figure 11.4.3(d) - Horizontal Filter Blanket to Lower Water Table

11
Figure 11.4.3(e) - Trenches to Intercept Flow through an Inclined Permeable Layer

January 2013

11-55

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.4.3(f) - Permeable Filter to Lower the Effect of Head from a Permeable Aquifer

(c) Standard Drains


Subsoil drains, described in Standard
Specification MRTS03 (TMR 2010c), are
shown in Figure 11.4.3(g). The depth of
these drains may be increased to suit the
particular installation.
Subsurface drain pipes may be surrounded
by a single stage filter, or by two stage
filters. Filter materials can consist of
aggregates (ranging in size from sand to
cobble size), geotextiles or combinations of
aggregates and geotextiles. The level of
filtering will be determined by the
prevailing soil types and any environmental
requirements on the discharge. In some
cases a second stage filtering may be
required and this can take the form of a
geotextile wrap either around the pipe or
around all the filter material.

11

A more recent form is the geocomposite


edge drain sometimes known as a strip filter
or fin drain. These are prefabricated with a
polymer core wrapped in a geotextile. They
can be installed in much narrower trenches
than traditional pipe-based drains.
Material requirements are contained in
MRTS03 (TMR 2010c) including those for
sheet filter drains, trench backfill, fibre
January 2013

11-56

reinforced concrete pipes, corrugated steel


pipes, polyvinylchloride pipes, and plastic
pipes (perforated and unperforated). See
also Section 11.4.4 on filter materials.
The departments Standard Specification
MRTS04 (TMR 2010c) is also relevant.
Figure 11.4.3(h) shows typical subsoil drain
outlets and cleanouts in an urban
environment. Standard Drawing No. 1116
(DMR 2009b) provides further details
including treatments for rural environments.
The outlets of subsoil drains discharging
into gully pits, manholes, or culvert
endwalls are preferred. Outlets discharging
to natural surface should be made
accessible for maintenance operations and a
concrete headwall should be constructed
together with a small area concreted or
rockpitched around it as shown on the
standard drawing. To aid finding the outlet
a timber marker post should be maintained.
Accurate records of the position, depth and
type of subsoil drains which are installed
should be maintained.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

the surrounding material, but at the same


time fine enough to keep that material in
place.
In addition the filter should be stable under
flow situations and should itself be
prevented from washing into perforations or
joints in drainage pipes.
These requirements can be satisfied in
various ways, usually by either granular
materials or synthetic filter fabrics
(geotextiles).

Notes: 1 Minimum cover for various compactors


unless approved otherwise:
Hand held compactors - 100 mm
Compactors < 15 tonnes 200 mm
Compactors > 15 tonnes 200 mm
2. All dimensions are in millimetres.

Figure 11.4.3(g) - Standard Subsoil


Drains

11.4.4 Requirements of Filter


Materials
This section contains extracts from the
Guide to the Control of Moisture in Roads
(NAASRA 1983) and applies to most
drains.
A filter material is required in any
permanent subsurface drainage system to
prevent fine soil particles from washing into
the system. For satisfactory performance a
filter material must be more permeable than

Filter materials are not usually necessary in


temporary drainage systems or where the
surrounding soil is known to be very stable.
Examples of stable material are fractured
rock, fissured or jointed heavy clays or
other weathered materials and naturally or
artificially cemented materials. Care should
be taken to determine whether fissured or
jointed materials are sufficiently stable
under adverse conditions to warrant
dispensing with a filter material. Water
flowing from joints should be examined for
suspended particles, and the susceptibility
of the material to erosion in the disturbed or
undisturbed state determined.
Standard Specification MRTS27 (TMR
2010c) describes the material requirements
and work to be carried out for the relevant
Geotextiles in drains and trenches and
drainage blankets, and Geotextiles under or
within embankments.
The design of granular filter material is
described in Subsurface Drainage of Road
Structures (ARRB 1987).

11

January 2013

11-57

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Note: The pavement base course may be more permeable than the subbase.
Relative permeabilities should be considered in locating the drains,

Figure 11.4.3(h) - Subsoil Drain Outlets and Cleanouts

11.4.5 Design Procedure


The steps involved in designing a
subsurface or subsoil drainage system are:
1.

2.

11

3.

Decide on an appropriate materials


testing and site investigation program
for the project.
After groundwater investigation, carry
out hydraulic design of cut-off drains
(Section 11.4.9).
At the road grading stage, ensure that
fills are high enough to inhibit
capillary rise (Section 11.4.17) and to
allow for subsurface drain outlets, and
that cuts can be properly drained.

January 2013

11-58

4.

5.
6.
7.

8.

Arrange for pavement design. The


pavement depth must be known to set
subsurface drain levels.
Obtain a copy of the road geometry.
Select the appropriate locations of
subsurface drains from Section 11.4.6.
If subsurface drains can be placed
parallel to the road surface in the
vertical plane, subsurface drain
detailing can follow stormwater
drainage design.
Where the longitudinal grade of the
road is very flat (less than 0.5%
grade), and there is a need for
independent grading of subsurface
drains, both drainage systems could be

Department of Transport and Main Roads


Road Drainage Manual

designed concurrently if appropriate to


do so.
9. Identify cut-to-fill lines and locate the
transverse drains.
10. Detail the locations of inlets and outlet
pits where these do not coincide with
stormwater pits.
Drainage facilities should be designed and
constructed recognising that periodic
inspection and repair will be required and
provide for the safety of maintenance
personnel as well as for road users.
Investigation of potentially cracked or
failed underground pipes should be carried
out using a remote television camera to
reduce the risk to inspection personnel.

11.4.6 Location of Subsoil Drains

Chapter 11
Road Surface and Subsurface Drainage Design

Subsurface drains should be placed along:


the low sides of pavements;
both sides of the pavement near any
cut-to-fill line;
both sides of a kerbed pavement;
the high side of pavement where
seepage is evident, or where water
may enter from batters, full-width
pavement, service trenches or
abutting properties; and
joins between an existing pavement
and a pavement widening where
pavement depths or permeabilities
could create a moisture trap.
In medians, subsurface drains should be
considered along the:

Where moisture ingress is unlikely, heavy


commercial traffic is light, and similar
pavement designs in the vicinity of the
project
have
already
performed
satisfactorily without subsurface drainage,
then subsoil drains for new projects may be
omitted. It is difficult to describe all
circumstances that warrant the installation
of subsurface drains, but where soils are not
free draining (i.e. clays, silts, loams) or
where there is a likelihood of water ponding
near the pavement, subsurface drains should
be considered.

low side of a dished median where


the median drain invert level is less
than 0.2 m below subgrade level of
the adjacent pavement;

While provision of subsurface drains


without design may appear excessive, it is
prudent to provide drains extensively on
main roads where soils are not free
draining. Omission of subsurface drains on
main roads has caused premature pavement
failure and considerable expense in
installing them afterward. The following
guidelines are offered where seepage is not
obvious.

Figure 11.4.6(a) shows typical locations of


subsoil drains in a divided road in a cutting.
The width and nature of the median
determines the number of subsoil drains
required.

low side of a kerbed median where


the cross-slope is 10% or more;
sides of a median with a fixed
watering system or wider than 6 m;
and
centre of flat grassed medians
without fixed watering systems and
less than 6 m wide.

Grassed medians can provide a means by


which water can enter the pavement or
subgrade. Medians should therefore be
constructed of a material of low
permeability (eg. a compacted soil
aggregate as recommended for shoulders)

January 2013

11-59

11

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

except for 100 mm of topsoil to grow grass.


Provision of an impermeable membrane
under the median should be considered
also.

entry of water from grassed medians, or


where there is a significant difference in
level between roadways, or where
permeable subsoil surface strata exists.

However, where median planting other than


grass is required for aesthetic or headlight
screening reasons, the low permeability
material or impermeable membrane will
inhibit growth and should not be used.

Figure 11.4.6(b) shows typical locations of


subsoil drains in a low embankment or
transition zone from embankment to cut.

Longitudinal subsoil drainage should be


provided where there is a possibility of

Figure 11.4.6(c) shows a typical example in


a cutting where subsoil drains are often
required.

See Note

Note: If invert of median drain is not much lower than pavement layers and/or the possibility of seepage from median
back under pavement exists, a subsoil drain should be considered here.

Figure 11.4.6(a) - Location of Subsoil Drains (Divided Road)

High Side

Low Side

11
Figure 11.4.6(b) - Subsoil Drains - Low Embankment or Transition from Embankment to
Cut.

January 2013

11-60

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

11

Figure 11.4.6(c) - Subsoil Drains in Cuttings

January 2013

11-61

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

11.4.7 Transverse Subsurface


Drains

a herringbone pattern if necessary to


achieve the minimum grade.

Transverse, strip filter subsurface drains, as


shown in Figure 11.4.7, should be
considered:
on the upstream side of cut-to-fill
lines;
along changes of pavement depth or
permeability; and
at both ends of bridge approach slabs.

11.4.8 Cut-off Drains


Subsurface cut-off
considered:

drains

should

be

along both sides of cuts where the


road is known to be below the water
table, or where seepage is
encountered during construction, or
where seepage is expected in wet
weather;
transversely at any seepage areas,
and further downgrade if required.
The transverse drains may be laid in

11.4.9 Design of Cut-off Drains


Simple analysis consisting of homogeneous
layers of differing permeability rarely
applies to natural conditions. Fissures,
joints, faults and bedding planes in soil or
rock structures can have large hydrostatic
head differences over short distances that
may vary rapidly. Strategic placement of
piezometers and standpipes is therefore of
the utmost importance. Theoretical models
can give good results only if the ground
conditions input during design, are close to
those in the field.
Road surfaces are more permeable than
generally imagined, and the quantity of
water entering a pavement (infiltration
rate), may be estimated by multiplying the
infiltration coefficient (refer Table 11.4.9)
by the two year, one hour rainfall intensity
over the surface area.

11
Figure 11.4.7 - Transverse, Strip Filter Subsurface Drain

January 2013

11-62

Department of Transport and Main Roads


Road Drainage Manual

Once the infiltration rate is estimated (in


m/s), and the coefficient of permeability has
been determined by laboratory testing, the
quantity of water entering the road or the
inflow is determined by applying Darcys
law:

qm kAi
Where:
qm = flow rate entering the surface
(m3/s);
k = permeability or infiltration rate
(m/s);
A = area of pavement (taken as one
metre square in this application); and
i = hydraulic gradient, i.e. head of
water divided by length of drainage
path (m/m).
A hydraulic gradient of unity is suggested
for rain falling on a surface. With a
hydraulic gradient of unity the inflow as
calculated from the above equation is equal
to the infiltration rate multiplied by the
surface area of the pavement.

Table 11.4.9 - Surface Infiltration


Coefficient

Chapter 11
Road Surface and Subsurface Drainage Design

However, to avoid the system failing due to


partial blocking of the drain, the drain
should be designed to carry at least three
times the expected flow. Generally, the
diameter of subsurface drains should be no
less than 100 mm. In some cases a
combined stormwater and subsurface
drainage system may be used.

11.4.11 Materials
Subsurface drains are manufactured from a
range of materials but all require some form
of perforation to allow subsurface water to
enter the pipe.
Corrugated polyethylene agricultural drain
is the cheapest material. Pipes with a
diameter of 90 mm are regarded as the
minimum for roadwork.
Smooth polyvinyl chloride (PVC) pipe is
used to convey flows across pavement, or
may be used where longitudinal gradients
are flatter than 0.5%. Since this material is
expensive, herringbone grading of
corrugated pipe is typically applied. Pipe
sizes generally range from 100 mm
diameter to 300 mm diameter.

11.4.10 Size of Drain

Prefabricated polyethylene (PE) drain, also


known as strip or fin drain, may be laid in
batters or parallel to the pavement to
intercept groundwater. It may be used
across a pavement if the trench is backfilled
with no-fines concrete. This material has
less
hydraulic
capacity
than
the
corresponding diameter of pipe, so this may
have to be checked. Depths of the fin
drains are typically 200 mm to 450 mm in
depth.

The grade line of the drain is generally


known. There are a number of design
charts available which can be used to size
the drain given the slope and the volume of
water that the drains are required to carry.

Concrete pipes are


groundwater flows
available in plastic
common use range
mm in diameter.

Surface type

Infiltration
coefficient

Sprayed seal

0.2 0.25

Asphalt

0.2 0.4

Cement concrete

0.3 0.4

Unsealed shoulders

0.4 0.6

typically used where


require diameters not
pipes. Pipe sizes in
from 300 mm to 750

January 2013

11-63

11

Department of Transport and Main Roads


Road Drainage Manual

Perforated corrugated steel is typically used


for deep cut-off drains where the soil and
groundwater are not highly corrosive. They
require specific structural design.

11.4.12 Access to Subsurface


Drains
Inlets and outlets for subsurface drains
should be located clear of the traffic lanes.
Where the inlet must be located in the
shoulder, a pit with a trafficable steel cover
should be used. The inlet should not be
located in a position where it would be
possible for stormwater to enter the
subsurface drainage system.
Figure 11.4.3(h) shows typical subsoil drain
outlets and cleanouts in an urban
environment. Standard Drawing No. 1116
(DMR 2009b) provides further details
including treatments for rural environments.
Pits for subsurface drainage should be
spaced not further than 150 m apart for ease
of inspection and cleaning of the pipes.
Maximum spacing between a cleanout and
an outlet should generally not exceed 120 m
to facilitate inspection and flushing. In
cuttings where groundwater is not present,
the distance to the outlet of a pavement
drain may be much greater, but
intermediate pits should generally be placed
at a maximum spacing of 120 m.
Where groundwater occurs in a cutting, the
seepage should be conveyed from the
subsurface drain into an impervious
collector pipe to minimise water penetration
of pavement remote from the problem area.

11

Outlets should be in areas that are easily


accessible and, where possible, visible to
personnel standing on the road surface. An
outlet should not hinder road maintenance

January 2013

11-64

Chapter 11
Road Surface and Subsurface Drainage Design

activities such as cleaning unlined table


drains or grass cutting.
Outlets should be provided with some form
of erosion protection commonly referred to
as a splash zone. Typically, this consists of
either:
a masonry or concrete apron; or
an area of large aggregate to dissipate
the outflow energy.

11.4.13 Lowering of Ground


Watertable
The lowering of a static watertable is
achieved through the use of a system of
vertical cut-off drains below the road
pavement as shown in Figure 11.4.13.
Table 11.4.13 is used for a preliminary
assessment of the effectiveness of a
proposed
trench
drainage
system.
However, if the subgrade permeability is
less than 100 nm/s then vertical drains on
both sides of the roadway are unlikely to be
effective in lowering a water table. An
alternative solution to subsoil drainage
should be adopted, such as pavement design
based on saturated subgrade strength.

11.4.14 Schilfgaardes Method


Schilfgaardes method (Austroads 2008a)
can be used to determine the drain spacing
that will lower the water table by mo m
(see equation below). The accuracy of the
solution is extremely dependent on the
accuracy of k and f (i.e. permeability and
drainable pore space of the subgrade). The
geometry of the drainage problem and the
effect of a typical subsurface pipe
arrangement is shown in Figure 11.4.14(a).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.4.13 - Typical Groundwater Drainage System

Table 11.4.13 - Effectiveness of Trench Drainage Systems


Amount of lowering in (m) after different periods midway between two 1
Coefficient
metre deep trenches
of
permeability
(initial watertable 1 m above bottom of trenches)

k (nm/s)

Trench spacing 3 m

Trench spacing 10 m

Trench spacing 20 m

3 months

3 months

3 months

1 year
1.00

1.00

1 year
1.00

1.00

1 year

10000

1.00

1.00

1000

1.00

1.00

0.93

1.00

0.48

0.93

100

0.94

1.00

0.23

0.65

0.06

0.23

10

0.25

0.68

0.03

0.10

0.00

0.03

1.0

0.03

0.11

0.00

0.00

0.00

0.00

0.1

0.00

0.00

0.00

0.00

0.00

0.00

Source: Adapted from Golden (1979) (as reproduced in Austroads 2008a)

The Schilfgaarde Equation is:


1

k d e m o d e m t 2

L 3 j
2 f m o m

Where:
L = spacing between drains (m);
j = geometrical factor (determined
from Figure 11.4.14(b);
k = saturated permeability (m/s);
de = equivalent depth of drain to
impervious barrier (m). Differs from
d because of convergence of the flow
lines;
d = height of drain above impervious
barrier (m);

mo = depth of drain below original


watertable (m);
m = depth of drain below lowered
watertable (m);
mo m = distance watertable is
lowered (m);
t = time to lower water table (sec);
and
f = drainable pore space, expressed as
a fraction of total volume drained at
600 mm tension (typically clays
range from 0.03 to 0.11, well
structured loams from 0.10 to 0.15
and sands range from 0.18 to 0.35).
The solution requires that a starting
estimate of L (drain spacing) be input with
known values of d, mo, m, k, t and f. The

January 2013

11-65

11

Department of Transport and Main Roads


Road Drainage Manual

equivalent depth, de is estimated from


Figure 11.4.14(b). This is then used to
calculate the following convergence factor:

d e d e mo

Chapter 11
Road Surface and Subsurface Drainage Design

recalculated and if different from the initial


guess, a further iteration of calculations is
commenced with the revised equivalent
depth.

The convergence factor is then used with


Figure 11.4.14(c) to estimate j. L is

Figure 11.4.14(a) - Geometry of the Drainage Problem and Effect of Subsurface Drains

11
Figure 11.4.14(b) Equivalent Depth for Convergence Correction

January 2013

11-66

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

Figure 11.4.14(c) Dependence of Factor j on Depth to Impervious Layer

11.4.15 Draining an Inclined


Aquifer
The design of a subsurface drainage system
which intercepts an inclined aquifer as
shown in Figure 11.4.15 is relatively
straightforward. Darcys law governs the
discharge from the aquifer.
Darcys law is:

qm kAi
Where:
qm = flow rate per unit length of
trench (m3/m);
k = permeability of the aquifer (m/s);
A = area of aquifer (taken as
thickness of aquifer in this
application, since the discharge
required is per unit metre of length);
and
i = slope of the aquifer (when the
piezometric heads within the aquifer
are equal) (m/m).

To ensure that the subsurface drainage


systems intercept all of the seepage, the
permeability of the filter material and the
width of the trench need to be checked.
This implies that the piezometric head must
drop to zero within the trench filter
material.
The principle is shown
mathematically below.
The ratio of the permeability of the aquifer
material divided by the permeability of the
filter material is:

tan B tan A k a k f
Where:
tan (B) = can be approximated by
W/T (width of the trench divided by
the thickness of the aquifer);
tan (A) = can be approximated by the
slope of the aquifer (shown as s in
Figure 11.4.15;
ka = permeability of aquifer material
(m/s); and
kf = permeability of filter material
(m/s).

January 2013

11-67

11

Department of Transport and Main Roads


Road Drainage Manual

The nomenclature used is shown in Figure


11.4.15.
There are two values, which can be altered
by the drainage designer (W or kf) to
balance the permeability ratio equation.
However, trench width is normally fixed to
a standard value (typically 300 mm) and so
it then becomes a case of selecting filter
material to ensure that the ratio tan (B)/tan
(A) is less than ka/kf.

11.4.16 Design of a Filter Blanket


to Lower a Water Table
Where a pavement is to be placed within or
below the natural groundwater level it may
be necessary to lower the water table. This
can be achieved by placing a horizontal
filter blanket below the pavement. The
design of the filter blanket should be
undertaken using analytical procedures such
as flownet procedures and finite element
methods. These analytical procedures are
beyond the scope of this document and are
usually undertaken by a geotechnical expert

11

Chapter 11
Road Surface and Subsurface Drainage Design

(refer to Guide to Road Design - Part 7:


Geotechnical Investigation and Design.
(Austroads 2008b).

11.4.17 Capillary Rise in Soils


Where a shallow formation is proposed
over saturated ground, or fine-grained
embankments crossing swamps, the height
of capillary rise of the groundwater should
be calculated to ensure that excess water
does not enter the pavement. The rise in
capillary water can be calculated using:

hc

10C
eD10

Where:
hc = capillary rise (mm);
C = an empirical constant that
depends on the shape of the grains
and varies from 0.1 to 0.5 cm2 (for
perfect spheres, C = 0.1 cm2); and

Source: Adapted from ARRB (1987)


Figure 11.4.15 - Trench Excavated Through an Inclined Aquifer

January 2013

11-68

Department of Transport and Main Roads


Road Drainage Manual

Chapter 11
Road Surface and Subsurface Drainage Design

D10 = Allen Hazens effective grain


size based on the sieve opening in cm
that 10% of the material passes. The
value is obtained from the grading
curve; and
e = void ratio.
The void ratio can be determined using:

VV
V VV
Where:
VV = total volume of voids (units).;
and
V = total volume (units).

11

January 2013

11-69

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

Appendix 11A
Worked Examples

11
A
March 2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

Appendix 11A Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

11
A

March 2010

ii

Reference
Section
-

Description of Revision

Authorised
by

Date

Initial Release of 2nd Ed of manual.

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

Aquaplaning Example
This example describes the process to determine the water film depth at key points along a
design pavement surface and to assess the aquaplaning potential.
The example commences after obtaining the road surface contours over a superelevation
transition section of the road and several drainage paths identified.
The tasks for this example is, given the data below, estimate the water film depth at key points
along the path and assess any aquaplaning potential.
Refer to Section Chapter 11.3.

Project Data
Design speed = 100 km/h;
Surfacing is Dense Graded Asphalt; and
Road surface contours and cross section - refer figures below.

DP3

DP2

DP4
Shoulder

DP1
Shoulder

11
A
March 2010

11A-1

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

Part of Type Cross Section

Step 1.
Review the contour plan. Four drainage paths have been identified:
For drainage path 1 (DP1), this path, while some 26m long, flows from one side to the
other and is not the critical path.
The remaining three paths (DP 2 to 4) all start on one side, travel towards the other side
and then turn to drain off the same side as they started.
The longest path is considered the critical path, therefore drainage path 2 (DP2) will be
analysed.
Extract drainage path profile for DP2:
Ht (m)

Distance
(m)

Start

6.68

0.0

start of flattest section

6.05

24.2

end of flattest section

5.75

17.7

Outer Wheel Path - inside lane

5.55

10.5

Outer Wheel Path - middle lane

5.40

6.4

Outer Wheel Path - outside lane

5.20

7.2

Point

Step 2.
Calculate the water film depth (D) for the longest drainage path (DP2).

11
A

March 2010

11A-2

0.103 T 0.11 L0.43 I 0.59


S 0.42

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

For Texture Depth T, refer Section and Table 11.3.7.3. For this example, the pavement is dense
graded asphalt, therefore use 0.5mm for T.
The Rainfall Intensity I, adopted for analysis is 50 mm/hr - refer Section 11.3.7.5.
For Drainage Path Length L, and Drainage Path slope S, refer Section 11.3.7.4. The slope to
each point assessed is the calculated Equal Area Slope (Se) as per the procedure shown in
Section 11.3.7.4. Summary of calculations is presented in the following table.

Chainage

Dist
(m)

Ht
(m)

Ht Diff.

Total A
(m2)

EAO
(m)

Se
(%)

6.68

24.2

24.2

6.05

0.63

7.62

0.63

2.60

41.9

17.7

5.75

.30

17.54

.837

2.00

52.4

10.5

5.55

.2

26.97

1.029

1.96

58.8

6.4

5.40

.15

35.31

1.201

2.04

66.0

7.2

5.20

.2

47.79

1.448

2.19

Where:
Total A is the total area under the profile.
EAO is the Equal Area Ordinate.
Se is the Equal Area Slope from the start of path to point of assessment.

With all variables determined, calculate the water film depth at each point:

0.103 0.5 0.11 66 0.43 50 0.59


0.5
2.19 0.42

(equation for last point)

= 3.68 mm
Summary of calculations for all points assessed:
Chainage

Water Film Depth (mm)

24.2

2.03

41.9

3.08

52.4

3.47

58.8

3.60

66.0

3.68

11
A
March 2010

11A-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 11A
Worked Examples

Step 3.
The assessment criteria for aquaplaning potential over drainage path 2 (DP2), is discussed in
Section 11.3.8.
The criteria that applies to this section is a maximum water film depth of 2.5 mm (desirable) to
4.0 mm (absolute).
It can be seen that all three outer wheel path points exceed the desirable limit, but are below the
absolute limit of 4 mm.

Step 4.
Every effort to comply with desirable limits should be made, therefore review of the geometrics
is required - refer Section 11.3.12 and the Road Planning & Design Manual for guidance.
It should be noted that on high speed, wide flat pavements, it can be difficult (near impossible)
to achieve the 2.5 mm desirable limit, however experience has shown that depths of about 3.25
mm are achievable.

End of Example

11
A

March 2010

11A-4

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Chapter 12
Basins

12
March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Chapter 12 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

12
March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Table of Contents
12.1

Overview

12-1

12.2

Detention Basins

12-1

12.3

12.4

12.2.1

Introduction

12-1

12.2.2

Advantages and Disadvantages of Detention Basins

12-2

12.2.3

Initial Design and Feasibility

12-3

12.2.4

Simple Hydrologic Method of Routing

12-4

12.2.5

Other Design Considerations

12-8

Extended Detention Basin

12-12

12.3.1

Description

12-12

12.3.2

Design Guidelines

12-12

Sediment Basins

12-13

12.4.1

Description

12-13

12.4.2

Design Procedure

12-14

12

12
March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

12
March 2010

iv

Chapter 12
Basins

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Chapter 12
Basins
12.1

Overview

In the context of this chapter, design


guidance is provided for basin systems,
which include detention, extended detention
and sediment basins. Retention basins are
rarely used on departmental projects and
their design is specialised, therefore they
not covered in detail in this manual.
Generally:
detention basins and extended
detention basins are designed to
reduce and delay peak flood flows;
sediment basins are designed to trap
and retain a range of sediment
particle sizes, thereby reducing both
coarse sediment and turbidity values
from the inflow; and
retention basins are designed to retain
some or all of the flow allowing it to
infiltrate into the soil.

12.2

Detention Basins

12.2.1 Introduction
Detention basins are basins or reservoirs
wherein water is temporarily stored for the
purpose of regulating a flood.
Water can be released from the basin
through its outlet which may be a weir,
culvert or pipe. The storage volume may be
recessed below natural ground level or
above natural ground level with the road
embankment acting as a dam wall.
In road design, a large basin is usually
located in a valley where the road crosses a

watercourse.
Smaller basins are often
located in those small triangular areas of
land isolated by on-ramps and off-ramps,
and occasionally in the centre of major
roundabouts.
There are basically two types of detention
basin: a dry basin and a wet basin.
Between flood events, a dry basin
generally has a dry bed which is often
achieved with the use of a low-flow channel
or pipe system. Thus, these basins can be
used for other land uses such as sporting
activities and open space. Detention basins
generally provide only minimal water
quality improvement.
A wet basin, however, incorporates a
permanent pond and retains some water in
the bed of the basin between flood events.
These basins provide flood attenuation
benefits during a range of flood events and
water quality benefits during more frequent
minor flood flows and regular storms. The
permanent pond within a wet basin may
consist of a lake, wetland or water quality
pond (refer Chapter 7).
Some basins may contain a temporary or
permanent sediment basin.
These are
discussed in detail in Section 12.4. In road
construction, most detention basins are
initially constructed and operated as
temporary sediment basins.
Upon
completion of the road works the basins are
converted into either a permanent, wet or
dry detention basins. Most sediment basins
will have substantial sediment trapping
capabilities, whilst the trapping ability of

March 2010

12-1

12

Department of Transport and Main Roads


Road Drainage Manual

detention basins will vary in accordance


with basin hydraulics.
The five important characteristics of basins
can be summarised as follows:
1. Flood Attenuation Features
Both detention and retention basins can
provide the benefits of flood attenuation.
The degree of attenuation depends on the
storage volume of the basin and the design
of the basins outlet system.
2. Pollution Control Features
In road construction, the primary role of a
detention or retention basin can shift from
flood attenuation to water quality
improvement and occasionally emergency
spill or pollution containment areas.
3. Ecological Features
Wet basins, with their incorporated
wetlands or lakes, can provide significant
stormwater treatment benefits as well as the
provision of wildlife habitat.
4. Potential Impact on Downstream
Channel Erosion.
The primary purpose of a flood control
system is to reduce the peak height of
overbank floods. To achieve this, detention
basins must extend the duration of inbank
flows.
It is noted that channel erosion can only be
caused by inbank flows, and the extent of
this erosion is directly related to the
velocity and duration of these flows.
Bankfull flow is usually used as the
indicator of potential channel erosion.

12

Thus, even though detention basins


marginally reduce the peak flood velocity,
they significantly increase the duration of
bankfull flows.
Therefore, increased
channel erosion is usually expected
downstream of flood control systems.

March 2010

12-2

Chapter 12
Basins

5. Required Land Area


The selection of detention basins as a
means for mitigation requires consideration
of how much land is required. This could
be an important factor for basins in both
urban and rural areas.

12.2.2 Advantages and


Disadvantages of
Detention Basins
The advantages and disadvantages of
detention basins need to be weighed against
the cost of constructing the basins.
Advantages of detention basins include:
reduced outflows which is important
where the downstream channel
capacity is limited, particularly in
urban areas;
decreased
downstream
channel
erosion (i.e. when the basin is
designed to significantly attenuate
the flood events smaller than the
bankfull capacity of the downstream
channel)
by
reducing
peak
discharges, though note below
disadvantages of longer duration of
flow;
sedimentation control;
pollutant control;
recharging of groundwater resources;
storage of water for irrigation of
public land; and/or
provision of wildlife habitat.
Disadvantages of detention basins include:
cost of obtaining land outside the
road boundaries;
a lower peaked but attenuated
may
produce
hydrograph
unacceptable times of submergence

Department of Transport and Main Roads


Road Drainage Manual

over water sensitive crops and


ecological systems (probably only a
consideration for very large basins);
longer duration of flow may saturate
banks downstream in some cases,
and erosion risk is increased;
the need to provide safety warning
signs and precautionary measures;
and/or
maintenance requirements including
possible weed and mosquito control.

12.2.3 Initial Design and


Feasibility
Preliminary design may be conducted using
a simple set of calculations. This will
determine if a detention basin is feasible, or
if there is any significant flood mitigation
effect to be obtained from such a basin.
With reference to Figure 12.2.3, the steps
are:
1. Convert the inflow hydrograph into a
simple triangular shape.
2. Estimate the allowable peak outflow
discharge, Qo and mark it on the
recession or falling limb of the
hydrograph.
3. Draw the outflow triangular hydrograph
(the recession limb is nominal only).
4. Initial sizing of a basin volume (Vs) can
be undertaken by a comparison of the
following
estimation
procedures
(QUDM 2008):

Chapter 12
Basins

Vs r 3 5r

Vi
8

(Carroll 1990)

Vs r 2 r

Vi
3

(Basha 1994)

Where:
r is reduction ratio calculated as:

Qi Qo
Qi

Vs = required storage volume (max.)


(m);
Vi = inflow volume (m);
Qi = peak inflow discharge (m/s);
and
Qo = peak outflow discharge (m/s).
If the rational method is used for the
determination of Qi, then the initial estimate
of the inflow volume (Vi) may be
determined as:

Vi

4t c rQi
3

Where:
tc = time of concentration (s)
The above equations are most appropriate
when it is necessary to limit the peak
discharge for only the nominated design
storm, such as the ARI 100 year event. In
those circumstances, where it is necessary
to ensure the post development peak
discharge for each tested storm duration is
not increased, then these equations are
likely to underestimate the required
detention volume.

Vs r 1 2r

Vi
3

(Culp 1948)

Vs
r
Vi

5. Calculate the required area and depth of


the retention basin.

(Boyd 1989)

If this assessment shows that the basin is


feasible, additional investigations should be
completed and flood routing as described in

March 2010

12-3

12

Department of Transport and Main Roads


Road Drainage Manual

Section 12.2.4 should be carried out to


further develop the design.
Additional investigations required to be
completed include:
Determination of groundwater level
within the proposed basin area;
Nominate the maximum basin depth;
and
Determination
of
maximum
allowable flood level within the
basin.

Chapter 12
Basins

In some locations, recessing a basin into the


ground and the consequential removal of
trees during basin construction may cause
long-term salt intrusion of the basin floor.
This can result in very poor grass cover. In
such cases, seek advice from the local
authority or regional Department of
Employment, Economic Development and
Innovation (formerly Department of
Primary Industries and Fisheries) office.
(b) Nominate Maximum Basin Depth
A maximum depth of 1.2 m during an ARI
20 year flood is generally recommended for
publicly accessible basins.
(c) Determine Maximum Allowable Flood
Level within the Basin
When determining the maximum water
level requirements for the detention basin
during various flood events, consideration
should be given to the following:
various
issues
raised
considerations (a) and (b) above;

in

elevation of the ARI 20 year flood


within the basin should be below the
road pavement box; and

Figure 12.2.3 - Triangular Shaped


Hydrographs

elevation of the ARI 50 year flood


within the basin should be 300 mm
below the road shoulder.

12.2.4 Simple Hydrologic Method


of Routing

(a) Determine Groundwater Level


If the basin is to be recessed into the
ground, then bore hole testing should be
used to determine the groundwater level.

12

Dry basins should be located above


groundwater level. The wet region of
basins may be located below groundwater
level.

March 2010

12-4

There are many computer programs which


will carry out flood routing procedures.
However, the basic inflow hydrograph, the
stage-discharge curve and the storage
curves described in detail for the Simple
Hydrologic Method of Routing in this
section, are all required as input into the
computer programs in one form or another.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Standard flood routing procedures are


incorporated into this method. However,
the modification of the inflow hydrograph
into a triangular shape simplifies the
calculations and this procedure should be
sufficiently accurate for the smaller
detention basins.
A brief description of the input data is given
before the design procedure.
(a) Storage Curve
This is a plot of water level (height) versus
volume of water stored in the detention
basin. A typical curve is shown in Figure
12.2.4(a).

Figure 12.2.4(b) - Typical Outflow Rating


Curve

Although usually under inlet control


initially, where there are culverts, it is also
necessary to have a rating curve for the
channel downstream of the basin. This is to
determine the tailwater level corresponding
to the outflow discharge and to allow
culverts to be checked against outlet control
as well as inlet control.
(c) Triangular Inflow Hydrograph

Figure 12.2.4(a) - Typical Storage Curve

(b) Outflow Rating Curve


This is a plot of the discharge from the
basin versus the water level (height) in the
basin.
The discharge from the basin may be the
total of flows over a spillway / weir plus the
flow through culverts, or simply the flow
from one of these structures only. A typical
outflow rating curve is shown in Figure
12.2.4(b).

The inflow hydrograph is a plot of


discharge versus time and the area under
the hydrograph curve represents the total
volume of runoff in the flood.
A simple triangular shape may be used to
represent the inflow hydrograph so that the
area of the triangle represents the same total
volume of runoff in the inflow flood.
Figure 12.2.4(c) gives a triangular
hydrograph where the base length tb, is
2.67tp, where tp is the time to peak. This is
a common shape for runoff hydrographs
and may be used if there is no better
calculated indication of the hydrograph
shape. The time of concentration, tc may

March 2010

12-5

12

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

also be adopted for the time to peak, tp, in


this simplified procedure.

Where:

I n,n 1

inflow

(m3)

volume

between points n at start of time


interval t, and n+1 at end of time
interval t on the time axis;
On = outflow discharge at point n
(m3/s); and
Sn = storage at point n (m3).
Step 1
Calculate the following curves:
Storage Curve (Figure 12.2.4(a));

Figure 12.2.4(c) - Division of Inflow


Hydrograph into Unit time Periods

Outflow Rating
12.2.4(b)); and

A curvilinear hydrograph may be


substituted for the triangular shape, but the
design procedure applies.
(d) Design Procedure
The fundamental storage relationship is that
in each time period:

S I O
Where:
S = increment of storage (min);
I = inflow during period (m3/min);
and
O = outflow
(m3/min).

during

period

For time period t:

S S n 1 S I n ,n 1

On t On 1t
2

This can be rearranged to give the flow /


storage equation:

O t
O t

I n,n 1 S n n S n 1 n 1
2
2

12
March 2010

12-6

Curve

Triangular
Inflow
(Figure 12.2.4(c)).

(Figure

Hydrograph

Step 2
Divide the inflow hydrograph into time
intervals as shown in Figure 12.2.4(c). In
smaller catchments, a time interval of 1
hour or less is often used. It would be rare
to use more than 2 to 4 hour intervals in
non dam size catchments. It is not essential
that the last time interval be equal to the
other intervals.
The routing periods should not exceed
about one quarter of the time to peak. The
larger the number of intervals that are used,
the more the accurate is the answers
obtained, particularly for larger basins.
Step 3
Calculate the volume of outflow from:
Outflow = Oo t
Where:
Oo = outflow discharge (m/s); and
t = unit time interval used in the
routing analysis (seconds).

Department of Transport and Main Roads


Road Drainage Manual

Plot the outflow volume versus height on


the storage curve previously calculated.
Both curves are illustrated in Figure
12.2.4(d).

Chapter 12
Basins

First Time Interval


During the first interval, where n and n+1
are the points along the time axis of the
inflow hydrograph and n = 0, the inflow
volume is equal to I0,1 (from Figure
12.2.4(c)).
At the beginning of the interval, i.e. at n =
0, the storage and outflow are both zero,
and the flow / storage equation reduces to:

I 0 ,1 S1

Figure 12.2.4(d) - Volumes of Storage


and Outflow for Varying Flow Depths

O1t
2

Plot these values as shown in Figure


12.2.4(f) (RL = reduced levels).
Second Time Interval

Step 4
For a range of heights, read off storage S,
and outflow and calculate and plot the
graphs shown in Figure 12.2.4(e) from the
data obtained in Step 3.

During the second time interval, where n


and n+1 are the points along the time axis
of the inflow hydrograph and n=1, the
inflow volume is equal to I 1, 2 .
At the beginning of this interval when n=1,
the flow / storage equation becomes:

Ot
Ot

I 1, 2 S1 1 S 2 2
2
2

Plot these values as shown in Figure


12.2.4(f).
Third and Subsequent Time Intervals

Figure 12.2.4(e) - Volumes of Storage


Adjusted by Half the Outflow Volume for
Varying Flow Depths

Step 5
Calculate the storage level / time curve
starting at the initial height and working
through the inflow hydrograph as described
below to determine the height at each time
interval using the flow / storage equation:

At the beginning of the third time interval,


n=2, the flow / storage equation becomes:

Ot
Ot

I 2, 3 S 2 1 S 3 3
2
2

Plot these values as shown in Figure


12.2.4(f).
This procedure is repeated until the total
inflow hydrograph has been put through the
routing procedure.

12
March 2010

12-7

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Figure 12.2.4(f) - Routing Procedure

Step 6
Plot the height versus time values from
Figure 12.2.4(f).
This becomes the outflow hydrograph
shown in Figure 12.2.4(g).

12.2.5 Other Design


Considerations
This section provides information on topics
that need to be considered in the design of
detention basins.
1.

Detention Embankment Design

All major fill embankments for detention


basins should be designed as dams and will
therefore require the same degree of
geotechnical and hydraulic scrutiny.
The minimum recommended embankment
crest widths are provided in Table 12.2.5(a)
Figure 12.2.4(g) - Outflow Hydrograph

Step 7
Convert the heights in Figure 12.2.4(g) to
discharges from the outflow rating curve.

12

A conventional outflow hydrograph of


discharge versus time may then be plotted.

March 2010

12-8

Internal batter gradients in detention basins


need to be consistent with the requirements
of personal safety and generally within the
following upper limits:
Where the permanent water depth is
less than 150 mm when surcharging,
1 on 2 to 1 on 4 on earth structures;
and vertical on rock gibber or gabion
basket structures.

Department of Transport and Main Roads


Road Drainage Manual

Where the permanent water depth is


between 150 mm and 1500 mm when
unfenced
and
surcharging,
a
maximum slope of 1 on 5.
Where the permanent water depth is
between 150 mm and 1500 mm when
fenced and surcharging or greater
than 1500 mm:
1 on 2 to 1 on 4, on earth structures;
1 on 1.5, on rock gibber structures;
4 on 1, on gabion basket structures;
and
4 on 1, on stacked (rough squared)
rock structures.

Chapter 12
Basins

2.

To allow for settlement, the design height


of the basin embankment should be
increased by 10% if hauling equipment will
be the sole means of compaction, or 5% if
compaction equipment is used.
3.

Fill embankment
height (m)

Minimum top width


(m)

<3

2.5

3-4.5

4.5-6

3.5

6-7.5

4.5

The actual bank gradient will depend on the


slipperiness of the saturated sediment, i.e.
whether or not a person can achieve a firm
footing and exit the basin.
Slippery
sediments should have less steep gradients,
in the order of 1 on 8 or even 1 on 10.
Otherwise, the basin should be fenced.
All batters that are accessible to the public
should have a maximum slope of 1 on 8.

Surface Drainage of Basin Floor

If the floor of a basin does not drain freely,


then grass cutting can be difficult during the
wet season. Some poorly drained areas
may even promote mosquito breeding.
In detention basins and in the dry areas of
retention basins, a minimum bed gradient of
2% or greater is recommended to allow for
efficient drainage of the basin.
4.

Table 12.2.5(a) - Minimum Embankment


Top Width for Detention Basins

Allowance for Settlement

Stormwater Treatment

If a detention basin is being designed, then


some degree of stormwater treatment and
pollution containment can still be achieved.
For the design of a wetland within a
detention basin, refer to Chapter 7.
In order to achieve pollution control
requirements, it may be necessary for the
surface area of the basin to be larger than
that determined for flood control.
Detention basins will usually require a
gross pollutant trap and/or trash rack
facility to be constructed at major inflow
points. For trash rack design, refer to
Chapter 7.
5.

Pollution Control

Some basins can be modified to act as


pollution traps to contain spills from a road
accident. A major chemical spill from a
truck accident can be fully contained within
a detention basin if the outlet has been
designed for temporary placement of
sandbags or stop boards.
In detention basins, an oil barrier should be
constructed around the primary outlet weir.

March 2010

12-9

12

Department of Transport and Main Roads


Road Drainage Manual

These oil barriers or guards usually consist


of a sheet of aluminium bolted around the
outlet at the weir level to force all flow to
pass under the barrier before passing over
the weir.
Consideration must be given to the
potential contamination of groundwater
supplies and the management of such a
problem.
6.

Design Primary Outlet Structure

The primary outlet is the outlet system that


will discharge water from the basin prior to
discharge over the emergency spillway.
Some basins will also contain under-ground
drainage
designed
to
bypass
a
predetermined flow rate past the basin.
The primary outlet may consist of an orifice
plate, siphon inlet, pipes or culverts.
Detailed hydraulic analysis of the outlet
structure is required to determine the
storage-discharge relationship of the basin.
This hydraulic analysis must make
allowance for issues such as inlet and outlet
control and possible debris blockage.
To achieve the desired hydraulic response,
a primary outlet may need to consist of a
multi-level outlet containing a range of
outlet pipes of different sizes and set at
different levels.
To minimise the risk of initiating or
adversely affecting downstream creek
erosion, the peak discharge from the
primary outlet should be less than the
bankfull discharge of the downstream
channel.
7.

12

Primary Inlet Protection

Where necessary, controls should be


installed to prevent or manage debris
blockage of the primary outlet. Debris
blockage should be considered when the
upstream catchment currently contains, or is
March 2010

12-10

Chapter 12
Basins

likely to contain in the future, significant


riparian vegetation or litter.
These outlets must also be designed to
minimise identified public safety hazards.
In many cases it will be desirable to
separate the screens that provide public
safety from those that are installed to
reduce debris blockage.
Protection can be achieved with the
installation of trash racks, bar screens or
fences such as pool fencing. In public
areas, a trash rack or bar screen is usually
surrounded by a pool safety fence.
Design criteria are provided in Table
12.2.5(b).
An anti-vortex plate should be attached to
the outlet if floating debris is likely to exist
in the basin and blockage of the outlet is
considered undesirable.
It should be noted that clearing a debrisblocked inlet while the basin is full of water
can be a very difficult and dangerous task.
8.

Outlet Pipe Design

To prevent piping failure, outlet pipes


should have spigot and socket rubber-ring
joints and lifting holes should be securely
sealed.
The bedding material should be carefully
specified to minimise permeability, and cutoff walls (anti-seep collars) should be
installed.
9.

Spillway Design

Where possible, locate the emergency


spillway in virgin soil at the side of the
basin or otherwise at a location least likely
to result in failure.

Department of Transport and Main Roads


Road Drainage Manual

Table 12.2.5(b) - Criteria for Outlet


Structures
Item

Criterion

Spacing of vertical
bars

125 mm (max)

Inclined spacing of
horizontal supports

600 mm (max)1

Net clear opening


area

3 times the
calculated outlet
area

Limiting velocity
through trash racks

0.6 m/s (not readily


accessible); 1.5 m/s
(accessible)

Chapter 12
Basins

from Guidelines on Acceptable Flood


Capacity for Dams (NR&W 2007).
Designers should refer to the local authority
to determine the desirable spillway
capacity. Also refer to ANCOLD (2000).
10. Outlet Protection Design
Scour protection and/or energy dissipation
should be placed on the outlets of the
primary outlet and the emergency spillway.
Refer to Section 9.15 for the design of
scour protection at the outlet of culverts.
11. Wildlife Migration

Source: QUDM (2008)

Consideration must be given to the


necessary migration of aquatic and
terrestrial wildlife through the basin (refer
Section 7.6 (Fauna Passage)).

Notes:

12. Buffer Zones

1. The maximum (inclined) spacing of


horizontal supports aims to allow a trapped
person to climb up the screen safely.
2. The calculated outlet area may depend upon
the level of the outlet relative to the water
surface. Where the outlet is contained in a drop
structure the outlet area used to determine the
net clear opening for the intake may need to be
adjusted to account for the level difference.
3. The limiting velocity through the trash rack
should be related to the accessibility of the
intake structure for cleaning purposes.

The spillway should be designed for an ARI


100 year design flood. If significant debris
blockage of the primary outlet is expected,
then the spillway should be designed to
pass the full ARI 100 year flow rate, i.e.
assuming full blockage of the primary
outlet.
If failure is expected to result in loss of life,
then the spillway should be designed for the
Probable
Maximum
Flood
(PMF).
Additional information may be obtained

Where space is available, surround the


basin with a minimum 20 m buffer zone.
To improve the aesthetics, this buffer zone
can be planted with screening plants (refer
Section 7.5.6 (Buffer Zones)).
13. Public Safety
A maximum depth of 1.2 m during a ARI
20 year flood is recommended for publicly
assessable basins.
All batters that are accessible to the public
should have a maximum slope of 1 on 8.
Provisions should always be make to allow
safe egress from the basin during wet
weather, i.e. when the banks are wet and
slippery. Clear warning signs should be
displayed prominently.
Consideration
should also be given to the placement of
depth indicators. For example, the inclusion
of a fence, when side batters down to the
waters edge are steeper than 1 in 5 and/or
water depths are more than 750 mm for
more than 24 hours.

March 2010

12-11

12

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

14. Maintenance Access


Maintenance access should aim to bring
trucks as close to the basin as possible to
allow direct placement of extracted material
from the basin into the trucks. Thus the
access ramp should ideally have a
maximum grade of 1 on 10. Absolute
maximum grade is 1 on 6.

Have potential for erosion and resuspension of deposited sediment;


Be prone to clogging of the outlet
structure; and
May require large land areas.

If maintenance trucks are located too far


from the basin, then during de-silting
operations the access track between the
basin and the trucks can become
contaminated with mud. This mud can
make the track very slippery and dangerous
if it is too steep.

Due to the absence of a permanent pool, resuspension of sediments may occur during
stormwater events. Therefore the overall
pollutant removal provided by an extended
detention basin is lower and less reliable
than that offered by a constructed wetland
or water quality pond.

12.3

12.3.2 Design Guidelines

Extended Detention
Basin

12.3.1 Description
An extended detention basin may be
defined as a basin that stores runoff for an
extended period of one to two days to
provide water quality improvement. In
between storm events the basin stays dry.
Most extended detention basins usually
incorporate a water retention embankment
and a water outlet structure that provides a
controlled discharge.
The advantages of extended detention
basins include:
Generally can be constructed at
steeper sites than constructed
wetlands;
Have multiple uses, particularly as
recreational area; and
Detain
flows
and
downstream flood peaks.

12

Provide only limited removal of fine


sediment or dissolved pollutants;

attenuate

The disadvantages of extended detention


basins include:

March 2010

12-12

Basin design should aim to maximise the


retention time for as broad a range of storm
sizes as possible. Ideally a continuous
simulation approach should be undertaken
and an appropriate storage volume should
be selected on the basis of long term
performance rather than performance for a
single event.
As with water quality ponds, extended
detention basins should be protected from
flood flows larger than the design storm
flow. The design features given for water
quality ponds in Chapter 7 also apply to
extended detention basins.
(a) Pretreatment
A sediment trap at the inlet can be provided
to enhance pollutant removal of coarse
particulates. Provision of a litter trap will
minimise litter pollution within the basin,
thereby improving ease of maintenance and
basin aesthetics.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

(b) Outlet Configurations


Outlet configurations may include orifice
type devices such as:
Reverse sloped pipe;
Perforated riser; and
Single orifice riser.
(c) Maintenance Access
An access track should be provided for the
maintenance of extended detention basins,
including:
Removal of debris and
pollutants after storm events;

gross

Removal of accumulated sediment;


and
Mowing.

Figure 12.4.1 Sediment Basin

Sediment basins generally perform two


functions: firstly to settle the sand and
coarse silt particles; and secondly to collect
and retain significant volume of sedimentladen runoff to allow time for gravitational
settlement or chemical flocculation of the
finer, clay-sized particles.

(d) Monitoring
Monitoring of an extended detention basin
should be undertaken after large storm
events or every six months to assess the
performance.
Design References
Additional design information may be
obtained from Best Practice Erosion and
Sediment Control (IECA 2008).

12.4

Sediment Basins

12.4.1 Description
A sediment basin is a purpose built dam
usually containing an inlet structure, a
settling pond, a controlled or free-draining
outlet structure, and an emergency spillway.
The settling pond consists of two levels, the
settling zone and the sediment storage zone.
An example of a sediment basin is shown in
Figure 12.4.1.

There are basically two types of sediment


basins:
A dry basin is designed to
commence draining the moment
water enters the basin (i.e. free
draining).
As a result, a large
percentage of the slow settling, claysized particles can pass through this
type of basin in the form of turbidity.
These basins are referred to as Type
C basins and are only suitable for
Coarse grained soils; and
A wet basin is designed to retain
the water for long periods allowing
extended time for the gravitational
settlement of finer soil particles or
for chemical flocculation. These
basins are not drained until a suitable
water quality is obtained within the
settling pond. There are two forms of
wet basins:

12
March 2010

12-13

Department of Transport and Main Roads


Road Drainage Manual

Type F basins are used for Fine soils


that do not require chemical
flocculation; and

When space is limited and the ideal basin


cannot be built, then the largest feasible
basin should be built.

Type D basins for Dispersible soils


that require flocculation.

Step 2: Select basin type


Three types of basins are considered:

Soils usually contain a range of particle


sizes from the very fine clays to the coarse
sands. Fine soil particles (e.g. less than
0.002 mm in diameter), generally pass
through most sediment control structures
including sediment fences, aggregate filters,
and non flocculated sediment basins.

Less than 33% of soil finer than 0.02


mm - Type C basin;

Most clay particles are negatively charged.


Due to the very small size of the clay
particles, this negative charge can have a
significant effect on their ability to settle
under gravity. As a result, some clay
particles take several hours to settle, while
others may never settle unless the water is
treated with a suitable flocculent.

Formal soil testing may be required as


outlined in Table 12.4.2(a).

A flocculent is intended to effectively break


down the dominance of these electromagnetic charges, allowing individual
particles to be drawn together and settle
under gravity.

12.4.2 Design Procedure


This section sets out 17 steps to be followed
in the design of sediment basins (IECA
2008).
Step 1: Assess the need for a sediment
basin

12

Chapter 12
Basins

Sediment basins are usually required when


the disturbed area within a catchment
boundary is greater than 1 ha; the disturbed
soils are dispersible; and/or there is a
recognised need for turbidity control within
the catchment. As a general rule, the
further upstream road works are within a
catchment, the greater the need for turbidity
control.

March 2010

12-14

More than 33% of soil finer than 0.02


mm - Type F basin;
More than 10% of soil dispersible Type D basin.

Soils that are considered dispersible have a


combined percentage of clay (< 0.002 mm)
plus half the percentage of silt (0.002-0.02
mm) multiplied by the dispersion
percentage (Richie 1963) equal to or greater
than 10.
Two simple tests exist to provide an
indication of the existence of dispersible
soils; the Field Emerson Aggregate Test
and the Aggregate Immersion Test (both
described in the departments Soils Manual
(2010)).
The Jar Settlement Test is a simple field test
that provides only an indication of the
suitability of a Type C, Type F or Type D
sediment basin for a given soil type. This
test is described the departments Soils
Manual (2010).
Step 3: Determine basin location
Locate basins to maximise the collection
and treatment of contaminated runoff.
Where practicable, locate basins above the
Q5 (i.e. ARI 5 year) flood level.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Table 12.4.2(a) - Summary Basin Types, Soil Testing and Selection Criteria

Basin Type

Indicator Tests Only

Type C

Jar Settlement Test:

Rapid settling
of coarse
sediments

Less than 33%


finer than 0.02
mm.

Type F

Basin Design Capacity

Basin Selection
Soil Tests

Setting Zone + Sediment Storage Zone

Settling Zone

Sediment storage
zone

Aggregate Immersion
or Field Emerson
Aggregate Test:

Surface area = 3,400 Q Equivalent to


100% of the
& Q = 0.25 Q1
settling volume.
Minimum depth =
Decant time > 24
0.6 metres.
hours.
Minimum length to
width ratio of 2:1
with baffles.

Clay particles
settle in less than
1 hour.

Soil does not


disperse, but may
slake.

Jar Settlement Test:

Slow settling of
fine sediments

Clay particles
settle in less than
5 days.

Greater than 33%


finer than 0.02
mm.

Aggregate Immersion
or Field Emerson
Aggregate Test:

Jar Settlement Test:

Flocculated
settlement of
dispersible
soils

Greater than 33%


finer than 0.02
Clay particles do
mm; and more
not settle in 5
than 10%
days or part of the dispersible
water remains
material.
cloudy.

Aggregate Immersion
or Field Emerson
Aggregate Test:

Volume [m3] = 10
R(Y%,5-day) Cv Ac

R(Y%,5-day) = Rainfall
over 5 days

Ac= Catchment
Area [ha]

Cv= Volumetric
runoff coefficient
(=1.0 for clayey or
compacted soils, or
= 0.5 for open soils)

Minimum depth =
0.6 metres.

Minimum depth =
0.6 metres.

Minimum length to
width ratio of 2:1.

Soil is not
dispersive.

Type D

Minimum length to
width ratio of 3:1
without baffles.

Equivalent to 50%
of the settling
volume.
Settlement time
typically 36-48
hours.

Soil is dispersive.

Note:
1. Basin selection and/or operation may need to be upgraded to a Type F or D basin if turbidity control is
judged necessary, or if the approved basin fails to achieve the desired water quality objectives.

12
March 2010

12-15

Department of Transport and Main Roads


Road Drainage Manual

Step 4: Divert up-slope clean water


Where possible and practicable, up-slope
clean water should be diverted around the
basin to decrease the required size and cost
of the basin and increase its efficiency.
Flow diversion may need to be altered
during the construction phase as new areas
of the road reserve are first exposed, then
rehabilitated.
Step 5: Size the basin
Type C basins:
The critical design parameters in the sizing
of Type C basins is the pond surface area,
As, and base of settling zone, Ab. Where
possible, the effective surface area of the
basin should be maximised.
Note: The effective surface area does not
include dead water areas that contain poor
circulation. It is also noted that several
small basins will not have the same settling
efficiency as a single basin of the same total
surface area.
Pond area at the base of the settling zone Ab
is determined as:

Ab 3400 0.25Q1
Where:

Chapter 12
Basins

replaced by the D30 grain size (i.e. such that


30% of particles are smaller than D30).
In dry basins, there will always be a
sediment grain size that is too small to
settle within the basin. The critical
sediment grain size is the minimum particle
size the sediment basin is theoretically
designed to trap. Hunt (1992) recommends
that the critical particle size should be
chosen such that at least 70% of the
sediment particles flowing into the basins
are larger than this critical particle size.
The minimum critical particle size is 0.02
mm.
If a critical sediment size larger than 0.02
mm is chosen, then the pond size maybe
determined from Table 12.4.2(b).

Table 12.4.2(b) - Particle Settling


Velocities
Particle size
(mm)

Settling
velocity
(m/s)

Basin
surface
area (m2)

0.1

0.007

140

0.05

0.0019

530

0.02

0.00029

3400

Q1 = ARI 1 year discharge.


When determining the critical storm
duration, special consideration should be
given to the determination of the time of
concentration tc. It is noted that the critical
storm duration is determined at the basins
inlet, not at the outlet (i.e. the basin is
assumed to be full).
Critical Sediment Size:

12

The above design formula is based on a


critical sediment size of 0.02 mm. In
areas where the soil has a uniform, coarsegrain size, the critical sediment size may be

March 2010

12-16

Types F & D basins:


To control the turbidity of contaminated
stormwater runoff, it is usually not feasible
to rely solely on gravitational settlement,
especially if the soils are dispersive. Both
Type F and Type D basins are designed to
trap and treat a specified volume of
stormwater runoff while allowing excess
water to pass through the basin and over the
spillway.
Thus, the critical design parameter for
Types F and D basins is the pond settling

Department of Transport and Main Roads


Road Drainage Manual

volume, not surface area. Where possible,


the effective settling volume of the basin
should be maximised, but at a minimum
depth. Adopting a shallow depth reduces
settlement time, which is a critical factor in
the operation of these basins.
Types F and D basins operate on the
principle of producing high quality effluent
from the more frequent storm events, while
still removing most of the coarse sediment
during the larger, less frequent storms. In
most cases, even when the basin is full,
contaminated water should still be directed
to the basin to allow for the settlement of
sand-sized particles.
The same methodology is used to size both
Types F and D basins. The only difference
between the two basins is the recommended
method of operation, i.e. the use of
flocculants in the Type D basin.
3

The settling volume Vs (m ) is determined


from the following equations:

Vs 10 R y %,5day ) C v Ac
Where:

R y %,5 day

I 1 yr ,5 day 0.28
0.2171 0.0021 Y

= the 5-day total rainfall depth


(mm) which is not exceeded in Y% of
rainfall events (Table 12.4.2(c));

I 1 yr ,5day = average rainfall intensity


for a 1 in 1 year, 120 hr (5-day)
design storm (mm/hr);

Y= design probability (%) of the 5day rainfall depth not exceeding the
calculated value RY %,5 day
Cv = volumetric runoff coefficient:
0.5 for open, loose soil; and

Chapter 12
Basins

1.0 for impervious surfaces, clayey


soils, or most soils compacted by
construction traffic.
Ac = catchment area (ha).
It is important to note that a 75th percentile,
5-day rainfall depth is recommended for
most design cases (Table 12.4.2(c)).
However, the nominated design value
should reflect local conditions and the
assessed environmental risk. For example,
in Brisbane City, the 85% rainfall depth is
currently (2001) recommended by the local
authority.
Also, the volumetric runoff coefficient, Cv
is not the same as the discharge runoff
coefficient, C that is used in the Rational
Method to calculate storm runoff rates.
All basins:
The minimum depth of the settling zone is
0.6 metres.
Where practicable, avoid
having settling zone depths greater than 1.0
metre as deeper pond depths can increase
settlement times, especially for the finer
clays.
Restrictions are also placed on the aerial
shape of the settling pond to reduce the risk
of short circuiting. To ensure the optimum
sediment trapping efficiency, the distance
between the inlet and outlet of the basin
should be the maximum practicable. These
restrictions apply to both wet and dry
basins.
Minimum length:width = 3:1 (single
inflow point)
Minimum length:width = 2:1
(multiple inflow points with baffles)
For basins longer than 120 m, the minimum
depth is L/200, where L = basin length (m).

12
March 2010

12-17

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Step 6: Determine sediment storage


depth
Type C basins: 100% of settling
volume.
Type F & D basins: 50% of settling
volume.
Step 7: Assess need for baffles
Inlet baffles:
If inflow velocities are high, settled
sediment can be remixed within the settling
pond, thus decreasing the efficiency of the

basin. Baffles can be used to form an inlet


chamber to dissipate any inflow jets
(Figure 12.4.2(a)).
Inlet chambers can be constructed from
perforated
sediment
fence
(Figure
12.4.2(b)). The perforations are needed to
allow sediment-laden water to flow evenly
across the full width and depth of the
sediment basin. These perforations are also
needed to prevent sediment blockage of the
fabric.

Cross Section

Plan View
Source: Catchments and Creeks Pty Ltd

12

Figure 12.4.2(a) - Inlet Chamber Baffle

March 2010

12-18

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Table 12.4.2(c) - 1 year, 5-day Rainfall Intensity and 75%, 5 Day Rainfall Depth

12
March 2010

12-19

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Source: Catchments and Creeks Pty Ltd


Figure 12.4.2(b) Inlet Chamber Baffle constructed from Perforated Sediment Fabric

Technical Note:
To construct an inlet chamber, perforate a
suitable length of woven sediment fabric
with approximately 50mm diameter holes at
300mm spacing. These holes should not be
placed within the area of fabric that will be
located directly in front of a piped inlet.
Install the sediment fence across the full
width of the basin approximately 1 to 2
metres from the inlet. The top of the fence
should be level with the crest of the primary
outlet spillway.
The spacing between
support posts should be 0.5 to 1.0 metres
depending on the expected hydraulic force
on the fence.
Source: Catchments and Creeks Pty Ltd

Internal baffles:
Internal baffles are used to increase the
effective length to width ratio of the basin
(Figure 12.4.2(c)). If internal baffles are
installed, then a check should be made on
the potential scour velocity.

12

Sediment scour velocities can


determined from Table 12.4.2(d).

March 2010

12-20

be

Table 12.4.2(d) - Sediment Scour


Velocities
Critical Particle
Diameter (mm)

Scour Velocity (m/s)

0.1

0.007

0.05

0.0019

0.02

0.00029

The crest of these baffles should be set


level with, or just below the spillway crest
level. This is to prevent the re-suspension
of settled sediment during severe storms.
Outlet baffles:
An outlet baffle is used to keep settled
sediment away from the primary outlet
system, particularly riser pipes. These
baffles can be used to reduce the time and
cost of basin desilting by keeping sediment
away from the easily damaged outlet
structure.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Step 8: Design the primary outlet system

Perforated riser pipe outlets:

The outlet for a Type C basin normally


consists of two flow systems, the primary
outlet and the emergency spillway. The
primary outlet may consist of a Riser Pipe
Outlet (Figure 12.4.2(d)), a Rock Filter
Dam, a Sediment Weir, or Gabion Wall
(Figure 12.4.2(e)).

The primary outlet on a riser pipe system


normally contains a low-flow aggregate
filter, and a medium-flow spillway located
around the crest of the vertical riser pipe.
The hydraulic capacity of the primary outlet
system must be sufficient to discharge the 1
in 1 year flow at a pond water elevation
level with the crest of the emergency outlet.

Types F & D basins have pumped outlet


systems that can discharge the entire basin
in less than 24 hours.
Appropriate rock protection, or similar,
should be placed at the end of the outlet
pipe and spillway to dissipate energy and
control undesirable soil erosion.

To prevent the buoyant displacement of a


riser pipe outlet system during the early
stages of decant, an anti-flotation weight
must be placed on the base of the pipe. The
anti-flotation weight placed on perforated
riser pipes must have a mass of at least 1.1
times the displaced water mass of the riser
pipe (including the anti-flotation weight).

Long Section

Plan View

12

Figure 12.4.2(c) Internal Baffles

March 2010

12-21

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Primary Outlet Detail (Example only)


Source: Developed from Department of Housing (NSW) (1998)
Figure 12.4.2(d) Type C Basin with Riser Pipe Outlet

Source: Developed from Department of Housing (NSW) (1998)


Figure 12.4.2(e) - Type C Basin with Gabion Outlet Source

For risers less than 1.5 m high, fitting an


anti-vortex type trash rack is advisable
(Figure 12.4.2(a)). A trash rack may also
need to be considered on the riser crest.

of the holes graded vertically. That is, one


hole is provided at the base level of the
settling zone, with progressively larger
holes provided higher up the riser.

The minimum size of the barrel for a pipe


outlet should be 250 mm diameter. At least
one anti-seep collar should be placed on the
riser pipe to prevent seepage along the outer
surface of the pipe.

When multiple holes are used throughout


the settling zone, it becomes necessary to
adjust the relative areas of each hole to
provide a suitable overall de-watering
period. This must take into account the
variation in head over the de-watering time.

When a single large hole is to be used for


de-watering, it should be located at the base
level of the settling zone.

12

To maximise sediment trapping efficiency,


several holes of different sizes could be
used within the settling zone, with the size

March 2010

12-22

De-watering holes in the settling zone


should be covered with wire mesh (25 to 50
mm opening) or coarse gravel to prevent
blocking by debris, but should generally not
be covered by a geotextile or a filter cloth.

Department of Transport and Main Roads


Road Drainage Manual

De-watering of the sediment storage zone


should be considered to facilitate basin
clean-out. De-watering can be achieved by
using filtered holes in the riser, ideally at
the base of the sediment storage zone.
It is generally not recommended for filter
cloth to be placed around an outlet riser
pipe. However, if filter cloth is to be used,
it should not be placed in close contact with
the riser. An air gap is essential between
the perforated riser and any geotextile to
allow free draining of the basin.

Chapter 12
Basins

aggregate. Filter cloth is the preferred liner


for rock filter dams that have a short
operational life.
Gabion wall outlets:
Gabion walls should be lined on the inside
with filter cloth, not aggregate. The filter
cloth should NOT be placed between
gabion baskets so that the cloth can be
replaced during maintenance (Figure
12.4.2(e)).
Sediment weir outlets:

Wire mesh should be wrapped around and


secured to the riser before attaching
geotextile filter cloth, to increase the rate of
water seepage into the riser. The fabric
must be replaced after each storm event.

The design and operation of sediment weir


outlets is similar to gabion walls.

Rock filter dam outlets:

Whichever method is used, reverse flush or


replace the filter cloth each time the
sediment is removed from the basin.

Rock can be used to form a rock filter dam


outlet system. In this outlet system, a
structural rock wall is constructed as the
primary outlet location of the basin. The
upstream face of the rock dam is lined
either with aggregate, or a layer of needlepunched filter cloth. It is the aggregate or
filter cloth that controls the decant rate and
prevents most of the finer sediment passing
through the dam.
An upstream aggregate layer has the
advantage that it can be placed by
machinery and can be readily replaced with
a backhoe if it becomes blocked with
sediment. However, some guidelines do
not recommend the use of aggregate filter
layers because of reported maintenance
difficulties.
Needle-punched filter cloth has the
advantage of being cheap, but its
replacement can be messy and leaks may
occur if the replaced filter cloth is not
installed properly.
The filter cloth is
usually placed over a thin layer of

For further information, refer to the Soil


Erosion and Sediment Control Engineering
Guidelines (IEAust 1996).

Step 9: Select internal and external bank


gradients
Internal batter gradients need to be
consistent with the requirements of personal
safety and generally within the following
upper limits:
Where water depth is less than 150
mm when surcharging, 1 on 2 to 1 in
4 on earth structures; and gabion
basket structures;
Where water depth is between 150
mm and 1500 mm when unfenced
and surcharging, a maximum slope of
1 on 5;
Where water depth is between 150
mm and 1500 mm when fenced and
surcharging or greater than 1500
mm:
1 on 2 to 1 on4, on earth structures;

March 2010

12-23

12

Department of Transport and Main Roads


Road Drainage Manual

1 on 4, on gabion basket structures;


and
1 on 4, on stacked (rough squared)
rock structures.
The actual bank gradient will depend on the
slipperiness of the saturated sediment, i.e.
whether or not a person can achieve a firm
footing and exit the basin.
Slippery
sediments should have less steep gradients,
in the order of 1 on 8 or even 1 on 10.
Otherwise, the basin should be fenced.
Step 10: Design emergency spillway
The capacity of the emergency spillway
should be designed in accordance with the
following:
Less than 3 months operation - Q10;
3-12 months operation - Q20;
Greater than 12 months - Q100;
If failure is expected to result in loss of life
Probable Maximum Flood.
Spillway crest should be at least 300 mm
above the primary outlet, 300 mm below a
bank formed in virgin soil, and at least 750
mm below a fill embankment (unless
supported by geotechnical investigations).
For reasons of safety and structural
stability, all reasonable and practicable
efforts should be taken to construct the
spillway in virgin soil, not in fill soil.

12

The spillways horizontal alignment can be


curved upstream of the crest, but must be
straight from the crest to the energy
dissipator (Figure 12.4.2(e)). Ensure that
the approach section has a slope towards
the impoundment area of not less than 2%
and is flared at its entrance, gradually
reducing to the design width at the spillway
crest.

March 2010

12-24

Chapter 12
Basins

The downstream face of the spillway


usually needs to be protected with rock or a
rock mattresses.
Appropriate rock protection, or similar,
should be placed at the end of the outlet
pipe and spillway to dissipate energy and
control soil erosion.
Step 11: Determine overall dimensions of
the basin
If a sediment basin is constructed with 1 on
3 side slopes, then a typical basin would be
7 to 10 m longer and wider than the length
and width calculated in Step 5 above (given
a 0.6 m settling zone depth, 0.3 m spillway
freeboard and 0.75 m freeboard for a fill
embankment).
Thus, it is clearly important to determine
the overall dimensions of the basin to make
sure the basin can fit into the designated
area.
The minimum recommended embankment
crest width is 2.5 m.
Step 12: Locate maintenance access
ramp for desilting
The basin width should allow for desilting
works, so that equipment can access the
basin for desilting. If trucks need to access
the basin, then a maximum 1 on 10 access
ramp will need to be constructed. It is
usually preferable to bring the trucks to the
sediment, then to try and carry the sediment
to the trucks. The more the sediment is
handled, the more sediment is spilt around
the work site.
If the sediment is to be removed from the
site, then a sediment drying area should
exist adjacent to the basin or somewhere on
site within the basins catchment area.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Step 13: Design the inlet system

Where possible, a floating inlet chamber


should be used to minimise the resuspension and discharge of fine sediment.

Basin inlet channels should be protected


against erosion and scour. This may be
achieved by using a rock or geotextile lined
chute.
When sediment is allowed to enter a basin
via a permanent stormwater pipe system,
efforts should be made to prevent sediment
settling within the stormwater pipes.
The use of a geotextile inlet chamber
should be considered if the expected inflow
velocity exceeds 1 m/s on Type C basins
(refer to Step 7).
Step 14: Assess the need for safety
fencing
Sediment basins located within urban areas
should be suitably fenced if:
The settled sediment depths exceed
300 mm; and/or
Permanent water depths exceed 150
mm; and/or
Public safety is at risk.
Step 15: Define basin operation.
A sediment desilting marker post shall be
installed in the basin. This marker post
should clearly show the top of the sediment
storage zone. The basin will need to be
desilted when the settled sediment reaches
the appropriately marked (i.e. painted) level
on this marker post.
Types F & D Basins:
Where possible, operation of the basin
should ensure that water has drained from
the settling zone of the basin, and
preferably from the sediment storage zone,
prior to the next rainfall event that is
expected to causes runoff.
Types F and D basins are usually decanted
36 to 48 hours after each storm event.

If a pump is used to decant the basin, then


the decanting must cease prior to settled
sediments being drawn into the intake pipe.
The intake pipe must not be allowed to rest
on or near the settled sediment. In some
cases it may be sufficient to place the foot
valve in a large perforated drum.
Types F & D basins should be flocculated if
the trapped sediment-laden water does not
achieve a desirable standard, usually 50
mg/L total suspended solids (TSS).
Sediment Flocculation:
Many flocculating agents exist, including
gypsum, alum, ferric chloride, ferric
sulphate,
polyelectrolytes
(long-chain
natural and synthetic organic polymers) and
salt (sodium chloride). Gypsum and alum
have traditionally been applied to captured
stormwater runoff.
Gypsum (calcium
sulphate) and alum (aluminium sulphate)
are suitable chemicals for this purpose and
are applied within 24 hours of the
conclusion of each storm event. However,
alum should always be used with caution
and only after the potential environmental
impacts have been assessed.
Application of Flocculating Agents:
In larger ponds, the agents should be mixed
into a slurry with water and then sprayed
over the pond surface.
In smaller ponds and tanks, add agents by
simply broadcasting it over the surface by
hand.
Whichever method is chosen, it is essential
that the flocculating agent is spread evenly
over the entire pond surface for proper
treatment of water unless local experience
or other criteria suggest differently.

March 2010

12-25

12

Department of Transport and Main Roads


Road Drainage Manual

Gypsum should be applied at a rate of


approximately 32 kilograms per 100 cubic
metres of stored water (actual dosing rates
depend on the soil properties). Alum
(where approved) should be applied at 1.5
to 8 kilograms per 100 cubic metres of
stored water (higher rates are more
effective but can influence water pH more).
Care should be taken with the choice of an
agent, its dosing rate and any special
conditions to ensure that toxic situations are
not created with consequent damage to the
ecology.
In areas where repeated high intensity
storms are likely, it is recommended that
gypsum dosage rates be increased to 70
kilograms per 100 cubic metres.
Depending on the clay mineralogy, this can
achieve flocculation within 24 hours
allowing discharge within two days from
the conclusion of a storm.
Effectiveness of agents:
When choosing a flocculating agent, note
that:
the trivalent positive aluminium
(Al3+) ion is 2,000 times more
effective than the monovalent
positive sodium (Na1+) ion; and
the bivalent position calcium (Ca 2+)
ion is only 50 times more effective
than sodium (Barnes 1981).
As such, alum produces a faster
flocculation rate than gypsum, which has
been shown for sediment basins in New
South Wales (Goldrick et al 1996). Table
12.4.2(e) summarises some characteristics
of common flocculating agents. Trials
should be conducted on samples to
determine the most appropriate dosing rate

12
March 2010

12-26

Chapter 12
Basins

to reduce the likelihood of excessive


dosing.
The use of alum as a flocculant is only
recommended when it is used under
controlled circumstances and by users that
are aware of the potential downstream risks
to the environment.
Details of the flocculation and decanting
procedures should be provided in the
Erosion and Sediment Control Program.
Settlement time:
Normally, sufficient sediment will have
flocculated and settled within about 36 to
72 hours in the case of gypsum.
Expected water quality:
Following flocculation, a total suspended
solid (TSS) content of less than about 50
milligrams per litre is typically achievable.
A practical field test that approximates this
level is to fill a clear plastic or glass 65 mm
diameter soft drink bottle with the water
and hold it up to the light. If seeing clearly
through the sample is not possible, it is
probably above about 50 milligrams per
litre and needs further treating (Freeman &
Howells 1995).
The pH of the flocculated water must be in
the range 6.5 to 8.5 prior for discharge. If
this is not achieved, pH adjustment may be
necessary (e.g. dosing with lime to raise
pH).
Site calibration
Despite the above comments, each pond
should be calibrated after the first two
storm events to assess the actual flocculent
application rate and settling time required.
Standard jar tests are the usual method
(Barnes 1981).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 12
Basins

Table 12.4.2(e) Characteristics of Various Flocculating Agents (adapted from Freeman


and Howells 1995)

Agent

Indicative
dosage

Comments

Precautions and
constraints

Gypsum

32 kg/100m

Little pH change, slight


increase in salinity.

Needs to be spread evenly


across pond, can cause
scum
deposits
in
equipment.

(calcium sulphate)

Alum (aluminium
sulphate)

1.5
8kg/100m

Produces a stable sludge


that
binds
pollutants,
optimum pH 6 to 7.4. Do
NOT overdose as pH will
be lowered.

Likely toxic impacts on


ecology at pH levels <5.5
due to release of dissolved
aluminium. Must only be
used
in
controlled
conditions. Note1

Ferric chloride

1 - 3kg/100m

pH greater than 5 is
required or it might lower
oxygen levels.

Is very corrosive, needs


rubber
or
glass
containment.
Do
NOT
overdose. Note1

Ferric sulphate

1 2.5kg
/100m

pH greater
required.

Stored
in
containers.

than

is

wooden

Do NOT overdose. Note1


Polyelectrolytes
(long
chained
polymers)

0.05 0.2kg
/100m

Careful
preparation
needed and adequate
mixing with water body
needed, little pH or salinity
change, might be toxic.

A few are banned for use


with potable water in some
countries due to possible
monomer impurities. Do
NOT overdose.

Salt
(sodium
chloride)

5.25m
seawater per
100m
of
fresh water

Flocculation is complete
for some clays with 2000
to 3000 mg/l, little extra
benefit is gained when the
salinity
is
above
10,000mg/L.

Only
used
when
the
sediment basin discharges
directly to sea water. Note:
sea
water
contains
approximately 35,000 mg/L
salt.

Notes:
1. The pH of the water in the basin must be in the range of 6.5 to 8.5 before release

12
March 2010

12-27

Department of Transport and Main Roads


Road Drainage Manual

In some situations it might be necessary to


test water samples in a laboratory before
discharge to prove that the suspended solid
content is, in fact, below recommended
values, e.g. where the receiving waters are
particularly sensitive.
In these cases,
sampling details should be clearly set out
on the sites Erosion and Sediment Control
Plan.
The final application rate should be
sufficiently high to permit sediment
flocculation and pond discharge within two
to four days from the conclusion of each
storm event, whilst maintaining other
required water quality criteria such as pH.
Water discharge
The water can be discharged from the basin
once the suspended solid load has been
lowered to an acceptable value.
The
discharge system should:
permit drainage of the pond in less
than 24 hours; and
has a floating inlet to prevent
flocculated sediments being removed
as well.
It is essential that materials from the
sediment layer are not discharged in the
pumping process.
Warnings

12

With use of alum, accurate


measurement if water pH must be
undertaken to ensure that values
remain in the range of 6.5 and 8.5.
Values lower than pH 5.5 will result
in
environmentally
toxic
concentrations of soluble aluminium
that can kill fish and other aquatic
life. Treated waters should not be
discharged if the pH is below 6.5
unless site-specific environmental

March 2010

12-28

Chapter 12
Basins

risk assessment shows that it is safe


to do so.
Excessive
dosing
polyelectrolytes can:

with

Result in the release of materials that


can kill fish and other aquatic life;
and
Reduce the effectiveness of the
flocculent.
Step 16: Define
location / method.

sediment

disposal

Trapped sediment can be mixed with onsite soils and buried, or removed from the
site. If sediment is removed from the site,
then it should first be de-watered.
Removed sediment should be disposed of
so as not to cause an erosion hazard.
Step 17: Define final rehabilitation of the
basin area.
Following completion of the road works,
sediment basins may either be back-filled
and revegetated, or retained and converted
into a permanent stormwater treatment
feature or similar beneficial attribute of the
road works.
In rural areas, sediment basins are often
constructed within adjoining properties and
remain as permanent farm dams.
Obviously, negotiations for the temporary
use of the adjoining property and the
retention of the basin must be made with
the land owner during the planning stage of
the road works.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

13

Chapter 13
Erosion and
Sediment Control

March 2010

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Chapter 13 Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table of Contents

13

13.1

Introduction

13-1

13.2

Erosion and Sedimentation

13-1

13.2.1

13.3

13.4

Understanding Erosion & Sedimentation

Soil Properties and Influence

13-2

13-2

13.3.1

Soil Permeability

13-3

13.3.2

Use of Soil Permeability Classes

13-3

13.3.3

Identifying Waterlogging and Seepage Zones

13-4

13.3.4

Other Factors Affecting Site Drainage

13-4

Soil Erodibility

13-4

13.4.1

Dispersion

13-5

13.4.2

Slaking

13-5

13.4.3

Sampling for Erodibility Tests

13-5

13.4.4

Field Observations and Dispersion Tests

13-5

13.4.5

Laboratory Testing

13-5

13.4.6

Dispersion in Departmental Soil Groups

13-6

13.4.7

Relevance to Weathered Bedrock

13-7

13.4.8

Interpretation of Erodibility Tests

13-7

13.5

Erosion Risk Assessment

13-8

13.6

Objective Rating System

13-9

13.7

Determining Erosion Risk Factors

13-10

13.7.1

Rainfall Erosivity

13-10

13.7.2

Soil Erodibility

13-11

13.7.3

Slope Gradient

13-14

13.7.4

Slope Length

13-15

13.7.5

Combined Effect of Slope Gradient and Length

13-15

13.7.6

Vegetation Cover

13-15

13.8

Implications of Risk Rating

13-15

13.9

Environmental Assessment - Review of Data

13-16

13.10 Environmental Management Plan (Construction) and Erosion and


Sediment Control Plans
13-17
13.10.1

An ESCP Overview

13-17
March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

13.10.2

Benefits of ESCPs

13-18

13.10.3

Timing and Acceptance

13-18

13.10.4

ESCP components

13-18

13.11

Preparing ESCPs

13-21

13.11.1

Step 1 - Review Site Characteristics

13-21

13.11.2

Step 2 - Identify Areas of Construction Disturbance

13-22

13.11.3

Step 3 - Selection of Controls

13-22

13.12 Design Standard and Selection of Erosion and Sediment Control


Measures
13-31
13.12.1

Select Controls

13-32

13.12.2

Treatment Train Approach

13-33

13.13

Installation of Control Measures

13-34

13.14

Maintenance of Control Measures

13-34

13.14.1

Regular Clean-Out

13-35

13.14.2

Repair and Replacement

13-35

13.15

Removing Control Measures

13-35

13.16

Monitoring, Auditing and Reporting ESC

13-36

13.17

Revegetation

13-36

13

March 2010

iv

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Chapter 13
Erosion and Sediment Control
13.1

Introduction

This chapter is designed to assist,


Site managers,
Project and contract managers,
Designers, and
Contractors
with the management of
drainage, erosion and sedimentation,
whilst planning for and undertaking road
projects. It should be read in conjunction
with the departments Soils Manual
(2010).
Construction
of
road
drainage
infrastructure has the potential to cause
environmental harm and subsequent
damage to infrastructure if not properly
managed.
The loss of soil from a construction site
can also be expensive in terms of
remediation, and replacement costs.
Implementing effective erosion prevention
or minimisation practices while managing
drainage on a site will reduce the impact of
sedimentation. This can translate into
project cost benefits, increased project
efficiency and productivity (due to less
time
managing
/
cleaning
up
sedimentation), and can minimise downslope adverse environmental implications.
This chapter focuses on identifying and
implementing the most suitable permanent
and temporary drainage, erosion and
sediment control measures in order to
minimise potential environmental harm
and damage to assets during construction.

Careful consideration should always be


given to soil types and behaviours when
planning erosion and sediment control,
therefore some sections dealing with soil
erodibility and risk assessment is provided
in this chapter as well.
It is important that the focus lies on
effectively managing drainage and
preventing or minimising erosion as well
as incorporating sediment controls. It is
extremely unlikely to completely prevent
onsite erosion from occurring at all.
Notwithstanding this, priority should
always be given to preventing or
minimising soil erosion wherever possible
and practical. Erosion and sedimentation
cannot be effectively managed at the end
of the pipe.
Controlling drainage and preventing
erosion could reduce the size, number,
maintenance, cost and reliance on downslope site sediment controls, but will not
generally eliminate the need for them.
Drainage, erosion and sediment controls
need to be appropriately integrated into all
phases of construction in order to work
effectively, as promoted in Best Practice
Erosion & Sediment Control published by
International Erosion Control Association
(IECA) (Australasia) in 2008.

13.2

Erosion and
Sedimentation

One of the most common environmental


impacts of road construction is the erosion
of soil and subsequent sedimentation of
watercourses,
roads
or
drainage
infrastructure.

March 2010

13-1

13

Department of Transport and Main Roads


Road Drainage Manual

13

Erosion and sediment control (or ESC), as


a practice relies on careful consideration of
a number of other site characteristics,
including soil types, local hydrology and
topography, vegetation types, climate and
more. These are addressed in the Erosion
and Sediment Control Plan component of
the Environmental Management Plan
(Construction).
Sediment suspension and deposition
within watercourses has direct impacts on
water quality and the health of aquatic
ecosystems.
This can include fauna
mortality, impact on and loss of habitat,
algal blooms and disturbance to ecological
breeding patterns. Soil loss or deposition
can also have impacts on adjacent
properties such as farms and residential
communities.
Figure 13.2 shows how vegetated slopes,
drainage channels, velocity reduction, and
sediment basins are used in conjunction
with one another to minimise erosion and
sediment deposition on this site.

Chapter 13
Erosion and Sediment Control

the different causes and types of erosion


that occurs on departmental projects.
The handbook explains that the mechanics
of erosion involve soil detachment, soil
transportation and soil deposition.
Soil detachment can be facilitated by the
physical forces of raindrop impact,
flowing water or shear stress from flowing
water.
Soil transportation occurs after soil
particles have been detached, and is also
facilitated by raindrop impact, flowing
water, scouring and wind.
Sedimentation will occur when eroded soil
particles have been deposited after being
transported from their origin.
Identifying these processes forms the
beginning of understanding how to prevent
/ minimise erosion and to successfully
control drainage and sedimentation.
The next sections explain the theory and
processes behind making risk assessments
through understanding and analysing soil
types and behaviour. Having a solid
understanding
of
the
site
soil
characteristics
through
undertaking
erodibility risk assessments, various field
and laboratory tests and developing a
knowledge of soil groups will assist
project staff with effective implementation
of erosion and sediment control measures.

13.3
Figure 13.2 - Interactive Remediation
Works Minimising Erosion and
Sediment Deposition

13.2.1 Understanding Erosion &


Sedimentation
The departments Road Construction
Erosion & Sediment Control Workshop
training handbook contains an outline of

March 2010

13-2

Soil Properties and


Influence

Erosion potential is one of the major issues


for road planning and design across much
of Queensland and it is strongly influenced
by the soils properties.
The next few sections provide sufficient
information to give an understanding of
soil properties and their effects on site
drainage (incl. soil permeability) and to

Department of Transport and Main Roads


Road Drainage Manual

enable
a
broad
assessment
and
determination of erosion potential for
planning and design purposes.
More
detailed, site specific assessment is
generally required for construction
purposes.

13.3.1

Soil Permeability

Soil permeability refers to the capacity of


soil to conduct gases and fluids, and more
particularly water. Soil permeability has
an impact on runoff and erodibility and
also determines the rate at which water
moves through the profile.
Soil
permeability is a major factor affecting site
drainage.
Soil permeability depends on both soil
texture and structure, as it is controlled by
the number of connected pores, the size of
their narrowest section, and the tortuosity
and windiness of the connected pathways.
A poorly structured soil is often
associated with chemical properties that
cause dispersion and thus many dispersive
soils are also relatively impermeable.
The capacity of the soil to shrink and swell
with changing moisture will also affect
permeability as such soil movement can
alter pore size and disrupt connected
pathways within the soil matrix.
Soil permeability will control how fast
water can infiltrate into the soil, and thus
will impact on runoff and erodibility.
Soil permeability can be quantified using
measurements of saturated hydraulic
conductivity but this property is difficult to
measure.
However, permeability can also be
inferred for a soil profile from the period
required to reach field capacity after
thorough wetting if there were no
obstructions to movement from the profile.
Field capacity represents the point at
which there is no significant rapid

Chapter 13
Erosion and Sediment Control

downward movement of water through the


soil in response to the hydraulic gradient,
mostly gravity.
Qualitative assessment of permeability can
be made as follows:
Very slow - the soil profile would take
periods of a month or more to reach field
capacity after thorough wetting. Soil
texture is usually clay and no pores are
visible (with a hand lens).
Slow - the soil profile would take periods
of a week or more to reach field capacity
after thorough wetting. Soil texture is
usually clay but a few pores are visible
(with a hand lens).
Moderate - the soil profile would take
periods of no more than a number (1-5) of
days to reach field capacity after thorough
wetting. Soil texture is variable but clay
loam or clay soils have at least moderate
structure with peds of fine to medium size.
High - the soil profile would take no more
than a number (1-12) of hours to reach
field capacity after thorough wetting. Soil
texture is generally sandy or sandy loams
but may include strongly structured sandy
clay loam to clay soils with peds of fine
size.

13.3.2 Use of Soil Permeability


Classes
A permeability class is usually assigned to
an entire soil profile and is determined by
the least permeable horizon within the
profile (McDonald et al. 2009). However,
permeability can also be considered for
individual soil horizons.
Indicative saturated hydraulic conductivity
ranges for the four permeability classes
are:
Very slow - less than 5 mm per day
drainage time = months

March 2010

13-3

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Slow 5 to 50 mm per day drainage


time = weeks

change in hydraulic conductivity


(soil type); or

Moderate 50 to 500 mm per day


drainage time = days

other
obstruction
to
water
movement (such as rock outcrop).

High - more than 500 mm per day


drainage time = hours.

Soils below a break in slope are often


heavier and less permeable than those
above and thus seepage water is
commonly forced close or on to the
surface at these landscape positions.

Assigning a permeability class and its


typical hydraulic conductivity range to
each soil horizon will indicate relative
rates of water movement through a profile
and thus reveal zones in which
waterlogging and seepage are likely.

13.3.3 Identifying Waterlogging


and Seepage Zones
All soils will become saturated if the rate
of water entry into a profile exceeds the
permeability of any soil horizon.
Depending upon the volume of incoming
water, saturation (waterlogging) may last
from a few hours to several months.
If the capacity of the surface horizon to
transmit water (hydraulic conductivity) is
exceeded by the rate of water entry, the
excess water will leave the site as surface
runoff.
For water actually entering the soil profile,
the rate of movement through the profile
will be determined by the permeability of
each horizon. Waterlogging is most likely
to occur wherever one soil horizon
overlies another of much lower
permeability.
If there is any slope at the site, water
ponding at this boundary water will seep
down slope along the top of the less
permeable horizon. This process is also
described as interflow.
Seepage (or interflow) will continue down
slope within this zone until it encounters
either a:
drainage line or water body that
intersects the transmission zone;
March 2010

13-4

Waterlogging and seepage can also occur


preferentially in any horizon that is more
permeable than its overlying soil horizon.
For example, a watertable is commonly
found in sandy buried horizons below clay
subsoil. Weathered rock is often more
permeable than the subsoil above and the
harder, little-weathered bedrock below.
If there is any slope, such horizons can
transmit water sideways rather than into
the less permeable horizon below.
Thus, there may be several seepage zones
within and below a soil profile.

13.3.4 Other Factors Affecting


Site Drainage
Though soil permeability is a major
determinant, drainage for any particular
site will also depend upon the volume and
quality of incoming water, presence of any
physical barriers, slope attributes and
position in the landscape.

13.4

Soil Erodibility

Soil erodibility refers to the inherent


potential for soil to erode when exposed to
raindrop splash or running water.
A soils erodibility is determined by the
rate of infiltration at the surface,
permeability of the soil profile and
coherence of the soil particles.
However, these properties are generally
not measured directly. Instead, these
properties can be estimated through the

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

identification of soil types using key soil


features such as horizons, texture, surface
condition, consistence, colour, structure
and dispersion capacity.

When a soil sample is placed in water,


slaking is demonstrated by subsidence of
the sample into a pile of loose material
with no clouding of the water.

The process of dispersion can markedly


increase soil erodibility.

13.4.3 Sampling for Erodibility


Tests

13.4.1 Dispersion

Field and laboratory tests for soil


erodibility are based upon predicting the
degree of clay dispersion. The tests
involve dispersive and slaking responses
to wetting and analysis of relevant soil
chemical properties.

Dispersion is the detachment of individual


clay particles from peds when placed in
water. The clay particles and move into
suspension, clouding the water.
Dispersion can damage soil structure by
destroying large flocculated peds and
filling the voids between these peds with
much smaller dispersed material. Thus,
the porosity and permeability of the soil
declines and the erodibility increases as
the small dispersed particles are easily
moved in water. Thus, dispersive soils are
very unstable when wet.
Dispersion can be a result of:
low soil organic matter;
high soil exchangeable sodium;
and/or
low soil exchangeable calcium /
exchangeable magnesium ratios.
Soil is composed of sand, silt, clay and
organic matter. The particles are bound
together as peds by the clay and organic
matter. The arrangement of the peds gives
the soil its structure.
Dispersive soils are often poorly
structured.
Erosion rates for poorly
structured soils can be many times higher
than for well structured soil.

13.4.2 Slaking
Slaking is the breakdown (sloughing off)
of the soil sample into smaller fragments
and is caused by the clay swelling and by
un-trapped air escaping. With slaking
individual soil particles do not detach.

For testing:
a representative soil sample of at
least 500 g should be collected from
each horizon of concern.
the sample should be dry when used
for testing. If collected wet, it will
need to be air dried before testing.

13.4.4 Field Observations and


Dispersion Tests
Field observations can indicate whether
soil is likely to be dispersive.
Severe rilling of exposed batters and
pronounced rill and gully erosion in drains
indicate possible dispersion caused by
raindrop splash and/or run-off. Dispersion
is highly likely where sink holes and
tunnels form, even on gently sloping
terrain.
Dispersion in the surface horizon is a
major factor in the formation of a surface
crust. A friable soil profile is usually non
dispersive whereas very strong to rigid
subsoil with either massive structure or
coarse peds, may be dispersive.

13.4.5 Laboratory Testing


One of the laboratory tests used for soil
dispersion is the Emerson Dispersion Test
(EDT). A modified dispersion test based
on EDT is used by the department as a
field test and is referred to as the Soil
March 2010

13-5

13

Department of Transport and Main Roads


Road Drainage Manual

13

Dispersion Indicator field test (SDI). The


test simply requires the placement of soil
aggregates (crumbs) into deionised water;
the reaction of the soil with the water is
observed for slaking or dispersion.
Slaking is particularly evident, and occurs
quicker, in soils with low organic matter
content.
Dispersion is usually caused by either:

Chapter 13
Erosion and Sediment Control

soil salinity as measured by


Electrical Conductivity (EC) - can
be both laboratory and field tested;
soil exchangeable cations (calcium,
magnesium, sodium, potassium and
aluminium) laboratory test only;
effective cation exchange capacity
(ECEC) laboratory test only

too much sodium attached to the


surface of clay particles; or

exchangeable Sodium Percentage


(ESP) derived from exchangeable
cations;

too much sodium attached to the


clay and a higher proportion of
magnesium than calcium attached to
clay surfaces.

ratio of the exchangeable cations


calcium and magnesium (Ca:Mg)
derived from exchangeable cations;
and

Soil salinity as measured by Electrical


Conductivity (EC), pH and behaviour of
the soil bolus whilst texturing are other
field tests that can be performed to
indicate the tendency for dispersion.

percent
aluminium
of
total
exchangeable cations (important in
acid soils with pH < 5.5) derived
from exchangeable cations.

The Emerson Dispersion Test (EDT) is a


simple laboratory test that can be used to
indicate dispersion.
However, a basic set of soil chemical
properties of the soil are essential to make
best possible prediction of soil dispersion
characteristics.
The MRTS16 Appendix B (TMR 2010c)
sub-suite analysis schedule has been
developed to provide a useful screening
tool for assessment of dispersive soil
properties.
The sub-suite analysis is designed to
provide a minimum dataset for assessing
soil erodibility, fertility, pH and salinity
effects on plant and infrastructure.
Briefly these tests involve:
soil pH - can be both laboratory and
field tested;

March 2010

13-6

Exchangeable Sodium Percentage, also


referred to as soil sodicity, and the Ca:Mg
ratio are derived from chemical analyses
of soil exchangeable cations and must be
interpreted with care. When the actual
measured
exchangeable
cation
concentrations are low to very low, any
small change in one value can cause a
disproportionate change in the percentage
or ratio calculation and thus alter the
dispersion rating.

13.4.6 Dispersion in
Departmental Soil
Groups
This section refers to the soil groups as
detailed in the departments Soils Manual
(2010).
Dispersion is not a common feature in
sandy and loamy textured surface horizons
and subsurface horizons but can occur in
the clayey surface horizon of the Uniform
cracking clays and similar, coarsely
structured Uniform non-cracking clays.

Department of Transport and Main Roads


Road Drainage Manual

Dispersion is a characteristic subsoil


feature of the Texture contrast soils
(dispersive) departmental soil group.
However, this is not the only departmental
soil group that may exhibit subsoil
dispersion.
Uniform cracking clays and similar,
coarsely structured Uniform non-cracking
clays can also have a tendency to
disperse, especially in profiles developed
on sedimentary rocks deposited in
estuarine and marine conditions and on
unconsolidated sediments derived from
these rocks.
Dispersion is occasionally found in subsoil
of Loamy gradational soils, especially
those with a bleached subsurface horizon
and/or coarsely structured subsoil.
Dispersion may also occur infrequently in
subsoil of Sandy gradational soils with
similar profile features.
Dispersion is rarely observed within the
Uniform sands and sandy loams,
Uniform loams and clay loams and in
Uniform non-cracking clays with fine
ped size.
Dispersion may be a feature of
Waterlogged soils (non saline) but not
the Waterlogged soils (saline), except
after the watertable is drained and the salts
are leached out.
Dispersion may also occur in buried
horizons below non dispersive soils and in
weathered rock.

13.4.7 Relevance to Weathered


Bedrock
Soils on hard rock landscapes overlie and
merge into weathered rock.
The same factors that determine soil
erodibility also govern inherent erodibility
of the soft and poorly consolidated,
weathered rock below.

Chapter 13
Erosion and Sediment Control

Thus, the overlying soil type can give


important information on its likely
behaviour. For example, the weathered
rock below highly erodible soil is likely to
be highly erodible as well.
However, examples of non dispersive
subsoil overlying dispersive weathered
rock are not uncommon.
Therefore, the dispersion tests described in
Sections 13.4.4 and 13.4.5 above should
also be applied to soft, weathered rock
wherever this material will be exposed,
such as in cut batters, table drains and fill
embankments.

13.4.8 Interpretation of
Erodibility Tests
The field and laboratory tests described in
Sections 13.4.4 and 13.4.5 are used as
independent estimates of soil dispersion.
The MRTS16 Appendix B (TMR 2010c)
sub-suite and SDI tests use different
surrogates as indicators of soil dispersion
and thus do not necessarily agree.
Soils should be rated as either definitely,
probably, possibly, unlikely or non
dispersive according to the combined
results of all tests.
A suggested framework for rating the soil
tendency for dispersion is provided in
Table 13.4.8. The framework uses the
results of four field and laboratory tests
ESP, Ca:Mg ratio, SDI and EDT.
This framework may need to be adjusted
to allow for disproportionate ESP and
Ca:Mg ratio calculations in (particularly
sandy) soils with very low exchangeable
cation levels (refer Section 13.4.5).
Other relevant field and laboratory tests to
be considered are pH, salinity and the field
texture test for sodicity. Using results
from these tests will help to confirm the
rating for a particular soil.

March 2010

13-7

13

Department of Transport and Main Roads


Road Drainage Manual

13

ESP 15

Finally, the dispersion rating should also


be compared with site and profile
conditions
observed
during
field
inspection to ensure a realistic assessment
has been achieved.

or
Ca:Mg ratio < 0.1

13.5

Table 13.4.8 - Suggested Dispersion


Rating
Rating

Definitely
dispersive

Criteria

and
2 of 4 tests indicate
dispersion
ESP 6-14

Probably
dispersive

or
Ca:Mg ratio 0.1-0.5
and
2 of 4 tests indicate
dispersion
ESP 15

Probably
dispersive

or
Ca:Mg ratio < 0.1
and
only 1 of 4 tests
indicate dispersion
ESP 6-14

Possibly
dispersive

or
Ca:Mg ratio 0.1-0.5
and
only 1 of 4 tests
indicate dispersion
ESP < 6

Possibly
dispersive

or
Ca:Mg ratio > 0.5
and
2 of 4 tests indicate
dispersion
ESP < 6

Unlikely to
disperse

Non dispersive

or
Ca:Mg ratio > 0.5
and
only 1 of 4 tests
indicate dispersion
No test results
indicate dispersion.

NOTES: Rating criteria are based on 4 tests ESP;


Ca:Mg ratio; SDI and EDT.
(Source: McDonald et al. 1990)

March 2010

13-8

Chapter 13
Erosion and Sediment Control

Erosion Risk
Assessment

The results of the site and erosion risk


assessment can provide a preliminary
indication of which control measures will
be most suitable for a given location.
Erosion risk refers to the risk of causing
significant soil movement within a site and
overall soil loss off site as a result of either
disturbance during the construction phase
or altered land management practices.
An assessment of the erosion risk both
during and after construction should be
used as an aid in designing control
measures and selecting appropriate site
management practices.
In all cases, erosion risk should be
considered as an additional design factor,
which provides an indication of whether a
high standard of erosion and sediment
control is required.
Risk assessment can be subjective,
objective or quantitative. A subjective
assessment is based on observed
relationships between land attributes and
erosion, and can be very accurate when
undertaken by experienced personnel.
However, such assessments can be
unreliable when applied by inexperienced
users and are often difficult to extrapolate
to environments that differ from those
originally used for their development.
A subjective assessment must therefore
only be undertaken by experienced
personnel and should be recorded as
subjective. In all cases where any risk is
indicated, assessment results should be

Department of Transport and Main Roads


Road Drainage Manual

confirmed by a specialist and or


quantitative data via MRTS16 Appendix B
(TMR 2010c) soil analyses.
An objective assessment (see Section
13.6) is based on specified criteria that can
be applied to a range of environments.
The criteria are not necessarily
quantitative but, provided that they are
clearly defined, they can generally be used
to produce reliable assessments by both
experienced and inexperienced personnel.
A quantitative assessment of soil risk
involves a mathematical calculation of
expected soil loss.
A mathematical
calculation of soil loss has the potential for
being the most accurate and reliable of all
assessments and may be used as a design
tool for sizing sediment basins and
determining the placement and frequency
of runoff control measures.
As the purpose of this manual is
essentially to provide design information
for control measures, an objective
assessment of relevant factors is to be used
and is described here.
This assessment produces a numerical
rating of erosion risk that can be reliably
applied at both planning and design stages.

13.6

Objective Rating
System

An objective rating system for classifying


erosion risk according to a number of key
environmental factors is presented in this
section with Figure 13.6 outlining the risk
assessment process.
The factors and criteria used in the
classification
have
been
designed

Chapter 13
Erosion and Sediment Control

specifically for Queensland conditions


with the following factors used to assess
erosion risk at a site:
Rainfall erosivity;
Soil erodibility;
Slope gradient and length; and
Vegetation cover.
A site is individually assessed for each of
the above factors and allocated a rating of
1 to 5 coinciding with the scale shown in
Table 13.6.
The peak score is used to guide the user as
to the level of erosion control required on
site, though note should also be taken of
the average value to ensure that a practical
approach is adopted.
For example, a site may have low ratings
(say 1 or 2) for slope, erosivity and
erodibility, but a high rating for vegetation
(e.g. in arid areas where there is little or no
vegetation). In this case, the high rating
arising from the lack of vegetation could
be tempered by a low average rating.

Table 13.6 - Scores for Erosion Risk


Factors
Description

Score for Erosion


Risk Factor

Very Low

Low

Moderate

High

Very High

(Source: McDonald et al. 1990)

March 2010

13-9

13

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

13

Figure 13.6 Process for Assessing Erosion Risk

In all cases, it is important to note that the


erosion risk rating is a guide to be used in
conjunction with other information
relevant to the site.

13.7

Determining Erosion
Risk Factors

13.7.1 Rainfall Erosivity


In Queensland, the erosive potential of
rainfall, termed erosivity, is highly
correlated with annual rainfall and rainfall
intensity, which in turn are dependant
upon the number of wet days, number of
thunder days, temperature, latitude and
rainfall seasonality (Rosenthal & White
1980).
The ability of rainfall to cause erosion is
termed rainfall erosivity.
Rainfall
erosivity has been found to be correlated
with the ARI 2 year, 6 hour rainfall events
March 2010

13-10

(Rosewell & Turner 1992). A formula has


been derived to estimate an annual average
erosivity value - R-factor. The formula for
annual average rainfall erosivity is:

R 164.741.1177 S 0.6444
S

Where:
S = ARI 2 year, 6 hour rainfall event
(mm/h).
This formula is as presented in Best
Practice Erosion & Sediment Control
(IECA 2008). It can be used to determine
a site specific R-factor anywhere in
Queensland.
Both the average annual R and the highest
monthly percentage occurring during the
period of proposed construction activities
are used as criteria for assessing erosion
risk.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table 13.7.1(a) provides the average


annual R values for 15 sites within
Queensland.
Rainfall intensity is generally higher
during the summer months (December to
March) than the winter months, resulting
in a seasonal distribution to erosivity. This
seasonality is shown in Table 13.7.1(a)
where the percentage monthly contribution
to the average annual R value is listed.
Note that no locations in Table 13.7.1(a)
have a monthly R contribution above 10%
in any month from the end of March to the
beginning of December. Clearly, there are
advantages to programming construction
between April and November, especially
where other risk factors are high.

permeability of the soil profile and


coherence of the soil particles. However,
these properties are generally not
measured directly. Instead, they can be
estimated through the identification of
departmental soil groups using soil
features previously described.
In addition, field and/or laboratory tests
may be undertaken where soil types cannot
easily be identified. Relevant field tests
include those for dispersion, pH and
sodicity with specific laboratory tests able
to confirm erodibility when necessary.
Each of the eleven departmental soil
groups can be placed into one of the five
erodibility categories (ratings) on the basis
of the above soil features.

Table 13.7.1(b) may be used to derive the


erosivity rating using values interpolated
from Table 13.7.1(a) or calculated for a
specific site.

13.7.2 Soil Erodibility


The erodibility of soil is determined by the
rate of infiltration at the surface,
Table 13.7.1(a) - Monthly Percentage and Annual Rainfall Erosivity (R-factor) Values
Location

Jan

Feb

Mar

Apr

May

Jun

Jul

Aug

Sep

Oct

Nov

Bowen
Brisbane
Bundaberg
Cairns
Emerald
Gympie
Mackay
Mount Isa
Normanton
Rockhampton
Roma
Southport
Toowoomba
Townsville
Weipa

31.9

28.5

15.4

4.2

2.2

1.9

1.0

0.9

0.9

1.1

2.4

9.7

Annual
R
value
4,600

18.1

18.5

13.8

7.7

4.8

4.3

3.0

2.1

2.1

5.2

8.1

12.2

3,705

25.2

18.2

12.6

6.5

4.3

4.0

2.8

1.4

1.7

4.0

6.3

13.1

2,979

24.7

27.1

23.6

6.9

2.1

0.7

0.3

0.4

0.6

1.0

3.8

8.7

19,118

21.0

19.5

11.4

4.7

3.4

3.8

2.7

1.8

2.4

4.8

8.6

15.9

1,804

18.2

21.6

14.1

7.1

3.6

3.4

2.3

1.4

2.8

4.4

6.7

14.3

3,646

28.9

25.3

17.2

6.0

2.6

1.7

0.8

0.6

0.7

1.8

3.2

11.2

7,411

26.8

23.4

15.4

3.4

2.9

1.9

0.8

0.3

0.7

2.7

4.6

17.2

1,219

32.7

27.8

16.1

2.2

0.3

0.5

0.1

0.1

0.2

0.6

3.7

15.7

4,707

20.9

21.9

13.5

3.9

4.5

2.5

1.8

1.9

1.6

4.6

7.5

15.5

3,684

17.4

15.3

13.6

5.3

4.4

3.8

3.9

2.4

3.4

7.3

9.6

13.5

1,697

15.3

17.9

15.9

9.8

7.4

5.2

3.2

2.2

2.2

4.8

6.2

10.0

3,909

19.1

15.7

10.8

6.1

4.8

3.8

3.4

2.2

3.0

6.3

9.4

15.5

2,642

28.7

29.5

17.3

4.8

1.6

0.7

0.4

0.7

0.4

1.2

3.8

11.0

9,790

26.9

27.9

18.8

4.5

0.4

0.1

0.0

0.0

0.2

1.1

5.1

15.1

4,786

Dec

Source: IECA (Australasia) (2008)

March 2010

13-11

13

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table 13.7.1(b) - Erosivity Criteria

13

Average Annual
R1

Highest Monthly
Proportion of
Average Annual R2,3

Erosivity Rating

<1,500

not applicable

Very low (1)

1,500-4,000

< 25%

Low (2)

1,500-4,000

25%

Moderate (3)

4,000-10,000

< 25%

Moderate (3)

4,000-10,000

>= 25%

High (4)

> 10,000

not applicable

Very High (5)

Notes:
1 & 2 Using Table 6.14.1(a) or site calculated.
3
During period of construction.

Typical erodibility ratings for each


departmental soil group may be
determined using Table 13.7.2. In order to
use this table, the departmental soil group
must first be determined from field
inspections, and in some cases, laboratory
analysis.
Two methods are provided for the
determination of soil erodibility at a
particular site.
13.7.2.1 Method 1
Step 1 - Using the results of the soil
features analysis, allocate the soil sample
to one of the eleven departmental soil
groups. Wherever possible, use should
also be made of available soil maps.
Step 2 - Apply Table 13.7.2 to determine
the typical erodibility rating.
13.7.2.2 Method 2
It may be quite difficult placing some
profiles into a particular departmental soil
group whilst other profiles may clearly fall
within one departmental soil group but
possess some properties that are not
typical for the group. Soil erodibility is
variable within some departmental soil
groups. In all these situations, it will be
necessary to:
March 2010

13-12

(a) conduct a field dispersion test; or


(b) collect samples from all obvious soil
horizons for laboratory analysis.
Field tests should be conducted first, and
will often provide a definitive answer
whether the sample is dispersive or non
dispersive.
However, for those samples where doubt
remains, laboratory analysis may also be
required and the following procedure
should be followed:
Step 1 - Determine the soil horizon most
likely to be exposed during construction.
In the majority of cases, topsoil is removed
and stockpiled prior to construction, and
hence the determination of erodibility
should focus on subsurface and subsoil
horizons.
Step 2 - Obtain a sample of soil from the
relevant horizon.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table 13.7.2 - Typical Erodibility Ratings for Queensland Soils


Departmental Soil
Group
Uniform sands and
sandy loams
Uniform loams and clay
loams
Uniform non-cracking
clays

Uniform cracking clays

Sandy gradational soils


Loamy gradational soils
Texture contrast soils
2
(non dispersive)
Texture contrast soils
(dispersive)2

Waterlogged soils
(nonsaline)
Waterlogged soils
(saline)

Features
Incoherent sand, loamy sand and clayey sand and
coherent sandy loam with single grained to massive
structure
Coherent loams, sandy clay loams and clay loams with
massive to strong structure
Light to heavy clays with strong structure
- fine aggregates
- coarse aggregates
Light medium to heavy clays that shrink and crack open
when dry and swell when wet; gilgai microrelief common
Texture gradually increases from a sandy surface to
sandy clay loam or sandy light clay with depth; single
grain to massive structure
Texture gradually increases from a loamy surface to
sandy clay loam or clay with depth; massive to strong
structure
Sandy or loamy surface abruptly overlying a non
dispersive and generally friable clay subsoil
Sandy or loamy surface abruptly overlying a hard,
dispersive clay subsoil with
ESP 6 and/or Ca:Mg <0.5
ESP 15 and/or Ca:Mg <0.1
Uniforms sands, uniform clays, gradational soils, and
texture contrast soils that are saturated with nonsaline
water for several months of the year
Uniforms sands, uniform clays and texture contrast soils
that are saturated with saline water for several months of
the year

Erodibility
Rating1
Moderate (3)
Very low (1)
Very Low (1)
Low (2) to
Moderate (3)
Low (2) to
Moderate (3)
Moderate (3)
Low (2)
Moderate (3)

High (4)
Very high (5)
Variable3
Very low4 (1)
5

Made land

Created by human activities with no natural soil horizons

Variable

Dispersive buried soil


layers2

Any soil material buried below non dispersive soils with


ESP 6 and/or Ca:Mg <0.5
ESP 15 and/or Ca:Mg <0.1

High (4)
Very high (5)

Notes:
1.
The erodibility rating is based on the typical soil features for each departmental soil group but individual
profiles may have soil particular horizons that result in a different erodibility.
2.
Refer to Soils Manual for definitions of dispersive and non dispersive soils.
3.
The erodibility of non-saline waterlogged soils is variable and dependant upon soil features at a particular
location. Laboratory testing may be required if soil is to be drained.
4.
If these soils are drained a moderate to high erodibility rating may result.
5.
Made land can consist of any combination of human-generated rubbish or fill with or without entire or
part profiles from natural soils. Hence, they cannot be readily classified in terms of erodibility. Each
human or naturally generated horizon needs to be assessed separately and, if necessary, sampled for
testing.

March 2010

13-13

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Step 3 - Conduct field dispersion tests.


Soil Dispersion Indicator Test (SDI);
pH;
EC; and
field texture test for sodicity.
Step 4 - Interpret field results.
Step 5 - Determine the need for laboratory
testing.
If the field dispersion tests indicate a soil
horizon may be dispersive then laboratory
tests should be conducted to confirm the
erodibility of that soil.

Table 13.7.2.2 - Erodibility Rating for


Emerson Classes
Emerson Class

Erodibility Rating

1
2(1)
2(2)
2(3)
3(1) and 3(2)
3(3) and 3(4)
4
5
6
7
8

4 or 5
3 or 4
3, 4 or 5
4 or 5
1 or 2
3 or 4
1 or 2
2 or 3
1 or 2
1 or 2
1 or 2

Simple laboratory tests that can be


performed to assess erodibility include:

If all horizons are non dispersive or


unlikely to be dispersive then that soil
should be rated as very low to moderate
erodibility, depending on other soil
features.

(a) Emerson Dispersion Test; and

13.7.3 Slope Gradient

(b) exchangeable cations and ECEC tests


(which provide an assessment of ESP
and Ca:Mg ratio).

Slope gradient is used to determine the


impact of topography on erosion. It is very
difficult to directly modify slope gradient
although land shaping may be cost effective
in small areas. However, most construction
activities increase rather than decrease
slope gradient.

Step 6 (if required)

Test results should be interpreted


Table 13.7.2.2, which provides a
indication of erodibility based on
modified Emerson Class has
identified.

using
broad
which
been

Step 7 (if required)


The results of laboratory tests should be
reviewed in line with discussion within the
departments Soils Manual (2010).
Section 13.4.8 provides a framework for
determining whether a soil horizon is likely
to be dispersive based on the results of field
and laboratory testing. A soil should be
considered as having a high to very high
erodibility rating if, by using the
framework, any horizon is described as
being possibly, probably or definitely
dispersive.

March 2010

13-14

The slope gradient for a particular landform


can be described in terms of modal slope,
mean slope, minimum slope or maximum
slope.
Modal slope is defined as the most common
slope gradient occurring in a landform
pattern. Modal slope classes, based on
equal increments of a logarithmic scale,
have been used to normalise frequency
distributions of observed slope (Speight
1990).
Although not intended to represent natural
clustering of slope values (which have not
been shown to occur), the classes do
separate significant Australian landform
patterns. The modal slope classes are listed

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

in Table 13.7.3. Anecdotal evidence from


across the state suggests there is a strong
relationship between the slope classes and
soil erosion hazard.

an increase in soil loss by a factor of


2.6 (i.e. more than double).

The modal slope classes described in Table


13.7.3 are used to set criteria for separating
significant categories of erosion risk.

The criteria for slope length shown in Table


13.7.4 have been established using the
NSW Department of Housing (1998)
comparative table for the slope length
factor at construction sites.

Table 13.7.3 - Erosion Rating for Modal


Slope Classes

Table 13.7.4 - Erosion Rating for Slope


Length

Modal slope

Erosion

Slope Length

Erosion Rating

Rating

<5m
5 - 25 m
25 - 50 m
50 - 100 m
>100 m

Very Low (1)


Low (2)
Moderate (3)
High (4)
Very High (5)

Class
(%)

(degree)

Level

<1

0.35

Very Low
(1)

Very gently
inclined

1-3

0.35-1.45

Low (2)

Gently
inclined

3-10

1.45-6.45

Moderate
(3)

Moderately
inclined

10-32

6.45-18

High (4)

Steep

32-56

18-30

Very High
(5)

Very steep

56100

30-45

Very High
(5)

Precipitous

100300

45-72

Very High
(5)

Cliffed

>300

>72

Very High
(5)

Source: Speight (1990)

13.7.4 Slope Length


Increasing the slope length increases the
capacity of runoff water to concentrate and
thus increases the potential for detachment
and loss of soil particles. It is therefore
important to obtain an estimate of the mean
and range of slope length (and range of
slope lengths) at a particular site.
For example, research in the USA has
shown that doubling the length of a 9%
slope will lead to:
an increase in runoff by a factor of
1.8 (i.e. less than double); and

Source: McDonald et al. (1990)

13.7.5 Combined Effect of Slope


Gradient and Length
The ratings determined in Tables 13.7.3 and
13.7.4 need to be averaged in order to
provide an overall erosion rating for slope.

13.7.6 Vegetation Cover


The vegetation cover on the ground can be
assessed as described here by an erosion
risk rating assigned using the criteria shown
in Table 13.7.6.
Table 13.7.6 - Erosion Rating for Ground
Cover
Ground Cover

Erosion Rating

> 80 % cover
60-80% cover
40-60% cover
20-40% cover
< 20%

Very low (1)


Low (2)
Moderate (3)
High (4)
Very high (5)

13.8

Implications of Risk
Rating

The potential impact of rainfall erosivity


can be minimised by planning construction

March 2010

13-15

13

Department of Transport and Main Roads


Road Drainage Manual

13

activity during the period with lowest R


values, which in Queensland is generally
between May and October.
The erodibility of most soils is inherent and
cannot be modified. However, erodibility
can be reduced in dispersive soils by
reducing their ESP and increasing their
Ca:Mg ratio. This involves the addition of
calcium products (such as agricultural
limestone and agricultural gypsum),
mechanical mixing and leaching. This
process will work but results may not be
evident for a considerable period,
depending on type of product, method of
incorporation, degree of compaction and
weather.

Chapter 13
Erosion and Sediment Control

13.9

Environmental
Assessment - Review
of Data

The need for erosion and sediment control


will have been identified in the
environmental
assessment
and
in
subsequent erosion risk assessment,
drainage and other overviews of hydrologic
process outlined in other chapters.
The information which should be available
will include details of areas of the
environment considered vulnerable to
environmental harm and information
pertaining to the nature of potential erosion
and sedimentation related impacts.

Although a key factor in soil erosion, slope


length is easily modified by placing runoff
control structures across the slope that will
effectively shorten the slope length and
therefore minimise its impact.

Depending on the scale of the project, the


project environmental assessment should
include:

Slope shape primarily affects the layout of


runoff control structures. A convex shape
spreads water across the slope thus
decreasing the potential for detachment and
loss of soil particles. A concave shape
concentrates water towards the centre of the
landform thus increasing the erosive
capacity.
A planar slope neither
concentrates nor disperses runoff water.

an erosion risk rating for specific


areas within the site;

The rating system has been developed to


apply to disturbed land at construction sites.
It can also be used to compare the erosion
risk prior to establishing erosion control
measures with the risk after selected
measures are in place.

a map of soil types and their erosion


potential;

climate and stream flows;


topography and natural geographical
features (e.g. is the site within a
floodplain?);
proposed changes to the site
topography for each stage of the
project including extent of cut and
fill batters;
a map of existing vegetation, limits
of proposed clearing, and identifying
areas to be retained;
details of areas of cleared land at
each stage of the development, and
the period of time that each area will
be exposed for;

March 2010

13-16

Department of Transport and Main Roads


Road Drainage Manual

calculation of stormwater flows


within micro-catchments within the
site, for each stage of the project; and
nature and location of works that will
occur in close proximity to natural
waterways or other sensitive
environmental areas (e.g. wetlands).

13.10 Environmental
Management Plan
(Construction) and
Erosion and Sediment
Control Plans
13.10.1 An ESCP Overview
Risks associated with drainage works and
activities that disturb soil can be managed
through the preparation and implementation
of Erosion and Sediment Control Plans
(ESCPs).
The ESCP makes up one component of the
Environmental
Management
Plan
(Construction), or EMP(C). The EMP(C) is
written to ensure the project environmental
assessment is carried out.
When preparing ESCPs or undertaking
construction activities, legislative approvals
relating to erosion and sediment control
may be required.
These may include permits for:
undertaking works in an area of
preservation or high preservation;
wild river area;
Great Barrier Reef catchment; and/or
interfering with the flow of a
watercourse, taking of water from a
watercourse or removing vegetation
from a watercourse.
Project environmental personnel should be
consulted regarding provisions for permits.

Chapter 13
Erosion and Sediment Control

These legislative requirements are outlined


in Chapter 1, Section 1.4, and are further
detailed on the Office of the Queensland
Parliamentary Councils legislation web
page <www.legislation.qld.gov.au>, correct
as of May 2009.
The departments Standard Specification
MRTS51 Environmental Management
(TMR 2010c) requires a contractor to take
all reasonable precautions to ensure
construction works do not result in erosion
and
sedimentation
that
causes
environmental nuisance or harm outside the
construction zone. To achieve this goal, the
specification requires the contractor to
submit a design for the erosion and
sediment control measures for the site (as
part of the EMP(C)) at least seven (7) days
prior to site disturbance. In a project, this
constitutes a formal Hold Point, which is an
identified point in the construction process
past which the Contractor shall not proceed
without a direction from the Superintendant
(MRTS01 (TMR 2010c)).
ESCPs detail how the potential impacts of
erosion and sedimentation will be managed
on a project.
ESCPs will vary in content and detail
depending on the scale of works and the
physical characteristics of the project site.
ESCPs are dynamic documents that should
be modified with changes in project scope
where different soil profiles or different soil
types are disturbed.
They should be
reviewed regularly, when new information
and products become available, changes in
resource availability, when ESC devices
fail, or when construction procedures
change.
ESCPs should be compiled for the design
phase of a project and also for the
construction phase.

March 2010

13-17

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

The ESCP should be a living document


that is able to be reviewed throughout
duration of the project or until land
disturbance has been stabilised on the
project site.

13.10.2 Benefits of ESCPs


ESCPs generate significant benefits for the
project and the surrounding environment by
providing:
(a) a holistic and systematic approach to
ESC;
(b) contractors with an understanding of
ESC requirements;
(c) cost effective methods and timing for
ESC;
(d) an ability to reduce downtime
associated with wet weather, clean-up
costs and delays in project completion;
(e) a reduction in off-site impacts (e.g.
environmental, economic and social
impacts);
(f) a demonstrable approach to compliance
with relevant legislation and contract
documentation;
(g) an aid for auditing, monitoring and
maintaining site performance in relation
to ESC;
(h) guidance on how to use site runoff as a
resource (e.g. irrigation), rather than a
hazard or waste;
(i) a record of site soil data; and
(j) a record of site conditions
environmental constraints.

and

13.10.3 Timing and Acceptance


The ESCP is a requirement of the Contract
Documentation.
ESCPs are prepared by the Contractor as
part of the EMP(C), which is submitted to

March 2010

13-18

the Project Superintendent for acceptance


following the award of contract.
The assessment of the EMP(C) for
suitability by the Superintendant (and
subsequently any further amendments made
by the Superintendant) is a Hold Point
under MRTS51 (TMR 2010c).
It is a requirement of MRTS51 (TMR
2010c) that the Contractor undertakes the
design of temporary ESC measures for the
contract. The Contractor identifies in the
EMP(C) whether the Principals design will
be adopted or not.
A checklist for the preparation of an
Erosion and Sediment Control Plan is
available in Appendix 13A.
The timing for ESCP preparation and
approval is provided in Figure 13.10.3. The
actual number of days will be determined
during detailed design and should be
specified in the contract.

13.10.4 ESCP components


All ESCPs should contain drawings as a
minimum requirement but some projects
may not need an ESCP Report. This is
determined in the documentation from the
environmental assessment as outlined in the
departments Road Project Environmental
Processes Manual (DMR 2004b).
A
Document Control Status and Revision
Number must also be inserted in the front
of the ESCP report.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

13

Tender documents
issued

Contractor submits
tender

Under MRS 11.51,


assessment of the
EMP(C) by the
Superintendant is a HOLD
POINT

Contract award

Contractor submits the EMP(C)


usually at the pre-start meeting,
but is required 28 days from
contract award

Project Superintendant notifies


acceptance (i.e. a letter in the form
of a Contract Plans Approval) or
otherwise of EMP(C)

Commence
construction

Revise Contractors
EMP(C)

Under MRS 11.51, any


amendment made to the
EMP(C) is a HOLD POINT

Figure 13.10.3 - EMP (Construction) and ESCP Acceptance Process.

March 2010

13-19

Department of Transport and Main Roads


Road Drainage Manual

13

ESCPs may
information:

contain

the

Chapter 13
Erosion and Sediment Control

following

(a) Site characteristics and features:


catchment description (including site
access point(s), landscape, receiving
catchments, catchment boundaries,
watercourses, drainage lines and
water bodies within and adjacent to
the site (from site data and
environmental assessment), etc;
topography
including
contours
(obtained
from
environmental
assessment);
soils, erosion risk assessment and
zones (obtained from environmental
assessment);
water quality (obtained
environmental assessment);

from

vegetation;
hydrology; and
weather and climatic patterns
expected over the construction
period.
(b) Construction Drainage Plans;
(c) Selection of Control Measures;
(d) Scheduling of works (proposed time
schedules),
design,
installation,
maintenance and removal of control
measures;
(e) Monitoring, Auditing and Reporting;
(f) Appendices:
ESCP Drawings (controls, drainage
path, analysis of contour plan of the
site, disturbance / non-disturbance
zones, location of stockpiles, borrow
pits, storage areas, site office,
amenity
blocks,
location
of
sidetracks, etc);

March 2010

13-20

Fact Sheets for ESC measures (see


Appendix 13B);
calculations (e.g. hydrology
sediment basin sizing);

or

ESCP Checklist (Appendix 13A);


Supporting data (aerial photos, construction
program, EMP requirements, revegetation
specification).
The ESCP should include drawings and
information detailing:
(g) features of the site including contours,
vegetation,
drainage
paths
and
receiving waterways;
(h) relevant construction details of all ESC
structures, compounds and temporary
works;
(i) all permanent and temporary ESC
measures, including the control
measures to be implemented in advance
of, or in conjunction with all
construction works; and
(j) an installation sequence for all ESC
measures that highlight such issues as
the construction and stabilisation of all
culverts and surface drainage works at
the earliest practical stage.
These drawings should be attached as an
appendix to the ESCP report or used as the
ESCP if no report is required and should:
include a North Point;
be to scale (preferred scale between
1:500 to 1:2,500);
include a Title Block (should include
Project Name, ESCP Title, Plan No.,
Date, Chainage); and
show a legend (should include
standard symbols for control
measures).

Department of Transport and Main Roads


Road Drainage Manual

For more detail on ESCP content, refer to


Appendices 13A and Book 1, Chapter 5:
Preparation of (ESC) Plans from IECA
(2008).

13.11 Preparing ESCPs


The ESCP should be written by a
professional with experience in drainage
and ESC.
It is presented to the
departments Principal as the contractors
commitment to drainage management and
ESC for the project. As such, it needs to be
prepared in a format and to a level that can
be easily understood and implemented by
the contractors staff.
It is advantageous to develop and
incorporate the ESCP in conjunction with
the Site Construction Program (including
EMP(C)).
This strategy assists more
efficient cohesion of site works with ESC
measures.
Preparation of an ESCP will also require a
review of the data obtained using Chapter 4
Data Collection. Reference to general
principles for drainage, ESC, and
revegetation should also be incorporated
into the ESCP. Furthermore it will require
the completion and documentation of a
detailed risk assessment.
ESCPs are generally prepared in a number
of sequential steps that have been ordered
in a logical way to reflect an understanding
of ESC requirements. This starts from
providing an understanding of the physical
features of the site. This will also ensure
relevant information is reviewed prior to
the selection and installation of appropriate
control measures, and the selection of
monitoring and maintenance procedures.
IECA (2008) suggests that for higher risk
sites, a conceptual ESCP should be written.
The purpose of this type of preliminary plan

Chapter 13
Erosion and Sediment Control

is to develop controls and sequence of


events before key site layout and design has
been confirmed.
High risk sites may also benefit from being
managed by a series of ESCPs (rather than
one single plan) that is applicable to each
stage of the project (IECA 2008).
Ultimately, the site characteristics will play
a major role when choosing appropriate
ESC measure selection, design and layout.
It is essential to note that one formula for
ESC will not apply for all regions in
Queensland.
Just as environmental
conditions will vary from region to region,
so will the selection, design and layout of
ESC measures.
Interpretation and
application of measures need to be tailored
from fundamental principles to suit each
individual project and site characteristics.

13.11.1 Step 1 - Review Site


Characteristics
Step 1 in the preparation of an ESCP is to
obtain an understanding of the sites
physical features and the identification of
physical constraints. The planning and
design of effective temporary ESC
measures involves assessing construction
options with a thorough understanding and
analysis of the proposed changes to the site.
Effectively, the planning and integration of
ESC should be incorporated into
construction and/or maintenance activities.
Much of the required information for Step 1
will have been obtained from the processes
undertaken using other chapters of this
manual. Relevant information will also be
available from the project Review of
Environmental Factors (Planning). The
EMP (Planning and Construction) may also
be a key source of information relevant to
mitigation measures to reduce potential

March 2010

13-21

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

ESC impacts of the project. More detail


and other principles and relevant
information should be available in the
contract documents and Book 1, Chapter 5:
Preparation of (ESC) Plans from IECA
(2008).

If applicable, mark on the overlay of the


construction drawings the following items:

Existing site characteristics should be


presented on an overlay of the construction
plans. If applicable, ensure the following
items (and relevant items listed in Section
13.10.4) are marked on an overlay of the
construction drawings:

locations of site access points,


parking areas, site compounds,
storage areas, and stockpile areas.

(a) existing and proposed


drainage structures;
(b) limits
of
site
documents); and

(from

permanent
contract

(c) special conditions (refer Chapters 7, 9,


12
and
obtained
from
the
environmental assessment);

13.11.2 Step 2 - Identify Areas of


Construction Disturbance
The Site Characteristics overlay in Step 1
should depict the major physical
characteristics of the site, which may affect
or be affected by land disturbance, erosion
and sedimentation.
The purpose of Step 2 is to obtain an
understanding of what physical landchanges may result from construction. A
construction information overlay may need
to be prepared to show staging of
construction, locations for stockpiles,
compounds, access tracks and ancillary
construction activities.
When this information is overlaid on the
construction drawings with the site
characteristics overlay, it should be possible
to visualise how and where construction
works may lead to erosion and
sedimentation.

March 2010

13-22

extent of clearing and grubbing;


cut and fill locations;
clearing and construction stages; and

13.11.3 Step 3 - Selection of


Controls
Step 3 involves using the information
prepared in Steps 1 and 2 to assist in the
selection of appropriate control measures
for the project. ESC selection tables from
Chapter 7 should also be referred to in this
step. Selected measures should be marked
on construction drawings that are suitable
for use in the field.
13.11.3.1 Standard Symbols
Standard symbols for control measures are
contained in standard drawings and should
be used where practical. When planning
each stage of works, standard symbols will
need to be marked on the construction
drawings for each stage. IECA (2008)
contains a comprehensive list of standard
symbols that should be referred to as well.
One of the first tasks in the ESC measure
selection is the establishment of the existing
and proposed catchment boundaries and to
identify those catchments channelling clean
(i.e. upslope unpolluted) runoff and those
carrying dirty (i.e. site-generated, sediment
laden) runoff.
This is critical to managing how water
enters, moves through, and exits the site
during site works.
Once the clean / dirty catchments have been
determined, the types of physical soil /
water interactions that will occur during the

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

construction phase of the project need to be


established. This means the contractor will
need to specify site-specific ESC measures
that best manage the site according to
erosion prevention and/or sediment
retention, or which erosion mechanisms
may affect the site (e.g. raindrop impact,
sheet flow, concentrated flow, rill and gully
erosion, etc). ESC measures need to be
adapted according to these and other
processes.
Prior to selecting ESC measures, several
factors should be considered. These factors
are:
opportunities (including
available site materials);

existing

constraints (including limitations of


ESC devices vs. site characteristics)
potential for environmental harm;
permanent and
measures; and

temporary

ESC

preferred order of controls (i.e.


drainage,
velocity,
erosion,
sedimentation) for each individual
project.
Chapter 4 of IECA (2008) lists factors that
are used for selecting devices most
appropriate for the site characteristics, type
of works involved and nature of the project.
These factors are listed below:

6. technical capabilities of site personnel


and potential training or specialist
installation; and
7. performance over time to a desired
standard.
13.11.3.2 Opportunities and
Constraints
The Contractor should identify ESC
opportunities and constraints relating to the
project. This may include consideration of:
available resources (e.g. surplus rock,
access for hydromulching);
topography (e.g. opportunities to
divert water);
receiving waterways; and
timing of construction.
The duration of construction and the
lifespan of control measures must be
considered in addition to the availability of
measures.
One of the key factors to be considered
when selecting ESC measures is the timing
of construction. Where construction is
scheduled to occur during peak rainfall
months, the potential for erosion will be at a
maximum. Hence, it will be necessary to:
(a) minimise the time of exposure of any
part of the site (e.g. staged clearing of
vegetation); and/or

1. applicability to all site conditions;

(b) reschedule construction.

2. availability of materials (local suppliers,


delivery timeframes, etc);

13.11.3.3 Potential for


Environmental Harm

3. community acceptance;

The contractor should identify and prepare


controls for the activities:

4. benefit / cost ratio analysis;


5. ongoing structural durability throughout
natural and anthropogenic processes and
incidents;

likely to cause most environmental


harm; and
the areas at most risk of experiencing
environmental harm.

March 2010

13-23

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

The extent of potential impacts of each


activity should be determined.

should be investigated as part of the design


process for all permanent measures.

The project environmental assessment will


assist in providing such information. For
example, disturbance to an area of
dispersive soil upstream of a sensitive
watercourse is considered a high-risk
activity. Such activities / areas / impacts
should be given highest priority in terms of
planning for, applying and maintaining
ESC. This will ensure that most effort is
given to the areas likely to cause most
environmental harm.

As a responsibility of the designer,


permanent ESC measures will typically
function during construction and after the
construction phase has finished.

13.11.3.4 Permanent and


Temporary ESC Measures
Contractors and designers need to identify
which ESC measures are best suited for the
needs of the road activity, nature of works,
surrounding environment, and timeframe
for which they will be applied for.
Environment and project-specific factors
will determine whether temporary or
permanent controls are required.
Implementation and usage of temporary
ESC measures usually would not exceed
the duration of the construction phase.
During construction, temporary control
measures are designed, implemented and
must undergo maintenance if not
performing to its intended capacity. They
provide temporary protection whilst the site
is exposed to erosive elements, until more
permanent protection is constructed.
Permanent measures provide permanent
drainage,
velocity,
erosion
and/or
sedimentation control post construction.
They often require formal design and are
normally more costly.
In many cases, the most effective (and
practical) approach will involve the siting
and design of temporary controls to allow
conversion to a permanent measure. This

March 2010

13-24

13.11.3.5 Erosion Control


Techniques
Temporary and permanent erosion control
techniques are used to prevent controllable
erosion and to minimize the adverse effects
of sedimentation. As mentioned, good
erosion control techniques will not only
decrease the dependence on sediment
controls, but will reduce the long term costs
associated with the site.
Appendix 13B comprises a set of fact
sheets that describe the temporary and
permanent control measures in terms of:
lining controls
drainage controls
velocity controls
erosion controls
sedimentation controls
(a) Lining and Drainage Controls
Although lining and drainage controls are
dealt with separately in Appendix13B, the
two controls are usually linked and are dealt
with together in this section.
Drainage paths where stormwater enters the
site will have been identified and mapped in
the EMP(C) and on the overlay during
ESCP preparation: Step 1 (see Section
13.10.1). This information should then be
used to select and locate appropriate
drainage controls.
Where practicable, clean, overland drainage
flow should be intercepted and conveyed
around or through the construction site

Department of Transport and Main Roads


Road Drainage Manual

without flowing across disturbed areas.


This means that clean run-on water needs to
be effectively intercepted, diverted, and
disposed of in a manner that doesnt cause
erosion or where it may be contaminated by
disturbed soil or dirty water. The separation
of clean run-on and dirty runoff water
should be the objective; that is, never to let
dirty water mix with clean water.
Figure 13.11.3.5(a) illustrates a concrete
lined diversion drain redirecting clean
water from an exposed soil slope.

Chapter 13
Erosion and Sediment Control

(b) Velocity Controls


The quantity of soil eroded is generally
proportional to soil erosivity, the velocity
and duration of runoff. That is, the longer
runoff occurs for, the greater the potential
for soil erosion for a given soil erosivity.
Reducing flow velocity is thus a key
method for minimising the potential for soil
erosion and subsequent sedimentation.
Figure 13.11.3.5(b) shows a rock-lined
channel which dissipates the erosive energy
of the runoff, while protecting exposed soils
and collecting mobilised soils.

Figure 13.11.3.5(a) - Concrete - Lined


Diversion Channel.

Use of existing stable flow paths (i.e.


watercourses or drains) is preferable
however in the absence of these, it will be
necessary to construct:
temporary drains or watercourses; or
permanent waterways prior to the
commencement of other works.
Measures available to control drainage with
different linings as temporary of permanent
controls are outlined in Table 13.11.3.5(a).

Figure 13.11.3.5(b) - Velocity Control


Rock-lined Channel.

Areas within the construction site likely to


be subject to erosive flow velocities should
be identified and protected by the
appropriate control measures.
Velocity control measures and devices are
outlined in Table 13.11.3.5(b). Most rely
on roughening the flow surface or
decreasing the longitudinal gradient of the
flow path.

March 2010

13-25

13

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Table 13.11.3.5(a) - Measures for Drainage Control


Practice

Velocity
Range

Application

Intended Duration of
Usage

Channels with
impervious linings

High

Protection against high


velocity.

Permanent

Chutes

High

Passage of flow down steep


slopes or embankments.

Permanent

Diversion Banks

Low Medium

Diversion of flow away from


disturbed areas

Permanent

Diversion Channel /
Drain

Medium

Large drainage areas

Permanent

Erosion Control Mats / Medium Geotextiles


High

Used to line channels where


medium to high velocities are
expected.

Permanent or
Temporary

Outlet Protection

Control of erosion at outlets.


Reduces velocity and
dissipates energy.

Permanent or
Temporary

Reinforced Turf Lined


Channels

Medium High

Alternative to hard channel


linings in urban environment.

Permanent or
Temporary

Rock Lined Channels

Medium

Temporary and permanent


channels

Permanent or
Temporary

Rock Mattress
Channels

High /
Turbulent

Areas with turbulent flow or


Permanent or
high velocity. Can be used as a
Temporary
permanent measure

Slope Drains

High

Conveyance of flow down long


or irregular grades

Temporary

Temporary
Watercourse
Crossings

Low-High

Most types of watercourses


subject to environmental
impacts

Temporary

Turf Lined Channels

Low -

Alternative to hard channel


linings in urban environment

Permanent or
Temporary

Medium

March 2010

13-26

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table 13.11.3.5(b) - Measures for Velocity Reduction


Practice

Application

Intended Duration of Usage

Check Dams
- Rock

Act as a physical barrier and


reduces gradient through ponding.

Permanent or Temporary

Drop Structures

Can be created using rock, timber


or geotextile. Reduce gradient of
channel, but require stabilisation at
base of drop

Permanent or Temporary

Lengthen Flowpath

Where sufficient room exists,


velocities may be reduced with
subsequent lowering of invert
Permanent or Temporary
grade, or through lengthening of the
flowpath. The latter may be
achieved using bunds or baffles

Surface Roughening

Increase roughness through


vegetation, gravel or other barriers
to shallow flow. However, this will
not always be cost effective.

- Recessed Rock
- Gravel / Sand Bag

(c) Erosion and Sedimentation Controls


Erosion control practices can protect
exposed soil or reduce the duration of soil
exposure. Sediment-trapping controls may
be effective for preventing medium to
coarse sediment like sand and silt from
entering watercourses and drainage lines.
However, erosion controls are designed to
prevent or minimise the amount of sand,
silt, clays and fine sediment from being
removed or transported from a site. Hence,
erosion control should be used to prevent
on-site degradation and the occurrence of
sedimentation, whilst sediment controls are
more useful for the prevention of off-site
transport and impacts associated with
deposition of already-eroded, coarser
particles.

Permanent or Temporary

Logically, sedimentation is a product of


erosion, therefore minimising erosion will
ideally assist in the efficacy of sediment
controls. Therefore, erosion prevention or
minimisation should be a primary focus of
onsite drainage and ESC.
A review of the soils data mapped in Step 1
(see Section 13.10.1) will help identify
those areas currently disturbed. Step 2
identified areas of the site that are planned
for disturbance (i.e. areas of cut and fill).
This knowledge of soil types and risk
ratings is very important in the selection of
erosion control measures. For example,
sodic soils are highly erosive and will
readily mobilise if not stabilised and
protected.
Hence, better practice
management requires that disturbance to
sodic and other erosive or dispersive soils is
minimised by practical and appropriate
March 2010

13-27

13

Department of Transport and Main Roads


Road Drainage Manual

13

ground protection and erosion prevention.


This will reduce the potential for sediment
mobilisation and deposition in watercourses
and drainage lines.

Chapter 13
Erosion and Sediment Control

Figure 13.11.3.5(d) shows a partly


revegetated side slope as an erosion control
measure.

Measures for controlling erosion are


outlined in Table 13.11.3.5(c).
These
practices are common for road projects in
Queensland. A combination of these below
mentioned control measures with flow
velocity control devices can effectively
decrease erosion and sedimentation
processes.
Some of the most effective methods of
erosion prevention and minimisation rely
on keeping soils covered (i.e. natural
vegetation coverage) for as long as
possible. This means that if possible and
practicable, clearing and construction
activities should be staged as to reduce the
amount of bare soil exposed for the shortest
time allowable.
Figure 13.11.3.5(c) is an example of a
failure to apply appropriate erosion
protection, and an incorrect choice of
sedimentation control. The consequences
of this are large amounts of erosion from
the slope and sediment deposition at the
bottom of the slope.

Figure 13.11.3.5(c) - Example of


Incorrect Erosion Control and
Sedimentation Control

March 2010

13-28

Figure 13.11.3.5(d) - Example of a Partly


Revegetated (left side) Slope

13.11.3.6 Sediment Retention


Techniques
Sedimentation management devices and
controls should be installed to retain
mobilised sediment on site and to prevent
sediment from entering waterways.
Sediment is typically exported as a result of
erosion, therefore, the most efficient form
of sediment control being the prevention of
erosion. Hence, sediment controls are
selected only after first choosing effective
drainage, velocity and erosion controls.
When soil is eroded, sediment particles are
suspended in runoff and are subject to
movement and deposition.
Sediment
deposition can have significant impacts on
drainage systems and aquatic ecosystems.
Erosion and sediment transport can result in
loss of soil and land degradation, which
imposes
ecological
and
economic
consequences for the landholder. The cost
of cleaning out drainage structures or
clearways can be significant as well.
Sedimentation management and control
measures are outlined in Table 13.11.3.6(a).

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

13

Table 13.11.3.5(c) - Measures for Erosion Control


Practice

Chemical Surface
Stabilisers

Application

Intended Duration of Usage

Dust control on large areas such as haul


roads and access tracks. Some products
can be used as an alternative to erosion
control blankets.

Three-dimensional honeycomb-shaped
mesh that protects unbound or exposed
Geocellular
steep cut and fill slopes or earth channels
Confinement System
from hydraulic flows, erosion and
downward slip of unbound materials.
Erosion Control
Blankets

Protect against raindrop impact erosion


and sheet flow down batters.

Permanent or Temporary
depending on construction
materials

Permanent or temporary.

Temporary

Mulching
- timber
- rock
- straw
- hydro

Protection against raindrop impact erosion


usually as a channel liner when in
conjunction with revegetation. Stabilises
exposed surfaces. Also reduces moisture Left in situ after application.
loss from the soil.

- bitumen emulsion
Revegetation for
Used on disturbed, cleared or graded
Erosion Control
areas and stockpiles. Provides surface
(refer Appendix 13D) protection and soil stability.
Surface Roughening

Increases surface infiltration, minimises


rutting and reduced wind erosion.

Temporary or Permanent.

Temporary

Figure 13.11.3.6 is an example of a


sediment fence that requires maintenance
for excessive sediment build up.
In addition to these sediment controls, the
techniques shown in Tables 13.11.3.6(b)
and 13.11.3.6(c) may be used for dewatering and in-stream sediment control.
With all in-stream sediment controls,
amounts of sediment that exceed regional
water quality objectives or parameters must
be prevented form entering the main body
of the watercourse. Clean and dirty water
must not be allowed to mix.

Figure 13.11.3.6 - Poorly Maintained


Staked Sediment Fence

March 2010

13-29

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Table 13.11.6.3(a) - Measures for Sedimentation Control


Practice

Application

Intended Duration of Use

Construction
Exits

Removal and trapping of sediment from vehicles


leaving site.

Temporary

Drop Inlet and


Pipe Inlet
Protection

Used to trap entrained sediment prior to entering drain


inlet.

Temporary

Vegetative
Buffer Zone

Includes buffer zones and grassed filter strips.


Involves filtering and trapping of sediment runoff
in areas of sheet flow. Most effective on sandy
soils

Permanent or Temporary

Rock Sediment
Trap

Placed in gullies and other flowpaths to trap


entrained sediment. Use to form small sediment
ponds.

Permanent or Temporary

Sediment
Basins

Used to pond water and settle sediment-laden


runoff.

Permanent or Temporary.

Sediment
Fences

Used to reduce velocity and pond runoff to


promote settlement of entrained sediment. Not to Temporary
be used in concentrated flowpaths

Table 13.11.3.6(b) - De-watering Sediment Control Techniques


Practice

Filter Bags and


Filter Tubes

Description

Large reinforced, non-woven,


geotextile filter bags attached
to the end of a hose or pump

Typically a temporary aboveground pond formed by and


earthen embankment,
Sediment Basin
however can be below ground
effectively making it a
sediment basin

In stream sediment controls must be


designed so that no more than 50% of the
channel width is obstructed by in-stream
sediment controls at any one time.
Unpolluted water flows are not to be
obstructed.

March 2010

13-30

Application

Only suitable for low to medium flow rates


depending on the sediment content of the
water and the total surface area of the filter.
The primary action is that of gravitational
settlement of the sediment, thus these ponds
are usually significantly larger than filter
ponds. They are normally used for dewatering large volumes of water at the
beginning of each day.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

Table 13.11.3.6(c) - In-Stream Sediment Control Techniques


Practice

13

Description and Application

Floating Silt
Curtain

Geotextile filter fabric (silt curtain) suspended vertically in a water body


between contaminated and uncontaminated water.

Rock Sediment
Trap

A rock dam constructed across a drainage channel or small watercourse.


The dam is formed primarily from coarse rock with a filter layer on the
upstream face. The upstream filter layer may consist of a geotextile filter
cloth or a layer of clean aggregate.

A standard U-shaped sediment fence staked across a channel. A wire mesh


Staked Sediment backing is normally placed behind the fabric to provide additional support.
Fence
Only used as a precaution in minor drains when flow is not expected but
sediment flow may still occur due to non-stormwater flow.

13.12 Design Standard and


Selection of Erosion
and Sediment Control
Measures
In many cases, erosion and sediment
controls are not designed to a specific
standard. Instead, they are located on site
in keeping with the size of the contributing
catchment, available materials, the physical
constraints of the site, climatic/seasonal
conditions and a maintenance schedule.
However, in particular circumstances, such
as when a site is considered highly erodible
or when works are being proposed in an
area of high environmental sensitivity, a
specific design standard may be required
and detailed in the EMP(C).
Detailed design can be undertaken using
Appendix 13B and the respective guidelines
provided in subsequent design chapters of
this manual.
Selection criteria, design standards and
technique selection are discussed further in
Best practice erosion and sediment control,
(IECA) (Australasia) (2008). Section 4 of
Book 1 in IECA (2008) contains tables that

list suitable ESC methods for various


gradient drainage, velocity control, outlet,
erosion control, lining, dust control and
sediment control measures.
Factors that can determine the required
design standard will include:
erosion risk rating;
environmental sensitivity; and
duration of construction.
Selection of the appropriate design standard
can be made by referring to Table 13.12, or
by use of the following equation (Hunt
1992):
ARI =

L
ln(1 P)

where:
ARI = design average recurrence
interval (yr);
L = design life (yr);
P = acceptable probability
exceedance (fraction); and

of

ln - denotes natural logarithm.

March 2010

13-31

Department of Transport and Main Roads


Road Drainage Manual

13

Chapter 13
Erosion and Sediment Control

Table 13.12 - Suggested Design Average Recurrence Intervals for Various Erosion and
Sediment Control Measures on Construction Sites.

Source: Hunt 1992

For selection of a design standard for


sediment basins, reference should be made
to Chapter 12 in this manual and Australian
Runoff Quality A guide to Water Sensitive
Urban Design (EA 2006).
Some control measures, such as temporary
sediment basins require formal planning
and design to ensure long term
effectiveness and sustainability. Whilst
ESC measures will not always require
formal design, relevant factors which need
to be considered are provided in the Fact
Sheets in Appendix 13B and in Book 1
Chapter 4 of IECA (2008).
Much of the information required to
undertake formal design will have already
been undertaken when applying earlier
chapters of the manual. For example,
estimates of rainfall and runoff needed for
sizing measures such as diversion banks
and culvert outlet dissipaters will have been
calculated. This may reduce the need for
additional calculations.
Chapter 4 of IECA (2008) can assist in this
process as it lists default Drainage, Design,

March 2010

13-32

Erosion Control and Sediment Control


standards.

13.12.1 Select Controls


In order to select erosion and sediment
controls that are most appropriate for the
site environmental conditions and the
proposed construction program, the user or
designer should identify each of the
following for the project site:
(a) erodibility rating / soil type;
(b) areas of suitability;
(c) flow type; and
(d) actual and intended life of control.
With an indication of the measures most
suitable for the site conditions, detailed
design may then proceed.
The most
appropriate control measures may not be
fully
identified
until
construction
commences and measures are implemented.
Therefore, it is important that contract
design documents are flexible and dynamic
documents.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 13
Erosion and Sediment Control

13.12.1.1 Erodibility Rating / Soil


Type

condition of the device and the space within


the road reserve.

Erodibility ratings and general soil types of


the project site are discussed in Sections
13.4 to 13.8 and in the departments Soils
Manual (2010). These principles should be
reviewed as well as soil / erodibility maps
for the site (should be available in the
EMP(C)).

13.12.2 Treatment Train Approach

13.12.1.2 Areas of Suitability


The user should assess the control measures
against the constraints imposed by the
surrounding environment to gain an
understanding of the most suitable
measures for the site conditions.
13.12.1.3 Flow Type
The type of flow present will influence both
the potential for erosion and the type of
measure to prevent erosion.
Section 13.10.3 briefly mentions different
types of erosion, touching on flow types.
Sheet flow occurs as a relatively uniform
depth of water in the form of a surface
sheet.
This type of flow is often
characteristic of flat to gently sloping sites
where flow would not tend to concentrate in
rills, channels or contours.
Concentrated
flow
occurs
where
depressions or small channels are eroded
and flow concentrates. It often forms after
sheet flow has travelled some distance.
Steep sites will tend to feature concentrated,
high velocity flow at most locations.

A recommended approach to maximise the


use of ESC measures is the treatment train
approach.
This technique uses a
combination of ESC measures in a
sequence to allow a more efficient means of
managing erosion and sedimentation on and
from a construction site.
The advantages of using treatment trains are
derived from the ability of each treatment
device to individually target drainage,
velocity, erosion or sedimentation more
effectively as a sequence; rather than using
a singular device for a site.
Therefore, the combined installation of a
series of connected treatment devices may
be able to better manage erosion and
sediment processes.
For example, mulched slopes, channelling,
rock sediment traps, and grassed swales
designed and installed specifically to be
used in conjunction with one another can
reduce velocity and a large percentage of
eroded sediment prior to discharging from a
site.
Site-specific characteristics (i.e. existing
environment and design plans) will largely
determine the appropriate treatment
sequence.

13.12.1.4 Lifespan of Controls

The efficacy of treatment train style ESC


will rely significantly on correctness of
installation, practicality, appropriateness,
and regular maintenance.

Correct
selection,
installation
and
maintenance will maximise the lifespan and
functional capability of erosion and
sediment controls.

Figure 13.12.2, taken from Austroads


(2003), illustrates a spectrum of typical
treatment devices and the size of pollutants
they are designed to remove.

The frequency and type of maintenance


activity will depend on the type of device,

March 2010

13-33

13

Department of Transport and Main Roads


Road Drainage Manual

13

Particle Size
Grading
Gross Solids
>5000 Um

Coarse to Mediumsized Particulates

Chapter 13
Erosion and Sediment Control

Treatment Mewasures
Treatment Measures

Gross
Pollutant
Traps

Sedimentation
Basin
Grass Swales
(Wet & Dry)
&
Filter Strips

5000 Um to 125 Um

Fine Particles
125Um - 10Um

Surface
Flow
Wetlands

Infiltration
Systems

Sub-Surface
Flow
Wetlands

Very Fine Colloidal


Particulates
10 Um - 0.45Um
Dissolved Particles
<0.45Um

Figure 13.12.2 Typical Treatment Devices and Pollutant Removal (Austroads 2003).

The figure is a comparative table showing


the intended target pollutant removal sizes
for each category of stormwater treatment
measure. The diagram demonstrates the
effect of the treatment train to remove a
broader variety of pollutant sizes before
releasing run-off into receiving waters.
When using the diagram, consider run-off
to be flowing in a downward gradient from
left to right. It is seen then, how each
treatment stage should remove a particular
range of pollutants, thus improving the
efficiency of run-off treatment techniques.

13.13 Installation of Control


Measures
The installation of control measures should
be coordinated with the projects
construction works. Measures should be
installed prior to and during clearing and
grubbing, topsoil stripping and stockpiling,
installation of access tracks and compounds
and cut and fill operations.

March 2010

13-34

The fact sheets in Appendix 13B contain


installation information for each listed
device.

13.14 Maintenance of
Control Measures
Regular maintenance of drainage, erosion
and sediment controls is as important as
correct installation. At all times during
their required operational life, ESC devices
must be maintained in proper working order
(IECA 2008). The ESCP should therefore
document proposed maintenance activities
for each control measure selected.
Requirements
for
self-auditing
the
effectiveness of measures and compliance
with the ESCP should also be documented.
This should include details of the frequency
of all maintenance and auditing activities.
For maintenance of ESCs during
construction, the defects liability period and
the on-maintenance phase, there are two
basic requirements. These are outlined in

Department of Transport and Main Roads


Road Drainage Manual

the following sub-sections, Regular CleanOut and Repair and Replacement.

13.14.1 Regular Clean-Out


The frequency of maintaining and/or
replacing control measures should be based
on the type of measure, type of soil, type of
vegetation used and expected runoff
characteristics.
The accessibility of a
control measure after a rainfall event must
also be taken into account when
determining the frequency of cleaning and
hence the required size of the control. As a
general rule, control measures should be
cleaned out when 60% full of sediment.
The sediment removed from control
measures should be properly disposed of in
accordance with the contract requirements.
Contract requirements should also state the
location for sediment disposal. Removed
sediment must be prevented from returning
to the ESC device or to other drainage lines
or watercourses during subsequent rain
events.

13.14.2 Repair and Replacement


The second key maintenance requirement
includes the repair and replacement of
deteriorated materials within the control
measure (e.g. sediment fence fabrics).
Controls that are continually damaged may
indicate the need for additional or
alternative controls. If a control measure is
to be replaced, its replacement should be
chosen based on the compatibility of the
objectives
outlined
in
the
EMP
(Construction).
Effective management of site access and
appropriate ESC practices can minimise the
need for repair and replacement.
All temporary control measures should be
inspected and maintained where necessary
immediately after a rainfall event (e.g.

Chapter 13
Erosion and Sediment Control

removing excess sediment that has


deposited behind a sediment fence). Areas
adjacent to creeks and drainage paths
should have priority with respect to
inspection and protection followed by
devices protecting stormwater inlets.

13.15 Removing Control


Measures
Before the removal of all temporary control
measures, the site should be inspected and
stabilised.
This should generally be
specified in the contract. In addition, any
specific requirements for removing control
measures should be documented in the
ESCP report or on the ESCP drawings.
Generic guidelines are provided below:
(a) check for evidence of erosion prior to
removal to ensure erosion will not
occur through the removal, and after
the removal of control measures;
(b) remove and dispose of any deposited
sediment in an appropriate manner;
(c) where practical, clean and store ESC
materials for re-use on other projects;
(d) dispose of waste materials collected by
control measures (i.e. damaged
sediment fence fabric) with nominated
waste disposal areas;
(e) ensure that any machinery used to
remove control measures does not
cause excessive ground disturbance;
(f) ensure any revegetation of the area has
stabilised prior to the removal of ESC
and after their removal; and
(g) stabilise all areas disturbed during the
removal process.
Where it is evident that a temporary control
should not be removed, the contractor will

March 2010

13-35

13

Department of Transport and Main Roads


Road Drainage Manual

13

need to seek
superintendent.

direction

from

Chapter 13
Erosion and Sediment Control

the

13.16 Monitoring, Auditing


and Reporting ESC
Monitoring, auditing and reporting of ESCP
implementation is a requirement of the
departments
standard
specification
MRTS51 (TMR 2010c). This is part of
departments
environmental
reporting
responsibilities under the Environmental
Protection Act 1994 to notify of actual or
potential environmental harm.
All ESCs are to be inspected weekly and
within 24 hours of a rain event or as often
as specified in the contract specifications.
All defects are to be rectified as soon as
practicable and within 7 days of detection.
A generic ESC Site Inspection Checklist is
contained in Appendix 13C. This can be
adapted to site specific project conditions
and the project ESCP.
The contractor shall record all nonconformances on the Site Inspection
Checklist. Where a non-conformance is
repeated, the contractor shall advise the
superintendent and provide a copy of the
relevant non-conformance reports.
All monitoring details and site inspection
checklists shall be reported in the Monthly
Environment Report.
This shall be
submitted to the superintendent no later
than 7 days after the end of the month.
The Contractor is also required to establish
a process of auditing the performance of the
ESCP. Such audits must be undertaken by
suitably
qualified
and
experienced
personnel. Details of the audit process (i.e.
audit frequency, audit criteria, and
Auditors details) should be documented in
the EMP.

March 2010

13-36

The ESCP should be reviewed monthly or


in light of non-conformances and
construction progress. Revised copies of
the ESCP shall be issued to the
superintendent for acceptance.

13.17 Revegetation
Vegetation is one of the most effective
means of controlling erosion and
subsequent sedimentation.
The link
between a Revegetation Plan and an ESCP
is therefore critical.
A Revegetation Plan should normally be
developed during the detailed design phase
by designers to address permanent
vegetative stabilisation of the site. ESCPs
however, should normally be developed
during the construction phase by the
contractor
and
should
reference
revegetation aspects.
Revegetation guidelines are provided in
Appendix 13D to assist in selecting
appropriate revegetation techniques for
ESC and to provide a higher level of
understanding to the ESCP author. For
further information and detail, refer to the
departments Road Landscape Manual.
Figure 13.17 illustrates a minor failure and
sediment loss from a newly vegetated slope.

Figure 13.17 - Minor Slope Failure and


Sediment Loss

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13A
Preparing Erosion & Sediment Control Plan

13
A

Appendix 13A
Preparing an
Erosion & Sediment
Control Plan

March 2010

Department of Transport and Main Roads


Road Drainage Manual

13
A

Appendix 13A
Preparing Erosion & Sediment Control Plan

Appendix 13A Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

March 2010

ii

Reference
Section
-

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by
Steering
Committee

Date

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13A
Preparing Erosion & Sediment Control Plan

Appendix 13A
Preparing an Erosion &
Sediment Control Plan
Region/District:
Project Name:
Contract / Project Number:

Timing and Acceptance


Is the ESCP a HOLDPOINT where natural surface disturbance shall not be undertaken until
the ESCP is accepted by the Superintendent?
Yes / No
Date of ESCP submission is

to superintendent.

ESCP Report
Has an ESCP Report been prepared for he project

Yes / No

If No, document reason.


___________________________________________________________________
___________________________________________________________________
If yes, check sections from list below.
Introduction

Background
Performance Objectives
Design, Installation, Maintenance
and Removal of Control Measures
ESCP Methodology
____________________________
____________________________

Selection of Control Measures


Project Description

Legal and Other Requirements

Monitoring, Auditing and Reporting


____________________________
____________________________

Site Characteristics

Catchment Description
Soils
Hydrology
Vegetation
Document Control Status
Supporting Data/Information (aerial
photos, construction program, EMP
requirements, revegetation
specification)

Appendices

ESCP Drawings
Fact Sheets for control measures
Calculations
Water Quality
Revision No.

_______________________________
_______________________________

March 2010

13A-1

13
A

Department of Transport and Main Roads


Road Drainage Manual

13
A

Appendix 13A
Preparing Erosion & Sediment Control Plan

ESCP Drawings
Check that the following has been included on the ESCP Drawings:

Controls and drainage path

Location of disturbance and non-disturbance zones

Location of stockpiles, borrow pit, storage areas

Legend (including Standard Symbols);

Location of control measures

Construction / maintenance / removal notes for ESCs

North Point

Scale (preferred scale between 1:500 to 1:2,500)

Title Block

Project Name

ESCP Title

Plan No.

Revision No.

Date

Chainage

Legend (including Standard Symbols);

_________________________________

_________________________________

_________________________________

Step 1 Review Site Characteristics


If applicable, check that the following items are marked on an overlay of the construction
drawings:
Check

Source

Existing & proposed permanent


drainage structures

Construction Drawings

Limits of site

Contract

Contours

Contract, REF or Part B

Catchment Boundaries

REF

Watercourses

Topographic Map/Survey/Site Inspection

Drainage Lines

As above

Waterbodies

As above, REF

Existing Vegetation

REF

Soil Types and Properties

REF & Part B

Special Conditions

Part C

Erosion Risk Zones

Part B

March 2010

13A-2

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13A
Preparing Erosion & Sediment Control Plan

Step 2 Identify Areas of Construction Disturbance


If applicable, check that the following items are marked on an overlay of the
construction drawings:

Cut and Fill Locations

Construction Stages

Location of site access points, parking areas, site compounds, storage areas, and
stockpile areas.

Step 3 Selection of Controls


Check that the following has been identifies:

Constraints and Opportunities

Priorities

Check the ESCs most appropriate to this project:


Drainage Controls

Velocity Controls

Erosion Controls

Sediment Controls

Diversion Banks

Rock Check Dams

Buffer Zones

Diversion
Channels / Drains

Recessed Rock
Check Dams

Chemical Surface
Stabilisers
Erosion Control
Blankets

Chutes

Gravel / Sand Bag


Check Dam

Grasses Filter
Strips

Mulching

Construction
Exists

Drop Pipes

Erosion Control
Mats

Sediment Fences

Geotextille Lined
Channels

Outlet Protection

Straw Bales

Rock Lined
Channels

Temporary
Revegetation

Rock Filter Dams

Temporary
Watercourse
Crossing

Surface
Roughening

Sediment basins

Sediment Weirs
Portable Sediment
Tanks
Drop Inlet/Pipe
Inlet Protection

Check that the following has been considered in the selection of ESCs:

Soil Type and Erosion Risk

Flow Velocity

Sensitivity of the receiving environment

Duration of construction and life of ESC measure

Access and Maintenance

March 2010

13A-3

13
A

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13A
Preparing Erosion & Sediment Control Plan

Have ESCs been designed in accordance with the ESC Fact Sheets?

13
A

Yes / No

If No, document reason.


___________________________________________________________________________
___________________________________________________________________________

If yes, check where the relevant calculations and designs are documented:

Appendix of ESCP Report

Environmental File

________________________________________

Check that ESC maintenance requirements have been documented:

in the ESCP Report

on the ESCP Drawings

________________________________________

Check that ESC removal requirements have been documented:

in the ESCP Report

on the ESCP Drawings

________________________________________

Check the ESC monitoring, auditing and reporting requirements have been
documented:

in the ESCP Report

on the ESCP Drawings

________________________________________

March 2010

13A-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

13
B

Appendix 13B
Fact Sheets for
ESC Measures

March 2010

Department of Transport and Main Roads


Road Drainage Manual

13
B

Appendix 13B
Fact Sheets for ESC Measures

Appendix 13B Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

March 2010

ii

Reference
Section
-

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

Appendix 13B
ESC Fact Sheets Index
Techniques commonly used to prevent
controllable erosion and to minimise the
adverse effects of sedimentation are
described in the following fact sheets. Most
control techniques presented in the fact
sheets are to be implemented during the
construction phase only, and are thus
temporary measures. However, some
permanent measures can be integrated into
the operational layout during construction.
Examples
include
buffer
zones,
revegetation, and lined channels.

Velocity Controls
Velocity Control 1 Check Dam

Erosion Controls
Erosion Control 1 Chemical Surface
Stabiliser
Erosion Control 2 Erosion Control
Blanket
Erosion Control 3 Mulching
Erosion Control 4 Temporary
Revegetation

Lining Controls

Erosion Control 5 Surface Roughening

Lining 1 Grass and turf


Lining 2 Reinforced Grass
Lining 3 Geotextile Lining
Lining 4 Erosion Control Mat
Lining 5 Rock Lining
Lining 6 Rock Mattress
Lining 7 Cellular Confinement
Lining 8 Permanent Impervious Lining
Lining 9 Temporary Impervious Lining

Drainage Controls

Sediment Control
Sediment Control 1- Buffer Zone
Sediment Control 2 Grassed Filter Strip
Sediment Control 3 Construction Exit
Sediment Control 4 Sediment Fence
Sediment Control 5 Rock Sediment Trap
Sediment Control 6 Sediment Basin
Sediment Control 7 Sediment Weir
Sediment Control 8 Portable Sediment
Tank

Drainage Control 1 Diversion Channel /


Catch Drain
Drainage Control 2 Diversion Bank
Drainage Control 3 Chute
Drainage Control 4- Drop Pipe
Drainage Control 5 Temporary
Watercourse Crossing

March 2010

13B-1

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-2

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 1

Description
Grass or turf refers to a layer of topsoil and
grass/turf harvested from the field by
specialist machinery.
Application and Function
Used in both sheet flow and concentrated
flow to provide erosion protection
Remove sediments and nutrients in periods
of low flow, thus improving the site overall
discharge water quality
Limitations
Deposited sediments can kill grass or turf.
Rill erosion along the edges of the grass or
turf may occur.

Alternatives
Impermeable sheeting: e.g. plastic sheeting.
Advantages
Quick Installation
Provide instant erosion protection
Disadvantages
Can cause sediment problems and flow
concentration if overtopped during a heavy
storm
Special Requirements
Minor erosion and saturation problems can
occur if subjected to a constant trickle flow.
In permanent drainage channels, a concrete
invert or underground low-flow pipe may be
required to alleviate mosquito and drainage
problems
Overview
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the engineer
or responsible on-site personnel for
assistance.

Appendix 13B
Fact Sheets for ESC Measures

Grass and Turf

13
B

Improve infiltration of stormwater, reduce


runoff volumes, and allow downstream
sediment basins to drain more effectively
between storm events.
Often used as softer alternative to hard
channel linings in urban situations.
Early failure of a channel resulting from grass
or turf being placed directly over a dispersible
soil.
Impermeable fabric should be considered on
highly dispersive soils
Cellular grid: e.g. turf reinforcing
Water quality improvements
Easily available product
Can restrict movement of equipment around
the site and access to stockpiles
Invert erosion caused by persistent trickle flow
that can saturate the soil and cause damage
to the root system and/or mowing problems
Displacement of unpegged turf strips (sod)
caused by high velocity or deep water flows
within the first few weeks after placement
Grass and turf lining can only withstand flow
velocities of 1.5 to 2 m/s.

March 2010

13B-3

Department of Transport and Main Roads


Road Drainage Manual

13
B

Construction
Prepare
a
smooth
seed
bed
of
approximately 75 mm of topsoil - clear away
trash and large stones, and grade smoothly
to eliminate footprints, tracks and ruts.
As deposited sediment can kill grass and
turf, upstream erosion protection and
sediment detention measures must be
installed before the grass or turf can be
installed.
Seed, fertilise, water and rake to remove
any remaining surface irregularities.
Maintenance
Requires mowing in urban environments.
Grass/turf need to be watered until
adequately established.
Removal
If the lining is temporary, it must be replaced
with permanent stabilisation measures as
specified in the approved plan.
Dispose of the lining properly.

March 2010

13B-4

Appendix 13B
Fact Sheets for ESC Measures

Preferably start at the downstream end and


align the grass or turf strips.
The edges of the grass or turf strip must be
installed flush with the existing soil surface to
avoid erosion along the interface between
grass or turf and soil surface. Where more
than one strip is used - overlap the ends by at
least 300 mm, with the upstream strip placed
over the downstream sheet and stapled.

All surface-laid linings should be inspected on


a regular basis.
Temporary stabilisation works must be
maintained until arrangements have been
made to install the permanent lining.

Department of Transport and Main Roads


Road Drainage Manual

Lining 2

Description
Reinforced grass refers to a layer grass/turf
grown on an artificial two-dimensional grid
to provide additional strength
Application and Function
Used in concentrated flow to provide
erosion protection.
Remove sediments and nutrients in periods
of low flow, thus improving the site overall
discharge water quality.
Limitations
Not suitable for in areas prone to grass
fires.
Not suitable in areas of high turbulence
Alternatives
Impermeable
sheeting:
e.g.
plastic
sheeting.
Advantages
Provide immediate and long-term erosion
protection.
Disadvantages
Some maintenance problems can occur in
permanent channels resulting from damage
induced by mower or slasher.
Deposited sediments can kill grass or turf.
Special Requirements
Particular attention should be given to the
crest, toe and sides to avoid erosion and
up-lifting.
Overview
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.

Appendix 13B
Fact Sheets for ESC Measures

Reinforced Grass

13
B

Improve infiltration of stormwater, reduce runoff


volumes, and allow downstream sediment
basins to drain more effectively between storm
events.
Often used as a softer alternative to hard
channel linings in urban situations.
Reinforced grass is generally not suitable in
areas frequented by grazing animals.

Water quality improvements.


If the channel is ever widened, reshaped or
relocated in the future, then it is likely that the
topsoil containing the permanent reinforcing
will need to be disposed of and replaced.
Damage by grass fires.
Should not be placed directly over a dispersible
soil. A minimum 100 mm layer of non
dispersible topsoil should be placed over the
in-situ dispersible subsoil prior to placement
Reinforced grass can withstand flow velocities
of up to 4 m/s.

March 2010

13B-5

Department of Transport and Main Roads


Road Drainage Manual

13
B

Construction
Prepare a smooth seed bed of
approximately 75 mm of topsoil - clear
away trash and large stones, and grade
smoothly to eliminate footprints, tracks and
ruts.
As deposited sediment can kill grass and
turf, upstream erosion protection and
sediment detention measures must be
installed before the grass or turf can be
installed.
Seed, fertilise, water and rake to remove
any remaining surface irregularities.
Maintenance
Requires mowing in urban environments.
Grass/turf need to be watered until
adequately established.
Removal
Temporary stabilisation works must be
maintained until arrangements have been
made to install the permanent lining.

March 2010

13B-6

Appendix 13B
Fact Sheets for ESC Measures

Preferably start at the downstream end and


align the grass or turf strips.
The edges of the grass or turf strip must be
installed flush with the existing soil surface to
avoid erosion along the interface between
grass or turf and soil surface. Where more than
one strip is used - overlap the ends by at least
300 mm, with the upstream strip placed over
the downstream sheet and stapled.

All surface-laid linings should be inspected on


a regular basis.

Department of Transport and Main Roads


Road Drainage Manual

Lining 3

Appendix 13B
Fact Sheets for ESC Measures

Geotextile Lining

13
B

No picture available
Description
Geotextile mats can be either surface-laid
or buried.
Surface-laid geotextile include impervious
sheets used to cover bare soil and pervious
mats used to enhance the role of
vegetation. Geotextile mats also act as a
scour prevention / suppressant.
Application and Function
Some geotextile linings provide temporary
protection to earth drains that are intended
to be removed or upgraded within 6 months
and to newly formed grassed channels
established with seed or runners.
Limitations
Geotextile linings are best used with
vegetation.
Non-biodegradable products have limited
use in fauna inhabited bushland areas.
Alternatives
Impermeable
sheeting:
e.g.
plastic
sheeting.
Permeable fabric: e.g. filter cloth.
Permeable netting: e.g. erosion control
mats.
Impermeable netting: e.g. bitumen coated
nets.
Advantages
Quick installation.
Wide variety of products and uses.
Most products provide instant erosion
protection.
Disadvantages
Some fabrics have a very limited working
life.
Environmental problems associated with
nonbiodegradable fabric used in bushland
areas.
Special Requirements
Four general requirements exist for
effective protection against erosion: good
contact must be achieved; seepage flow
under the channel liner should be
discouraged;
surface
irregularities
removed; and good anchorage must be
provided.
Particular attention should be given to the
crest, toe and sides to avoid erosion and
up-lifting.

Buried geotextile can be used to improve


the long term performance of vegetation
and to provide root reinforcement.

In high velocity areas, buried geotextile


mats can be used to reinforce turfed
channels.

Biodegradable mats generally have a lower


allowable velocity limit.
Impermeable fabric should be considered
on highly dispersive soils.
Permeable mats: e.g. buried root
reinforcing.
Cellular grid: e.g. turf reinforcing.
Pocket fabrics: e.g. grout-filled mattresses.

Can be used for emergency repairs.


Products are available for short and long
term use.
Maintenance problems with some buried
mats.
Bitumen-based products can release
phosphorus to receiving waters.
Geotextiles should not be placed directly
over a dispersible soil. A minimum 100 mm
layer of non dispersible topsoil should be
placed over the in-situ dispersible subsoil
prior to placement
The outer edges of the geotextile needs to
be buried and stapled into a 200 mm deep
trench. The trench should then be
backfilled to allow the free flow of water
laterally into the channel.

March 2010

13B-7

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Unless otherwise specified, geotextiles
shall be jointed with either; a sewn 100 mm
(min.) face to face or lapped seam, or
stapled (75 mm min. spacing) lapped
seam, or overlapped 300 mm (min.).
Construction
The method of installation varies with the
type of geotextile. Installation procedures
may be provided by the manufacturer or
distributor. A typical installation procedure
is described below, but should be
confirmed with the product manufacturer or
distributor.
Prepare a smooth bed of approximately 75
mm of topsoil - clear away trash and large
stones, and grade smoothly to eliminate
footprints, tracks and ruts.
Preferably start at the downstream end and
align the geotextile strips, or unroll the
geotextile, upstream against the dominant
direction of flow. In situations where the
bank is subject to both concentrated inbank
(stream) flow and transverse overland
sheet flow, align fabric with the dominant
flow - usually the concentrated stream flow.
Bury the downstream end of the fabric in a
300 mm trench. When spreading the fabric,
avoid stretching the textile mesh - the fabric
should be in good contact with the ground
at all points.
Staple the fabric within the trench at 200 to
250 mm spacing using 100 mm (wide) x
150 mm (penetration length) U shaped, 8
- 11 gauge wire staples. Narrower U
sections may easily tear the matting when
placed under stress.
Maintenance
Inspect the geotextile to see if construction
activity, or falling rocks or trees have
damaged it, or if runoff is undermining the
fabric.
All surface-laid fabrics should be inspected
on a regular basis.
Removal
If the fabric is temporary, it must be
replaced with permanent stabilisation
measures as specified in the approved
plan.
Dispose of the fabric properly.

March 2010

13B-8

Appendix 13B
Fact Sheets for ESC Measures

Until it is placed in the channel, geotextiles


shall be stored away from direct sunlight or
covered with ultraviolet light protective
sheeting.
Vehicles and construction equipment shall
not be permitted to manoeuvre over the
geotextile unless it has been covered with
a layer of soil or gravel at least 150 mm
thick. The fill material shall not be mixed
over the geotextile.
Damaged geotextile shall be repaired or
replaced.
Staple the exposed fabric surface at 1000
mm centres.
Where more than one sheet is used overlap the ends by at least 300 mm, with
the upstream sheet placed over the
downstream sheet and stapled. Where high
flow velocities are anticipated, the
upstream end of all sheets should be
buried in a 300 mm deep trench and
stapled, with the upstream sheet lapped
over the downstream sheet and stapled.
Overlap the sides by 100 mm (depends on
the type of geotextile seek appropriate
advice). If the bank is subject to both
concentrated inbank (stream) flow and
transverse overland sheet flow, the upper
bank sheet should overlap the lower sheet
and the upstream sheet should overlap the
downstream sheet.
In such cases, it may be preferable to start
laying the sheets at the downstream invert,
then moving first up the bank, then up the
drainage line.
Anchor the outer edge and top of the
protected channel in a 300 mm deep trench
and staple at 200 to 250 mm centres.
Additional staples will be required wherever
the mat does not contact the ground
surface.
If damaged, repair or replace the damaged
section. If water is undermining the fabric,
repair any holes or joints or rebury the
upper ends of the damaged sections.

Temporary stabilisation works must be


maintained until arrangements have been
made to install the permanent lining.

Department of Transport and Main Roads


Road Drainage Manual

Lining 4

Description
Control mats are made of durable materials
such as jute, coconut fibre, nylon and polypropylene.
Control mats can be either surface-laid or
buried.
Application and Function
Used in concentrated flow environments to
protect soil from raindrop impacts and
erosion, and to control seed germination,
weed growth, soil temperature fluctuations
and soil moisture loss.
Some mats provide temporary protection to
earth drains that are intended to be
removed or upgraded within 6 months and
to newly formed grassed channels
established with seed or runners.
Limitations
Erosion control mats are best used with
vegetation.
Non-biodegradable products have limited
use in fauna inhabited bushland areas.
Synthetic products should be used with
extreme caution in areas used by grazing
animals.
Alternatives
Impermeable
sheeting:
e.g.
plastic
sheeting.
Impermeable netting: e.g. bitumen coated
nets.
Advantages
Quick installation.
Wide variety of products and uses.
Most products provide instant erosion
protection.
Disadvantages
Some fabrics have a very limited working
life.
Environmental problems associated with
nonbiodegradable fabric used in bushland
areas.
Erosion control mats overlapped in the
wrong direction may be lifted by flowing
water.

Appendix 13B
Fact Sheets for ESC Measures

Erosion Control Mat

13
B

Surface-laid erosion control mats include


impervious sheets used to cover bare soil
and pervious mats used to enhance the
role of vegetation
Erosion control mats also act as a scour
prevention and suppression device.
In high velocity areas, buried erosion
control mats can be used to reinforce turfed
channels. Buried erosion control mats can
be used to improve the long term
performance of vegetation and to provide
root reinforcement

Biodegradable mats generally have limited


shear strength and their resistance to
concentrated flow deteriorates rapidly over
a period of a few months.
Impermeable fabric should be considered
on highly dispersive soils.
Cellular grid: e.g. turf reinforcing.
Rock lining.

Can be used for emergency repairs.


Products are available for short and long
term use.
Maintenance problems with some buried
mats.
Tenting may occur when blankets are
placed over irregular surfaces, i.e. as a
result of poor surface preparation or
blankets being used in inappropriate
environments.
March 2010

13B-9

Department of Transport and Main Roads


Road Drainage Manual

13
B

Special Requirements
Four general requirements exist for
effective protection against erosion: good
contact must be achieved; seepage flow
under the channel liner should be
discouraged;
surface
irregularities
removed; and good anchorage must be
provided.
Particular attention should be given to the
crest, toe and sides to avoid erosion and
up-lifting.
Overview
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Unless otherwise specified, control mats
shall be jointed with either; a sewn 100 mm
(min.) face to face or lapped seam, or
stapled (75 mm min. spacing) lapped
seam, or overlapped 300 mm (min.).
Construction
The method of installation varies with the
type of mat. Installation procedures may be
provided by the manufacturer or distributor.
A typical installation procedure is described
below, but should be confirmed with the
product manufacturer or distributor.
Prepare a smooth seed bed of
approximately 75 mm of topsoil - clear
away trash and large stones, and grade
smoothly to eliminate footprints, tracks and
ruts.
Seed, fertilise, water and rake to remove
any remaining surface irregularities.
Preferably start at the downstream end and
align the mats, or unroll the mats, upstream
against the dominant direction of flow. In
situations where the bank is subject to both
concentrated inbank (stream) flow and
transverse overland sheet flow, align fabric
with the dominant flow - usually the
concentrated stream flow.
Bury the downstream end of the fabric in a
300 mm trench. When spreading the fabric,
avoid stretching the textile mesh - the mats
should be in good contact with the ground
at all points.
Staple the mat within the trench at 200 to
250 mm spacing using 100 mm (wide) x
150 mm (penetration length) U shaped, 8
- 11 gauge wire staples. Narrower U
sections may easily tear the matting when
placed under stress.

March 2010

13B-10

Appendix 13B
Fact Sheets for ESC Measures

Erosion control mats should not be placed


directly over a dispersible soil. A minimum
100-mm layer of non dispersible topsoil
should be placed over the insitu dispersible
subsoil prior to placement
The outer edges of the mat needs to be
buried and stapled into a 200 mm deep
trench. The trench should then be
backfilled to allow the free flow of water
laterally into the channel.
Until it is placed in the channel, a control
mat shall be stored away from direct
sunlight or covered with ultraviolet light
protective sheeting.
Vehicles and construction equipment shall
not be permitted to manoeuvre over the
control mat unless it has been covered with
a layer of soil or gravel at least 150 mm
thick. The fill material shall not be mixed
over the mat.
Damaged mats shall be repaired or
replaced.
Staple the exposed fabric surface at 1000
mm centres.
Where more than one sheet is used overlap the ends by at least 300 mm, with
the upstream sheet placed over the
downstream sheet and stapled. Where high
flow velocities are anticipated, the
upstream end of all sheets should be
buried in a 300 mm deep trench and
stapled, with the upstream sheet lapped
over the downstream sheet and stapled.
Overlap the sides by 100 mm (depends on
the type of mat seek appropriate advice).
If the bank is subject to both concentrated
inbank (stream) flow and transverse
overland sheet flow, the upper bank sheet
should overlap the lower sheet and the
upstream sheet should overlap the
downstream sheet.
In such cases, it may be preferable to start
laying the sheets at the downstream invert,
then moving first up the bank, then up the
drainage line.
Anchor the outer edge and top of the
protected channel in a 300 mm deep trench
and staple at 200 to 250 mm centres.
Additional staples will be required wherever
the mat does not contact the ground
surface.
Mats, once fixed, may be rolled with a roller
weighing 60 to 90 kg/m, then watered if
grass seeding has occurred.

Department of Transport and Main Roads


Road Drainage Manual

Maintenance
Biodegradable mats should be inspected
after the first few runoff producing storms.
Inspect the mat to see if construction
activity, or falling rocks or trees have
damaged it, or if runoff is undermining the
fabric.
Removal
If the fabric is temporary, it must be
replaced with permanent stabilisation
measures as specified in the approved
plan.
Dispose of the fabric properly.

Appendix 13B
Fact Sheets for ESC Measures

All surface-laid fabrics should be inspected


on a regular basis.
If damaged, repair or replace the damaged
section. If water is undermining the fabric,
repair any holes or joints or rebury the
upper ends of the damaged sections.
Temporary stabilisation works must be
maintained until arrangements have been
made to install the permanent lining.

March 2010

13B-11

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-12

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 5

Description
Sized and graded rock is placed along the
bed and banks of a diversion channel, drop
chute, channel bend or spillway.
Application and Function
Rock lining of drains and channels is one of
the simplest kinds of surface treatment.
It is particularly useful in critical sections of
a channel such as bends and stormwater
outlets.
Limitations
There is often only a limited range of rock
sizes available, or the calculated rock size
may be too large in proportion to the drain
size.
Rock may not be readily available in certain
areas.
Alternatives
One of the most common and inexpensive
channel lining materials.
Advantages
One of the most common and inexpensive
channel lining materials.
The porous nature of rock protects the
channel from uplift and allows for
revegetation.
Disadvantages
Problems with unsightly weeds.
Undersized rocks can migrate downstream
and cause further erosion as they move
along the creek bed during periods of flood.

Special Requirements
An underlying geotextile or rock layer is
generally required unless the thickness of
the rock layer is at least three times the
nominal rock diameter, however, if the
voids between the rocks are filled with soil
and planted, then a filter layer is not
required.

Appendix 13B
Fact Sheets for ESC Measures

Rock Lining

13
B

Rock is generally placed over a geotextile


or rock filter layer, however, if the voids
between the rocks are filled with soil and
planted, then a filter layer is not required.
Used on channel slopes steeper than 5%
and in heavy traffic (vehicular or human)
areas.

Cannot be placed directly over dispersible


soils.
Success often depends on the introduction
of suitable vegetation to anchor the rocks.

The porous nature of rock protects the


channel from uplift and allows for
revegetation.
Does not require a well formed channel
cross section.

Often difficult to de-silt rock lined channels.


Some rock types may cause pH problems
in low flows.
Structural failure of rock chutes and steep
rock-lined channels is common in the first
few years.
Particular attention must be given to the
edges of the rock protection to avoid
undermining and erosion to adjacent bank
areas.
Rock lining should blend with surrounding
land (i.e. boxed into the channel) to allow
adjacent runoff to freely enter the channel.
March 2010

13B-13

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Geometry:
o
o

Side slopes should not exceed 2(H):1(V).


The channel should be over-excavated such
that the top surface of the rock is level with
the surrounding land.
Minimum thickness should be twice the
nominal rock size (or 1.5 times the
maximum rock size) if a filter layer (filter
cloth or graded rock) is used. Otherwise,
the thickness of the armour rock layer
should be 3 times the nominal rock size.

Rock type:
o
o

Crushed rock is generally more stable than


rounded stone.
The rock should be durable and resistant to
weathering, and should be proportioned so
that neither the breadth nor the thickness of
a single rock is less than one-third its
length.

Construction
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel.
Clear the proposed channel area of trees,
stumps, roots, loose rock etc.
Subgrade preparation:
o

Excavate the cross section to the lines and


grades of the foundation of the liner as
shown on the plans. Bring over-excavated
areas to grade by increasing the thickness
of the liner or by backfilling with moist soil
compacted to the density of the surrounding
material.
Cut the subgrade sufficiently deep that the
finished grade of the rock will be at the
elevation of the surrounding area and the
finished inside dimensions and grade of the
channel meet the design requirements.
Sand and gravel filter blankets:
Place the specified filter blanket or graded
filter layer immediately after the foundation
is prepared. Spread the filter in a uniform
layer to the specified depth (minimum 150
mm). Where more than one layer of filter
material is used, spread the layers with
minimal mixing between layers.

Synthetic filter fabric underlay:


o

Place the filter fabric directly on the prepared


foundation. Overlap the edges by at least
300 mm, and place anchor pins at one (1)
metre (min.) spacing along the overlap.
Bury the upstream end of the cloth 300 mm
(min.) below ground and where necessary,
bury the lower end of the fabric or overlap
300 mm (min.) over the next downstream
section.

March 2010

13B-14

Appendix 13B
Fact Sheets for ESC Measures

Grouted rock may be used. The grout must


be placed such that it penetrates the rock
voids. Weep holes and proper filter layers
must be provided.

Rock sizing:
o
o

o
o
o

Minimum nominal size is 200 mm.


Rock size is either determined in relation to
the channel velocity, or more accurately to
the boundary shear stress.
Maximum rock size should not exceed 1.5
times the nominal rock size.
Rock size on the outside of bends should be
increased.
Rocks should be evenly-graded.

Rock underlay and filter layer:


o

Rock should be placed over a geotextile


filter cloth or suitably graded, rock-gravel
filter layer a minimum 150 mm thick.

Take care not to damage the cloth during or


after placement. If damage occurs, remove
the rock and repair the sheet by adding
another layer of filter material with a
minimum overlap of 300 mm around the
damaged area. If extensive damage is
suspected, remove and replace the entire
sheet.
Where large rock is used, or machine
placement is difficult, a 100 mm layer of fine
gravel or sand may be needed to protect
the fabric.

Rock placement:
o

Placement of rock should occur immediately


after placement of the filter. Place rock so
that it forms a dense, well-graded mass of
rock with a minimum of voids. The desired
distribution of rock throughout the mass
may be obtained by selective loading at the
quarry and controlled dumping during final
placement.
Place rock to its full thickness in one
operation. Do not place rock by dumping
through chutes or other methods that cause
segregation of rock sizes. Take care not to
dislodge the underlying base or filter when
placing the rocks.
The finished slope should be free of pockets
of small rock or clusters of large rocks.
Hand placing may be necessary to achieve
the proper distribution of rock sizes to
produce a relatively smooth, uniform
surface. The finished grade of the rock
should blend with the surrounding area. No
overfall or protrusion of rock should be
apparent.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

Perform all channel construction to keep


erosion and water pollution to a minimum.
Immediately upon completion of the
channel, vegetate all disturbed areas or
otherwise protect them against soil erosion.
Topsoil and vegetate the rock as specified
in the approved plan. Where channel
construction will take longer than 30 days,
stabilise channels by reaches.
Maintenance
Desirable vegetation should generally be
promoted, especially in high flow velocity
areas.
Rock-lined channels should be inspected
periodically and after major rains for scour
or dislodged rock. Repair damaged areas
immediately. Special attention should be
given to the outlet and inlet sections and
other points of flow entry.
Removal
Temporary structures may only be
removed when an alternative, stable
drainage path or surface is available.

Carefully check stability of the rock lining


looking for indications of piping, scour
holes, bank failures, or rill erosion along the
edges of the rock fill. In general, once a
rock installation has been properly
designed and installed it requires little
maintenance other than the removal or
slashing of weeds.
Stabilise the area as specified on the
approved plans.
Grade the area and smooth it out in
preparation for stabilisation.

March 2010

13B-15

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-16

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 6

Appendix 13B
Fact Sheets for ESC Measures

Rock Mattress

13
B

No picture available
Description
Rock mattresses consist of multi-celled,
rectangular,
PVC
coated,
heavily
galvanised wire cages filled with wellgraded rock. PVC coated cages should be
used in all watercourse situations.
Application and Function
Used for channel linings, spillway
protection and energy dissipation areas
downstream of high-flow check dams.
Limitations
If poorly constructed, they can be
expensive to rectify.
Water-transported sediment (sand and
gravel) can reduce the service life of the
wire.
Often ineffective on dispersive soils.

Alternatives
Rock mattresses are available in
thicknesses of 0.17, 0.23, 0.3 and 0.5
metres, at a length of 6 metres and width of
2 metres.
Advantages
Rock mattresses are very useful for small
localised
drainage
problems
where
immediate lining and/or protection is
required.
Disadvantages
Unsightly weed infestation can be a
maintenance problem in urban areas.
Sediment is difficult to remove from the
wire cages without damage to the wire,
especially if the location or shape of the
cage structure is difficult to distinguish
under the sediment load.
Special Requirements
On delivery, wire cages should be labelled
and stockpiled separately according to
cage size and thickness.

Can be used to line open channel chutes.

The turbulent transportation of bed load


(gravel and debris) can break the wire
through a hammering action. This can
shorten the service life of the mattresses
considerably.
Failure may result if runoff is not controlled
and scour occurs behind the gabion (i.e. if
the gabion is installed in a dynamic
environment).

Are a well-proven protection measure.

Some rock mattress and gabion structures


have experienced durability problems
caused by trickle flows, high sediment
loads and debris.
Often difficult to obtain efficient channel
depth.
Labour intensive construction requiring
good supervision.

March 2010

13B-17

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Mattress details:
o

o
o

Mattresses are available in thicknesses of


170, 230, 300 and 500 mm, at a length of 6
metres and width of 2 metres.
Heavily galvanised, PVC coated cages
should be used in watercourses.
Mattresses should be laid over a filter fabric
or properly designed gravel filter.

Appendix 13B
Fact Sheets for ESC Measures

General design notes:


o

Rock-fill details:
o
o
o
o

Rock-fill should be angular and blockshaped.


Minimum rock size around a 1/3 the basket
depth.
Maximum rock size around 2/3 the basket
depth.
The rock should be uniformly graded with
80% by number greater than 100 mm in
size.
Nominal rock size should be 200 to 300 mm
when used within the splash zone of weirs
and drop structures.

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel.
Mattresses of different thicknesses should
be stored on-site in separate piles and
clearly labelled.
Flatten out each mattress on a hard, flat
surface, and stamp out any unnecessary
creases. Edge creases will need to be
stamped into the bottom of the 2nd and 4th
internal diaphragms.
Ensure that each diaphragm is vertical and
the correct height. Fold the sides and ends
of the mattress to meet the top of the
diaphragms. Fold the side panel flaps to lie
adjacent to the diaphragms. Tack
temporarily either by using short lengths of
binding wire, or alternatively by twisting the
top diaphragm wire over the flap selvedge
wire.
The ends of the diaphragms must now be
permanently laced to the sides of the
mattress. At the four corners, bend the
projected lengths of the end panels to
overlap the sides, and lace up with binding
wire.
When the mattress is placed over a
geotextile, care must be taken to ensure
that projecting ends of wire are bent
upwards to avoid puncturing or tearing the
cloth. Geotextile should be placed
according to specifications.

March 2010

13B-18

Mattresses can be anchored to adjoining


fixed structures with the use of I-bolts set
into the wall, and galvanised rods threaded
through the I-bolts after installing of the
mattress cages.
When used as the lining of embankment
chutes, filter cloth and a subsoil drain
should be laid under the mattress. Two
layers of 300 mm (min.) thick mattresses
when used within the splash zone of weirs
and drop structures.
Channel bed and stepped mattresses should
be aligned such that the 2 m wide cells are
transverse
to
the
flow
direction.
Embankment mattresses should be aligned
with the cells parallel to the embankment
contour.

On slopes, the mattress should generally


be laid with the diaphragm across the slope
rather than up and down the slope. On
stream beds, the mattress should generally
be laid with the diaphragm at right angles
to the main direction of water flow.
Place the fill material, by hand or
mechanically, in the compartments, starting
at the bottom if on a slope. The fill should
be a hard, durable stone, in size between
80 mm and two-thirds the thickness of the
mattress, but generally no greater than 200
mm.
Filling can be done unit by unit, but several
units should be ready for filling at any one
time.
For units with PVC coated wire mesh,
particular care shall be taken to ensure that
sharp edges of quarry stone are not placed
against the mesh in order to avoid causing
unnecessary abrasion.
Slightly overfill each mattress to allow for
settlement. Tack the lid to the corners of
the mattress, and then securely wire it to
the tops of the sides, ends and
diaphragms, using alternate single and
double loops as specified above.
With more than one mattress filled, the
edges of adjacent lids can be wired down
in the same operation, saving both time
and binding wire.
When the mattress is laid on a slope
steeper than 1.5(H):1(V), it should be
secured by star pickets or hardwood pegs
driven into the ground just inside the upper
end panel at 2 m centres or as necessary.

Department of Transport and Main Roads


Road Drainage Manual

Carry the wired-up mattress to its final


position, and wire it securely to the
adjacent mattresses. Mattresses should be
placed and wired together empty as it is
difficult to wire mattresses together when
both are full of stone. (Continued overleaf)
All hand wiring must be done as a
continuous lacing operation. Begin wiring
by securing the binding wire to the corner
of the panels to be joined by looping it
through and twisting it together. Then lace
with single loops and double loops in turn
at 100 mm intervals. Finally poke the loose
end inside the mattress. Tight mesh and
wiring is essential at all times.
Maintenance
Mattresses should be inspected at least
annually for damage and if damaged
repaired as soon as possible.
Extreme care should be taken when
slashing or mowing vegetation to avoid
damage. All damage should be repaired
immediately.
Removal
None

Appendix 13B
Fact Sheets for ESC Measures

On soft or sandy slopes, pegs may be used


to hold the mattress in position during
filling.
Mattresses maybe shortened where
necessary, by cutting along the fold at the
top of a diaphragm and removing the
bottom spiral connections. (Continued
overleaf)
Always
consult
manufacturers
specifications and assembly instructions
before modifying the shape of the mattress
or wiring deformed mattress shapes.

The long-term success of mattress and


gabion lined channels usually depends on
the successful integration of desirable
vegetation and the minimisation of instream debris and bed load (sand, gravel
and rock).

March 2010

13B-19

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-20

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 7

Description
Cellular confinement products are threedimensional honeycomb HDPE mesh
developed originally for the Gulf War to
allow heavy trucks to traverse sand dunes.
Application and Function
The three-dimensional mesh confines infill
material to resist flows, minimise erosion,
and prevent downward migration of
embankment materials.
May be used to contain sand, mud and
other unbound soil material for temporary
roads or creeks crossings.
Limitations
Product is expensive compared to alternate
techniques.
The mesh can lift in concentrated flow due
to the memory action of HDPE and
insufficient pegging.
Alternatives
All other channel linings.
Advantages
Easy to transport to site.
Disadvantages

Appendix 13B
Fact Sheets for ESC Measures

Cellular Confinement

13
B

May be used as an alternative to traditional


rock armouring in areas that have limited
sources of suitable rock.

Can promote water-logging of the ground


unless adequate drainage exists either
below the matting or within the sidewalls of
the material.

Special Requirements
Overview
Refer to manufacturers guidelines.
Construction
Refer to manufacturers guidelines.
Maintenance
Refer to manufacturers guidelines.
Removal
Refer to manufacturers guidelines.

March 2010

13B-21

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-22

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 8

Description
Open channel lined with concrete
(permanent
structure)
for
erosion
protection.
Application and Function
Permanent linings are suitable for use on
in-situ soils where protection from high
velocity flows is required.
Limitations
Concrete-lined channels may overtop
during high flows.
Access for concrete trucks can be difficult
in remote or steep locations.
Can be expensive.
Alternatives
Advantages
Designs usually have a lower risk of failure
compared to vegetated channel liners.
Disadvantages
If concrete lining sits proud of the soil
surface, erosion along the soil/concrete
interface may occur.
When used on non-insitu soils, differential
settlement can occur, causing the concrete
to crack.
Special Requirements
Overview
Channel cross-sectional area and capacity
should be designed to safely convey the
peak discharge from the design storm.

Appendix 13B
Fact Sheets for ESC Measures

Permanent Impervious Lining

13
B

Used when non-porous erosion protection


is required such as on dispersible soils.
Usually results in a near uniform channel
cross section which gives the appearance
of a drain rather than a stream or creek.
Usually less aesthetically pleasing than a
vegetated channel.

Design usually have close to 100%


strength immediately after placement.
Higher flow velocities may result than if
softer alternatives like grass linings are
used.

Some form of energy dissipator is required


at the outlet to the lined drain.
Cut-off trenches are required to prevent
water entry under the concrete.

Construction
Maintenance
Removal

March 2010

13B-23

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-24

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Lining 9

Description
Short channels or chutes lined with plastic
sheeting (temporary structure) for erosion
protection.
Application and Function
Temporary linings are suitable for use on
short channels or chutes.
Limitations
Can be expensive.

Appendix 13B
Fact Sheets for ESC Measures

Temporary Impervious Lining

13
B

Used when non-porous erosion protection


is required such as on dispersible soils.
Usually less aesthetically pleasing than a
vegetated channel.

Alternatives
Advantages
Design usually have close to 100%
strength immediately after placement.
Disadvantages
Problems can arise from water eroding
under the plastic.
Special Requirements
Overview
Channel cross-sectional area and capacity
should be designed to safely convey the
peak discharge from the design storm.

Higher flow velocities may result.

Some form of energy dissipator is required


at the outlet to the lined drain.
Cut-off trenches are required to prevent
water entry under the lining.

Construction
Maintenance
Removal

March 2010

13B-25

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-26

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Drainage Control 1

Appendix 13B
Fact Sheets for ESC Measures

Diversion Channel / Catch Drain

No picture available
Description
Channel or catch drains are usually
excavated with a grader blade and should
be at least 300 mm deep and 1000 mm
wide.
Application and Function
To limit the flow path length down steep
grades on unstable or newly grassed
areas.
To divert runoff around disturbed areas
To direct contaminated flow within
disturbed areas to sediment traps.
Limitations
Typical gradient is 0.5%, may be as low as
0.25% or as high as 0.75%.
Alternatives
Diversion banks
Perimeter banks
Backpush banks
Straw bale perimeter banks (if used for
longer than 1 day, bales must be wrapped
in geotextile)
Advantages
Quick to establish or re-establish if
disturbed.
Inexpensive to construct.
Disadvantages
Can cause sediment problems and flow
concentration if overtopped during a heavy
storm.
Special Requirements
The erodible nature of the subsoil should
be investigated before planning any
excavated drains.
Straw bales or other sediment traps should
not be placed within channels as they
usually cause the flow to be diverted from
the channel.
Overview
Minimum 300 mm deep and 1000 mm
wide.
Bank slopes should not be steeper than
2(H):1(V).
Unless side slopes are 6(H):1(V) or flatter,
U shaped or parabolic sections are
preferred where practicable.
Gradient:
o
o

Non-eroding flow velocity during a discharge


of 0.25 times the critical 1 in 1 year storm.
Typical gradient is 0.5%, however it may be
as low as 0.25% or as high as 0.75%.

Bank slopes should not be steeper than


2(H):1(V).

To divert runoff around stockpiles.


To function as a permanent drainage
channel located above a cut batter.

In areas of dispersive subsoils, perimeter


banks or backpush banks should be used
in preference to channels.

Usually do not require any formal design.

Can restrict movement of equipment


around the site and access to stockpiles.
Channels must have a stable outlet.
The outlet may require an energy dissipator
or level spreader. Level spreaders are used
to convert the concentrated flow back into
sheet flow before it is released down an
even slope.
Outlet:
o
o

Must be stabilised and not discharge to fill


slopes.
May consist of a drop pipe, chute, rock-lined
channel, turfed level spreader or stabilised
diversion channel.

Location:
o

A series of drains may be formed along the


contours to try and maintain sheet flow and
increase infiltration.

March 2010

13B-27

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

Appendix 13B
Fact Sheets for ESC Measures

The deeper the flow depth, the lower the


maximum allowable gradient.
Maximum gradient of around 0.4% for open
earth drains.
Erosion Protection:
o

To provide erosion protection, the channel or


drain should be covered with appropriate
lining (refer lining fact sheets)

Slope

Spacing

1%

90m

2%

60m

4%

40m

6%

32m

8%

28m

10%

25m

Bank above must not outlet onto the back


batter of next bank.
Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
When temporary channels or drains are
required and their locations are not shown
on the approved plans, determine their
location on the ground taking into
consideration the following:
o

Location of down-slope sediment trap. The


channel or catch drain should drain to a
sediment trap if diverted water is expected
to contain sediment. Clean water should be
diverted around these devices.

The channel or drain must have positive


drainage with a maximum grade of 0.4%
for open earth drains.
Allow sufficient space between the toe of a
fill slope and channel or drain for
maintenance.
Provide access for personnel and
equipment for installation.
Maintenance
Materials and equipment must be available
for maintenance at all times.
If sediment accumulates in the channel or
drain, remove it so that its capacity is not
reduced.
Do not dispose of the sediment in a
manner that will create an erosion or
siltation hazard.
Removal
When the construction work above the
channel or drain is finished and the area is
stabilised, the measure should be
removed.
Remove accumulated sediment.

March 2010

13B-28

The channel or drain should be covered


with appropriate erosion protection, unless
it will operate for less than 30 days or if
significant rainfall is not expected during
the life of the measure. Refer to the lining
fact sheets for selecting an appropriate
cover.
Do not clear above the drainage area until
the channel and down-slope sedimenttrapping devices are in place.
Grade the channel or drain. The sides of
the channel and the bank must be no
steeper than a 2:1 slope. The completed
flow diversion must be at least 300 mm
deep.
Check the bed slope to ensure positive
drainage in the desired direction.
Check that the channel or drain has a
stable outlet and does not discharge to an
unstable fill slope. Where a stable outlet
does not exist, a drop pipe or other suitable
drainage control structure may be required.
Remove roots, stumps, and other debris
and dispose of them properly.
The channel or drain must be inspected:
o

during construction: to check if machinery,


falling trees, etc. have damaged it; if
damaged, make repairs. In addition, record
and remove signs of blockage.
after each rainfall event: to ensure that
runoff is flowing to and not around
sediment-trapping devices, and modify the
arrangement if required.

Dispose of accumulated sediment properly.


Grade the area and smooth it out in
preparation for stabilisation by grassing or
as specified in the approved plan.

Department of Transport and Main Roads


Road Drainage Manual

Drainage Control 2

Description
An excavated drain and bank formed by
pushing soil down the slope.
A diversion bank constructed slightly off the
contour is often referred to as perimeter
bank.
Application and Function
Used above batters, borrow pits and
exposed surfaces to protect them from upslope stormwater runoff.
Limitations
Typical gradient is 0.5%, may be as low as
0.2% or as high as 0.6%.

Appendix 13B
Fact Sheets for ESC Measures

Diversion Bank

13
B

A diversion bank formed by pushing in-situ


soil up the slope is often referred to as
backpush bank.

Alternatives
Diversion channels and catch drains.
Advantages
On larger catchments, the cost savings
resulting
from
the
diversion
of
uncontaminated flow and the resulting
reduction in erosion can be significant.
Disadvantages
Can limit vehicular movement around a
site.
Channels down from the bank may be
subject to erosive flows
Special Requirements
Channels downstream of bank must have a
stable outlet.
Overview
Dimensions:
o
o
o
o

Minimum bank height of 500 mm.


Minimum 150 mm freeboard.
Bank slopes should not exceed 2(H):1(V).
Allow embankment settlement of 10% of fill
height.

Gradient:
o
o

Non-eroding flow velocity during a discharge


of 0.25 time the critical 1 in 1 year storm.
Provide sufficient hydraulic capacity for Q1
peak design discharge with minimum 150
mm freeboard.
Typical gradient is 0.5%, however it may be
as low as 0.2% or as high as 0.6%.

Low maintenance requirements.

Outlet flows are concentrated and may


require energy dissipation and/or a flow
spreader.

Design notes:
o

The diversion of uncontaminated flow


around exposed or contaminated areas is
recommended on all sites independent of
the upstream catchment area. If grass lined
channels are used, they should be
established and stabilised prior to
commencing earthworks.

March 2010

13B-29

Department of Transport and Main Roads


Road Drainage Manual

13
B

Consideration should be given to the


bypassing of flows greater than the design
Q1 discharge.
Construction
Refer to approved plans for location,
extent, and details. if there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
When banks are required and their location
is not shown on the approved plans,
determine the location on the ground taking
into consideration the locations of downslope sediment-trapping devices. The
diversion bank should drain to a sediment
trap if diverted water is expected to contain
sediment.
Clear the location for the bank, clearing
only what is needed to provide access for
personnel and equipment for installation
and maintenance.
Maintenance
Diversion banks should be regularly
inspected.
Check for settlement of banks and reestablish required freeboard if excessive
settlement occurs.
Check for vehicular damage to banks.
Removal
When the construction work above the
bank is finished and the area is stabilised,
temporary diversion banks should be
removed.
Remove accumulated sediment.

March 2010

13B-30

Appendix 13B
Fact Sheets for ESC Measures

Do not clear the catchment area above


until the bank and down-slope sedimenttrapping devices are in place.
Remove roots, stumps, and other debris
and dispose of them properly. Do not use
debris to build the bank.
Build the bank with compacted fill. The
sides of the channel and the back of the
bank must be no steeper than a 2(H):1(V)
slope. The completed flow diversion must
be at least 500 mm deep.
Vegetate bank immediately, unless it will
operate for less than 30 days.
Check that the end of the bank has a stable
outlet and does not discharge to an
unstable fill slope. Where a stable outlet
does not exist, a drop pipe or chute may be
required in association with an appropriate
energy dissipator.
Damaged or eroded areas should be
repaired and where possible, flow should
be directed around damaged areas until
the channel is stable.

Dispose of accumulated sediment properly.


Grade the area and smooth it out in
preparation for stabilisation by grassing or
as specified in the approved plan.

Department of Transport and Main Roads


Road Drainage Manual

Drainage Control 3

Description
Chutes consist of a relatively steep
impervious open channel with an erosion
resistant lining.
Inflow is directed to the chute either by
diversion banks or from a stormwater
outlet. Energy may be dissipated down the
chute face and/or within a specially
constructed outlet.
Application and Function
Transportation of concentrated flow down
an embankment.
Diverts 'clean' stormwater around a work
site.
Limitations
Local topography must allow safe collection
of water at the inlet.

Alternatives
Open channel chutes can consist of half
rounded corrugated metal or concrete
pipes,
prefabricated
concrete
units,
shotcrete, grouted rock, asphaltic concrete,
grout filled mattresses, plastic sheeting,
filter cloth, turf or old conveyor belt rubber.
Also refer to Channel Linings Fact Sheet.
Advantages
Economical for low flows and high, steep
drops.
Some chute types are quick to construct.
Disadvantages
Usually have a defined service life.
May be damaged by overtopping flows.
Special Requirements
Chutes should have a minimum depth of
300 mm and bends should be avoided.

Appendix 13B
Fact Sheets for ESC Measures

13
B

Chute

Temporary chutes are usually constructed


from filter cloth. Permanent chutes may be
constructed from materials such as turf,
gabion mattresses, concrete or rock.
Permanent rock chutes are normally
covered with a thin layer of soil and
seeded.
Directs contaminated flow to a sediment
basin/trap.

Bitumen is generally not suitable as a


permanent chute liner.

Some
structures
requires
earthworks and construction skills.

limited

The chute may be subject to slippage


caused by poor foundations.
An energy dissipater outlet is usually
required.
Good subsoil drainage and foundations are
required.
March 2010

13B-31

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Dimensions:
o

Minimum depth 300 mm

Foundations:
o

The chute lining should be adequately


anchored to the foundations to avoid
slippage, with a maximum distance of 3 m
between anchorage points.
Prefabricated units may need to be bolted
together to avoid slippage. It is important
that all bolt holes are sealed with a flexible
sealant.

Inlet design:
o

The inlet area should be protected against


possible scour resulting from accelerating
flow velocities.
In rock lined chutes, the rock should be
extended at least 1 m upstream of the crest
of the chute, and the rock must be suitably
recessed into the ground to allow the free
entry of water into the centre of the chute.
This is to prevent water from either
undermining the top of the chute, or being
diverting along the edge of the rock lining.
In some cases sand/gravel bags can be
used to direct the inflow towards the centre
of the chute.

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Construct the subgrade to the elevations
shown on the plans. Remove all unsuitable
material and replace with stable material.
Compact the subgrade thoroughly and
shape it to a smooth, uniform surface. On
fill slopes, ensure that the soil adjacent to
the Chute for at least 1 metre is wellcompacted.
Maintenance
Inspect chutes and flumes after each runoff
event and repair all damage noted in
inspections immediately.
After Chute slopes are permanently
stabilised, periodic inspections are required
only after major storm events.

Removal
Temporary structures may only be
removed when an alternative, stable,
drainage path is available.
Remove accumulated sediment and
suitably dispose of in a manner that
minimises environmental harm.

March 2010

13B-32

Appendix 13B
Fact Sheets for ESC Measures

Outlet Design:
An energy dissipater outlet similar to that
used on drop pipes will be required.
o

The outlet structure may consist of a bed of


nominal 150 mm rock (minimum) placed
with a minimum bed thickness of 250 mm
or at least 1.5 times the maximum rock
size.

Typical dimensions of the rock bed are:


o
o

Length (L) = 6De (m) (minimum);


Width = T + 0.6 (m) at the chute outlet, to =
T + 0.5L + 0.3 m at the end of the
dissipator;
where: De = equivalent pipe diameter (m) of
the chute flow area, and, T = top width (m)
of flow in the open chute channel.

Design notes:
o

Each side of the chute should be lined with a


minimum 300 mm wide grassed filter strip
to control side erosion caused by splash.
This turf should be suitably maintained
while the chute is in use. In bushland areas
where the total disturbance is to be
minimised, or where introducing turf is
undesirable, other forms of erosion control
such as geotextiles may be preferred.

If concrete is used, keep the subgrade


moist at the time concrete is placed. Form,
reinforce, and place integrally cutoff walls,
anchor blocks and channel lining.
Place filters and foundation drains, when
required, in the manner specified and
protect them from contamination when
pouring the concrete chute.
Properly stabilise all disturbed areas
immediately after construction.

The inlet and outlet should be checked


regularly to ensure they are clear of
obstructions.
Channel should be checked for movement
or surface cracking.
Erosion control turf around the Chute
should be maintained in a healthy and
vigorous condition.
Grade the area and smooth it out in
preparation for stabilisation.
Stabilise the area as specified in the
approved plan.

Department of Transport and Main Roads


Road Drainage Manual

Drainage Control 4

Appendix 13B
Fact Sheets for ESC Measures

Drop Pipe

13
B

No picture available
Description
Flows are usually directed to the inlet by
the use of perimeter banks and then
directed through the pipe to an energy
dissipator outlet.
Application and Function
Limited application.
Transportation of concentrated flow down
embankments usually greater than 3 m in
height.
Limitations
Local topography must allow collection of
surface water at the pipe inlet.
Usually only economical for low flows.
Alternatives
Drop pipes can be formed from flexible
pipes or solid pipes with flexible joints at
the top and base.

Advantages
Requires
limited
earthworks
and
construction skills.
Can be relocated with relative ease.
Disadvantages
Pipe entry subject to blockage by sediment
and debris.
Usually only suitable as a temporary
structure.
Special Requirements
Trash racks/bars may need to be
considered at the entrance of some drop
pipes to avoid debris blockage.
Soil around the inlet must be well
compacted.
Overview
Sizing:
o

Hydraulic design capacity (typically 0.25


times Q1) usually based on inlet-control
culvert hydraulics with maximum water level
150 mm below the embankment crest.

Details of Catch Bank at Inlet:


o
o
o

Minimum 500 mm high, 1200 mm wide and


3(H):1(V) slopes.
Minimum 300 mm above inlet pipe obvert.
Well-compacted, at least by hand (around
pipe) and in 100 mm layers.

Pipe geometry:
o

Bends should be avoided down the slope


and
anchor
points
provided
at
approximately 3 m intervals.

The pipe is placed and appropriately


anchored on the embankment surface
allowing the pipe to be moved with relative
ease as construction progresses.
Diverts clean stormwater around a work
site.
Directs contaminated flow to a sediment
basin/trap.
Commercially available lay-flat pipe sizes
are limited to a maximum of around 300
mm diameter.
The drop pipe may consist of a collapsible
lay- flat lightweight pipe attached to a solid
inlet pipe.
Alternatives include open channel chutes
which may be cheaper.
Economical for low flows and high, irregular
drops.
Reusable.
Pipes may also be subject to damage by
corrosion, high flows, or vandalism.
A bypass spillway may be required.
Anchorage points must be provided along
the drop pipe at 3 m maximum spacing.
Embankment height over the inlet should
be at least 300 mm above the pipe obvert.

Minimum bed thickness of 250 mm, but


typically 1.5 times dMAX.

Typical geometry of outlet rock bed as


follows:
o
o

Length: = 6D m;
Width: = 3D + 0.6 m;
diameter in metres.

where: D = pipe

Design notes:
o

Debris collection bars or trash racks may


need to be considered on the entrance of
some drop pipes to avoid blockage of the
inlet. These typically should be placed at
least three pipe diameters away from the
pipe entrance, with scour protection placed
between the bars and pipe entrance.
March 2010

13B-33

Department of Transport and Main Roads


Road Drainage Manual

13
B

o
o
o

Inlet section laid at a minimum 3% slope.


Outlets should be at least 1.5 m long on a
grade no steeper than 1%.
Drop pipes should not discharge onto a fill
slope or unstable ground.

Outlet energy dissipator:


o

Level bed of 150 mm rock (minimum). For


non-level outlet beds, the rock size should
be increased.

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Place drop pipes on undisturbed soil or
wellcompacted fill at locations and
elevations shown on the approved plan.
Slightly slope (minimum 3%) the section of
pipe under the embankment constructed
around the pipe inlet.
Anchor the inlet at the top of the slope.
Drive stakes on both sides of the inlet a
minimum of 450 mm into the ground and
secure the pipe to the stakes with wire or
cord.
Hand tamp the soil under and around the
entrance section in lifts not to exceed 100
mm.
Ensure that the embankment over the inlet
has minimum dimensions of 500 mm
height, 1200 mm width, 300 mm over pipe
obvert and 3(H):1(V) slopes.
Ensure that all drop pipe connections are
watertight.
Ensure that all fill material is wellcompacted.
Maintenance
The inlet and outlet must be checked
regularly to ensure they are clear of
obstructions or damage. Sediment and
debris should be removed from the inlet
and outlet.
Joints should be checked to ensure they
are watertight.
Removal
Drop pipes should be removed only when
an alternative stable drainage path is
available.
Remove accumulated sediment.

March 2010

13B-34

Appendix 13B
Fact Sheets for ESC Measures

A sediment trap consisting of a small


excavated pit may be located at the inlet of
the drop pipe.
Collapsible or lay-flat pipes should be
securely attached to solid (corrugated)
pipes embedded within the inlet control
embankment. They are then laid down the
slope and secured at regular intervals with
a wire loop. These pipes can usually be
extended
for
long
distances
over
precipitous and uneven terrain without any
further support. At the outlet, the collapsible
pipe may be knotted or otherwise sealed,
and small holes cut in the pipe to promote
erosion free discharge.

Securely fasten the exposed section of the


drop pipe with anchors spaced no more
than 3 m apart.
Extend the drain beyond the toe of the
slope and adequately protect the outlet
from erosion. Do not direct the outlet to a
fill slope or unstable ground.
Make sure the crest of the compacted, flow
diversion or perimeter bank is no less than
300 mm above the top of the pipe at every
point.
Immediately stabilise all disturbed areas
following construction.
If the drop pipe is not installed correctly, it
shall be immediately made good or
replaced.
A common failure of drop pipes is caused
by water from saturated soil seeping along
the outside of the pipe. The resulting soil
erosion creates voids, piping failure and
washouts. Proper backfilling around and
under the pipe with stable soil material and
hand compaction in 100 mm lifts to achieve
firm contact between the pipe and soil at all
points will generally eliminate this type of
failure.

Inspect all drop pipes and supporting


diversion banks after every runoff event.
Promptly make all necessary repairs to
observed damage and potential problem
areas.

Accumulated sediment should be disposed


of properly
Grade the area and smooth it out in
preparation for stabilisation as specified in
the approved plan.

Department of Transport and Main Roads


Road Drainage Manual

Drainage Control 5

Appendix 13B
Fact Sheets for ESC Measures

Temporary Watercourse Crossing

No picture available
Description
A temporary access culvert consists of one
or more sections of pipe placed over filter
cloth and covered with a suitably graded
aggregate layer.
Application and Function
Temporary crossings provide safe access
for construction equipment with minimal
disturbance to the watercourse.
To help keep sediment generated by
construction traffic out of a watercourse.
Limitations
Generally restricted to minor streams or
creeks.
Culverts and bridges may be subject to
flood damage.
Alternatives
Alternatives include pipe and box culverts
(based on a solid watercourse bed), fords
and bridges.
In shallow bank and rock-based streams,
fords may be used if the road is not
expected to be subjected to regular
sediment-laden traffic.
Advantages
Minimises disturbance to the watercourse.
Improves access to a construction site.
Disadvantages
They can be a direct source of water
pollution.
They may aggravate flooding and/or create
a safety hazards.
Special Requirements
When the crossing is no longer needed, all
materials including pipes, aggregate and
filter cloth should be removed. Final clean
up also includes restoration of the
watercourse to the original channel cross
section and stabilisation of the banks.

To restrict construction traffic to stabilised


watercourse crossings.

Fords require a rock bed and should not be


used where the base flow exceeds 75mm
depth.
In high-bank streams, culvert crossings are
preferred to fords as they usually cause
less disturbance to the watercourse.

Temporary ford crossings are less likely to


be damaged by flood waters.
Expensive to construct and costly to repair
if damaged by flood waters.

Approval to construct a temporary culvert


may be required from the Department of
Natural Resources and Mines Water
Resources Section.

March 2010

13B-35

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
All designs should be appropriate for the
local watercourse conditions.
Culvert crossings:
o
o

Location
Preferably located on a straight section of a
watercourse, well downstream of a sharp
bend.
The approach road should be straight for at
least 10 metres each side of the crossing
and should desirably cross the watercourse
at right angles.
In any case, crossings should be located in
an area that will cause the least overall
disturbance, especially to those areas that
are required to remain in a 'natural' state.

Sizing:
o

o
o
o

Hydraulic capacity of the culvert generally


should be equivalent to the in-bank
hydraulic capacity of the watercourse below
the deck level.
The culvert should consist of the largest
diameter and greatest number of cells that
will fit into the existing channel.
Minimum pipe size, 450 mm.
Minimum spacing between pipe cells, 300
mm.
Culvert length to extend at least 300 mm
beyond rockfill.

Construction
Minimise disturbance to stream banks.
Culvert cells should be placed on a
geotextile filter overlaid with a bed of
washed rock.
Cells should be covered with a sufficient
layer of compacted 150 mm rock to prevent
their damage when subjected to the
expected traffic loads.
Maintenance
Temporary access crossings should be
inspected regularly, and always after heavy
rainfall.
Sediment and debris trapped upstream or
within the culvert should be removed.
Removal
Avoid deposition of fill material in the
watercourse.
Downstream
bunding
or
temporary
damming may be required if there is a
likelihood of sediment re-suspension and
potential environmental harm. Provision for
low flows should be maintained in any
case.

March 2010

13B-36

Appendix 13B
Fact Sheets for ESC Measures

Stabilised fords:
o
o
o

o
o
o
o
o
o
o
o
o

The watercourse must have a solid rock


bed.
The watercourse should not have a base
flow greater than 75 mm depth.
Access ramps should be stabilised with
geotextile overlaid with 150 mm (min.) size
rock.
Department of Environment and Resource
Management permit requirements for
'defined' watercourses typically include the
following:
Maximum 6 m width of disturbance.
Maximum of 0.6 m of cut and fill.
Constructed in a stable stream section.
Crossing to take the shortest practical route.
Crossing shall not store or divert water.
Access down the banks should be directed
downstream.
Natural stream bank levees should not be
breached.
Erosion protection works provided.
Drainage diversions installed on access
ramps.

Culvert fill should be raised a minimum of


300 mm above the adjoining approach
sections to prevent erosion from surface
runoff and to allow flood flows to pass
around the structure.
No earth or soil material should be used as
fill within the waterway channel.
Structures should be removed as soon as
possible after alternative site access is
achieved.

Consultation with the Department of


Natural
Resources
and
Mines
is
recommended when constructing and
removing in-stream works.

Department of Transport and Main Roads


Road Drainage Manual

Velocity Control 1

Description
Check dams may be constructed from
either semipervious or impervious materials
typically, medium size rock or sand/gravel
filled bags. Check dams should not be
constructed from straw bales.
Application and Function
Check dams are normally used to control
flow velocities in newly formed channels
that have a grade of less than say 1 in 10.
To control minor gully erosion.
Limitations
Check Dams should be limited in height to
around 0.5 m, or 1 m if formal design.
Catchment area generally limited to 1 ha.
Check dams are not normally placed in
defined watercourses. In circumstances
where they are necessary (i.e. during major
channel works) then sediment weirs, rock
filter dams or formally designed drop
structures should be used.
Alternatives
At the principal discharge point on major
construction sites other structures such as
rock filter dams can be used as the final
check dam.
Advantages
Quick and inexpensive to install.
May be used to avoid the need to turf of
new drainage channels.
Disadvantages
Damage can occur to grass cutting
equipment if check dams are not removed
from the channel or table drain.
Special Requirements
A marker post may need to be placed
adjacent to all rock check dams to make
their existence obvious to grass cutting
maintenance personnel. These posts
should only be removed once the rock
check dams have been removed.

Appendix 13B
Fact Sheets for ESC Measures

Check Dam

13
B

Check dams may be recessed into the


channel bed to limit the crest height of the
dams.

To control flow velocity in drainage


channels, especially during the early
stages of revegetation.
Entrapment of sediment is only a
secondary (minor) function.
In large, shallow overland flow paths, rock
check dams may need to be recessed into
the channel bed to avoid flow diversion
from the channel.
Check dams cannot be used to control
invert erosion in channels constructed in
dispersible soils. In such cases, a 200 mm
layer of non dispersible soil should be
placed over the in-situ dispersible soil
before placement of the check dams.

Can be a quick maintenance option on


eroding channels during or after the
establishment period.
Permeable dams are free-draining.
Check
dams
may
cause
erosion
downstream of the dam if poorly designed,
maintained, or if subjected to high flows.
Care must be taken to prevent failure
caused by water undermining or side
cutting the structure.

March 2010

13B-37

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Check dams should only be used in
channels conveying concentrated flows.
Rock check dams should be used in deep
(>500 mm) channels. If the channel is less
than 500 mm deep, then sand or gravel
filled bags should be used to form the
check dam.
Dimensions:
o

Maximum height of around 0.5 m or 1 m if


formally designed.
Crest invert to be at least 150 mm lower
than outer edges.
Maximum bank slope of 2(H):1(V).

Typically 200 mm.

Appendix 13B
Fact Sheets for ESC Measures

Downstream erosion control:


o

Downstream erosion control:


o

Rock Size:
Spacing:
o

Check dams are located such that the toe of


the upstream dam is at the same elevation
as the crest of the downstream dam.

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Place rock, on a filter fabric foundation, to
the lines and dimensions shown in the
approved plan.
Keep the centre rock section at least 150
mm below natural ground level where the
dam abuts the channel banks.
Maintenance
Inspect check dams and channels for
damage after each runoff event. Correct all
damage immediately. If significant erosion
occurs between dams, check spacing of
dams and/or install a protective rock liner in
that portion of the channel.
Maintain the dams so that runoff will flow
through the washed rock or over the dams
spillway. Keep the spillway shaped so that
runoff will go over the centre of the dam or
over the ends where it can wash out the
more erodible stream banks. If necessary,
line the bottom and sides of the channel
below the spillway so that flow through the
spillway is on rock, not the erodible
channel.

March 2010

13B-38

Erosion downstream of each dam will be


minimised if the dams are correctly spaced
such that the crest of
Erosion downstream of each dam will be
minimised if the dams are correctly spaced
such that the crest of the downstream dam
is level with the toe of the upstream dam.
Where necessary, erosion control at the toe
of the dam may consist of a rock apron
extended downstream of the toe a distance
of at least twice the height of the check
dam. Erosion control may also consist of a
suitable geotextile filter cloth or erosion
control mat anchored under the check dam.

Design notes:
o

A U-shaped weir crest is normally centred


over the centre line of the drainage channel
to provide the greatest length of energy
dissipation down the back face of the dam.
Rock
protection
may
be
required
downstream of the structure to dissipate
discharge energy.

The rock shall be extended up the channel


bank (where practical) to a level at least
150 mm above the crest level of the dam.
Set spacing between dams to assure that
the elevation at the top of the lower dam is
the same as the toe elevation of the upper
dam.
Protect the channel downstream of the
check dams (especially downstream of the
lowest dam), noting that water will flow over
and around the dams.
Ensure that the channel reach above the
most upstream dam is stable.
Remove sediment accumulated behind the
dams as required to prevent damage to
channel vegetation. Allow the channel to
drain through the check dam, and prevent
large flows from carrying sediment over the
dam. Add rocks as needed to maintain
design height and cross section.
Ensure that channel structures, such as
culvert entrances below check dams, are
not subject to damage or blockage from
displaced stones from the check dam/s.
Place sediment in a disposal area or, if
appropriate, mix it with dry soil on the site.
Do not dispose of sediment in a manner
that will create an erosion hazard.

Department of Transport and Main Roads


Road Drainage Manual

Removal
When construction work within the
drainage area above the check dams has
been completed and disturbed areas
sufficiently stabilised to restrain erosion,
the dams must be removed.

Appendix 13B
Fact Sheets for ESC Measures

Remove accumulated sediment from


behind the filter and dispose of
appropriately.
Remove the rock and any filter cloth.

March 2010

13B-39

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-40

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Erosion Control 1

Appendix 13B
Fact Sheets for ESC Measures

13
B

Chemical Surface Stabiliser

No picture available
Description
Most chemical stabilisers are soil binders
that provide a thin surface crust.
Application and Function
Generally effective for dust control or the
control of erosion caused by raindrop
impact.
Limitations
Products have a limited life and
consideration should be given to the use of
geotextiles if the exposed surfaces need to
be protected for extended periods or during
the wet season.
Alternatives
Some products are used on their own as
temporary surface stabilisers, while other
products may be used with vegetation or
geotextile matting.
Advantages
Provide instant protection.
Disadvantages
Usually less effective than mulches.
Bitumen products can breakdown and
release pollutants to receiving waters.
Some products can reduce water
infiltration.
Special Requirements
For many products, the protective layer
must remain intact to be effective and
therefore no traffic, of any kind, is
permitted.
Overview
Bitumen-based products:
o

A mixture of bitumen, emulsifying agent,


stabilising agent and water will have a low
viscosity and may be applied without
heating. This mixture does not create any
chemical change in the bitumen. However,
the physical state is altered so that it is no
longer a homogeneous mass but a series of
minute particles, each coated with the
emulsifying agent to keep them apart and
suspended in the water (i.e. an emulsion).
When applied, typically by spraying, the
emulsion breaks and the bitumen particles
merge as the water is absorbed or
evaporates. The rate of break can be
varied, depending on the emulsifying and
stabilising agents.

Also used for tacking organic mulches.

Many products will be damaged by traffic


movement over the treated surface,
therefore, light protective fencing maybe
required.
Alternatives include: mulches, Erosion
Control Blankets and temporary grassing

Suitable for temporary stabilisation while


construction is in progress.
Vegetation may also root through cracks
causing crumbling of the surface.
The established surface crust must remain
intact to be effective.
Specialised machinery is generally needed
for their application.

Two types
recognised:
o

of

bitumen

emulsion

are

Cationic bitumen emulsion which is acid in


nature and is favoured for road sealing due
to the attraction of the positive charge
particles to road base aggregates. They are
rarely used in Australia for soil conservation
purposes.
Anionic bitumen emulsion which is alkaline
in nature; slow-breaking anionic bitumen
emulsions are commonly used in soil and
batter stabilisation programs.

Others:
o

These
include
proprietary
chemical
stabilisers including resins derived from
timber.
March 2010

13B-41

Department of Transport and Main Roads


Road Drainage Manual

13
B

Construction
Application
as
specifications.

per

manufacturers

Maintenance
Areas should be revegetated as soon as
possible otherwise reapplication may be
required as the protective surface
deteriorates.
Disturbance to the surface crust by
machinery, vehicles, or otherwise post
application, must be avoided if possible, as
this breaks up and disturbs the surface
crust reducing the effectiveness of this
method. Temporary barricading, or fencing
off the area treated would reduce the
likelihood of this occurring.
Removal
The surface crust is easily removed by
applying a light grade across the treated
area. If this is not possible (due to terrain)
then
other
removal/rehabilitation
techniques could be tried.

March 2010

13B-42

Appendix 13B
Fact Sheets for ESC Measures

If used as an erosion control method for the


medium to long term works, repeat
applications may be necessary to ensure
the integrity of the surface crust.
Root growth of plants can grow through the
surface crust in the medium to long term.
This is most likely to occur if vegetation is
adjacent to the area treated, or the area
treated was not devoid of vegetation
(ground cover) prior to treatment, and if
follow up applications are not undertaken.

Department of Transport and Main Roads


Road Drainage Manual

Erosion Control 2

Description
Erosion
Control
Blankets
include
impervious sheets used to cover bare soil
and pervious fabrics utilised to facilitate
and promote plant growth.
Application and Function
Fabrics provide only temporary erosion
control (usually no more than 6 months) on
exposed earthern surfaces, newly seeded
grass channels or newly grassed channels.
Limitations
Blankets are best used in conjunction with
vegetation buffer zones (eg. grassed rows).
Limited versatility for non biodegradable
products in bushland sites.

Alternatives
Biodegradable fabrics.
Synthetic fabrics.
Composite
materials
(biodegradable
materials reinforced with synthetic mesh).
Permeable
mats:
eg.
buried
root
reinforcing.
Advantages
Quick installation.
Wide product range and associated uses.
Most products instantly reduce erosion
potential if installed correctly.
Disadvantages
Some fabrics have a very limited life span.
Environmental problems associated with
non biodegradable fabric used in bushland
areas.
Special Requirements
Four general requirements must be met in
order to provide effective erosion control:
o
o
o
o

1. good soil-blanket contact;


2. seepage flow under the fabric should be
avoided;
3. soil surface irregularities removed; and
4. the blanket must be securely anchored.

Appendix 13B
Fact Sheets for ESC Measures

Erosion Control Blanket

13
B

Erosion blankets can be effectively used on


batters while erosion mats are more
appropriately used as channel liners.
Biodegradable products generally have a
lower allowable velocity limit.
Impermeable fabric should be considered
on highly dispersive soils.

Cellular grid: eg. turf reinforcing.


Pocket fabrics: eg. grout filled mattresses.
Also refer to Channel Linings Fact Sheet.

Can be installed for temporary emergency


repairs.
Products are available to suit short and
long term requirements in the field.
Undermining of fabric may occur after peak
flow events which requires ongoing
maintenance attention.
The crest, toe and sides of the fabric
should be anchored to avoid up-lifting,
undermining and subsequent erosion.

March 2010

13B-43

Department of Transport and Main Roads


Road Drainage Manual

13
B

Appendix 13B
Fact Sheets for ESC Measures

Overview
Refer to manufacturers specifications.
Construction
The method of installation varies with the
type of material used and the intended
purpose being performed by the blanket.
Methods of installation of these products
are normally provided with the materials.
However, a typical laying procedure is
described below:
Refer to approved plans for location,
extent, and specifications. If there are
questions or problems with the location,
extent, or methods of installation, contact
the engineer or responsible on-site
personnel for assistance.
Prepare
a
smooth
seedbed
of
approximately 75 mm of topsoil - clear
away trash and large stones, and grade
smoothly to eliminate footprints, tracks and
ruts.
Seed, fertilise, water and rake to remove
any remaining surface irregularities.
Preferably starting at the downstream end,
align fabric strips/rolls in the direction of
flow. In situations where the bank is subject
to both concentrated in-bank stream flow
and transverse overland sheet flow, align
fabric length with the dominant flow usually the concentrated stream flow.
Bury the downstream end of the first roll in
a 300 mm deep trench.
When spreading the fabric, avoid stretching
the textile mesh - the blankets should be in
good contact with the ground at all points.
Maintenance
All surface-laid fabrics should be inspected
on a regular basis.
Inspect the fabric to see if construction
activity or fallen rocks and trees have
damaged it, or if runoff is undermining the
fabric.
Removal
If the fabric is a temporary measure, it must
be replaced with permanent stabilisation at
an appropriate time as specified in the
plan.

March 2010

13B-44

Staple the fabric within the trench at 200 to


250 mm spacing using 100 mm wide, 150
mm penetration length, "U" shape, 8 - 11
gauge wire staples. Narrower "U" sections
may easily tear the blanket.
Staple the exposed fabric surface at 1.0 m
centres.
Where more than one sheet is used overlap the ends by 300 mm, with the
upstream
sheet
placed
over
the
downstream sheet. In high velocity areas,
the upstream end of all sheets should be
buried in 300 mm deep trenches and
stapled.
Overlap the sides by 100 mm. If the bank is
subject to both, concentrated inbank
stream flow and transverse overland sheet
flow, the upper bank sheet should overlap
the lower sheet as well as the upstream
sheet overlapping the downstream sheet. It
may be preferable in these cases to start
laying the sheets at the downstream invert
and then moving first up the bank, then up
the drainage line.
Anchor the outer edge and top within a 300
mm deep trench and stapled at 200 to 250
mm centres. Additional staples will be
required wherever the blanket does not
contact the ground surface.
Blankets, once fixed, may be rolled with a
roller weighing 60 to 90 kg/m length, then
watered.
If damaged, repair or replace the damaged
section. If water is undermining the fabric,
repair any holes or joints or rebury the
upper ends of the damaged sections.
Biodegradable
blankets
should
be
inspected after the first few storms that
cause runoff.
The temporary stabilisation must be
maintained until arrangements have been
made to install the permanent lining.
Dispose of the fabric properly.

Department of Transport and Main Roads


Road Drainage Manual

Erosion Control 3

13
B

Mulching

Description
Mulching is the application of a protective
blanket of straw or other plant residue,
gravel, or synthetic material to the soil
surface.
Application and Function
Applied to clay-soil surfaces to limit runoff
turbidity caused by raindrop impact.
Applied to mild slopes to control raindrop
impact as well as erosion caused by sheet
flow.
Used to control soil temperature and
moisture loss.
Limitations
Some mulches are not suitable in bushland
areas due to possible introduction of
unwanted seeds.
Not suitable for areas subjected to
concentrated flow unless a suitably sized
gravel mulch is used.
Alternatives
Mulch materials can
wood chip, sugarcane
of sawdust and wood
newspaper, straw,
fibreglass.

Appendix 13B
Fact Sheets for ESC Measures

include pine bark,


bagasse, a mixture
shavings, shredded
paper pulp and

Advantages
Most effective and practical means of
controlling erosion prior to vegetation
establishment.
Can be applied on irregular and steep
terrain.

In windy areas, straw mulch is held in place


by crimping the mulch into the soil with
agricultural machinery, by covering with a
surface netting, or by applying a tackifier (a
spray-on chemical binder).
Applied to steep slopes (> 20%) to control
erosion, but usually reinforced with netting
or tackifiers.
Mulch can be used to aid or inhibit seed
germination, and to control weed growth.

Mulch should cover 80-100% (minimum) of


the soil surface to give adequate protection
against erosion.
Is effective in controlling raindrop erosion,
however sub-surface flows can result in
erosion occurring underneath the mulch
blanket.
Where long-term protection is required
fibreglass mulching can be used, however,
such mulch is not be considered
environmentally acceptable.
Alternatives include: biodegradable erosion
control blankets.
Particularly useful in higher rainfall areas to
protect against raindrop impact; they also
restrict moisture loss, increase infiltration
rates,
and
minimise
temperature
fluctuations.

March 2010

13B-45

Department of Transport and Main Roads


Road Drainage Manual

13
B

Disadvantages
Decomposition of some wood products can
tie-up significant amounts of soil nitrogen,
thus requiring modification to the fertiliser
application rates.
Special Requirements
Mulch should be spread evenly with a
maximum depth of 50 mm. Thicker layers
(75 to 100 mm) inhibit germination and can
be used to control weed growth.
Overview
Coverage:
o

Appendix 13B
Fact Sheets for ESC Measures

Associated bitumen-based fixers can


release pollutants to receiving waters.
May be displaced if subjected to flooding or
concentrated overland flow.
On steep slopes, mulch is normally kept in
place by spraying with a bitumen emulsion
or other chemical tackifier, or by covering
with a fine geosynthetic netting.
o

Generally spread evenly to a maximum


depth of 50 mm over a minimum 80-100%
coverage of the soil surface.

Weed control:
o

Spread evenly to a minimum thickness of 75


to 100 mm.

Tackifiers and mechanical anchoring:


o

Chemical tackifiers or mechanical anchoring


may be used to hold the mulch in position
to reduce loss by wind and water,
especially on sloping ground.

Application:
o

The necessity for mulching should reflect the


length of the expected construction delay,
topography, soil and weather conditions,
and the overall level of environmental risk.

Suggested application rates (guide only):


o

Straw Mulches: The most commonly used


mulch, particularly in conjunction with
seeding. Straw mulch is effective on sites
with a high soil erosion hazard; or where
soil moisture is likely to be inadequate for
successful plant establishment. Other than
in bushland areas, wheat or oat straw is
suitable. The mulch should be dry when
applied and should have a low leaf content.
Straw mulch may be applied to small areas
such as earth embankments and soil
stockpiles. Application rate: 250 bales per
hectare, or 3.4 to 4.5 tonnes per hectare
(machine placement); 1 bale per 25m2
(hand spread).
Brush Mulch: This mulch is preferred on
areas where regeneration with native plants
is desired. Brush mulch should be applied
parallel to the contours, and should be
stockpiled with care since spontaneous
combustion is reported to occur. In
bushland areas the mulch should comprise
native species where available and be free
of non-endemic seed.

March 2010

13B-46

Wood Cellulose Fibre Mulch: Such mulches


consist of short cellulose fibres applied by
hydroseeders. Wood fibre does not require
tacking, although tacking agents or soil
binders can easily be added to the slurry.
Wood fibre hydroseeder slurries may be
used to tack straw mulch on steep slopes,
critical areas, and where harsh climate
conditions exist. Wood fibre mulch does not
provide sufficient erosion protection to be
used alone (North Carolina SCC& Wood
Cellulose Fibre Mulch: Such mulches
consist of short cellulose fibres applied by
hydroseeders. Wood fibre does not require
tacking, although tacking agents or soil
binders can easily be added to the slurry.
Wood fibre hydroseeder slurries may be
used to tack
Wood Cellulose Fibre Mulch: Such mulches
consist of short cellulose fibres applied by
hydroseeders. Wood fibre does not require
tacking, although tacking agents or soil
binders can easily be added to the slurry.
Wood fibre hydroseeder slurries may be
used to tack straw mulch on steep slopes,
critical areas, and where harsh climate
conditions exist. Wood fibre mulch does not
provide sufficient erosion protection to be
used alone (North Carolina SCC&DEHNR,
1993). Application rate: 1 to 2 tonnes of
wood pulp per hectare, typically 1.6 t/ha.
Wood Chip Mulches: Developed from
mulched tree loppings, pine flake, or
processed hardwood, these mulches are
useful for weed control and mulching small
areas that are not closely mowed.
Application of a nitrogen rich fertiliser may
be required. Application rate: 11 to 14
tonnes per hectare.
Bark Chips and Shredded Bark: A byproduct of timber processing often used in
landscape planting. Unlike wood chip, the
use of bark does not require additional
nitrogen fertiliser. Pine bark lowers the pH
of the soil, so should not be used on low pH
soils. Application rate: 2.5 cubic metres per
hectare.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
If the mulching is not effective in containing
soil erosion it should be replaced or an
alternative erosion control procedure
adopted.
Remove rocks, stumps, roots, and other
debris that will interfere with the mulch
contacting the soil and make maintenance
difficult.
Maintenance
Inspect all mulches periodically, and after
storm events to check for rill erosion,
dislodgment or failure.
Any places where the mulching is
dislodged and the soil exposed must be
repaired with additional mulch. Tacking or
netting shall be applied where necessary.

Removal
N|A

Hydroseeding,
and
Hydromulching
(increased mulch content), involves the
mixing of seed, fertiliser, paper pulp (5
tonnes per hectare) or wood pulp (not less
than 2.5 tonnes per hectare) and a bitumen
emulsion or tackifier with water to form a
slurry that is sprayed over the area to be
revegetated. The seed generally sticks to
the pulp which improves the microclimate
for
germination
and
establishment.
Application rate: should be between
30,000-40,000 litres per hectare. The slow
curing anionic bituminous emulsion is
applied at a rate of around 1500 to 2000
litres per hectare.

Where practicable, mulch material shall be


free of weed species especially prohibited
noxious weeds.
Spread enough mulch to completely cover
the surface of the soil at the density
specified in the approved plans.
To prevent wind or water dislodging the
mulch, hold the mulch in place by tacking it
with slow setting anionic bitumen (AS
1160), covering it with netting, or other
specified product. On flat or gentle slopes,
straw mulch may be fixed to the soil by
running a straight-set disk over it. The disk
should not be sharp enough to cut the
straw.
If washout occurs, repair the slope, reseed
and reinstall mulch cover.
Continue inspections until vegetation is
suitably established.

March 2010

13B-47

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-48

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Erosion Control 4

Appendix 13B
Fact Sheets for ESC Measures

13
B

Temporary Revegetation

No picture available

Description
Revegetation is the process of restoring
vegetation to disturbed areas by application
of vegetative cover of various types.
Application and Function
Soil
surface
protection
and
soil
reinforcement.
Stabilisation of shallow land slips.
Interception and retention of stormwater
runoff.
Reduce rainfall impact energy.
Increase
soil
permeability
and
evapotranspiration, thus reducing the
volume of total annual runoff.
Limitations
There are limits to the role vegetation alone
can play in controlling erosion. Both soil
strength and vegetative cover (including
root system) can take years to develop to
the required condition.
Alternatives
Vegetation options include temporary
seeding for erosion control during
construction delays.
Advantages
In terms of soil surface protection,
vegetation is the best long-term defence
that can be used to protect soil against
wind and water induced erosion.
Disadvantages
Long establishment time.
Subject to damage in heavy traffic areas.
Conflicts can exist between the choice of
native and exotic species.
Special Requirements
Usually requires guidance from local
experts.
At least 70% ground cover is considered
necessary to provide a satisfactory level of
erosion control.
A mulch cover layer is usually required to
control short-term erosion and provide
good growing conditions.

Established vegetation (shrubs and trees)


can be used to control traffic routes
(vehicular, human, or stock) and thus
reduce the extent of surface damage and
soil compaction.
Reduce wind-induced soil erosion.
Filter sediment from sheet flow.

Usually not suitable in heavy traffic areas


or for long slopes greater than 2(H):1(V).

Alternatives include mulching and erosion


control blankets.
Also see buffer zones and grassed filter
strips.
Environmentally sound and an inexpensive
long-term erosion control measure.
Self regenerating.
Reduces the long-term loss of topsoil.
Introduced non native species can migrate
from the local area, block downstream
watercourses, indirectly increase bank
erosion, etc.
Requires suitable soil and soil conditioning.
Requires a reliable water supply.
May require a geosynthetic surface cover
on slopes steeper than 4(H):1(V).

March 2010

13B-49

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Selecting
the
most suitable
plant
establishment techniques, appropriate
species, seeding rates, planting densities,
fertiliser types, watering rates, and
maintenance techniques, requires the
guidance of experts such as local bushland
groups,
landscape
consultants
and
government bodies.
Construction
Seedbed preparation:
o
o

Refer to approved plans for location, extent,


and details.
Remove rocks, stumps, roots, etc. to avoid
their interference with seeding and
maintenance procedures.
Prepare the seed bed by applying soil
conditioners and fertiliser as specified on
the approved plans. Rip the soil 100 to 150
mm to mix the components into the soil and
to loosen and roughen the soil surface
before seeding.
Where possible, there should be sufficient
soil depth to provide an adequate root
zone. Soil pH should be within the specified
range.

Seeding:
o

Appendix 13B
Fact Sheets for ESC Measures

Mulching:
o
o

Tree protection:
o

o
o

Apply seed uniformly by hand or with a


cyclone seeder, drop-type spreader, drill,
hydroseeder, hydromulcher, or other
suitable equipment.
Apply seed at the recommended rate, and
disc or otherwise mechanically treat the
surface to bring the seed into contact with
the soil.

Maintenance
Areas must be reseeded and mulched
where the vegetation fails to establish or is
damaged by runoff or construction
activities.
Following significant storm events, monitor
that vegetation is controlling erosion and
stabilising soil slopes.
If the temporary vegetation fails before the
permanent stabilisation is provided, it must
be replaced with an appropriate type of
cover sufficient to restrain erosion.
If the permanent vegetation fails to
establish or to adequately restrain erosion
for any reason during the construction or
maintenance period, the area should be
revegetated or protected with other erosion
control measures as appropriate.
In areas where the established vegetation
is considered inadequate for erosion
control, over-seed and fertilise the area
using half the originally specified rates, or
otherwise replant following advice from an
appropriate specialist.
March 2010

13B-50

The seeded area should be mulched as


specified in the approved plan.
Where appropriate, apply a tackifier, netting,
or crimp the mulch into the soil with a
straight-set disk. The disk should not be
sharp enough to cut the mulch.
Place barriers to prevent the approach of
equipment within the drip line of trees that
are to be retained.
Do not spike or nail boards to retained trees
during the building operation.
The cutting of tree roots inside the tree drip
line shall be avoided wherever reasonably
possible. (Continued overleaf)
Do not place equipment, construction
materials, topsoil, or fill within the limit of
the drip line of retained trees.

Inspect and where necessary repair


protective fencing at maintenance periods
not exceeding one month.
Re-firm plants loosened by wind.
At the regular maintenance inspections and
at the end of the maintenance periods,
prune any plants of dead and diseased
parts.
Dispose of cleared vegetation in an
appropriate manner such as chipping or
mulching, on site burial, or off site disposal.
Watering:
o

o
o
o

Watering should start immediately after


planting. Watering should comply with
specifications supplied with the approved
plants. Generally watering should vary
depending on weather and soil conditions,
but may well consist of the following:
25mm every second day for the first three
waterings;
25mm twice a week for the next three
weeks;
25mm once weekly for a further two weeks.

Department of Transport and Main Roads


Road Drainage Manual

Maintain grass strand length to a minimum


50 mm within high velocity drainage areas,
and 20 to 50 mm within low velocity
overland flow paths.
Where necessary, or as directed by the site
supervisor, slash the temporary crop/grass
cover to allow the successful growth of the
underlying permanent vegetation cover.
Control weed growth within 1 m of
immature trees for 6 to 12 months for fast
growing species such as eucalypts, and 18
to 20 months for slower growing species, or
until the end of the specified maintenance
period.
Where mulch is used to control weed
growth, inspect and where necessary,
renew at maintenance periods not
exceeding 4-6 months.
Removal
N|A

Appendix 13B
Fact Sheets for ESC Measures

Retained trees:
o

Repair damaged tree roots by cutting off the


damaged areas and sealing them with an
approved product. Spread moist topsoil
over exposed roots.
Cut off all damaged tree limbs above the
tree collar at the trunk or main branch. Use
several cuts including undercutting to avoid
peeling bark from the healthy areas of the
tree.

March 2010

13B-51

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-52

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Erosion Control 5

Appendix 13B
Fact Sheets for ESC Measures

13
B

Surface Roughening

No picture available
Description
A technique that leaves the soil surface in a
roughened state to increase water
infiltration, decrease and slow down run-off
and to encourage sediment retention and
vegetation establishment.
Track marks left by a dozer working up and
down the slope (sandy soils only) is one
surface roughening technique.
Application and Function
On recently seeded or exposed earth
surfaces, erosion protection can be
improved by roughening the soil surface to
increase infiltration and delay the formation
of rutting.
Limitations
Surface roughening or ripping is not
effective during major rainfall events where
concentrated runoff willbreak through the
furrows and cause rill erosion.
On uncompacted soils, track-walking can
concentrate runoff in a vertical direction,
thereby increasing the risk of rill and gully
erosion.
Alternatives
Surface roughening may be carried out
before or immediately after topsoil
placement and seeding.
Advantages
Increases infiltration and reduces runoff.
Aids the establishment of vegetation.
Disadvantages
Of only limited value during periods of
heavy rainfall.

The soil surface can also be roughened by


horizontal grooves produced by an
excavator's toothed bucket, or a chisel
plough, scarifier or ripper.
On gentle slopes only intermittent grooves
are required.

Limits wind-induced soil erosion.

On dispersive soils, contour ripping may


increase the risk of erosion. Gypsum or
lime should be applied to dispersive soils.
Tracking is generally not as effective as the
other surface roughening methods as it can
compact heavier soils.

Alternatives include: catch drains, erosion


control
blankets,
chemical
surface
stabilisers and mulching.
Can improve the stabilisation of topsoil
when surface roughening has been applied
to the subsoil.
Questionable benefit for the outlaid time
and money.

Special Requirements
Existing rutting and gullies should be filled
or suitably contoured.

March 2010

13B-53

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Coppin and Richards (1990) guidelines:
o
o
o
o
o

slopes <18% (3H:1V) should be left


roughened and undulating
slopes >18% but <27% can be step graded,
grooved or roughened
slopes >27% (2H:1V) can be step graded.
Roughening or grooving is done along the
contours:
Contour ripping usually penetrates at least
300 mm and is performed on hard packed
soils to improve water infiltration, or on
overburden immediately prior to topsoiling
to assist bonding between the two soil
layers. Rip lines vary from 1 m to 10 m
apart.
Contour furrowing or grooving involves the
construction of small furrows, the distance
between the furrows depending on the soil
type and slope. Contour furrowing aims to
pond water, allowing it to soak into the soil.
This technique is normally restricted to
grazing lands.

Contour Ripping:
o
o

Penetrates at least 300 mm


Rip lines vary between 1 m to 10 m apart

Operate tracked machinery up and down the


slope to leave horizontal depressions in the
soil. Do not back-blade during the final
grading operation.
Limit roughening with tracked machinery to
sandy soils to avoid undue compaction of
the soil surface.

Roughening with tracked machinery:

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel
for assistance.
Maintenance
Periodically check the seeded slopes for
rills and/or washouts.
Removal
N|A

March 2010

13B-54

Appendix 13B
Fact Sheets for ESC Measures

Cut slope steeper than 3(H):1(V)


o
o

Stair grade or groove slopes.


Use stair-step grading on any erodible
material soft enough to be ripped with a
bulldozer. Slopes consisting of soft rock
with some subsoils are particularly suited to
step grading.
Make the vertical cut distance less than the
horizontal distance, and slightly slope the
horizontal surface of the step in towards the
vertical face.
Do not make individual vertical cuts more
than 600 mm deep in soft materials or more
than 1000 mm deep in rocky material.
Grooves should consist of a series of ridges
and depressions running along the contour
and created by any appropriate implement
that can be safely operated on the slope.
Grooves are not to be less than 75 mm
deep nor more than 400 mm apart.

Fill slope steeper than 3(H):1(V)


o

Ensure that the face of the slope consists of


loose, uncompacted fill 100 to 150 mm
deep. Otherwise, use grooving to roughen
the face of the slopes, if necessary.
Do not blade or scrape the final slope face.

Cuts, fills, and graded areas not steeper


than 3(H):1(V)
o

Roughen areas with shallow grooves by


normal tilling, discing, harrowing, or other
suitable means. Make the final pass along
the contour.
Grooves shall be spaced at no less than 250
mm and shall be not less than 25 mm deep.
Excessive grooving is undesirable where
mowing is planned. Areas should generally
be mulched.

Fill these areas slightly above the original


grade, then as soon as possible reseed
and mulch.

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 1

Description
Buffer zones are corridors of vegetation
which have been retained or installed to
decrease the velocity of overland flow and
facilitate settling of suspended sediments.
Application and Function
Generally, the wider the corridor the more
effective the buffer zone will be in achieving
a reduction in overland flow velocity and
settling of suspended particles.
Limitations
Buffer zones are only capable of trapping
coarse sediments. Clays and fine silts will
pass through a buffer zone during periods
of heavy rainfall.
Alternatives
Alternatives include sediment fences.
Advantages
If buffer zones are appropriately retained or
properly established then they reduce the
need for on-site sediment controls required
during construction.
A reduction in transportation of total
sediments downslope is achieved.
Disadvantages
Limited sediment trapping performance
under high flows.

Appendix 13B
Fact Sheets for ESC Measures

Buffer Zone

13
B

Generally, the wider the corridor the more


effective the buffer zone will be in achieving
a reduction in overland flow velocity and
settling of suspended particles.
Primary function is to control sediment
runoff from access roads, stockpiles, and in
general construction areas on site.
Buffer zones are suitable
between 1% and 10%

on

slopes

Cost savings are realised if existing


vegetation corridors are retained as this will
reduce the need to revegetate the whole
site.
Impact of construction on surrounding land
is reduced.
Buffer zones can be easily disturbed or
destroyed by poor on-site management

Special Requirements
Buffer zones should be fenced off to
exclude traffic and prevent damage to
vegetation and wheel rut formation.
Rills and gullies have the capacity to
concentrate flow, therefore decreasing the
effectiveness of the buffer zone.

March 2010

13B-55

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Dimensions:
o

As a general guide, the width of the buffer


zones in metres should be five times the
percentage slope (i.e. 50 m buffer width on
a 10% slope). It is also recommended that
the buffer width should be 1.5 times the
width of the disturbance. The latter
generally applies to grassed buffers
adjacent to strip construction such as roads
and utility installations. In some locations
these design rules may not need to be
applied.

General guide:
o
o
o

(i) Width [m] = 5 x [%] slope, for broad


disturbances; or
(ii) Width = 1.5 x width of upslope
disturbance, for narrow disturbances.
Minimum recommended width is 6 metres.

Grassed buffer zones:


o

For maximum efficiency, the maximum


depth of sheet flow should not exceed the
minimum maintained grass strand length.
Grassed areas should be maintained with a
minimum grass strand length of 50 mm
(well trimmed lawns are less effective).

Construction
Refer to approved plans for location,
extent, and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel.
The buffer zone must remain undisturbed
throughout the time it is used to trap
sediment.
Runoff from the disturbance must be able
to flow to, and be dispersed as sheet flow
through the buffer zone during all stages of
construction as the grades and drainage of
the disturbance change.
Maintenance
Inspect the buffer zone after each runoff
event. Check for evidence of concentrated
flow or any flow bypassing. Take necessary
steps to correct the passage of sheet flow.
Remove excessive accumulations of
sediment that may cause the concentration
of flow.
The sources of any excessive sediment
should be investigated.

Removal
Buffer zones are usually installed as
permanent
fixtures
and
are
not
recommended for removal.

March 2010

13B-56

Appendix 13B
Fact Sheets for ESC Measures

General design notes:


o
o

Recommended for the control of sheet flow


on slopes between 1% and 10%.
Fencing can be used to exclude traffic from
buffer zones thus preventing damage to the
vegetation and surface rutting.
Buffer zones should be incorporated into the
final development landscaping plan and
should be constructed (or retained) early in
the development program.

Performance monitoring:
o

Buffer zones adjacent to high quality


watercourses and environmentally sensitive
areas should be wide enough to trap all
visible sediment within the first quarter of
the buffer zone width.
If buffer zones are to be effective, at least
75% of the ground should be covered by
vegetation and weed growth should be
controlled.

Overland flow must not be allowed to


concentrate in rills or gullies but must
remain as sheet flow through the buffer.
If these requirements cannot be met, it will
be necessary to use another kind of
sediment-trapping device, possibly in
conjunction with the buffer zone.
As clearing upslope of the buffer zone
progresses, flow diversions should be
constructed to capture and direct runoff to
the buffer zone.

Perimeter fences should be regularly


inspected for damage and repaired
immediately.
Ensure that at least 75 percent of the soil
surface is protected by groundcover
vegetation.
Where monitoring demonstrates that this is
not being achieved, install additional
sediment traps within the site and/or install
a sediment fence upslope of the buffer
zone.

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 2

Description
Grassed filter strips are narrow strips of
vegetation which have been installed to
decrease the velocity of overland flow and
facilitate settling of suspended sediments.
Application and Function
Grassed filter strips are placed around
impervious surfaces (typically roads or
pavements etc in new developments) to
control run-off from entering the stormwater
system.
Vegetation strips downslope of construction
activities can provide a simple method of
trapping sediment.
Limitations
Filter strips are only capable of trapping
coarse sediments. Clays and fine silts will
pass through a buffer zone during periods
of heavy rainfall.
Generally, an increase in strip width equals
an increase in strip effectiveness.
Alternatives
Variations include retention of existing
grass prior to construction disturbance
upslope.
In new areas subject to disturbance (kerbs,
and road verges) turf products can be
utilised.
Advantages
Easy to install
Efficient during regular storm events, and
can reduce the total sediment load
downstream.
Disadvantages
Ineffective during heavy rainfall events
Easily disturbed during construction
activities (i.e. vehicles, stockpiled materials,
and establishment of services).
Special Requirements
In high velocity areas turf should be stapled
to a topsoil layer and/or over laid on
geofabric enforcement.

Appendix 13B
Fact Sheets for ESC Measures

Grassed Filter Strip

13
B

Generally, the wider the strip the more


effective the filter strip will be in achieving a
reduction in overland flow velocity and
settling of suspended particles.
Grassed filter strips provide a similar
function to buffer zones and stiff grass
hedges.

Grass
seeding
is
not
considered
appropriate as time taken to achieve total
ground cover may be too long and is
largely dependent on prevailing weather
conditions. Adverse conditions are not
conducive to quick establishment
Reduces on-site and off-site cleanup work
after storms.

Implementation is an additional cost.

March 2010

13B-57

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Strips of turf placed along the edge of
newly formed roads, pathways and around
other
impervious
surfaces.
It
is
recommended that on all subdivisions and
urban road developments, turf strips at
least 300 mm wide and preferably 600 mm
are laid adjacent to all such surfaces.
The flatter and wider (relative to the flow
direction) the grassed filter strips are, the
more effective the filtering process.
Construction
Refer to approved plans for location, extent
and details. If there are questions or
problems with the location, extent, or
methods of installation, contact the
engineer or responsible on-site personnel.
When laying turf, care must be taken to
ensure that edges are not raised and that
adjoining strips are level. Concentration of
turf may otherwise occur.
As clearing progresses, construct any
diversions needed to capture and direct
runoff evenly through the filter strips.
Maintenance
Inspect the filter strips after each runoff
event. Check for evidence of concentrated
rill-forming flow along the upper edge of the
turf.
The source of excessive sediment should
be investigated and resolved.

Removal
Grassed filter strips are usually installed as
permanent
fixtures
and
are
not
recommended for removal.

March 2010

13B-58

Appendix 13B
Fact Sheets for ESC Measures

Care should be taken to prevent


stormwater flowing down the up-slope edge
of the filter strips rather than flowing
through the strips.
Grassed filter strips can also be used to
control sheet flow and sediment runoff on
small, steep embankments. Typically, turf
strips (minimum 300 mm width) are placed
in continuous rows along the contour, and
at a spacing of 1 to 2 metres.
Runoff from the disturbance must be able
to flow to, and be dispersed through the
filter strips, and not flow in concentration
along the upper edge of the turf. When
placed across the contour, regular diagonal
turf strips may be required to force low
flows through the turf, thus avoiding rill
erosion along the upper edge.

Additional turf strips or small earth banks


may need to be installed diagonally to the
existing filter strips to direct flow through
the turf.
Where new turf is installed watering may
be required for an initial period until the
filter strip is established.
Vehicular traffic must be prohibited on
recently installed turf as damage to the
product is otherwise likely to occur.

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 3

Description
Crushed rock pads, vibration grids or washdown areas located at vehicle exit points.
Dry systems: crushed rock pads and
vibration grids.
Wet systems: drive-through wash racks
and manual wash bays.
Application and Function
To remove soil, mud, clods, dust and
debris from tyres of vehicles leaving the
construction site.
Limitations
May only result in limited removal of
sediment from vehicles unless a wash bay,
or suitably designed drive-through wash
rack is used.
Alternatives
"Cattle" grids; metal or timber sleepers
(100-mm high at 200-mm spacing),
constructed at least 3.5 (preferably 7)
metres long -- sandy soils only.
Crushed rock pads: 75-mm rock pad
(minimum 40- mm) at least 15 metres long
and 150-mm thick.
Advantages
Automatic process (except manual wash
bays).
Vibration grid systems can be readily
moved.

Appendix 13B
Fact Sheets for ESC Measures

Construction Exit

13
B

Rock pads normally contain a cross bank


(i.e. speed bump) to precent overland flow
washing deposited sediment onto the
public road. Sediment washed from the
rock pad should be directed into an
adjacent sediment trap, usually a U-shaped
sediment fence.
To prevent the tracking of such material
onto public streets.
To minimise off-site road safety hazards.

Supplementary to vibration type systems


can be the use of drive-through wash
racks, or wash bays.
Alternatively,
use
gravelled
roads
throughout the site.

Prefabricated steel vibration grids can be


constructed then reused for several years.
Can
reduce
community
complaints
regarding the tracking of sediment onto
public roads.

Disadvantages
Requires regular maintenance (i.e. further
placement of rock).
Special Requirements
Surface water flowing to the construction
exit sediment trap must be piped under the
trap, or a perimeter bank should be
constructed to direct surface flow away
from the trap.

Wash-off from these areas should be


directed to a suitable sediment trap or
buffer zone.
Should not be located within an overland
flow path.
Avoid site exits on steep grades.
Minimise the number of site entry/exit
points.
March 2010

13B-59

Department of Transport and Main Roads


Road Drainage Manual

13
B

Overview
Location:
o
o
o

Avoid steep grades.


Avoid locations of concentrated overland
flow.
Locate the sediment control device so that
contaminated vehicles cannot bypass it
when exiting the site.
The construction exit may not necessarily be
located at the permanent site entry/exit
point.
Gravel pads, vibration grids, wash racks and
wash bays should be set back from the
public roadway to allow maximum
deposition of sediment.
Where possible, allow sediment from the exit
point to wash into the main on-site
sediment basin.
The number of site exit points should be
minimised, preferably to one.

Drainage control:
o
o

Crushed rock pads:


o
o
o

Width: 2.4 m wide per lane, 3 m per lane


desirable.
Length: 15 m long.
Material type: 75-mm crushed rock, 150-mm
minimum thickness.

Vibration grids:
o

Only suitable for sandy soils. "Cattle" grids:


metal or timber sleepers (100-mm high at
200-mm spacing), constructed at least 3.5
(preferably 7) metres long.

Wash racks and bays:


o

Drive-through wash racks should be


designed with sufficient length to allow the
wheels to be washed through at least one
complete rotation and preferably more.
Construction needs to be of concrete or
other suitable non-erodible material.
Alternatively, wash bays incorporating
manual hosing of the vehicle may be used.

All drainage from construction exits should


be directed to a suitable sediment trap
Surface water flowing to the construction exit
must be piped under the roadway, or a
perimeter bank or catch drain constructed
upslope to direct surface flow away from
the road.

Construction
Refer to approved plans for location,
extent, and details.
If the location of the construction exit is not
clearly shown on the approved plan,
determine
its
location
taking
into
consideration
o

Appendix 13B
Fact Sheets for ESC Measures

(i) The construction exit must be in place


during all phases of construction
(ii) If a large number of vehicles will use the
exit, it is advisable to have a divided
entrance that directs entering traffic through
a separate roadway parallel to the
construction exit to reduce the number of
trips over the rock pad.
(iii) Wash bay should be constructed so that
only exiting traffic passes through the bay.

Where practicable, runoff and sediment


from the site must be directed away from
the exit; choose a location for the exit that
will facilitate the diversion of any runoff to
adjacent
sediment-trapping
devices.
Alternatively, an angled cross bank or
speed bump may be installed in the rock
pad to divert runoff laterally off the pad and
into a suitable sediment trap, such as a Ushaped sediment fence.
The number of entry/exit points should be
minimised, preferably to one, especially
during wet weather.

March 2010

13B-60

If rock is to be used to line the construction


exit, clear the location of the exit, removing
stumps, roots, and other vegetation to
provide a firm foundation so that the rock is
not pressed into soft ground. Clear
sufficient width to allow passage of large
vehicles, but clear only that necessary for
the exit. Do not clear adjacent areas until
the required erosion and sediment control
devices are in place.
Grade the location of the exit so that run-off
from the exit will not flow into the street but
will flow towards an appropriate sediment
trapping device.
If the soil at the location is soft, place a
sub-base of crushed rock under the rock to
provide a firm foundation and prevent the
surface layer of rock being pressed into the
ground.
A square-edged shovel and broom with stiff
bristles must be provided at the exit for
removing any mud that may be tracked
onto the street.
If a mechanical wash unit is installed
provide a source of pressurised water to
clean tires.
If the vehicles are to be washed by manual
hosing, then a hose long enough to reach
around any vehicle leaving the site must be
connected to pressurised water source.

Department of Transport and Main Roads


Road Drainage Manual

Maintenance
If the construction exit does not function to
keep soil and debris on-site, then any soil
or debris tracked from the site must be
physically removed from these areas by
first using a shovel and broom and then by
a mechanical vacuum unit, if available. The
road shall only be washed clean after all
reasonable efforts have be taken to sweep
the sediment from the road surface.
Diversions used to direct runoff away from
the exit to sediment-trapping devices must
be maintained according to specifications
for those devices.
Removal
When the construction exit is no longer
needed, when access roads have been
stabilised and the potential for tracking soil
and debris onto the street has ceased, the
construction exit may be removed.

Appendix 13B
Fact Sheets for ESC Measures

When the rock in the exit becomes


contaminated with soil and its function is
reduced to where sediment is being
tracked into the street, a 100 mm layer of
clean, graded rock must be added and/or
the rock pad must be extended.
The construction exit must be maintained
so that mud and dust are kept on-site.
Problems must be anticipated and
preventative
maintenance
must
be
performed.

The rock and any sediment should be


removed and properly disposed of where
they will not create an erosion or siltation
hazard.
The area must then be graded and properly
stabilised.

March 2010

13B-61

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-62

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 4

Description
Special sediment fence fabric is attached to
a wire and picket fence at a maximum
height of 700mm with an additional 200mm
(min) buried and compacted into a trench
on the upslope side of the fence. The
sediment fence may also be supported
(downslope side) with straw bales.
Application and Function
Temporarily reduces the velocity of sheet
flow containing suspended sediments there
by trapping sediment.
Control of sediment runoff from exposed
land, unsealed roads and stockpiled
materials.
Limitations
Have little ability to trap fine silts
Design flows limited to around 40 litres per
second in areas of concentrated flow.
Should not be used in concentrated flow.

Alternatives
Sediment fences are manufactured from
either woven fabric or non-woven, needle
punched fabric. Woven fabrics are
preferred on large sites when the service
life exceeds 1 month during the wet
season. Nonwoven, needle punched
fabrics are preferred on small disturbances
such a building sites.
Advantages
Easy to install.
Controls sediment runoff close to the
source of the erosion.
Highly visible sediment control measure.
Disadvantages
Easily damaged by construction equipment
and stockpiles.
Can cause concentration of sheet flow if
poorly located, installed or maintained.

Appendix 13B
Fact Sheets for ESC Measures

13
B

Sediment Fence

Sediment fences should be installed along


the contour.

Sediment fences can be used at regular


intervals (depending on slope angle and
length) to limit rutting caused by
concentrated flow.

Drainage area limited to 0.6 hectares per


100 metres of fence, or a maximum slope
length of 60m.
Depending on intensity and frequency of
rainfall, service life is approximately 6
months.

Cannot introduce weeds/seeds - a problem


experienced with straw bales.
Generally more efficient than straw bales.
Limited service life of around 6 months or
less during the wet season.
Often incorrectly installed and maintained.

March 2010

13B-63

Department of Transport and Main Roads


Road Drainage Manual

13
B

Special Requirements
Limited service life of around 6 months or
less during the wet season.
Often incorrectly installed and maintained.
Overview
Materials
Use a synthetic filter fabric or a pervious
sheet of polypropylene, nylon, polyester, or
polyethylene yarn, with a minimum of 6
months of usable construction life.
Stakes should consist of 1500 sq. Mm
hardwood (min.), Or 1.5 kg/m (min.) Steel
star pickets suitable for attaching fabric.
Fabric reinforcement should consist of wire
or steel mesh minimum 14 gauge with a
maximum mesh opening of 200 mm.
Location
Their location should be determined taking
into consideration space requirements for
maintenance, grading, filling, and other
construction activities, required spacing
between the fence and adjacent buildings.
Locate the fence on relatively flat land in an
area where temporary ponding and
settlement can occur upslope of the fence.
Construction
Refer to approved plans for location,
extent, and details.
Sediment fences should be located as
shown on the approved plan or as directed
by the site supervisor.
Excavate a 200 mm wide by 200 mm deep
trench along the proposed fence line,
placing the excavated material on the
upslope side.
Along the lower side of the trench, install
the stakes securely into the undisturbed
ground. Stakes should be spaced no
greater than 3 m if supported with mesh, or
not greater than 2 m without support mesh,
or at 0.5 m in areas of minor concentrated
flow.
Construct the sediment fence from a
continuous roll, avoiding joints wherever
possible. To join fabric, attach each end to
individual stakes, holding the stakes
together, rotate the stakes 180 degrees,
then drive the two stakes into the ground.
Securely attach necessary support mesh to
the upslope side of the stakes with the
mesh extending at least 200 mm into the
excavated trench.

March 2010

13B-64

Appendix 13B
Fact Sheets for ESC Measures

Sediment fences should not be installed in


locations subject to concentrated flow.
It is permissible to place the sediment
fence along the edge of existing trees if
care is taken to protect these trees during
installation, maintenance, and removal. Do
not attach the filter fabric to the trees.
Within the region of the root zone, the toe
of the fence should not be trenched, but
should buried under a layer of aggregate.
Determine the exact location of the
sediment fence spill-through outlet before
completing installation of the sediment
fence, taking into consideration that the
outlet must be installed at the lowest
point(s) in the fence where water will pond.
where it is accessible for installation,
maintenance, and removal.

Backfill the trench and tamp the fill to firmly


anchor the bottom of the fabric and wire
mesh to prevent water from flowing under
the fence.
Securely attach the fabric to the mesh and
stakes (25 mm staples or tie wire at max.
300 mm spacing) with the fabric extended
at least 200 mm into the trench. The
complete fence should be at least 450 mm,
but not more than 700 mm high.
Just below the gap in the fence, place a
layer of filter fabric on the ground to protect
the soil from erosion by discharge from the
sediment fence spill-through outlet. Place
150 mm of the upper edge of this fabric in
the sediment fence trench. Stake the other
edges of the fabric to hold the fabric in
place.
Along the gap where the outlet is to be
installed, secure a horizontal steel or
hardwood support between adjacent stakes
at the required level to control the
maximum height of ponding upslope of the
fence.
Where necessary, construct a splash pad
or chute downslope of the spill-through weir
outlet to control soil erosion.

Department of Transport and Main Roads


Road Drainage Manual

Maintenance
Inspect the sediment fence:
o

(i) During construction: to see if machinery


or falling trees have damaged the sediment
fence; if damaged, repair it. Also check that
adjacent fill material has not accumulated
against the fence; if it has, remove the
material, repair the fence, and move the
fence or fill so that it does not happen
again.
(ii) After each rainfall event: to see that
runoff is not flowing under the fence; if it is,
bury the bottom of the fence correctly. Also
check that runoff has not topped the fence
in low points; if it has, a sediment fence
outlet structure may need to be installed.

Removal
When disturbed areas upslope of the
sediment fence are sufficiently stabilised to
restrain erosion, the fence and any outlets
must be removed.
Remove accumulated sediment and
dispose of it properly.
Remove stakes, fence and fabric, and
dispose of them properly.

Appendix 13B
Fact Sheets for ESC Measures

Clean out accumulated sediment when it


reaches a depth of 300 mm or one-half the
height of the filter fabric whichever is the
lesser.
Place the sediment in a disposal area or, if
appropriate, mix it with dry soil on the site.
Do not dispose of sediment in a manner
that will create an erosion hazard. Do not
erect a new sediment fence on top of
accumulated sediment behind the fence.
If the fence is sagging between stakes,
install additional stakes.
When making repairs, always restore the
sediment fence to its original configuration.
Stabilise the disturbed area where the
fence was located.
Biodegradable sediment fence material is
available and does not necessarily have to
be decommissioned or removed from the
site.

March 2010

13B-65

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-66

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 5

Description
Rock filter dams may be described as
oversized rock check dams.
A rock-lined spillway usually forms an
integral part of the dam.

Application and Function


Rock filter dams are free standing rock
structures used to trap sediment in welldefined gullies, excavated pits and
overland flow paths.
Typically used on temporary construction
works, while sediment weirs have a greater
use in the control of rural gully erosion.
Generally used to intercept concentrated
flow.
Limitations
Limited control over runoff turbidity and the
trapping of fine sediments, except during
low flows.
Alternatives
Alternatives include: sediment basins.
Advantages
When used on sediment basins, they are
cheaper than the traditional piped outlet
system.
Disadvantages
The gravel filter layer on the upstream face
of a dam can clog with sediment and
require regular maintenance if the structure
is used for more than one wet season.
Special Requirements
In grassed channels, small rock filter dams
should be embedded at least 200-mm into
the soil to prevent water tunnelling beneath
them.

Appendix 13B
Fact Sheets for ESC Measures

13
B

Rock Sediment Trap

Consist of a full rock embankment and


spillway constructed between two existing
earth embankments; alternatively the rock
filter dam may form part of an earth
embankment of a purpose-built sediment
basin or settling pond.
Used to form a small sediment pond where
road construction crosses a minor drainage
path.
Maybe used as the outlet structure on
small sediment basins.
Can be used as large velocity control check
dams.

Where control of upstream flow velocity is


required, refer to discussion on check
dams.

Rock filter dams require little maintenance


if used on short term construction work.
Less effective at controlling fine sediments
and turbidity than a chemically treated
sediment basin.
Access will be required for maintenance.
Fences with warning signs may be required
if public access is possible.

March 2010

13B-67

Department of Transport and Main Roads


Road Drainage Manual

13
B

Appendix 13B
Fact Sheets for ESC Measures

Overview
Rock embankments:
Min. top width = 1.5 m
Max. height at centre line of embankment
=1.5 m
Side slopes: Max. 2(H):1(V) upslope face
and 3(H):1(V) downslope face.
Rock abutments should extend at least 450
mm above the spillway.
The entire upslope face of the rock
structure should be covered with fine
crushed rock
As a guide, the full storage volume should
discharge over no less than 8 hours.
Rock fill:
Rock should be well graded, hard, erosion
resistant stone.
Minimum D50 of 225 mm and a maximum
size of 350 mm (preferred sizing).
Protection from "piping":
Filter fabric with minimum 600-mm overlap
must cover the entire foundation including
any earth abutments.
Construction
Refer to approved plans for location,
extent, and details.
Clear the area of woody vegetation and
organic matter.
Establish an access track for ongoing
maintenance. (Continued overleaf).
Cover the foundation area including the
abutments with extra-strength filter fabric
before backfilling with rock. Excavate a 600
x 600 mm cutoff trench, with 1(H):1(V) side
slopes, along the centre line of the dam,
extending for the full extent of the earth
abutments.
Apply filter fabric under the rockfill
embankment, from the upstream edge of
the dam to the downstream edge of the
apron. Overlap filter material a minimum of
600 mm at all joints.
Overfill earth embankments 150 mm to
allow for settlement.
Material used in the rock section of the
embankment (not spillway) should be wellgraded mixture of rock with a D50 size of
225 mm and a maximum of350 mm. The
minimum rock size used within the spillway
section should be that shown on the
approved plan.

March 2010

13B-68

Spillway:
Typically 1 in 2 year design discharge
hydraulic capacity.
Maximum longitudinal grade of downslope
face 3(H):1(V).
Maximum 2(H):1(V) side slope of spillway
section.
Minimum depth of the rock-lined spillway
chute is 300 mm.
Minimum thickness of rock lining is 500
mm, or twice the nominal rock size,
whichever is larger.
Outlet protection:
A rock apron at least 500 mm thick should
extend downstream from the toe of the
dam, on zero grade, to prevent channel
erosion.
Safety:
Large sediment traps are considered
dangerous because they can attract
children, thus safety aspects should be
considered. Steep side slopes should be
avoided. Fences with warning signs may
be required if public access is possible.

Construct the rock spillway section as


above
Ensure the rock spillway outlet section
extends downstream past the toe of the
embankment or at a distance equal to the
height of the dam, whichever is greater.
Direct emergency bypass flows to natural,
stable areas. Locate bypass outlets so that
flow will not damage the embankment.
Excavate the storage pit as necessary to
create the required volume. Wherever
possible, storage pits should be free
draining. Excavated pits should have a side
slope of 2(H):1(V) or flatter for safety.
Stabilise the earth embankment and all
disturbed areas above the sediment
storage area and downstream of the trap
immediately after construction.
Drainage:
Establish upslope drainage to ensure that
runoff from the disturbed area is directed
into the sediment trap.

Department of Transport and Main Roads


Road Drainage Manual

Maintenance
Check structures after each runoff event.
The effectiveness of this type of structure is
dependent upon the permeability, type and
placement of rock or other filter media. The
sides of the structure need to be carefully
inspected, refilled and compacted if
undercutting or piping occurs.
Check the structure for damage from
erosion or piping.
If the permeability is reduced to an
unacceptable level, the 300 mm layer of
upslope filter rock/aggregate should be
removed and replaced with clean
rock/aggregate. If a greater degree of
filtration is required, geotextile fabric can be
placed in front of the structure.
Removal
Following permanent stabilisation of the
upstream construction area, remove the
structure and any unstable sediment.
All water and sediment should be removed
from the basin prior to dam or weir
removal.

Appendix 13B
Fact Sheets for ESC Measures

As a guide, the basin should discharge its


full capacity in no less than 8 hours.
Periodically check the depth of the spillway
below the embankment level and check the
embankment for settlement.
Any rock displaced from the spillway must
be replaced immediately.
Remove sediment and restore original
sediment storage volume when sediment
accumulates to about 10% of the design
volume. Replace the contaminated part of
the aggregate facing.

Dispose of sediment in a manner that will


not create an erosion hazard.
Smooth the basin to blend with the
surrounding area and stabilise.

March 2010

13B-69

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-70

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 6

Description
Sediment basins consist of an excavated
pit, stabilised flow entry points, low-flow
filtered outlet and high-flow emergency
spillway.
Sediment basins require a formal design
based on catchment hydrology, the
expected sediment transport rate, and the
required trapping efficiency.
Application and Function
Sediment basins can be permanent or
temporary structures that allow ponding
and settlement of sediment-laden runoff.
Basin operation may involve chemical
dosing to improve capture of fine sediment
particles.
Limitations
Generally used on catchments greater than
1 ha.
Sediment basins do not replace other onsite control measures such as drainage
controls, sediment fences and sediment
barriers.
Alternatives
Alternatives include: rock sediment traps,
however, there are few alternatives to
chemical dosing if the control of fine
sediments and turbidity is required.
Advantages
Very effective for coarse sediment removal.
Can be an effective control of fine sediment
runoff if suitably operated.

Disadvantages
The dosing process is difficult to automate.
Basins are difficult to relocate if the
construction or drainage layout changes.

Appendix 13B
Fact Sheets for ESC Measures

13
B

Sediment Basin

Sediment basins should be fenced if settled


sediment depths exceed 300 mm at any
point in the basin and public safety is at
risk.

Basins are located upstream of water


bodies, significant bushland areas and
major stormwater systems.
The installation of a sediment basins does
not excuse poor on-site drainage and
erosion control.
Limited control over fine silts or clays if
chemical dosing is not used.
The sides of sediment basins are often
unstable and unprotected and may be a
potential sediment source.
Basins may be fully
constructed by cut and fill.

excavated

or

Generally more effective than rock


sediment traps and other types of sediment
traps.
May be converted into a permanent
wetland for stormwater treatment after the
construction phase.
Decommissioned and backfilled sediment
basins generally attract lower land values
and are best integrated into open space
areas or the sites permanent stormwater
treatment system.
March 2010

13B-71

Department of Transport and Main Roads


Road Drainage Manual

13
B

Special Requirements
Early integration into the construction
layout
Basins should be free draining wherever
possible.
Avoid construction in dispersive soils.
Overland flow inlets should be stabilised
and should not cause erosion down the
basin embankment.
Ideally should be located above the 5 yr
ARI flood level.
Discharge water into the basin in a manner
that will not cause soil erosion.
Overview
Location:
Where access for construction and
maintenance activities is available.
Adjacent to level land where possible.
Not in dispersive soils, otherwise a non
dispersive lining may be required.
Preferably above the 1 in 5 year flood level
of a major watercourse.
Dimensional constraints:
Maximum earth batter slopes 3(H):1(V)
Minimum width at top of embankment 2.5
metres
Maximum embankment height 3.0 metres
Basin length:width ratio of 3:1 minimum,
otherwise baffles are required.
Primary outlet:
Minimum conduit diameter 250 mm.
Sized for full drainage within 24 hours.
Discharge conduits located within the
embankment should be constructed to
avoid piping failure.
The area around the outlet should be
protected with rock fill or similar.
Adequate anti-floatation weights must be
securely attached to the outlet conduit.
Attach an anti-vortex device and trash
guard to the riser as required.

March 2010

13B-72

Appendix 13B
Fact Sheets for ESC Measures

Safety aspects:
Sediment basins may attract children and
can be dangerous. Avoid steep, smooth
internal slopes. Appropriately fence basins
and post warning signs if unsupervised
public access is likely and public safety is
at risk.
If public safety is a concern, and if the
basin banks are steeper than 3(H):1(V),
then at least one bank should be turfed a
width of at least two (2) metres from the top
of bank to the toe of bank to allow ease of
exit during wet weather.

Inlets:
Inlets should be appropriately stabilised
with rock or other suitable product to avoid
erosion of the basin embankments.
Outlet seepage collars:
Minimum of two cutoff collars should be
installed around all discharge conduits.
Collars spacing no greater than 15 times
the vertical projection of each collar.
Spillway:
Minimum freeboard of 750 mm.
Preferably located on natural ground rather
than compacted fill.
Structurally sound during the 1 in 20 year
ARI critical basin discharge.
Performance criteria:
In the absence of other criteria, the
recommended discharge water quality for
chemically treated Sediment Basins is 20
mg/L suspended solids.
The performance criterion for nonchemically treated basins is usually based
on the gravitational settlement of the
expected D30 runoff soil particle size.
Chemical dosing:
If the water does not meet the prescribed
standards, it must be treated with a
flocculant (eg. gypsum) to settle out
suspended solids.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

Construction
Refer to approved plans for location,
extent, and
details.
Cutoff trench:
Excavate a cutoff trench along the
centerline of the earth fill embankment, no
less than 600 mm deep. The cutoff trench
must extend into both abutments to at least
the elevation of the riser pipe crest.
Side slopes of the trench are to be no
steeper than 1:1.
Any water that accumulates in the trench
must be removed.
Embankment:
Take fill material from the approved areas
shown on the plans.
Place fill material in 150 to 250 mm
continuous layers over the entire length of
the fill area and then compact it.
Construct the embankment to an elevation
10% higher than the design height to allow
for settling. Do not use the embankment as
a dump for debris from building the settling
pool.
Maintenance
Inspect the sediment basin during
construction and after each runoff event to
determine whether any components has
been damaged. If damage has occurred,
repair it.
Inspect the basin after each runoff event to
check whether sediment has accumulated
to a level where it must be removed (top of
the
indicator
post),
and
remove
accumulated sediment when necessary
Removal
When all disturbed areas are adequately
stabilised, the sediment basin must be
removed or otherwise converted to a
permanent pond, wetland, stormwater
detention or treatment structure. In either
case sediment should be cleared and
properly disposed of and the basin area
stabilised.

Conduit outlet:
Drill dewatering holes in the riser as
specified on the plan.
Securely attach the riser to the conduit or
conduit stub to make a watertight structural
connection. Attach the anti-seep collars to
the conduit. Place the conduit and riser on
a firm, smooth foundation of impervious
soil.
Place fill material around the conduit in 100
mm layers and compact it under and
around the pipe to at least the same
density as the adjacent embankment.
Place a minimum depth of 600 mm of
lightly compacted backfill over the conduit
before crossing it with construction
equipment. Anchor the riser in place to
prevent floatation.
Attach anti-floatation weights, anti-vortex
device and trash guard to riser.
Sediment settling pool area:
Place a post or stake to indicate clearly the
depth at which accumulated sediment must
be removed. The top of the stake must
indicate the elevation of the top of the
sediment storage volume.
Do not dispose of sediment in a manner
that will create an erosion hazard.
Check all pipe connections for leaks, and
repair as necessary.
Check fill material in the dam for excessive
settlement, slumping or piping; and make
all necessary repairs.
Remove all debris from the basin and riser.

Once the area that formed the basin is


stabilised to the point where erosion is
restrained, the embankment and outlet
structures can be removed and properly
disposed.

March 2010

13B-73

13
B

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-74

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 7

Appendix 13B
Fact Sheets for ESC Measures

13
B

Sediment Weir

No picture available
Description
Sediment weirs are rock filters contained
by solid, permeable walls.
Sediment weirs may be constructed of
pervious materials such as straw bales,
washed stone, gravel-filled bags or rockfilled gabions.

Application and Function


Sediment weirs are rock filters used to trap
sediment in well-defined gullies.
Sediment weirs are typically used to control
rural gully erosion, but they can also be
used as an alternative to rock sediment
traps.
Generally used to intercept concentrated
flow.
Limitations
Limited control over runoff turbidity or the
trapping of fine sediments, except during
low flows.
Alternatives
Alternatives include: rock sediment traps,
sediment basins.
Advantages
Sediment weirs can be very effective
temporary or permanent sediment traps for
perennial streams, and are particularly
applicable where work is to be done in the
stream itself.
Disadvantages
Less effective at controlling fine sediments
and turbidity than a chemically treated
sediment basin.
Special Requirements
Any straw bales should be well stacked
and securely anchored. Gravel-filled bags
should be stacked in an interlocking
fashion. To control turbidity, the gravel
bags can be half filled with gypsum.

In gullies or large drainage channels,


sediment weirs consisting of timber and
straw, or timber and stone may be used.
These structures comprise a core of baled
straw or stone, contained within timber or
log barriers.
Sediment weirs may also be constructed of
three steel post and mesh walls, backfilled
with rock.
A modified sediment weir may be
constructed around a field inlet structure to
act as an outlet for a sediment pond. These
outlet structures may be used when a
permanent detention basin is used as a
sediment pond during construction.

Where control of upstream flow velocity is


required, refer to discussion on check
dams.
On small catchments use sediment
barriers.
In sheet flow areas, use sediment fences.

Access is required for maintenance.

March 2010

13B-75

Department of Transport and Main Roads


Road Drainage Manual

13
B

Appendix 13B
Fact Sheets for ESC Measures

Overview
Location:
Where the weir is to be built with fence
posts and steel reinforcing mesh, locate the
filter where the gully banks are high
enough so that runoff will flow through the
spillway in the centre of the weir instead of
flowing around the rock filter and over the
erodible gully banks.
When possible, where gully banks are
protected with rock or trees, which will help
stabilise the banks and hold the completed
filter in place.
Outlet protection:
A rock apron at least 500 mm thick should
extend downstream from the toe of the weir
to a distance equal to the height of the
weir.
Structure:
Maximum 600 mm spacing of steel posts.
Minimum structure thickness of 1 metre.
Minimum 1 m key into the sides of the
channel or gully.
Construction
Refer to approved plans for location,
extent, and details.
Clear the area of woody vegetation,
organic matter etc. Delay clearing the
upstream storage area until the dam or
weir is in place.
When constructed in a gully or channel, the
weir should to be cut into the sides of the
gully or channel to a minimum depth of 1
metre. Steel pickets are spaced at 600 mm
centres and mesh attached to the upstream
side of the posts. Additional "fences" are
constructed as required with the spacing
usually being determined by the width of
the bucket of the excavator.

Maintenance
Check structures after each runoff event as
effectiveness is dependent upon the
permeability, type and placement of rock or
other filter medium. The sides of the
structure need to be carefully inspected,
refilled and compacted if erosion,
undercutting or piping occurs.
Any rock displaced from the spillway must
be replaced immediately.
Trash and debris are to be removed as
necessary
Removal
After the upstream construction area has
been permanently stabilised, remove the
structure.
All water and unstable sediment should be
March 2010

13B-76

Sediment Weirs used as Sediment Basin


outlet structures:
Dimensional details are not available for
the design of Sediment Weir outlet
structures.
However,
the
following
information is supplied as a guide:
o
o

(i) Minimum base width of 1 metre.


(ii) Maximum height of rock fill should be 300
mm below the spillway crest. Wire mesh
may extend above this level to act as a
debris collector.

Wire mesh should be securely tied to the


outside of the support posts, and the whole
structure should be bound with horizontal
wire ties at a maximum spacing of 300 mm.
This will enable the mesh and sedimentladen rock to be removed from the
structure during maintenance. The mesh is
then replaced, and the outlet structure
refilled with clean rock.

Clean aggregate of varying particle size is


then placed to a depth of 400 mm. Using
plain fencing ire, lace the posts together to
improve stability. he filling of the weir is
continued until the top of the posts is
reached.
To ensure that flows are maintained in the
centre of he channel, the middle section of
the weir can be left slightly lower than the
sides.
The sides are then backfilled and
compacted as necessary.
Construct a splash pad, chute or energy
dissipater downstream of the sediment weir
outlet as necessary to control soil erosion.
Establish an access track to the sediment
weir for ongoing maintenance.
Periodically check the depth of the spillway
below the embankment level and check the
embankment for settlement.
Maintain upslope drainage so that
sediment-laden runoff does not bypass the
structure.
If the sediment weir is used as the low-flow
outlet for a sediment basin, remove
accumulated sediment when 10% of the
storage volume has been lost.
Dispose of sediment in a manner that will
not create an erosion hazard.
Smooth the ground to blend with the
surrounding area and stabilise.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13B
Fact Sheets for ESC Measures

removed from sediment storage area prior


to weir removal.

13
B

March 2010

13B-77

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-78

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Sedimentation Control 8

Appendix 13B
Fact Sheets for ESC Measures

13
B

Portable Sediment Tank

No picture available
Description
The tank contains several compartments
that trap sediment and allow chemical
dosing.

Application and Function


Sediment-laden water is filtered through a
portable tank. Treatment usually involves
the use of chemical dosing to control
turbidity.
Limitations
Limited control over silt and turbidity unless
chemical dosing is incorporated into the
process.
Alternatives
Variations include free draining surface
units or buried tanks with a pumped
outflow.
Advantages
Small size.
Disadvantages
More expensive than one-off sediment
basins, but may be cost effective if reused
on several sites.
Special Requirements
Portable sediment tanks should be located
so that trapped sediments can be readily
removed
without
interference
to
construction activities and pedestrian
traffic.
Overview
Scott & Furphy [1988] suggest a tank
storage volume of:
o

V[m3] = 20 x pumped discharge rate [L/s].

Tanks may also be sized and operated


according to the specifications supplied for
sediment ponds, sediment basins or Gross
Pollutant Traps.
Construction
N|A
Maintenance
Provisions should be made for the regular
cleaning out of sediment.

Tanks may be sized according to the


sediment basin design rules, with due
consideration given to the soil type, area of
disturbance, slope and the anticipated
rainfall.
In congested areas, the portable sediment
tank may offer a practical alternative to
sediment basins.

Alternatives include: portable (pre-cast)


Gross Pollutant Traps and sump pits.
Reusable.

Can be difficult to clean out.

Tanks may be sized according to the


sediment pond design rules, with due
consideration given to the soil type, area of
disturbance, slope and the anticipated
rainfall.

The sediment tank should be completely


cleaned when it is more than one third full
of sediment.

Removal
N|A
March 2010

13B-79

Department of Transport and Main Roads


Road Drainage Manual

13
B

March 2010

13B-80

Appendix 13B
Fact Sheets for ESC Measures

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13C
Inspection of Erosion & Sediment Controls

13
C

Appendix 13C
Inspection of
Erosion & Sediment
Controls

March 2010

Department of Transport and Main Roads


Road Drainage Manual

13
C

Appendix 13C
Inspection of Erosion & Sediment Controls

Appendix 13C Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

13C-ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13C
Inspection of Erosion & Sediment Controls

Appendix 13C
Inspection of Erosion &
Sediment Controls
Region/District:
Project Name:
Contract / Project Number:

N/A

Conforms, no action required


Non Conformation, requires action
Not applicable

Drainage Control

Item

Check

Action Required

Catch Drains
Located at top of batter faces?
Is bed stable?
Diversion Channels
Channel lining stable?
Banks are stable?
Outlet is stable?
Drop Pipes & Chutes
Inlet stabilised?
Outlet stabilised?
Sediment trap located downslope or upslope of
structure?
Is structure well anchored?
Is there any active erosion around the structure?
Geosynthetic Lined Channels
Lining well anchored on all sides & upstream edge?
Fabric overlaps in direction of flow?
Rock Lined/Rock Mattress Channels
Geosynthetic fabric is installed under the rock
Temporary Watercourse Crossing
Do culverts have a flow similar to inbank area of
watercourse?
Is soil/gravel surface well compacted and stable?
Are inlets and outlets stabilised?

March 2010

13C-1

13
C

Department of Transport and Main Roads


Road Drainage Manual

13
C

Appendix 13C
Inspection of Erosion & Sediment Controls

Velocity Control

Item

Check

Action Required

Check

Action Required

Check Dams (Rock, Gravel and Sand Bag)


Outer edges higher than centre of crest?
Erosion protection downstream?
Trapping capacity at least 60%?
Runoff diverted away from graded slopes?
Is there any bypassing of the dam?

Erosion Control

Item
Chemical Surface Stabilisers
Are dust clouds visible?
Is runoff from surface causing erosion?
Erosion Control Blankets and Mats
Are blankets and mats anchored?
Are there signs of undermining of blankets & mats?
Do blankets & mats overlap in direction of flow?
Mulching
Have exposed surfaces been covered if work on a
particular area has been delayed?
Does mulch on steep slopes require anchoring?
Can mulch be removed by concentrated flow?
Outlet Protection
Are outlets stabilised?
Is outlet protection anchored/stable enough for
concentrated flow?
Temporary Revegetation
Soil preparation as per approved schedule?
Mulch applied?
Has watering of vegetation been occurring?
Has maintenance (eg. weed removal) been
undertaken?
Graded areas seeded and mulched as per approved
schedule?
Surface Roughening
Trimmed areas with no further earthworks to be
roughened.

March 2010

13C-2

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13C
Inspection of Erosion & Sediment Controls

Sediment Control

Item

Check

Action Required

Buffer Zones
Fencing of buffer zones and No Go Zones
Is sediment contained within first quarter of buffer?
Grassed Filter Strip
Are grass strips recessed and level with upslope
surface?
Are grass strips being undermined?
Construction Exits
Entry/Exit points stabilsed?
Is sediment runoff directed away from exit point and
adequately controlled?
Is sediment visible on roads?
Does the exit contain enough coarse rock/cattle grid to
shake sediment from vehicles?
Sediment Fences
Is fabric buried in 200mm (approx.) deep trench?
Do sediment fences have at least 60% trapping
capacity?
Are stakes spaced 2m (approx.) without wire mesh
backing or 3m (approx.) with wire mesh?
Constructed along contour?
Not subject to concentrated flow?
Ends of sediment fence returned to avoid bypassing?
Located at least 2m from toe of batter/stockpile?
Straw Bales
Not subject to concentrated flow?
Recessed in 100mm (approx.) trench?
Staked with minimum of 2 stakes?
Sediment Basins
Spillway stabilised?
Adequate trapping capacity with regular sediment
removal?
Is basin accessible for maintenance?
Is exclusion fencing required for safety?
Field Inlets
Bypass flows directed to stable areas?
Is fabric trenched to avoid undercutting?
Is fabric/gravel regularly maintained to allow adequate
filtering/ponding?

March 2010

13C-3

13
C

Department of Transport and Main Roads


Road Drainage Manual

13
C

Appendix 13C
Inspection of Erosion & Sediment Controls

General Site Management


Item

Check

Action Required

Item

Check

Action Required

Sediment Control

March 2010

13C-4

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13D
Revegetation Guidelines

13
D

Appendix 13D
Revegetation
Guidelines

March 2010

Department of Main Roads


Road Drainage Manual

13
D

Appendix 13D
Revegetation Guidelines

Appendix 13D Amendments Mar 2010


Revision Register
Issue/
Rev
No.
1

March 2008

ii

Reference
Section
-

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13D
Revegetation Guidelines

Appendix 13D
Revegetation Guidelines
13D.1. Introduction
Developing and implementing a site
specific revegetation program on any site
requires careful planning, appropriate
expertise and resources. In order to achieve
this, the process of collecting and
interpreting relevant information is
fundamental to the outcome.
In the majority of cases, it is recommended
that advice be sought from suitably
experienced professionals, such as a
landscape architect, who will be required to
balance issues of: budget, maintenance
requirements,
site
constraints,
environmental values, ecological stability,
sediment control and risk of failure.
Projects differ widely in scale, design
objectives and site conditions. Meeting
specific outcomes will often require
specialist expertise and skills in soil
conservation
/science,
environmental
science, landscape revegetation, road
construction, botany, and/or zoology to aid
in the interpretation and evaluation of
information in order to achieve the desired
environmental and functional outcomes.

13D.2. Purpose
This guideline is designed to provide a
broad overview of the methodology or
framework used in developing a
revegetation program. This includes
important principles and techniques to be
considered during the planning and
development of such programs. For further
detail on these principles, techniques and

considerations refer to the Road Landscape


Manual.

13D.3. Data Collection and


Interpretation
Data collection and interpretation is the
initial step required in the planning stages
of designing a revegetation program.
Design is in turn dependant upon the
quality of information collected and
evaluated during the environmental
assessment process described in Chapter 7.
Preparation of a thorough environmental
assessment during the planning stage will
ensure much of the information required to
design and implement an effective program
is available.
Relevant environmental and engineering
criteria to be assessed during the planning
and design stages are outlined in Chapter 2
other relevant chapters. Data collection for
the
purposes
of
designing
and
implementing a revegetation program is
essentially, an extension of data gathering
for other purposes.
Utilising existing
information and information sources during
the planning and design stage of a
revegetation program will reduce the cost
of collecting relevant information and will
lead to an increase in time efficiency.

13D.3.1. Bio-physical
Characteristics
Much of the bio-physical information
required to develop a revegetation program
should already be available in the

March 2010

13D-1

13
D

Department of Main Roads


Road Drainage Manual

13
D

Appendix 13D
Revegetation Guidelines

environmental assessment documentation


for the project. If these documents have
been appropriately compiled, then the
additional work and cost involved in data
collection and evaluation of information for
a particular site should be greatly reduced.
If additional data is required (such as a
detailed botanical or soil surveys of the site
prior to construction), then collection of this
information should be organised as early as
possible within the project schedule.
Information required for each site needs to
adequately
describe
the
following
biophysical conditions:
climate;
existing vegetation;
topography;
soils;
hydrology; and
current land use and future intended
land use.
13D.3.1.1. Climate
Climatic information will provide the
designer with an indication of the seasonal
weather conditions likely to affect plant
establishment and potential erosion risk.
The designer can then incorporate measures
to maximise the chance of successful plant
establishment and minimise erosion, given
certain seasonal conditions.
Climate issues to be determined include:
seasonal
rainfall;

(monthly)

variations

in

corresponding rainfall intensity;


temperature; and
humidity.
For further considerations related to
climate, refer to the Environmental

March 2010

13D-2

Processes Manual or the Road Landscape


Manual.
13D.3.1.2. Existing vegetation
A detailed botanical survey of the site or
study area will be required for the project in
accordance with the environmentally
sensitivity of the project and the
implications for the project surrounds.
Reliance on Regional Mapping alone can
not substitute field work; however it should
be used in conjunction with the exercise. If
available, species inventories for the area
will also provide valuable information on
the site and regional flora.
Knowledge of existing and surrounding
vegetation
characteristics
prior
to
construction will enable the designer to
prepare a revegetation program compatible
with the climatic and environmental
conditions on site. Collecting data about
the existing vegetation will assist in
determining the appropriate rehabilitation
technique, species to utilise, the method of
establishment and the opportune timing for
implementation.
13D.3.1.3. Soils
The collection and assessment of topsoil
and subsoil data will assist in determining
appropriate soil correction and plant
selection, accounting for site conditions and
engineering design requirements.
The
minimum data requirements for soils are
outlined within the departments MRTS16.
A Planting Media Management Plan
(PMMP) should be prepared for all sites
where soils are disturbed.
Collected
samples should be tested and analysed by
an accredited soil agronomist, to provide
recommendation for soil amelioration.

Department of Transport and Main Roads


Road Drainage Manual

For further considerations and detail refer


to the Road Landscape Manual, MRTS16
and any user specific guidelines.
13D.3.1.4. Topography
The designer should assemble and evaluate
basic landform information. The data
should be evaluated to determine:
existing slope angles, proposed
(temporary and permanent) slope
angles and their respective lengths;
basic site drainage (temporary and
permanent), existing catchment area
and proposed catchment boundaries;
and
erosion risk due to slope.
13D.3.1.5. Hydrology
Hydrological data relevant to the site will
assist the designer when addressing
possible constraints of a revegetation
program. Relevant information includes:
flood frequency (and relevant time of
year);
the presence of springs or seeps;
water table levels/depth and quality;
and
tidal influences.
This information may assist in designing
measures into the project that help with
stability and reduced maintenance costs.
13D.3.1.6. Historic, Current Land
Use and Future Intended
Land Use
Whilst data is collected under current land
use conditions, the designer should consider
past and future land use implications when
calculating and designing drainage systems
and site stability.

Appendix 13D
Revegetation Guidelines

Some issues to consider are:


previous land use/site contamination
or artificial fertilisation;
altering existing catchment area and
the impact on adjoining land
(particularly in agricultural areas);
and
species compatibility with adjoining
landuses, particularly agricultural and
waterways.
For further considerations and detail refer
to the Road Landscape Manual.

13D.3.2. Design Objectives


It is important to know whether a
revegetation program is to be designed for
permanent cover or as a temporary
stabilisation
measure
until
further
construction works are undertaken. These
end point objectives must be clarified
prior to the development of any
revegetation program. Reducing the area
disturbed across a site reduces the costs of
revegetation, erosion and sediment control.
Staged removal of vegetation can also
provide environmental and operational
advantages.
Standards for soils, recycled water and the
minimum
installation
standards
of
revegetation techniques are setout within
MRTS16 and any user specific guidelines.
Maintenance standards and intervention
levels are setout within the Road
Maintenance
Performance
Contract
Manual.

13D.3.3. Environmentally
Sensitive Areas
The data from the environmental
assessment and related documentation will
identify environmentally sensitive sites and

March 2010

13D-3

13
D

Department of Main Roads


Road Drainage Manual

13
D

areas which require special treatment or


protection measures.
Furthermore,
depending upon the scale, location and
nature of the project, it may be a
requirement to adhere to referral agency
input from the Department of Environment
and Resource Management (DERM)
concerning
revegetation
and
soil
stabilisation or management of adjoining
vegetation.
13D.3.3.1. Species Selection
Considerations
Where rare or threatened species,
communities, or regional ecosystems occur
in the immediate vicinity of the site, a
revegetation program designed to maintain
the sites genetic integrity should be
adopted where possible. Collection of local
seed stock from adjacent vegetation
communities may need to occur up to 12
months prior to installation. It should be
noted that adjacent vegetation communities
may not always have appropriate species
that will meet the functional requirements
required, particularly for drainage devices.
It should be noted that not all species seeds
remain viable for extended periods and/or
may only germinate at a particular time of
year and/or may not be cost effective to be
installed by seed due to poor strike rates.
Not all species are commercially and/or
readily available as container stock and
may require being contract grown, if they
are required to be used within a
revegetation program.
For further considerations and detail related
to species selection refer to the Road
Landscape Manual, MRTS16 and any user
specific guidelines.

March 2010

13D-4

Appendix 13D
Revegetation Guidelines

13D.4. Design: Project


Requirements
The revegetation program developed will
be
dependant
on
conditions
and
requirements stated in the contract/brief or
stipulated by relevant referral agencies such
as the DERM.
Works within the
departments road corridor should comply
with the MRTS16 unless specified
otherwise. The design should seek to fulfil
the relevant aspects of the Road Landscape
Design Objectives as stated in the Road
Landscape Manual.

13D.4.1. Development of a
Revegetation Program
Developing a revegetation program requires
competent management that accounts for
the implications of information gathered
during the planning stage and accounts for
and incorporates any site specific
peculiarities, or constraints into the design
of the program. In addition to biophysical
constraints, agency requirements, and
contract standards, the designer must also
consider the following aspects:
project schedule;
risk;
species selection;
access considerations;
site preparation requirements; and
maintenance
programs.

and

monitoring

13D.4.1.1. Project Schedule


The designer needs to consider the timing
of the project in order to address temporary
and permanent revegetation requirements,
and the associated techniques to be
employed. The order of works needs to be

Department of Transport and Main Roads


Road Drainage Manual

planned for disturbed areas, so that an


efficient revegetation program can be
implemented over a staged basis in
conjunction with construction.
Projects which are implemented during the
wet season (high risk period) will require
additional works and soil conservation
measures to achieve a similar standard of
erosion
control
and
environmental
protection.
Erosion risk is directly
proportional to the length of time a
disturbed area is exposed. Where a high
risk is determined on the basis of climatic
conditions, temporary revegetation or soil
conservation works will be required.
Watercourses and drainage lines may need
to
be
stabilised
and
revegetated
permanently throughout the construction,
rather than implementing temporary
measures prior to final rehabilitation.
13D.4.1.2. Revegetation Techniques
A range of techniques are available to the
designer and may comprise a combination
of these treatments to achieve a desired
outcome within a revegetation program.
Table
13D.4.1.2(a)
discusses
each
techniques advantages, disadvantages and
access considerations. Table 13D.4.1.2(b)
provides a qualitative risk assessment /
suitability for each of the revegetation
strategies described in Table 13D.4.1.2 (a).
A rating of 5 suggests that the
corresponding technique is appropriate
given the potential for erosion risk is
relatively minor, consequently a broader
range of techniques can be utilised in these
instances.
For further detail and considerations on
these revegetation techniques refer to the
Road Landscape Manual, MRTS16 and any
user specific guidelines.

Appendix 13D
Revegetation Guidelines

13D.4.1.3. Risk
Achieving revegetation standards may
require significant inputs owing to the high
risks associated with the environmental
conditions on a works site. As site and
climatic conditions vary widely, the factors
affecting success will vary accordingly.
During planning and design it is important
that the consequences of not achieving the
standards required are taken into account.
The
potential
cost
implications
(replacement of treatment, and potential
damage to pavement and structures etc)
of undertaking remedial works because
of failure of a low cost revegetation
technique implemented in a high risk
area must always be considered.
13D.4.1.4. Species Selection
Use of detailed botanical survey results, in
conjunction with site specific bio-physical
information, will enable the designer to
select those species best suited to the site.
The species selected must be capable of
providing the required ground cover and
soil stabilisation measures necessary for
appropriate erosion and sediment control.
Slope length, angle, potential disturbance
and predicted water flow must be
considered when choosing plants for
specific locations, particularly within a road
drainage system.
Other considerations
include, but are not limited to: longevity,
growth rates, water logging tolerance,
drought tolerance, salinity tolerance,
humidity tolerance, fire tolerance; and
maintenance requirements. For further
detail and considerations on species
selection refer to the Road Landscape
Manual.

March 2010

13D-5

13
D

Department of Main Roads


Road Drainage Manual

13
D

Appendix 13D
Revegetation Guidelines

Table 13D.4.1.2(a) Revegetation Techniques and Access Limitations


Revegetation
Technique

Advantages

Broadcast
Seeding

Tractor drawn
broadcasting is efficient
and economical.
Grass species are
typically easily applied.

Drill Seeding

Placement of the seed


and fertiliser into the soil
assists seed germination
and plant establishment.
This method reduces
fertiliser losses and
runoff.
Seed losses due to
insects and birds theft
are also reduced.

Hydroseeding

Quick and effective


technique for seeding
flatter profiles and drains.

March 2010

13D-6

Disadvantages
Broadcast seeding does not
provide any soil surface
protection against erosion.
Different seed types (sizes) are
generally not spread evenly and
no erosion control protection is
afforded by the technique.
Some seed types (i.e. fluffy/hairy
seeds) cannot be spread by
spinners.
Not suited to high rainfall areas
with prolonged wet seasons or
slopes >1:3.
May require temporary erosion
and sediment control measures
to protect drains/culverts/
waterways until surface
coverage is obtained.
May not ensure compliance with
safety and CPTED principles,
resulting in additional
maintenance requirements
(depending on species used).
Drill seeding does not provide
any soil surface protection
against erosion.
Not suited to high rainfall areas
with prolonged wet seasons or
slopes >1:3.
May require temporary erosion
and sediment control measures
to protect drains/culverts/
waterways until surface
coverage is obtained.
May not ensure compliance with
safety and CPTED principles,
resulting in additional
maintenance requirements
(depending on species used).
Hydroseeding does not provide
any soil surface protection
against erosion.
May require additional measures
such as jute mesh.
Not suited to high rainfall areas
with prolonged wet seasons or
slopes >1:3.
Water truck access for
establishment of vegetation may
require traffic control depending
on staging of construction
works.
May not ensure compliance with
safety and CPTED principles,
resulting in additional
maintenance requirements
(depending on species used).

Access Limitation
Factors
Requires access for either
4WD tractor or for
individuals to undertake
hand seeding. Machine
broadcast seeding is
limited by machines
capacity to access steep
gradients. Slopes <1:3 are
accessible slopes, >1:3
may require 4WD
capability or manual
seeding.
Access for water
truck/maintenance needs
to be planned where
perimeter access limits
watering.

Requires access for either


4WD tractor or for
individuals to undertake
hand seeding. Machine
broadcast seeding is
limited by machines
capacity to access steep
gradients. Slopes <1:3 are
accessible slopes, >1:3
may require 4WD
capability or manual
seeding.
Access for water
truck/maintenance needs
to be planned where
perimeter access limits
watering.
Access by truck within 40
metres of the treatment
area.
Access by truck with hose
within 200 metres of the
treatment area.
Treatment is suitable for
slopes up to 1:3 (1:4 if
maintenance/mowing is
required).

Department of Transport and Main Roads


Road Drainage Manual

Revegetation
Technique

Advantages

Appendix 13D
Revegetation Guidelines

Disadvantages

Access Limitation
Factors

Hydromulching

Provides a quick and


efficient soil protective
cover (including seed)
onto otherwise
inaccessible sites and
slopes.

Hydromulches are best applied


in moist environments, and
require maintenance watering.
Light mulches (<2000kg/ha) are
unsuitable for tropical conditions
or intense storm events.
May require multiple
applications in steep terrain
areas with high rainfall.
Water truck access for
establishment of vegetation may
require traffic control depending
on staging of construction
works.
May not ensure compliance with
safety and CPTED principles,
resulting in additional
maintenance requirements
(depending on species used).

Access by truck within 40


metres of the treatment
area.
Access by truck with hose
within 200 metres of the
treatment area.
Treatment is suitable for
slopes up to 1:3 (1:4 if
maintenance/mowing is
required).

Straw Mulching

Provides a protective soil


cover, reduces
evaporation and reduces
soil temperature
fluctuations.
Relatively low cost of
materials and is usually
readily available.

The technique can bring in new


weed seed if the straw is
contaminated with agricultural
weeds. The technique is not
suitable in urban environments
with the use of air blown
emulsions and tackifiers, which
can coat adjacent facilities.
Water truck access for
establishment of vegetation may
require traffic control depending
on staging of construction
works.
May not ensure compliance with
safety and CPTED principles,
resulting in additional
maintenance requirements
(depending on species used).

Access by truck within 25


metres of the treatment
area.

Trash
Blanket/Brush
Matting

Provides a soil protective


layer insulating the soil
from temperature
fluctuation, reduces
evaporation and provides
potential source of native
seed.

The technique relies on


appropriate waste vegetation
being available.
It is labour intensive and
expensive over large areas.
Uncomposted material may
consume available soil nitrogen
as material decomposes and
raise acidity levels.
Care needs to be exercised in
ensuring that weeds and
undesirable plants are not
included in imported material.
Placement and thickness of
material needs to consider the
fuel load potential of this waste
product.
Low return (seedling
establishment) for input costs,
best suited as a soil building
treatment.

Requires access for bobcat


and foot traffic within 10
metres of treatment area.

March 2010

13D-7

13
D

Department of Main Roads


Road Drainage Manual

13
D

Revegetation
Technique

Appendix 13D
Revegetation Guidelines

Access Limitation
Factors

Advantages

Disadvantages

Erosion Control
Blankets

Best utilised where


overland flow is high and
instant protection is
required or where
dispersive materials are
encountered.

Organic
Blankets

Low water requirements


after the initial 2-3 weeks
of regular watering.
Can be sprayed onto
steep slopes.
Slope stabilisation
established quickly.

Bonded Fibre
Matrix (BMFs)

These spray-on mats do


not suffer with tenting
and can be cheaper than
erosion control blankets.
These mats can also be
sprayed to form a weed
mat which can be planted
prior to spraying.

Erosion control blankets require


very good surface preparation
and are unsuitable on rough or
rocky surfaces as mats 'tent'
leaving space for vegetation
between the mat and soil.
Erosion can occur under the
blanket.
Requires blanket to soil contact
to be successful.
Required planting or seeding for
long term success.
Can be expensive dependant on
the size of area treated.
Cost can be prohibitive,
particularly in flatter areas,
where other methods are more
cost effective.
Availability of suitable mulch.
Highly susceptible to foot traffic
and disturbance, reducing its
functional capability.
Requires careful application to
perform to the standards. Not
all hydroseeders are capable of
applying the BFM. Hydroseeder
pumps have to be specially
modified or purpose built to
operate with the high viscosity
mixes and long fibred mulches.

Container Stock

Provides an opportunity
to establish plants that
have specific
characteristics.
Planting using container
stock provides a specific
vegetation cover in an
economical and
environmentally sensitive
manner.
Can ensure species
diversity from installation.
Ensures compliance with
safety and CPTED
principles.

Planting vegetation whilst


providing an instant cover is
more costly than seeding with
most techniques listed.
Requires mulch or other
methods of ground surface
protection in most situations.
Requires establishment
watering.

Foot traffic access


required.
Additional costs where
steepness and length of
batter are effected by WHS
requirements (for example,
harnessing).

Turfing

Turfing provides an
instant vegetative cover
and erosion protection
where velocities are less
than 2m/sec.

Turf is not appropriate in many


areas owing to the species
incompatibility with the
surrounding vegetation and/or
land use.
Turf can be expensive to install
over large areas.
High maintenance and nutrient
demand.

Requires access for truck


or bobcat adjacent to
treatment area.

March 2010

13D-8

Foot traffic access


required.
Additional costs where
steepness and length of
batter are effected by WHS
requirements (for example,
harnessing).

Access within 200m from


truck.

Access for tandem truck


within 50 metres of site.
Additional costs where
steepness and length of
batter are effected by WHS
requirements (for example,
harnessing).

Department of Transport and Main Roads


Road Drainage Manual

Revegetation
Technique
Stiff Grass
Barrier

Reinforced Turf

Willowing

Appendix 13D
Revegetation Guidelines

Access Limitation
Factors

Advantages

Disadvantages

Stiff grass barriers are


very effective in repairing
eroded gullies, directing
water or acting as a weir.
This plant can handle
extremely high flows
without detriment.
Only sterile species
should be selected.
Primarily used in
drainage swales or batter
chute lining. The root
protective matting assists
the grass in maintaining
the integrity of the grass
under overland flow
conditions, where
velocities are greater
than 2m/sec.
This technique provides
some immediate
protection and develops
local provenance
species.

Initial installation of the stiff


grass barrier is labour intensive,
and therefore expensive.
Problematic in dispersive soils.

Requires access for truck


or bobcat adjacent to
treatment area.

Reinforced turfs are more


expensive than turf and require
significant amounts of
maintenance water to ensure
deep root establishment.
The varieties of farm grown turf
types are limited to specific
conditions.

Foot traffic access


required.

The technique is effective with


only a few native species and
requires good horticultural skills
to be successful.

Foot traffic access


required.
Only applicable to creek or
river embankments.

13D.4.1.5. Access Considerations


Access can be a major constraint to
implementing
and
maintaining
a
revegetation program. Access issues can
generally be avoided with appropriate
planning and design. Where access is
restricted, the revegetation technique
available for use will be limited. Table
13D.4.1.2 (a) highlights some of the issues
which may impede a revegetation program
if access to the site proves difficult.
Access During Construction
Access to a total disturbed site can be
restricted when access ways are removed
during construction of a staged process.
Issues to consider include, but are not
limited to:
the height, angle and length of
embankments or cuts and the access
to the tops of these areas;

access across boggy or water logged


areas;
access to revegetation areas where
restricted by stockpiles, equipment or
materials on site;
access in and out of site; and
traffic control implications.
Access for Maintenance
Access for maintenance purposes may be
necessary for 12 months or longer once
construction activity ceases. This may
require the creation/retention of access
ways. Where access can not be provided,
the designer needs to consider low
maintenance approaches to revegetation
and possible staging of works so that only
permanent works are installed and
maintained throughout the construction
phase.

March 2010

13D-9

13
D

Department of Main Roads


Road Drainage Manual

13
D

Appendix 13D
Revegetation Guidelines

Table 13D.4.1.2(b) Erosion Risk Suitability

Erosion Risk Suitability


Revegetation Techniques

Rating

Broadcast Seeding

1,2

Drill Seeding

1,2

Hydroseeding

1,2

Hydromulching

1,2,3

Straw Mulching

1,2,3,4 +
1,2,3

Trash Blanket/Brush Matting


Erosion Control Blankets

1,2,3,4,5 *

Organic Blankets

1,2,3,4,5 *

Bonded Fibre Matrix (BMFs)

1,2,3,4,5 *

Container Stock (Virocells, Tubestock)

1,2,3

Turfing

1,2,3
1,2,3,4

Stiff Grass Barrier

1,2,3

Reinforced Turf

1,2,3,4,5

Willowing
+ = variable depending on strength of glue used.
* = variable depending on type of product used.

Rating System
Class

1
2
3
4
5

Negligible risk of erosion


Minor risk of erosion
Moderate risk of erosion
Severe risk of erosion
Extreme risk of erosion

13D.4.1.6. Site Preparation


Choosing appropriate site preparation
techniques is fundamental to implementing
a successful revegetation program. The
method chosen will differ according to site
conditions and the chosen techniques(s) of
revegetation. Preparation techniques may
relate to works undertaken within a:

March 2010

13D-10

Environmental Management
(Construction) (MRTS51);

Plan

Vegetation
(MRTS16);

Plan

Protection

Planting Media Management Plan


(MRTS16); and
Weed, Pest and Disease Control Plan
(MRTS16).

Department of Transport and Main Roads


Road Drainage Manual

For further considerations and detail refer to


the Road Landscape Manual, MRTS16, any
user specific guidelines, and MRTS51.
13D.4.1.7. Maintenance Programs
All revegetation projects require a
maintenance program to ensure the
establishment of vegetation and the
fulfilment of their respective roles in
conservation of environmental values, soil
and amenity.
The crucial stages of
establishing vegetation are in the first two
years when seedling and plants are young
and fragile to environmental extremes.
Each revegetation program will require its
own maintenance program, tailored to the
vegetation techniques used, local climatic
conditions and consequences associated
with failure (replacement of treatment, and
potential damage to pavement and
structures etc).
Techniques to be
considered are:
irrigation;
weed control;
fertilising;
mowing;
failure replacement; and
monitoring programs.
For further considerations and detail in
relation to maintenance programs refer to
the Road Landscape Manual and the Road
Maintenance
Performance
Contract
Manual.
Irrigation
Permanent irrigation systems are not
installed within the context of a
revegetation program. Soil corrections and
appropriate plant species selection will
negate the requirement of a permanent
irrigation system. Temporary irrigation

Appendix 13D
Revegetation Guidelines

may be applied during the maintenance or


establishment periods of a revegetation
program. As a general rule, irrigation is
backed off from the initial quantities and
frequencies over a designated period of
time, to allow the vegetation to establish
and adjust to local climatic conditions and
be independent of additional watering. This
irrigation is generally undertaken by water
truck or temporary above ground sprinklers
with water supplied by a water truck. For
further considerations and detail, in relation
to irrigation, refer to the Road Landscape
Manual.
Weed Control
Weed control can be undertaken manually,
mechanically or chemically. Weed control
is essential during the establishment phase
of a revegetation program, as weeds are
often more vigorous than nominated
species, and will compete for available
nutrients, water and light.
Without
management, infestations may occur and
compromise the success of establishing
long term vegetation and stabilisation
within a site. Where a known problem
exists on site a Weed, Pest and Disease
Control Plan should be prepared to aid in
management.
For further considerations and detail in
relation to weed management refer to the
Road Landscape Manual.
Pest Control
Pest control is generally undertaken using
targeted pesticides. Without management,
infestations may occur and compromise the
success of establishing long term vegetation
and soil stabilisation. Where a known
problem exists on site a Weed, Pest and
Disease Control Plan should be prepared to
aid in management.

March 2010

13D-11

13
D

Department of Main Roads


Road Drainage Manual

13
D

For further considerations and detail in


relation to pest management refer to the
Road Landscape Manual.
Fertilising
Fertiliser application is generally not
undertaken after installation of a
revegetation program. If appropriate soil
management, installation specification and
species selection has occurred during the
planning, design and construction phase,
fertilisation will not be required post
installation. There is a high risk that
fertilising post installation will result in
high nutrient levels being leached into
waterways. If required post installation,
slow release fertilisers are preferred as they
reduce the risk of contaminating waterways.
For further considerations and detail in
relation to fertilising refer to the Road
Landscape Manual.

Appendix 13D
Revegetation Guidelines

been installed.
Replacement generally
relates to container stock plantings, but may
also be applied to seeding treatments.
Replacement will aid in achieving adequate
coverage by the desired species, which aids
in the suppression of weeds and the
stabilisation of soils.
Replacement may also be required after the
establishment and monitoring periods.
Failure to replace stabilisation plants (for
example, Vetiver grass planted within a
hedge row) can lead to concentrated flows
and result in significant damage and repair
costs, and should therefore be replaced
immediately. Replacement may also be
necessary when failure of larger plants,
which are central to a corridor or aesthetic /
planting program, occurs.
For further considerations and detail in
relation to failure replacement refer to the
Road Landscape Manual.

Mowing
Mowing frequency is dependant upon the
species of grass planted and the type of use
the area is subject to. Mowing is generally
limited to drainage areas or areas adjoining
publicly accessible spaces.
Mowing
frequency is determined by intervention
levels, and varies from urban to rural areas.
It is important to select an appropriate
species of grass in relation to the intended
use and/or function of the area, as this will
reduce the level of mowing required
(determined by intervention level) for
maintaining functionality.
For further considerations and detail in
relation to mowing refer to the Road
Landscape Manual.
Failure Replacement
Failure replacement is generally undertaken
during the establishment and monitoring
periods after a revegetation program has

March 2010

13D-12

13D.4.1.8. Monitoring Programs


A monitoring program provides important
feedback as to the type of maintenance
required, the success of techniques
previously used whether any further work is
required and what extent the program is
achieving desired environmental objectives.
By monitoring the full width of the
corridor, clues may be provided in regards
to offsite influences affecting species
diversity and drainage structures.
Monitoring programs may be one or a
combination of the following techniques:
photo monitoring;
botanical
sampling;

transects

water sampling; and


general visual assessment

and

grid

Department of Transport and Main Roads


Road Drainage Manual

Appendix 13D
Revegetation Guidelines

13
D

For further considerations and detail, in


relation to monitoring program techniques,
refer to the Road Landscape Manual.

March 2010

13D-13

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

14

Chapter 14
Operation,
Maintenance &
Remediation

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Chapter 14 Amendments Mar 2010

14

Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Table of Contents
14.1.

Introduction

14-1

14.2.

Legal Aspects

14-1

14.3.

Operation

14-2

14.4.

14.5.

14.6.

14.3.1.

Period of Inspection

14-2

14.3.2.

Performance

14-3

14.3.3.

Reporting of Deficiencies

14-3

14.3.4.

Metal Culverts

14-3

Maintenance

14-3

14.4.1.

The Maintenance Process

14-3

14.4.2.

Types of Maintenance

14-4

14.4.3.

Metal Culverts

14-4

Drainage Failures

14-4

14.5.1.

Introduction

14-4

14.5.2.

Causes of Failure

14-4

14.5.3.

Types of Failure

14-5

14.5.4.

Environmental Impacts of Failures

14-7

14.5.5.

Identifying Failures

14-7

14.5.6.

Reporting of Failures

Remediation

14-10

14-10

14.6.1.

Introduction

14-10

14.6.2.

Remediation Options

14-10

14.6.3.

Evaluation

14-13

14.

March 2010

iii

14

Department of Transport and Main Roads


Road Drainage Manual

14

March 2010

iv

Chapter 14
Operation, Maintenance & Remediation

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Chapter 14
Operation, Maintenance &
Remediation
14.1. Introduction

14.2. Legal Aspects

Appropriate maintenance of drainage


infrastructure plays a crucial part in its
effective operation. This also minimises
environmental harm and provides a level of
safety to users of the road corridor.

Applicable requirements of key legislation


such as the Workplace Health and Safety
Act 1995, Environmental Protection Act
1994 and Environment Protection and
Biodiversity Conservation Act 1999
(Commonwealth Act) apply to the
departments operation and maintenance
activities with respect to road drainage
(refer Section 1.4).
Furthermore, the
department has a legal responsibility / duty
of care to ensure that the road under its
jurisdiction is maintained to provide an
acceptable level of safety to the public, road
users and to protect the environment from
harm. It is important that supervisory staff
overseeing these activities understand the
applicable requirements of the legislation to
ensure compliance.

This chapter of the manual is targeted


primarily at managers of road maintenance
operations
and
road
maintenance
contractors. It outlines the maintenance
process and uses examples of drainage
failures to illustrate the need for effective
maintenance operations. It also provides
steps for the remediation of problems or
deficiencies.
The process outlined in this chapter relies
on the design process undertaken in
previous chapters and reference to the
design criteria, assumptions, calculations
and assessments may be required.
Reference should also be made to the Road
Maintenance
Performance
Contracts
Manuals, Volumes 1 to 3 and Part 8 of the
Asset Maintenance Guidelines (DMR
2002a).
The philosophy of this chapter is to use the
maintenance process for identifying failures
in the drainage system and to assist learning
from these failures to prevent future
failures.

To ensure appropriate and timely


maintenance, it is important that regular
inspections, followed by appropriate
remediation works (where required), be
conducted. With respect to road drainage,
it is recommended that inspection of this
infrastructure should be conducted shortly
after significant rainfall / flood events when
failures are more likely to occur. Any
remediation work would depend on the
severity of any damage / failure identified.
Furthermore, the department must also
ensure prompt response to emergency
situations (such as water over the road or
subsidence of the roadway occurs) where
rapid remediation works are required.
March 2010

14-1

14

Department of Transport and Main Roads


Road Drainage Manual

In both situations presented above, failure


to act appropriately exposes the department,
and its officers, to increased risk of
investigation and/or legal action.

14

As an example, the following case study


outlines some details regarding an inquest,
held in NSW, where a Coroner found
Gosford City Councils negligent road
maintenance responsible for the deaths of a
family of five.
On 8 June 2007, a family drowned when
their vehicle plunged into the flooded Piles
Creek during heavy storms. The accident
occurred because a section of the Old
Pacific Highway had collapsed due to the
failure of a steel culvert.
During the inquest, the court was told that
the council knew the highway culvert had
been rusting for several years but did
nothing to repair it. The court was also told
that the council could have reinforced the
rusted culvert with concrete for $277,000
but the engineering advice got lost in the
system.
In his findings, the Coroner was quoted as
saying, I am satisfied that had Gosford
City Council dealt with the information it
had available to it concerning the Piles
Creek culvert in a competent, professional
and timely manner, the tragedy that
occurred on 8 June 2007 would not have
occurred.
The Coroner also recommended that the
NSW Local Government Minister (at that
time) initiate an independent review into the
councils operations and the competency of
its officers.

14.3. Operation
The operation of the road and inturn the
drainage system commences immediately
after the road is opened to traffic. This
March 2010

14-2

Chapter 14
Operation, Maintenance & Remediation

section deals with the period immediately


after construction while Section 14.4
discusses the ongoing maintenance
considerations and activities which span the
life of the road.
An important function or activity that
should be conducted in the period after
construction (refer Section 14.3.1) is the
inspection / check of the drainage systems
actual operation or performance against the
design intent. This can only happen after a
reasonable rainfall / storm event and will
either validate the design or identify
deficiencies. This performance check is
particularly important for drainage devices
protecting / maintaining water quality (refer
Section 14.3.2).
Depending on the
deficiency, remedial works (refer Section
14.6) may be covered under the defects
liability component of the construction
project.
It is important to note that the inspection
period for the site should be extended to
check the performance of any remedial
work.

14.3.1. Period of Inspection


It is difficult to define an actual time period
for inspection(s) / check(s) to be
undertaken, as rainfall events are
unpredictable and therefore it is impractical
to define a set time period. As stated in the
previous paragraph, the purpose of the
inspection is to check the operation or
performance of the drainage system and this
requires at least one significant storm or
rainfall event to test the system. It is
recommended that several rainfalls events
should be monitored / checked to ensure the
on-going, successful operation of the road
drainage system.

Department of Transport and Main Roads


Road Drainage Manual

It is also recommended that inspections


should occur as soon as possible after a
rainfall event so that actual stormwater
flows can be observed / tested.

14.3.2. Performance
Drainage infrastructure is constructed to
primarily carry / transfer and possibly treat
stormwater. These devices are designed for
a certain discharge and/or capability. With
respect to drainage devices constructed to
protect and/or maintain water quality, they
have been designed to meet specific water
quality requirements. It is important that
the performance of these devices be
checked to ensure that the devices are
achieving the design requirements. If it is
found that the requirements are not being
met, the site must be:
fully investigated to determine the
reasons why the device is or devices
are not achieving required targets;
appropriate remedial action planned /
designed to correct deficiency; and
remedial work undertaken.
This investigation and remedial work must
be undertaken as soon as possible after the
deficiency has been identified as the risk of
causing harm or damage will remain
elevated until the work is completed.

14.3.3. Reporting of Deficiencies


It is important that deficiencies in the
drainage system (when compared to design
intent) be identified, investigated and
reported. Reports must identify the reason
or cause for the deficiency.
While reports are processed and subsequent
remedial works actioned by regional /
district offices, it is requested that copies of
reports and determined remedial action be

Chapter 14
Operation, Maintenance & Remediation

forwarded to Principal Engineer (Road


Design Standards), Road Planning &
Design Section, Engineering & Technology
Division. This will enable a statewide
review of drainage failure / deficiencies
which may lead to changes in design /
construction
methodology,
and/or
departmental documents and policy.

14.3.4. Metal Culverts


The operation and inspection of metal
culverts is more specialised than other
culvert types, therefore inspectors and
maintenance engineers are referred to
Appendix 14A for further guidance.

14.4. Maintenance
14.4.1. The Maintenance Process
The Department of Transport and Main
Roads is the steward of the state controlled
road network. Part of this role is to
maintain the road network to a standard
which ensures the safety and efficiency of
the travelling public and protection of the
environment.
Road drainage infrastructure (or system) is
designed for a certain discharge and/or
capability and needs to be properly
maintained
to
ensure
continued
performance. Poor maintenance reduces
the performance (capacity) of the drainage
device or system and this inturn can
increase the risk of:
upstream flooding;
failure of the device / system and
potentially the road;
accelerated deterioration of the road
asset;
accidents (such as unexpected water
on the road surface); and

March 2010

14-3

14

Department of Transport and Main Roads


Road Drainage Manual

damage to the environment.

14

Maintenance works on the state controlled


road network is undertaken in accordance
with the departments Road Maintenance
Performance Contracts (RMPC).

14.4.2. Types of Maintenance


With respect to drainage the RMPC covers
predominantly two maintenance types. The
first, and most dominate type, is routine
maintenance.

Chapter 14
Operation, Maintenance & Remediation

are all common failures. Figure 14.5.1


illustrates a drainage failure that has
occurred during construction.
While flooding of the road corridor is the
most common problem as a result of
insufficient drainage capacity or a blockage
of cross drainage, there are other issues that
can occur as a result of failure (reduced
performance / capability).

Routine maintenance work includes those


activities that keep the road corridor in
good order, such as the cleaning and repair
of drainage systems (refer RMPC Vol. 3
(DMR 2004a)).
Emergency maintenance work relates
primarily to work performed immediately
following an emergency (e.g. vehicle
accident, natural event) to ensure the safety
of motorists and/or pedestrians using the
corridor. Other routine maintenance work
may be necessary after making the situation
safe.

14.4.3. Metal Culverts


The maintenance of metal culverts can be
difficult
and
specialised,
therefore
maintenance engineers are referred to
Appendix 14A for further guidance.

14.5. Drainage Failures

Figure 14.5.1 - Drainage Failure.

14.5.2. Causes of Failure


Failures in the drainage system or device /
component may result from a number of
situations. Some of the more prominent
causes of failure are:
inadequate / inappropriate design;
poor construction practice / post
construction inspection;
changes to the sites physical
conditions
(e.g.
alteration
to
landform,
vegetation,
and
surrounding land use).

14.5.1. Introduction

poor maintenance
practice;

A failure in road drainage may be caused by


any number of problems or combination of
problems and can occur during the
construction or operation of the road.
Erosion at culvert outlets, undermining of
pavement and drainage structures, soil loss
on steep batters and sedimentation of drains

an extreme rainfall event which


delivers a storm much greater than
the design storm the drainage
infrastructure was designed to
handle; and

March 2010

14-4

inspection

out of specification materials.

Department of Transport and Main Roads


Road Drainage Manual

Failure may also be caused by site


maintenance operations. For example, the
re-grading or re-cutting of table drains (to
clean out silt and so on) may accidentally
knock out table drain blocks. These devices
are typically used on downgrades to dam
stormwater flow and direct it into culverts
to carry the stormwater across the road. If
these blocks are removed, the stormwater
will bypass the culvert and continue
downgrade until it reaches the last or lowest
point culvert. This culvert will now be
expected to carry much more stormwater
than it was designed for which could result
in higher outlet velocities or even water
overtopping the road.
Where water
overtops the road, the risk of an accident
occurring is greatly increased.

Chapter 14
Operation, Maintenance & Remediation

sedimentation;
debris accumulation (e.g. litter and
vegetation); and
structural failure of the drainage
device / component.
These are discussed below in further detail.
14.5.3.1.

Erosion

Erosion is the most common failure. Roads


tend to concentrate stormwater flows which
inturn increases flow velocities and energy.
This combination increases the risk of
erosion and scour. Figure 14.5.3.1 shows
an example of erosion.

Another example could be where a slasher,


mowing a narrow grassed footpath at the
toe of a steep batter, damages the batter toe.
During a storm event, runoff over the
exposed batter face could lead to erosion
and undermining of the batter toe which
could lead to slippage of the batter slope.
An actual or potential failure in the road
drainage infrastructure is often not evident
until it has been subject to storm events or
similar conditions.
For example, the
scouring of a creek bed downstream of a
culvert outlet may not occur or be evident
until the first storm event.
It is recognised that maintenance of the
drainage infrastructure plays a crucial part
in its effective operation.

Figure 14.5.3.1 - Erosion of a Highly


Dispersive Soil

Erosion can occur at or in:


culverts (inlets and outlets);
bridges;
floodways;
diversion channels;
catch banks and drains;
table drains; and

14.5.3. Types of Failure


A number of types of failures are evident in
the road drainage system.
The more
common failures include:
surface / slope / bank erosion;
undermining or piping;

unsealed
batters.
14.5.3.2.

shoulders,

verges

and

Undermining or Piping

Undermining refers to the loss of soil from


underneath some part of the road

March 2010

14-5

14

Department of Transport and Main Roads


Road Drainage Manual

14

infrastructure (e.g. pavement surface,


concrete lined drain, culvert apron). This
can result in direct damage to the road
infrastructure such as cracking or slumping.
Figure 14.5.3.2 shows and example of
undermining.

Chapter 14
Operation, Maintenance & Remediation

shows an example of sediment deposition


in a culvert. This failure is also termed
blockage and reduces the capacity of the
culvert, which inturn can increase flooding
(afflux) upstream.

Figure 14.5.3.3 - Sedimentation


Figure 14.5.3.2 - Undermining
(Dispersive soil)

Piping is the term used to describe the


mode of embankment failure that involves
the washing out of the smaller soil particles
from a section of the road embankment by
water leaking through a weak point in its
structure. The weak point can occur either
from the side on the embankment, through
the top surface of the embankment via
cracks in surfacing or unsealed shoulders or
through gaps in disjointed culverts. The
progressive removal of fine soil particles
will further increase the rate of flow of
water and the rate of removal of the soil
particles.
As the rate of leakage
accelerates, larger soil particles can be
transported. Eventually localised collapse
of the embankment will occur.
14.5.3.3.

14-6

Debris Accumulation

Debris
accumulation
includes
the
accumulation of vegetation, litter and other
gross pollutants in the drainage system.
This may be caused by insufficient
hydraulic capacity of the drainage structure,
the size of debris entering the drainage
system or lack of maintenance. Figure
14.5.3.4 shows debris accumulation via the
growth of vegetation in the outlet of a
culvert.

Sedimentation

Sedimentation is another common cause of


failure and is the deposition of soil that has
been transported by flowing water. Soil
particles settle once the flowing water has
slowed or stopped. This often occurs in
culvert inlets and outlets as well as creeks
and other watercourses. Figure 14.5.3.3
March 2010

14.5.3.4.

Figure 14.5.3.4 - Blockage by Vegetation

14.5.3.5.

Structural

Structural failure is the failure of a drainage


structure either by separation of units
making up a single structure (for example,

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

the disjointing of culvert pipes or box units


or the headwall separating from the barrel)
or the actual structural failure of a unit (for
example, the collapse of a concrete pipe
unit due to excessive loading or the collapse
of a steel culvert due to weakening by rust).

14

14.5.4. Environmental Impacts of


Failures
Drainage failures have the potential to
cause environmental harm. Some common
impacts include:
(a) Disruption to vegetation (direct
terrestrial and aquatic habitat loss);
(b) Soil erosion and sedimentation;
(c) Altered stream hydrology;
(d) Altered overland flow paths causing
soil moisture changes; and
(e) Weed invasion.

14.5.5. Identifying Failures


As discussed, routine maintenance of the
road corridor is generally carried out by a
maintenance contractor.
Part of the
contractors role is to undertake regular
surveys and inspections of the corridor to
identify and prioritise maintenance works.
Significant failures, such as that shown in
Figure 14.5.5, need to be identified quickly
so that appropriate remedial works can be
determined, approved and undertaken.

Figure 14.5.5 - Culvert Headwall


Scouring

The inspection and reporting process of


failures could form part of the Maintenance
Contractors Quality Plan. Any failure
reports could also be submitted as part of
the contractors quarterly progress report, if
required.
Reports should include / detail:
the identified drainage failure(s);
relevant field notes / photographs;
and
identification
of
remediation types.

potential

This information can then be used to assist


in the development of an appropriate
remediation solution.
Following the completion of the report, the
maintenance contractor, in consultation
with the department, should prioritise
remediation
works
and/or
report
maintenance requirements.
Descriptions of remediation options are
provided in the following section.
Table 14.5.5 summarises the key drainage
failures, causes and the potential resultant
impacts experienced with Queensland
roads.

March 2010

14-7

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Table 14.5.5 - Common Drainage Failures, Causes and Impacts in Queensland


Failure
Type

Failure

Cause

Examples of Potential
Impact

Erosion

Scouring at culvert
outlet

Lack of downstream energy


dissipation/scour protection.

Sedimentation of
downstream waterways.

High velocity through culvert.

Decrease in water
quality.

14

Underestimation of catchment size.


Scour of bridge piers

Obstruction to flow.

Damage to bridge.

Shape of piers (poor design).

Scour of stream.

Poor construction method.

Compromise of public
safety.

Concentration of floodplain flow.


Velocities exceed threshold values.
Slips on cuttings

Erosion of creek
banks

Cutting too steep.

Blockage of road.

Poor drainage.

Loss of soil.

Skewed culverts discharging high


flows into creek bank.

Loss of riparian habitat


for semi-aquatic wildlife
such as Platypus, Water
Rat.

Bridge piers constructed on creek


banks.
Erosion of road
embankments/
batters

Flood overtopping road and


inadequate batter protection.

Weed invasion.

No catch drains provided.

Undermining of road
surface.

Batters too steep for revegetation.

Blockage of table drains.

Lack of permanent erosion protection.


Pavement overlay often raised and
shoulders left at lower level causing
runoff to drop onto exposed
embankment.
Excessive mowing/slashing of road
shoulders.
Scouring of sodic
soils

Scouring around
culvert
headwalls.(Refer
Figure 14.3.5)

Concentration of runoff.
Lack of controlled drainage.

Sedimentation of
downstream waterways.

Lack of surface protection.

Decrease in water
quality.

Runoff drains behind headwalls and


causes scour.

Sedimentation of
downstream waterways.

Lack of vegetation cover owing to


weed spraying.

Decrease in water
quality.

Insufficient culvert lengths resulting in


steep embankments above headwall.
Scouring of
channels and table
drains

Inadequate scour protection.


Maintenance activities stripping table
drains of vegetation.
Channels and drains excavated to
carry much larger capacities.

March 2010

14-8

Loss of vegetation and


fauna habitat.
Loss of vegetation for
water quality filtering
purposes.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Failure
Type

Failure

Cause

Examples of Potential
Impact

Erosion

Scouring of Bridge
Abutments

Blockage of bridge waterway.

Sedimentation of
downstream waterways.

Design underestimates degree of


construction disturbance around
abutment thus resulting in exposed
areas subject to erosion.

Decrease in water quality.

Abutment surfaces difficult to


revegetate.
Lack of controlled drainage.
Erosion at cut batter
toe

Undermining

Poor maintenance practices such as


over spraying for weeds and
undercutting by mowers and slashers.

Changes in table drain


flow velocity.

Erosion at cut and fill


transitions

Lack of sub surface drainage.

Sedimentation of
downstream waterways

Undermining of
culvert aprons

Highly dispersive soils with minimal


protection.

Loss of drainage
structure.

Erosion at Outlet.

Further scour likely to


occur within waterway.

Sedimentation of
downstream waterways.

Transport of sediment to
downstream locations.
Undermining of rock
gabions or
mattresses

Design of rock gabion or mattress


structures on highly dispersive soils.

Loss of drainage
structure.

Geofabric filter not placed under rock


mattresses.

Further scour likely to


occur within waterway.
Transport of sediment to
downstream locations.

Undercutting of
batter

Maintenance practices such as


slashing/mowing undercutting batter.

Changes in table drain


flow velocity.
Sedimentation of
downstream waterways.

Sedime
ntation

Debris
Accum
ulation

Sediment. deposition
in table drains and
channels and
culverts

Insufficient gradient.

Debris or litter buildup at culvert inlets or


bridge waterways

Lack of regular maintenance.

Flooding of upslope drain.

Vegetation growth at culvert inlet.

Road blockage from


flooding and hazard to
pedestrians and motorists.

Ponded outlet.

Changes in soil storage


capacity.
Vegetation growth at
culverts inlet/outlet.

Culverts or bridge spans inadequate in


size.

Unacceptable afflux.
Scouring at Bridge
Abutments and/or road
approaches.

March 2010

14-9

14

Department of Transport and Main Roads


Road Drainage Manual

14.5.6. Reporting of Failures

14

It is important that failures of drainage


devices
or
system
be
identified,
investigated and reported. Reports must
identify the reason or cause for the failure.
While reports are processed and subsequent
remedial works actioned by regional /
district offices, it is requested that copies of
reports and determined remedial action be
forwarded to Principal Engineer (Road
Design Standards), Road Planning &
Design Section, Engineering & Technology
Division. This will enable a statewide
review of drainage failure / deficiencies
which may lead to changes in design /
construction
methodology,
and/or
departmental documents and policy.

14.6. Remediation
14.6.1. Introduction
The purpose of remediation is to restore the
drainage device or system to the level of
performance / capability it was designed
for.
After inspection / checks have been carried
out and deficiencies and/or failures have
been
identified
and
investigated,
appropriate remedial work options need to
be developed and the best solution
determined.

14.6.2. Remediation Options


The remediation option may include one or
a number of the following:
One-off repair work;
Augmentation of existing works ;
Re-design
and/or

(and

re-construction);

Alteration to maintenance practices.


March 2010

14-10

Chapter 14
Operation, Maintenance & Remediation

Field inspection, to review site conditions,


failures and remediation options, should be
undertaken to confirm the most appropriate
option / solution.
14.6.2.1.

Repair

A failure may be attributed to a one-off


event such as a severe or extreme storm
event, vehicle accident, or vandalism. Such
instances may require repair work. General
wear and tear on the drainage system will
also require repair.
Volume 3 of the RMPC (DMR 2004a)
provides guidelines for the repair of
drainage works. These guidelines can be
used by the contractor in preparing the
Contractors Quality Plan work procedures.
Repair work should provide for the
drainage system to operate to its intended
design function and capacity. Such repair
work would require that the failure does not
continue.
Changes to routine maintenance activities
may be required following repair works.
These changes in maintenance activities
would need to be specific to the repair
works undertaken.
The maintenance
contractors routine maintenance practices
should be amended as necessary to
accommodate these changes.
14.6.2.2.

Augmentation

Augmentation or retrofitting of existing


drainage works may be required to avoid
the continuation of an existing failure.
A review of the site conditions that
influence the operation of the drainage
structure is required. This may include a
review of:
(a) Flow velocity;
(b) Soil type / erosion risk rating;

Department of Transport and Main Roads


Road Drainage Manual

(c) Catchment area; and


(d) Receiving environment.
The type of augmentation works required
will depend on the failure identified. If
formal design of augmentation works is
required, reference should be made to the
relevant design procedures in Chapter 7.
14.6.2.3.

Re-Design

Where simple repair or augmentation works


are not likely to or will not correct the
problem, a full review of the failure and
original design intent followed by the redesign process will be required to develop a
suitable solution. It is important to review
the original design to identify possible
deficiencies in the design (which lead to
failure) and to ensure any re-design work is
compatible with the rest of the drainage
system. Reference should be made to the
relevant design procedures in Chapter 7 for
any re-design work.
The focus of re-design is to ensure that the
new design of a drainage structure avoids a
recurrence of the drainage failure caused by
the same factors.
14.6.2.4.

Maintenance Practice

Where an existing maintenance practice or


activity is causing a drainage failure, it
should be reviewed and modified. New
practices may also need to be devised and
implemented.

Chapter 14
Operation, Maintenance & Remediation

(b) Weed spraying resulting in vegetation


loss and subsequent soil erosion;
(c) Clearing of vegetation from headwalls
and at culvert outlets resulting in soil
exposure and subsequent erosion; and
(d) Machinery hitting and damaging
drainage infrastructure which is hidden
from view by vegetation and so on.
Damage could be structural or exposure
of backfill material supporting the
structure (then subject to erosion).

Figure 14.6.2.4 - Grassed Table Drain


Damage

If a failure is attributed to maintenance


activities, a review of existing practices and
implementation of new practices should be
undertaken. Table 14.6.2.4 and Appendix 3
of RMPC Volume 3 (DMR 2004a) should
be reviewed in relation to appropriate
maintenance practices.

Some common failures of the drainage


system resulting from maintenance
activities include:
(a) Over-clearing or mowing of grassed
table drains (refer Figure 14.6.2.4),
batter toes and other areas resulting in
soil exposure and subsequent erosion;

March 2010

14-11

14

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

Table 14.6.2.4 - Summary of Sound Maintenance Practice

14

Drainage
Failure

Required
Maintenance

Best Practice Guidelines

Erosion

Backfilling

Determine cause of erosion (e.g. concentration of high


velocity flow).

Mulching
Revegetation
Reprofiling

Identify actual and potential sediment sources during


routine maintenance.
Plan to protect exposed areas and sediment sources as
soon as practicable.
Minimise the amount of disturbance to vegetated and other
sensitive (e.g. waterway) areas.
Minimise the amount of disturbance to unstable areas (e.g.
drains, banks, batters)
Plant only indigenous species native to the area.

Undermining

Backfilling
Installation of
protection measures
(e.g. rock, concrete)

Sedimentation

Hand removal of
sediment
Mechanical removal
of sediment
Disposal of sediment

Minimise the amount of disturbance to unstable areas (e.g.


drains, banks, batters).
Simulate natural conditions where practicable.

Remove and dispose of sediment in an area where it is


unlikely to wash into a drainage line or waterway (i.e. flat area
as far away as practicable from the nearest drainage line).
De-water sediment prior to disposal to reduce the potential
for sediment-laden runoff draining from the disposal site.
Identify any potential for sediment to be contaminated from
industrial or other sources. Where sediment is likely to be
contaminated, contact should be made with the local
authority to determine the most suitable disposal location and
method of transport.
Identify opportunities for re-use of uncontaminated sediment
(e.g. top dress landscaping, use in backfilling).
Clean drains in a manner that does not result in soil
disturbance or exposure to erosion.

Debris
Accumulation

Hand removal of
debris
Mechanical removal
of debris
Disposal of debris
material

Ensure that any potential changes or inadequacies in


hydraulic capacity of the drain, culvert inlet or similar are
identified and managed to minimise potential environmental
harm.
Ensure access is sufficient to allow the full removal of all
debris
Continual debris build-up may indicate culverts or bridge
spans of insufficient size.

Weed
Removal

March 2010

14-12

Removal of
vegetation

Ensure control agents are appropriate for the purpose and


that only the necessary rates of application are used.

Herbicide application

Ensure that chemical spill containment and clean-up


equipment is available when chemicals are being used and
staff is trained in the use of such equipment.

Department of Transport and Main Roads


Road Drainage Manual

Chapter 14
Operation, Maintenance & Remediation

14.6.3. Evaluation
In order to evaluate the effectiveness of
maintenance practices, and reduce the
potential for future drainage failures,
information
recorded
during
the
maintenance process should be reviewed.
This information can then be used for the
planning,
design,
construction
and
maintenance of future drainage systems.

14

As part of RMPC, the contractor is required


to measure and record all maintenance work
undertaken. This is undertaken to provide a
basis for progress payments.
This
information provides an insight into the
type and frequency of failures, the
frequency of maintenance works and types
of remediation.
To supplement this information, the
following would also assist in evaluating
maintenance activities and remediation
techniques:
(a) Observations and records of recurring
problems and failures and any resultant
environmental harm; and
(b) Practicable and effective remediation
options to suit site specific failures.
Review of the above information should be
undertaken by road drainage planners and
designers to minimise in the future the
potential for recurring failures in the
drainage system.

March 2010

14-13

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Appendix 14A
Management of
Corroded Metal
Culverts

January 2013

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

Appendix 14A Amendments Jan 2013

14
A

Revision Register
Issue/
Rev
No.

Reference
Section

Whole

January 2013

ii

Description of Revision

Authorised
by

Initial Release of 2nd Ed of manual.

Steering
Committee

Change of title better reflects intention of


appendix

M
Whitehead

Date

Mar
2010
Jan
2013

Department of Transport and Main Roads


Road Drainage Manual

14A.1 Introduction
14A.1.1 Purpose
This appendix is intended to be a guide for
the inspection and treatment of corrugated
metal culverts. It describes the information
that should be recorded, identifies typical
situations, the associated risk to the
travelling public and the urgent action that
may be required to provide a safe and
efficient road network.
This Appendix has been written by the
departments Bridge Design Branch located
within the Engineering & Technology
Group. Figure 14A.1.1 provides a flow
chart summarising the information given in
this appendix.

14A.1.2 Background
The Department of Transport and Main
Roads has installed many corrugated metal
culverts made from galvanised steel during
the 1970s and 1980s. Many of these
culverts now show significant corrosion,
and several have structurally failed. One of
these failures caused the closure of the
Bruce Highway south of Gin Gin a few
years ago. Another culvert failure in New
South Wales, where a family of five
drowned when their car was swept away,
was subject to a coroner's enquiry which
found the local authority was responsible
for their deaths by failing to maintain the
corrugated metal culvert. This enquiry has
highlighted the severity of the problem
associated with metal culverts, and the
importance
of
rehabilitating
and
maintaining culverts.

Appendix 14A
Management of Corroded Metal Culverts

Managing the Risk (approval date


21/02/2008) which describes the failure
mechanisms of the culverts and information
to be gathered during the inspection
process. It also states that any steel culvert
with a vertical opening in excess of 1.2 m
should be added to the inspection
programme.

14A.2 Inspection
Inspections should be conducted in
accordance with the Bridge Inspection
Manual and appropriately released Advice
Notes. If the structure is rated in Condition
State 4, a Level 3 Detailed Structural
Engineering Inspection should be requested
with the Principal Engineer of BAM.

14A.2.1 Information
Gathered

to

be

When inspecting culverts onsite, the


following information, as detailed in
Sections 14A.2.1.1 to 14A.2.1.9, should be
collected to aid in selecting a suitable
design process and repair method.

14A.2.1.1 Type of Culvert


Determine whether the culvert is one of the
two main types of metal culverts, Multiplate and Helical Corrugated Culverts that
have been used on Queensland roads.
Multi-plate Culverts
These were the first type of corrugated
metal culverts used in Queensland and were
constructed by hand bolting a number of
circular plate segments together to form a
ring (refer to Figure 14A.2.1.1(a)). Metal
thickness was 4 5 mm or more.

In response to the risk imposed by these


structures, the departments Bridge Asset
Management Branch (BAM) has issued an
Advice Note 88: Corrugated Steel Culverts

January 2013

14A-1

14
A

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

Identify
Metal Culvert
Gather Basic Information
- Culvert Type
- Size and Shape
- Corrugations
- Height of Fill
- Material Thickness
- Extent of Corrosion
- Max possible OD (if relining)
- Voids Present in Fill
- Other Defects

14
A

Determine Typical Situation

Situation 1
- Standing water
in culvert

- Allow culvert to drain


- Assess culvert condition
when dry

Very Low
Risk

- Plan to install concrete invert


or paint system
- Repair within 2 years

Low
Risk

- Inspect culvert after rain for erosion


of back fill through invert perforations
- Install concrete invert within 1 year

Medium
Risk

-Prop and repair ASAP


-Check road surface levels fortnightly, after
short or during long rainfall events
-Speed restriction and hazard signs

High
Risk

- Prop and repair ASAP


- Check road surface levels weekly, after
short or during long rainfall events
- Speed restriction and hazard signs

Very High
Risk

- Prop immediately
-Repair urgently
- Speed restriction and hazard signs

Situation 2
- Corrosion appearing
in the invert
- Substantial metal
thickness remains

Situation 3
- Considerable level
of corrosion in the invert
- Small holes appearing

Treatment

Treatment

Situation 4
- Culvert Invert rusted
through in large sections.
- Culvert remaining circular
-No significant flows
- Constant HV traffic volume

Treatment

Situation 5
- Culvert Invert rusted
through in large sections.
- Culvert remaining circular
-Local flooding occurs
-HV traffic volume increases

Treatment

Situation 6
- Culvert invert rusted
through in large sections.
- Culvert ring movement evident
-No road surface settlement

Treatment

Unknown
Risk

Treatment

Situation 7
- Culvert invert rusted
through in large sections.
- Culvert ring movement evident
-Road surface settlement

Figure 14A.1.1 Appendix Outline

January 2013

14A-2

Extreme
Risk

Treatment
- Close road immediately
- Repair urgently

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Figure 14A.2.1.1(a) - Multi-plate Culvert.

Helical Corrugated Culverts


These culverts are helically wound by
machinery which decreases installation time
(refer Figure 14A.2.1.1(b)).
Metal
thickness is thinner than the plate culverts,
and is typically 3 mm or less. They are
therefore more susceptible to corrosion and
to distortion during backfilling operations
in construction.

14A.2.1.2 Size and Shape


Mark a one metre square grid with spots of
paint on a side wall starting at the upstream
end. Measure and record the horizontal and
vertical diameter at ten or more points to
determine its shape. If there are particular

defects such as open joints, extra diameters


should be measured.

14A.2.1.3 Corrugations Pitch


and Depth
The pitch (P) and depth (D) of corrugations
should be measured at a few places to
determine which shape was used in the
culvert.
Table 14A.2.1.3 and Figure
14A.2.1.3 detail the standard corrugation
dimensions in millimetres.
Additionally, non-sinusoidal corrugations
may be present and the shape of these
corrugations should be recorded or obtained
from the manufacturer.

January 2013

14A-3

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Figure 14A.2.1.1(b) - Helically Wound Culvert.


Table 14A.2.1.3 - Standard Corrugation Dimensions

Standard Culvert Corrugations


Type

Helical

38

6.5

68

13

75

25

125

25

68

13

200

55

Plated

Figure 14A.2.1.3 - Corrugations for Steel Pipes (Source: AS1762-1984).

January 2013

14A-4

Department of Transport and Main Roads


Road Drainage Manual

14A.2.1.4 Height of Fill Material


Measure the height of fill material above
the culvert, as information may be required
to determine the suitable repairs and in the
design process for determining the earth
vehicle loads.

14A.2.1.5 Material Thickness


Where an edge piece can be found at the
ends or joints, measure the thickness using
a micrometer or other thickness gauge to an
accuracy of 0.1 mm. If this can be
measured at a relatively clean uncorroded
point, it will indicate the original thickness.
A non-destructive thickness gauge is useful
in measuring residual metal when extensive
corrosion is obvious (refer to BAM Advice
Note 87- Ultrasonic Testing).
It is
important to determine whether sufficient
metal thickness is left in the bulk of the
culvert to justify a repair, even if the invert
is seriously corroded. If metal thickness
measurements indicate severe corrosion on
the outside of the culvert, a concrete invert
repair is not a suitable treatment.

14A.2.1.6 Maximum
Diameter

Outside

If relining is a possible repair method, then


measure the straightness of the culvert with
string lines, or with appropriate levelling
equipment (refer Attachment 14A.A at end
of this Appendix). The straightness of the
culvert will determine the largest possible
diameter that will fit through the entire
culvert length.

14A.2.1.7 Voids Present in Fill


If the culvert has been perforated by
corrosion, typically by standing water in the
invert, then some of the granular backfill
around the culvert may have been flushed
out in flood flows. Determine by first
tapping around the culvert, starting from

Appendix 14A
Management of Corroded Metal Culverts

near the invert where a void is most likely.


If voids are detected, then paint a line
between the hollow and solid sounding
sections and map them using the culvert
dimension grid. The depth of cavities can
be checked by drilling small holes and
measuring the depth of the void. Holes
should be sealed with an appropriate
galvanised screw.

14A.2.1.8 Defect Identification


Construction or In
Service
The defects that affect corrugated metal
culverts
are
joint
defects,
invert
deterioration, corrosion, shape distortion
and soil migration (DIB 83 CDOT). The
cause of the defect can be a result of the
construction process or in service loading,
as explained below.
Joint Defects - Helical
The most common problems associated
with helical steel joints are misalignment,
water exfiltration, backfill infiltration and
joint separation.
Misalignment of the joints may be a sign of
settlement in the supporting soil structure.
This settlement may have occurred during
construction and stabilised. The more
serious problem is if progressive settlement
is continuing to occur while in service.
Misalignment can also lead to undermining
of the culvert, water exfiltration or
infiltration of backfill material.
Exfiltration occurs when leaking joints
allow water flowing through the culvert to
leak into the supporting material.
Exfiltration can result in piping where
supporting soil material is easily eroded.
Infiltration is the opposite problem to
exfiltration and occurs where water from
the backfill material is seeping through the

January 2013

14A-5

14
A

Department of Transport and Main Roads


Road Drainage Manual

14
A

culvert joints.
Infiltration can cause
settlement and misalignment problems if
the water carries fine grained soil particles
from the backfill material.
Joint separation may also occur due to
external loads and changing soil conditions
and this allows backfill infiltration and
water exfiltration.
Joint Defects - Multi-Plate
Joint defects for multi-plate culverts occur
at the bolt lines typically from construction
damage. The bolt lines are weaker than the
plate itself, and some construction
specifications called for the joint to be
offset in each ring to avoid a line of
weakness.
If this was not done, the
backfilling operation could put high
bending moments on the bolted joint lines,
causing local cracking in the plates where
excessive tension occurs.
Figure
14A.2.1.8(a) shows such cracking, probably
caused by excessive side pressures during
backfill.
This is a defect that should be avoided in
construction, however once the culvert is
completed and backfilled, the cracked joint
should go into compression and not be a
long term failure initiator.
It is essential, though, that the cause of such
defects be determined. If they are due to
continuing vertical loading they may
indicate the start of structural failure.
Invert Deterioration
Invert deterioration is usually due to a
combination of corrosion and abrasion.
Once the galvanising layer is abraded from
material carried by the flow of the water,
corrosion then attacks the bare steel and is
accelerated by further abrasion that
constantly removes the protective oxide
layer formed by corrosion.
The

January 2013

14A-6

Appendix 14A
Management of Corroded Metal Culverts

continuation of this action will ultimately


lead to the loss of the invert and the
creation of scour holes under the culvert.
For metal culverts to withstand significant
fill loads and heavy repetitive live loads, an
effective soil-structure interaction is
necessary. This composite behaviour uses
the compressive strength of the culvert
wall, with the compressive or bearing
strength of the well-compacted soil
surrounding the structure. As loads are
applied to the culvert, the flexible structure
attempts to deflect, with the vertical
diameter decreasing and the horizontal
diameter increasing.
The change in
horizontal diameter is resisted by the lateral
soil pressure and results in the relatively
uniform radial pressure around the culvert
that creates a compressive thrust in the
culvert walls, hence Ring Compression
Theory (ConnDOT Drainage Manual).
Figure 14A.2.1.8(b) illustrates this theory
whereby overburden and live load stresses
are evenly distributed to surrounding soil.
As the culvert behaviour is reliant on
uniform radial pressure around the pipe,
loss of the invert may result in severe
distortion and collapse of the culvert.
Corrosion
Corrosion of the culvert can ultimately
cause failure by reducing the material
thickness. This can occur at the invert level
due to the removal of the galvanising
through abrasion and standing water or on
the external face of the culvert from the
chloride and sulphate content of the backfill
material. As the external face of the culvert
is not visible, the determination of metal
thickness is important to establish the
aggressiveness of the environment and the
necessary repairs.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A
Figure 14A.2.1.8(a) - Cracks in Metal Plate.

Figure 14A.2.1.8(b) - Ring Compression Theory (Source: ConnDOT Drainage Manual).

Shape Distortion
The construction process for metal culverts
requires care and attention to the backfilling
procedure (refer Attachment 14A.B at end
of this Appendix). During construction,
culverts are flexible and will distort if
excessive earth pressure or construction
loads are applied. In extreme cases, the
culverts can collapse, or be severely
distorted during construction.
Figure
14A.2.1.8(c) illustrates a joint damaged

during
backfill,
probably
construction overload.

due

to

In assessing a culvert in service, it is


essential to consider the cause of any
distortion (out of roundness). Construction
damage should be assessed separately to
overloading, distortions or soil movements
under current service conditions.

January 2013

14A-7

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A
Figure 14A.2.1.8(c) - New Culvert Damage
.

Significant distortions occurring during


construction should be recorded in
construction files or earlier inspection
reports.
Unfortunately, construction
records may be difficult to access after
several decades.
Determining when
damage occurred may require structural
engineering advice.

Soil migration occurs when there is a loss


of backfill support due to water eroding fine
material from the trench side walls. For
migration to occur the backfill material
must be erodible and there must be a flow
path for the water.
Because granular
material is used as culvert bedding, any
perforation of the culvert by corrosion can
lead to flood water washing the bedding
sand out. Any significant flow of water
behind the metal culvert can lead to severe
loss of backfill followed by embankment
collapse, so repairs to culvert perforations
should be to undertaken as quickly as
possible before the next wet season.

14A-8

Data Collected

Culvert Type
(Multi-plate/Helical)
Size and Shape
Corrugation (pitch x depth)
Height of Fill Material (m)
Material Thickness
Extent of Corrosion
(invert/full height)
Maximum OD (if relining)
Voids Present in Fill

for

Figure 14A.2.1.9 shows an example table


that can be used to summarise the
information that may be gathered on site to

January 2013

Additionally a sketch containing this


information can also be useful to record
defect locations (refer Attachment 14A.C at
end of this Appendix).

Culvert Property

Soil Migration

14A.2.1.9 Example Table


Data Collection

determine a suitable repair method and aid


the design process.

Other Defects and Cause


(Construction/In Service)
Figure 14A.2.1.9 Example Summary
Table

Department of Transport and Main Roads


Road Drainage Manual

14A.3 Treatment
14A.3.1 Repair methods
The following is a range of repair
treatments that should be considered, taking
into account:
Cost
Remaining life of the non-corroded
portion of the culvert
Durability and expected service life of
the repair
Size of embankment over culvert
Traffic volumes
Water flow
Ability to detour traffic
The products detailed in the repair methods
below must be obtained from approved
suppliers.
Emergency Propping
Install emergency propping for immediate
temporary stabilising of the structure when
a safety issues become evident. Propping
of the culvert can only slow the complete
collapse of the structure and there is still a
future risk to the travelling public if a
failure is allowed to occur. Since most

Appendix 14A
Management of Corroded Metal Culverts

conditions that require propping are caused


by excessive corrosion of the invert,
vertical propping is often not an option.
However, propping at 45 degrees (refer to
Figure 14A.3.1(a)) is effective.
For safety, 6mm self-tapping screws can be
inserted in the culvert to hold timber
spreader beams with heavy wire ties until
the props are installed. Each sleeper should
have a minimum of two props. Timber
sleepers should be used to distribute the
load from the props across numerous
culvert corrugations. The timber sleepers
will be held in place against the wall by
friction, but a suitable connection of the
props to the sleepers should be considered
to prevent the props from being washed out.
Seek advice from Bridge Design Branch for
the required safe working load and
arrangement of the props.
Joint Repairs
Separated joints can be repaired using a
chemical grout such as polyurethane foams.
This will stop joints from leaking and
prevent the erosion of backfill material
behind the culvert. Where backfill material
has been lost, the voids can be grouted
using a low pressure cement based grout.

Figure 14A.3.1(a) - Example of Emergency Propping.


Note: This culvert had not failed and props were installed pending a full structural assessment.

January 2013

14A-9

14
A

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

Repair using a Paint System

14
A

Use this repair method where the culvert is


not extensively corroded and significant
metal thickness remains. The type of paint
system used will depend on the abrasive
conditions the culvert is subject to. If a
significant amount of debris flows through
the culvert regularly, high abrasion resistant
systems can be used. Alternatively, paint
systems with lower abrasion resistance can
be used in culverts which do not flow
regularly or if the flow does not contain
debris.
Repair the Corroded
Concrete Liner

Invert

with

This is the most common repair for larger


culverts with sufficient working space and
high embankments where it is difficult to
remove the culvert.
When using this treatment it is essential that
the concrete invert liner forms a structural
bond with the uncorroded metal above the
invert. Thin layers of concrete that are not
structurally fixed to the metal culvert are
not effective repairs. Figure 14A.3.1(b))
shows a thin concrete invert lining which

Figure 14A.3.1(b) - Failed Thin Concrete Lining.

January 2013

14A-10

has separated from the culvert and washed


away in flood. The invert has not been
protected and is badly corroded.
Voids that may be present beneath the
invert or in the backfill material will need to
be grouted. Bridge Design Branch can be
consulted for a site specific design.
Also, the rest of the metal culvert
(excluding the corroded invert) must be in a
reasonable condition.
The inside face
should
have
adequate
remaining
galvanising metal. Thickness should be
checked with a non-destructive thickness
gauge or by drilling a small number of
holes and checking with a thickness gauge
reading to 0.2 mm or better and holes must
be plugged with a galvanised screw
afterwards.
If there is significant thickness loss due to
corrosion of the external face of the culvert,
a concrete invert may not be effective as a
long term repair and relining should be
considered.

Department of Transport and Main Roads


Road Drainage Manual

Relining the Culvert


Where significant metal thickness has been
lost from the entire circumference of the
culvert, a repair option is to reline with a
new culvert of smaller diameter and grout
the anulus between the new and old culvert.
It is recommended that the new culvert be a
more durable material such as concrete or
steel
reinforced
polyethylene.
Alternatively, for larger diameter culverts,
conduct
relining
by
shotcreting
reinforcement cages that form a new
internal pipe. This can be an effective long
term repair, but it is considerably more
expensive than a concrete invert.
The culvert must be closest to its theoretical
shape for efficient relining. If there is
significant distortion, then a much smaller
pipe liner would be required for relining
which can cause several problems such as:
The void between the new liner and
the old culvert or in the backfill
material can be very large and
expensive to grout.

Appendix 14A
Management of Corroded Metal Culverts

The smaller diameter culvert may not


carry the design flood flow, resulting
in the road embankment overtopping
and washing away.
Figure 14A.3.1(c) shows an example of the
relining process. Relining needs some
specialist consideration and both Bridge
Design and Hydraulics Branch within the
departments Engineering & Technology
Group should be consulted.
Remove and Replace with a More Durable
Pipe
Removing and replacing a culvert should be
considered when culverts:
Have large distortion from original
shape (greater than 5%)
Significant reduction in waterway area
due to the culvert shape not being
circular (they can be oval etc in shape)
or not having a consistent diameter
along the length.

Figure 14A.3.1(c) - Example of Relining Process.

January 2013

14A-11

14
A

Department of Transport and Main Roads


Road Drainage Manual

14
A

Appendix 14A
Management of Corroded Metal Culverts

Have significant voids in the


embankment material and grouting is
not practical.

jacking is a specialist contracting skill and


advice should be sought from the Bridge
Design Branch.

Are considered to have structurally


failed.

14A.3.2 Typical Situations

Removing and replacing a culvert is


most practical option when:
Culverts are smaller than 1.5 m
diameter making it difficult to work
inside,
Embankment height is low and traffic
can be easily diverted, or
Larger culverts have failed in ring
compression, are badly corroded or
have distorted significantly, so they
can not be easily relined.
Pipe Jacking Around the Existing Culvert
Pipe jacking around the existing culvert is
an expensive option, but may be necessary
when there are no other alternatives.
Examples include failed culverts that are
unsafe to work in and culverts under high
fills or major roads where traffic diversion
and excavation are impractical.
Pipe

Figure 14A.3.2(a) - Standing Water in Culvert

January 2013

14A-12

The following is a description of typical


conditions that will be encountered when
inspecting metal culverts and an indication
of the asset management conditions state,
along with a risk ranking (depending on a
number of factors) and the action required
to ensure the protection of the asset and
travelling public.

Situation 1
Unknown Risk Condition
Standing water in the culvert, typically
caused by a blockage on the
downstream end
(refer to Figure
14A.3.2(a)); and
Structure will rapidly corrode and it is
difficult to inspect its condition. A
decision must be made quickly on how
to make the culvert durable, easy to
inspect
and
safe.

Department of Transport and Main Roads


Road Drainage Manual

Immediate Treatment
Survey downstream levels and the consult
land owner. If practical, cut a low flow
channel that will allow the culvert to drain.
If a blockage is caused by cattle or vehicles
pushing the bank into the stream bed,
negotiate a preventative strategy such as:
Constructing an alternate crossings
further downstream;
Installing a low flow pipe under a low
level concrete ford; or
Putting up a fence around the stream
for sufficient length to maintain exit
drainage.
Hydraulics Branch can be consulted if there
are concerns regarding this.
Later Treatment
Assess culverts when clean and dry to
determine whether additional repairs
are needed; and
Check culverts after each wet season to
ensure they remain self draining.
Note: If a culvert has deep standing water
and downstream surveys show it is not
practical to drain the culvert, Regional /
District Management must be informed and
a management plan developed.

Appendix 14A
Management of Corroded Metal Culverts

Situation 2
Very Low Risk Condition
Significant corrosion appearing in the
invert of the culvert, but substantial
thickness of metal remains (refer to
Figure 14A.3.2(b));
Culvert is retaining circular shape or
may have distorted during construction
process but is remaining stable (see
notes on defect identification); and
The culvert is assessed as being Asset
Management Condition State 3.
Treatment
Monitor culvert annually;
Plan to install a concrete invert or use a
paint system, while the culvert still has
adequate metal in the invert and before
the culvert perforates due to corrosion;
and
Complete repairs within 2 years.

Situation 3
Low Risk Condition
Considerable level of corrosion in the
invert and small holes appearing in the
invert of the culvert (refer to Figures
14A.3.2(c) and 14A.3.2(d));
No evidence of material loss from soil
behind culvert and no soil cavities
evident;

January 2013

14A-13

14
A

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Figure 14A.3.2(b) - Culvert Showing significant Corrosion in Invert


Note: Propping as shown is not necessary in this condition.

Culvert is retaining circular shape or


may have distorted during construction
process but is remaining stable (see
notes on defect identification) ; and
Culvert is assessed as being Asset
Management Condition State 4.
Treatment
Monitor and plan for concrete inverts
within the next 12 months.
If
corrosion extends above the level

Figure 14A.3.2(c) - Heavy Corrosion in Invert.


Note: Observe some small perforations to metal culvert.

January 2013

14A-14

where it is practical to install a


concrete invert consider relining; and
After a significant rainfall event, check
the culvert for erosion of backfill
material
through
the
invert
perforations.

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Figure 14A.3.2(d) - Heavy Corrosion.


Note: There is considerable loss of metal thickness.

Situation 4
Medium Risk Condition
Culvert invert rusted completely
through over a large portion of the
length;
Losses of backfill material below
invert (refer to Figure 14A.3.2(e));
Culvert is retaining circular shape
(loads are being carried by soil arch) or
may have distorted during construction
process but is remaining stable (see
notes on defect identification);
Culvert has no significant flows; and
Heavy vehicle traffic loading remains
constant.
Treatment
Prop immediately and repair as soon as
possible within two months or before
the next wet season starts;
Check culvert and pavement levels
fortnightly, after short rainfall events
and during extended rainfall periods
for any signs of ring compression
failure or settlement of the soil arch
over the culvert; and

Put speed restrictions and hazard


identification signage in place to
ensure adequate stopping sight
distance should a hazard develop.
Safety
If a dip in the pavement occurs the road
is to be closed immediately.

Situation 5
High Risk Condition
Culvert invert rusted completely
through over a large portion of the
length;
Losses of backfill material below the
invert (refer to Figure 14A.3.2(e));
Culvert is retaining circular shape
(loads are being carried by soil arch) or
may have distorted during construction
process but is remaining stable (see
notes on defect identification);
Local flooding event occurs creating
the risk that the culvert and
embankment could be washed out; or
Heavy
vehicle
increases.

traffic

loading

January 2013

14A-15

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

14
A

Figure 14A.3.2(e) - Culvert Invert Corroded Away.


Note: Observe loss of granular bedding material in invert.

road is not available for extended


closure.

Treatment
Culvert must be repaired as soon as
practical (within one month);
Props should be installed immediately;
Culvert and pavement surface should
be checked weekly, after short rainfall
events and during extended rainfall
periods for any signs of ring
compression failure or dips in the
pavement caused by soil arch
settlement;
If the replacement culvert cannot be
obtained immediately from the
manufacturer alternative options that
may be considered are:

Temporarily replace the culvert with


a readily available low flow
reinforced concrete pipe. The low
flow pipe can then be replaced with
the appropriately sized pipe when it
is available from manufacture.
Completing the works under full
road closure if sidetracks or
diversions are an issue.
Back filling with concrete can be
undertaken if necessary to decrease
the time required for installation if

January 2013

14A-16

Put speed restrictions and hazard


identification signage in place to
ensure adequate stopping sight
distance should a hazard develop.
Safety
If a dip in the pavement occurs the road
is to be closed immediately.
With reference to Figure 14A.3.2(e), it was
interesting to note that the helical lock
seams of the culvert had not distorted and
the culvert remained circular with no
evidence of ring compression failure

Situation 6
Very High Risk Condition
Culvert invert rusted through over
large portion of the length;
Compression ring movement is
obvious (metal has buckled or is
overlapping), but soil arch is mostly
intact (refer to Figure 14A.3.2(f));
Voids possibly present behind culvert
lining but no dip in road;
Culvert
structure
has
failed,
embankment soil arch at high risk of

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

imminent failure (Asset Management


Condition State 5); and
No visible settlement of road pavement
over culvert.
Treatment
Inform
Regional
/
Management immediately;

District

Seek structural engineering advice


send photos of culvert;
Prop the culvert immediately if safe to
do so;
Put speed restrictions and hazard
identification signage in place to
ensure adequate stopping sight
distance should a hazard develop; and
Urgent repairs will be necessary.
Safety
Culvert is structurally unsafe and can
fail under heavy traffic loading or in a
flood due to embankment erosion; and if
a dip in the pavement occurs the road is
to be closed immediately.

Situation 7
Extreme Risk Condition
Culvert invert rusted through over
large portion of the length;
Compression ring movement is
obvious (metal has buckled or is
overlapping) but soil arch is mostly
intact (refer to Figure 14A.3.2(f));
Voids possibly present behind culvert
lining but no dip in road;
Asset Management Condition State 5,
culvert
structure
has
failed,
embankment soil arch at high risk of
imminent failure; and
Road pavement above culvert shows
obvious signs of settlement (refer to
Figure 14A.3.2(g)).
Treatment
As given for Very High Risk.
Safety
Culvert has structurally failed; and road
is to be closed immediately.

Figure 14A.3.2(f) - Culvert Ring Movement.

January 2013

14A-17

14
A

Department of Transport and Main Roads


Road Drainage Manual

14
A

Figure 14A.3.2(g) - Culvert Soil Arch Failure.

14A.4 Conclusion
The objective of this Appendix is to act as a
guide for the inspection and treatment of
corrugated metal culverts to provide a safe
and efficient road network. The situations
given are typical and if there is any
uncertainty in the condition and treatment
of a culvert further advice can be obtained
from the departments Bridge Design
Branch located within the Engineering &
Technology Group.

January 2013

14A-18

Appendix 14A
Management of Corroded Metal Culverts

Department of Transport and Main Roads


Road Drainage Manual

Attachment 14A.A - Selecting the


Largest Possible Liner Size
(Outside Diameter)
At either end of the culvert, establish the
horizontal and vertical centre point with a
level and straight edge of suitable length to
fit inside culvert.
Put self-tapping screws through the culvert
and establish horizontal and vertical

Appendix 14A
Management of Corroded Metal Culverts

stringlines at each end with a laser at one


end and aim it at the centre of other end.
At 2 m centres, measure the horizontal and
vertical diameters and the laser centreline
intercept for both.
The dimensions can be plotted, and the
largest straight internal tube can be
measured.

Plot all horizontal and vertical diameters and laser intercepts.


Measure the largest included circle.
Choose a liner based on suitable construction clearances and grout thickness requirements.

January 2013

14A-19

14
A

Department of Transport and Main Roads


Road Drainage Manual

14
A

Appendix 14A
Management of Corroded Metal Culverts

Attachment 14A.B - Backfilling


Procedure

tolerance, while the embankment above


the pipe is compacted.

When completed, a metal culvert operates


in ring compression and is supported by the
soil all round to prevent buckling (and
subsequent failure in bending).
It is
therefore essential to have the backfill
placed and compacted correctly.
The
following was the backfilling sequence as
described in AS1762-1984:

5. After sufficient fill is placed over the


culvert, full compactive effort can be
used on the rest of the fill. The aim is
to end up with a culvert which is
circular (within the allowed tolerances).

1. The base support is placed to obtain


correct line and level, the culvert is
assembled or placed on a layer of loose
granular material of 12 mm maximum
aggregate size.
2. The lower third of the circumference is
then filled, normally with a granular
material as it is difficult to gain access
to compact a clayey fill under the sides
of the pipe.
3. The both sides of the culvert are then
compacted simultaneously in layers not
more than 150 mm thick. The back fill
must be placed evenly to keep it at the
same elevation on both sides of the
structure at all times. The horizontal
diameter of the culvert was checked to
ensure it stays within tolerance
(typically 5% of diameter). Horizontal
props may assist in controlling
distortion, at this stage, to within this
tolerance.
4. The top third of the culvert
circumference has to be very carefully
placed and compacted with lighter more
controllable equipment. At this stage
the vertical diameter will decrease and
the horizontal diameter increases as the
metal
culvert
goes
into
ring
compression. Vertical supports may be
used to limit deflections to within

January 2013

14A-20

Department of Transport and Main Roads


Road Drainage Manual

Appendix 14A
Management of Corroded Metal Culverts

Attachment 14A.C Example Sketch

Janaury 2013

14A-21

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Glossary

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Glossary Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Glossary
Glossary based on: (1) AS 1348 - 2002 Australian Standard for Road and Traffic Engineering;
(2) QUDM 2008; (3) American Association of State Highway and Transportation Officials,
Model Drainage Manual 1991; and (4) Queensland Water Act 2000.

Afflux - The rise in water level on the upstream side of a constriction such as a road, weir or
bridge in a stream or channel relative to the water level at that point without the constriction.
Aggregate - A soil aggregate consists of a cluster of primary soil particles bound together into a
clearly defined unit by electrostatic forces such as charges on clay particles, cementing
substances from organic matter or hydroxides of iron, or binding strands of roots.
Allotment Drainage - A system of field gullies, manhole chambers and underground pipes
constructed within private property to convey flows through and from allotments.
Alluvial Plain - A landform with extremely low relief formed by the accumulation of alluvium
from overbank stream flow over a considerable period of time. This accumulation may still be
occurring (flood plain) or may have ceased (terrace).
Amelioration - Refers to efforts made to minimise adverse effects of an activity (e.g. road
construction) after the event.
Anabranch - A branch of a stream which leaves and later re-enters the stream.
Annual Flood - The highest peak discharge in a water year consisting of 1 September to 31
August.
Apron - A floor or lining at either the inlet or outlet of a hydraulic structure such as a culvert to
protect the waterway channel from erosion.
Asphalt - A mixture of bituminous binder and aggregate with or without mineral filler produced
hot in a mixing plant. It is delivered, spread and compacted while hot.
Australian Height Datum (AHD) - A level datum, uniform throughout Australia, based on an
origin determined from observations of mean sea level at tide gauge stations, located at more
than 30 points along the Australian coastline.
Australian Height Datum Derived (AHDD) - The level datum based on a direct connection to
the Australian height datum.
Average Annual Time of Closure (AATOC) - The times of closure for every flood are totalled
and the sum is divided by the years of record.
Average Annual Time of Submergence (AATOS) - The times of submergence for every flood
exceeding a selected level (deck level) are totalled and the sum is divided by the years of record.
Average Recurrence Interval (ARI) - The average or expected value of the period between
exceedances of a given rainfall intensity or discharge.
March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Backwater - That part of a stream, the water level of which is kept above normal due to some
controlling influence downstream.
Backwater Curve Analysis - A procedure for determining water surface levels in open
channels under gradually varied flow conditions.
Basin - A hollow or depression within which water can be contained.
Batter - The side slope of an embankment or cutting.
Batter Chute - A temporary or permanent structure designed to convey concentrated storm
runoff down a cut or fill embankment, or over a change in grade, without causing erosion.
Bed and bank - With reference to a watercourse or lake, means land over which the water of
that watercourse or lake normally flows or that is normally covered by the water whether
permanently or intermittently, but does not include land adjoining or adjacent to the bed or
banks that is from time to time covered by floodwater. [Water Act 2002, Queensland.]
Bench (benching) - (1) A ledge cut or formed in the batter of a cutting or bank, to provide
greater security against slips; (2) A ledge constructed in a batter or natural slope to segment the
slope length to reduce erosion potential.
Berm - Typically used in reference to slope barrier measures designed to break the continuity of
slopes to reduce runoff velocities.
Biodiversity - (short for biological diversity) Number of species of vegetation and wildlife in a
given habitat; e.g. rainforests, [which typically have a large number of different plant and
animal species] are said to be high in biodiversity.
Borrow Pit - An excavation outside the formation limits for obtaining fill.
Broadcast Seeding - Any method of planting seed which scatters the seed in random pattern on
the surface of the soil.
Buffer Zone - Zone adjoining a sensitive area that is required for protecting stream banks from
erosion, providing habitat along stream corridors, and treating overland flow before it enters the
drainage network.
Building - A habitable room; retail or commercial space; factory or warehouse; basement
providing carparking space, building services or equipment; or enclosed carpark or enclosed
garage.
Bypass Flow - That portion of the flow in a road or in a channel which is not collected by a
gully inlet or field inlet, and which is redirected out of the system or to another inlet in the
system.
Capillary Water - Water drawn upwards into soil pores and held by surface tension.
Catch Bank - A small levee constructed to intercept and divert runoff away from cut slopes,
embankments, disturbed areas, and stockpiles. An alternative to, or can be combined with, a
Catch Drain. See also Std Dwg: 1178.
Catch Drain - A small surface channel constructed to intercept and divert runoff away from cut
slopes, embankments, disturbed areas, and stockpiles. See also Std Dwg: 1178.

March 2010

iv

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Catchment Area - That area determined by topographical or equivalent features, upon any part
of which rain falling will contribute to the discharge of the stream at the point under
consideration.
Causeway - A raised carriageway across wet or low areas or across tidal water.
Channel - (1) The bed of a stream or river; (2) A course or passage through which something
may move or be directed.
Channel Freeboard - Vertical distance between the design water surface elevation in an open
channel and the level of the top of the channel bank.
Channel Lining - Material placed on the surface of a channel or chute to protect it from
erosion. Materials include grass and turf, reinforced grass, erosion control mats, rock lining,
rock mattress, cellular confinement and impervious liners.
Channel Stabilisation - Materials used to stabilise the channel surface. Examples include soil
retention blankets (with appropriate seeding mixture) for non-structural cover, and concrete or
riprap (rock) for structural cover.
Check Dam - Check dams are typically used in channels conveying concentrated flows to
control flow velocity and minor gully erosion. They may be constructed from semi-pervious or
impervious materials such as medium-size rock or sand and gravel filled bags.
Chute - Used to convey water down slopes and are constructed with materials suited to the
expected life of the chute (i.e. concrete for permanent chutes).
Clearing - The removal of vegetation, structures or other objects.
Coastal Plain - A complex, level to very gently inclined landform pattern adjacent to the coast.
The plain was formed by the deposition of material from overbank stream flow, overland sheet
flow and marine inundation.
Coefficient of Runoff (Discharge) - Dimensionless coefficient used in the Rational Method for
the calculation of peak runoff discharge.
Coefficient of Runoff (Volumetric) - The ratio of the amount of water that runs off to the
amount that falls in a catchment area.
Cofferdam - A temporary enclosure formed to exclude water from an area in which
construction is to take place. Cofferdams can take a variety of forms and are constructed from
materials such as driven sheet piling, rock, earth or concrete.
Concentrated Flow - Water, usually storm runoff, flowing in a confined feature such as a
channel, ditch, swale, river, etc.
Contour Ploughing - Ploughing horizontally along the contour.
Cover Crop - Plants, particularly cereals, grown mainly to protect the soil on a temporary basis
during or prior to the establishment of more protective plant cover.
Cover Ground - Any vegetation producing a mat on or just above the soil surface. In forests,
this may be formed by low-growing shrubs, vines, and herbaceous plants under the trees.

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Crest - The summit or top edge.


Critical Depth - The depth occurring in a channel or conduit at a condition of flow such that
the specific energy is a minimum for the particular flow. Froude number equals 1.
Critical Flow - Flow condition such that the specific energy (of the mean flow) is minimum.
Froude number equals 1. When the flow is critical, small changes in specific energy cause large
changes in flow depth. In practice, critical flow over a long reach of a channel is unstable,
resulting in a wavy or undulating surface.
Critical Velocity - The average velocity of flow in a section of a channel or conduit when the
flow is at critical depth.
Cross Drainage - A system of pipes or culverts which convey storm flows transversely across
or under a roadway.
Crossfall - The slope, at right angles to the alignment, of the surface of any part of a
carriageway.
Crown - (1) The highest part of an arch; (2) The highest point on the cross-section of a
carriageway with a two-way crossfall.
Culvert - One or more adjacent pipes or enclosed channels for conveying a watercourse or
stream below formation level.
Dam - A barrier or embankment which confines water.
Design Discharge - The volume of water in cubic metres passing during one second, through a
drainage section to be designed, which determines the size the drainage section needs to be.
Detention Basin (Retention/Retarding/ Compensating Basin) - A storage pond, basin or tank
used to reduce and attenuate the peak discharge within a drainage system. In Australia,
detention basins may have an outlet pipe and/or a spillway and the term is interchangeable with
retention basin. It is also interchangeable with sedimentation basin when sediment control is the
main purpose of the basin.
Development Category - Refers to land use within a catchment (see Land Use).
Diversion Block - A small block constructed for the purpose of diverting water from the table
drain to a culvert or side drain. Can be combined with a Diversion Drain to divert water out of
the table drain and away from the road formation. See also Std Dwg: 1178.
Diversion Channel - A hydraulically designed, open channel that diverts or redirects a given
water flow from its current, natural flow path.
Diversion Drain - A drain leading water away from an existing small channel (for example, a
table drain). See also Std Dwg: 1178.
Drainage - Natural or artificial means of intercepting and removing surface or subsurface water
(usually by gravity).
Drainage Catchment - The area of land contributing stormwater runoff to the point under
consideration.

March 2010

vi

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Drainage System - A system of gully inlets, pipes, overland flow paths, open channels, culverts
and detention basins used to convey runoff to its receiving waters.
Dyke - See Levee.
Ecologically Sustainable Development (ESD) - Protection of the environment while allowing
for the development that improves the total quality of life, both now and in the future, in a way
that maintains the ecological processes on which life depends.
Ecology - The interrelationships between plants, animals and humans which compete and
depend on each other for existence in the physical environment.
EMC - Event Mean Concentration is defined as pollutant load washed off by a storm event
divided by the runoff volume.
Energy dissipator - Any means to reduce the total energy of flowing water especially highvelocity flows. In stormwater design, they are usually mechanisms that reduce velocity prior to,
or at, discharge from an outfall in order to prevent erosion. Materials used include gabions,
concrete splash pads, drop structures, riprap, and boulders.
Environment - Refers to the conditions in which an organism lives and survives or the
conditions in which an object or organism resides. These conditions can be described as aspects
of a physical, social or an economic environment, depending on the perspective perceived
by the observer. The Environmental Protection Act (1994) describes the environment as
including: (a) ecosystems and their constituent parts, including people and communities; (b) all
natural and physical resources; (c) the qualities and characteristics of locations, places and
areas, however large or small that contribute to their biological diversity and integrity, intrinsic
or attributed scientific value or interest, amenity, harmony and sense of community; and (d) the
social, economic, aesthetic and cultural conditions that affect, or are affected by things
mentioned in paragraphs (a) to (c).
Environmental Design Report (EDR) - Outlines the ways in which environmental
requirements outlined in the EMP (Planning) have been incorporated into the detailed design
and contract documentation. It is an audit report and not a contract document.
Environmental Management Plan (EMP) (Construction) - Prepared by the Construction
Contractor, considering the management of environmental impacts of activities in a given road
project during the Construction period. It identifies risks to the environment for the project and
the environmental requirements contained within the Contract documentation and outlines key
strategies for managing these risks and minimising undesirable environmental impacts.
Environmental Management Plan (EMP) (Maintenance) - Prepared by the Maintenance
Contractor. It considers the management of environmental impacts of activities in a given
maintenance project as well as environmental requirements contained within the Contract
documentation. It focuses on the minimisation of adverse effects (and the management of these
effects) on the environment by maintenance activities required.
Environmental Management Plan (EMP) (Planning) - Sets out the overall requirements,
outcomes and performance indicators for the design, construction and maintenance of a road
project.

March 2010

vii

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Environmental Monitoring - Includes activities which gather and evaluate information used
for the assessment of environmental performance.
Erosion - (1) The wearing away of the land surface by moving water, wind, ice or other
geological agents, including such processes as gravitational creep; (2) Detachment and
movement of soil or rock fragments by water, wind, ice or gravity. (i.e. accelerated, geological,
gully, natural, rill, sheet, splash, or impact, etc).
Erosion Control - Includes the protection of soil from dislocation by water, wind or other
agents.
Erosivity - The erosive potential of rainfall expressed as the product of total storm energy and
the maximum 30 minute intensity of each storm.
Extreme Flood - The rare flood event for which the performance of a detention basin or similar
structure should be checked in order to assess the economic and social risk that could be
associated with overtopping or failure of that structure.
Fan - A level to very gently inclined landform associated with rapidly migrating stream
channels. The landform occurs as a typical fan shape below uplands on hard rock and is formed
by deposition of alluvial material through overbank stream flow and overland sheet flow.
Filter Fabric - See Geotextile.
Filter Material - Granular material with the grading selected so that it will allow water to pass
through it, while retarding the movement of soil particles.
Filter Strip - Typically a long, relatively narrow area of undisturbed or planted vegetation used
to retard or filter sediment for the protection of watercourses, drainage basins, diversions,
reservoirs, or adjacent properties.
Flocculant - A chemical agent used to enhance the flocculation process (see Flocculation).
Flocculation - The process of combining individual soil particles to form larger aggregates to
facilitate settling.
Flood Boundary Line - A line defining the edge of the area submerged at the height of a flood.
Usually related to a given recurrence interval.
Floodgate - (1) A gate placed in a channel, drain or culvert to control the passage of flood or
tidal water; (2) A gate placed in a fence where it crosses a watercourse or drain, which permits
the passage of stormwater or floodwater but acts otherwise as an integral part of the fence.
Flood Immunity - Average recurrence interval (ARI) of a flood at the point of overtopping the
crown level or highest point of the road if superelevated.
Floodplain - The relatively flat area adjoining the channel of a natural stream which has been or
may be inundated with flood waters.
Floodway - Longitudinal depression in a carriageway specially constructed to allow the passage
of floodwater across it without damage.
Fraction Impervious - That part of a catchment which is impervious, expressed as a decimal or
percentage.

March 2010

viii

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Freeboard - The height between a given water level and the underside of a bridge, top of a
channel or embankment, or floor of a building to give a factor of safety against calculated
design flood levels.
Friction Slope - Sometimes referred to as the hydraulic gradient or pressure gradient and is the
slope of the line representing the pressure head, or piezometric head in a pipeline.
Froude Number - Froude number is proportional to the square root of the ratio of the inertial
forces over the weight of fluid.
For an irregular section F V

A
B

Where V is the mean velocity, A is the cross sectional area, B is the free surface width and g is
acceleration due to gravity.
When Fr < 1, then flow is subcritical, when Fr = 1, flow is critical and when Fr > 1, flow is
supercritical.
Gabion (reno mattress) - (1) Wire mesh basket or cage filled with rock and used to retain earth
or as a protecting agent against erosion. Used as energy dissipators, channel liners, steep-slope
protectors, and retaining walls; (2) Geotextiles filled with soil, in which cuttings (brush) are
placed used for streambank stabilisation and termed soft gabions.
Geofabric - See geotextile.
Geotextile (filter fabric, geofabric) - (1) A synthetic fabric, woven or non-woven, used for
various purposes including embankment reinforcing and stabilization, as a filter layer between
dissimilar materials and as a strain alleviating membrane; (2) Channel lining - to protect
channels against erosion and to filter fine particles from runoff.
Grading - Any stripping, cutting, filling, stockpiling, or combination thereof which modifies
the land surface.
Ground Water - The water below the water table.
Grubbing - The removal of roots and stumps below ground level.
Head - The difference in water level upstream and downstream of a structure.
Head Loss Coefficient (Pressure) - A dimensionless coefficient which, when multiplied by the
velocity head in the outlet pipe, gives the difference in the level of the hydraulic grade line
(HGL) between the inlet and outlet pipes. It may be positive (indicating that the HGL rises
upstream) or negative (indicating that the HGL is less upstream).
Headwater - The height of water above the invert of a culvert measured at the inlet of the
culvert.
High Level Basin Outlet - The outlet of a detention/retention basin from which flows greater
than those handled by the low level outlet will be discharged (usually a weir type or glory hole
spillway).

March 2010

ix

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Hydraulic Design - In relation to stormwater drainage, this involves the determination of


velocities, the hydraulic grade line and water levels as storm runoff passes through the drainage
system.
Hydraulic Grade Line (HGL) - A line representing the pressure head along a pipeline,
corresponding to the effective water surface elevation in the piped portions of the stormwater
drainage systems.
Hydraulic Gradient - The slope of the hydraulic grade line - see also Friction Slope.
Hydrograph - A graph of stream height or volume rate of flow past a specific point against
time.
Hydrology - Prediction of runoff based on an assessment of rainfall.
Hydromulching - A mechanical method of applying seed, lime, fertiliser and mulch in a water
slurry to which has been added fixatives for soil stabilisation.
Impact Assessment Study (IAS) - Is a detailed study of the likely effects (both positive and
negative) on the environment and of the ameliorative strategies proposed for a particular project.
Impermeable - Cannot be penetrated by a fluid such as air or water, but commonly refers to
water penetration.
Impervious Area - The area within a drainage catchment that is impermeable.
Infiltration - The slow movement of water into or through a soil or drainage system.
Inlet Control - A condition where discharge through a culvert is dictated by the depth of
headwater and entrance geometry at the inlet.
Intensity-Frequency-Duration Data (I.F.D) - Rainfall data used in the calculation of rainfall
runoff rates.
Invert - The lowest portion of the internal surface of a drain or culvert.
Junction Structure - A pit or chamber constructed at the junction of two or more pipes, or at a
change of grade.
Lake - Includes (1) A lagoon, swamp, or other natural collection of water, whether permanent
or intermittent; and (2) The bed and banks, and any other element, confining or containing the
water. [Water Act 2002, Queensland.]
Land Use (Development Category) - The particular use or uses of land within a catchment
such as central business, commercial, industrial, residential, open space and parks, major and
minor roads.
Leaching - The removal of soluble material and colloids by percolating water.
Legal Point of Discharge - A point of discharge which is either under the control of a Local
Authority or Statutory Authority, or at which discharge rights have been granted by registered
easement in favour of the Local Authority or Statutory Authority, and at which discharge from a
development will not create a worse situation for downstream property owners than that which
existed prior to the development.

March 2010

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Levee - An earth or rock embankment constructed to: prevent flooding of low lying land (for
example, along the banks of a stream or river), or control the level or direction of flow of water
at or into a structure.
Level Spreader - A device to convert channel or pipe flow to sheet flow to prevent
concentrated, erosive flows from occurring, and to enhance filtration.
Loss Rate - The rate at which rainfall is lost through processes such as infiltration, evaporation
and local storage, and therefore does not contribute to surface runoff.
Major Drainage System - The major drainage system is that part of the overall drainage
system which is designed to convey a specified rare flood event. This system may comprise: (a)
Open space floodway channels, road reserves, pavement expanses and other flow paths that can
act as overland flow paths for flows in excess of the capacity of the Minor Drainage System; (b)
Detention Basins and Lagoons; and (c) Major underground piped systems installed where
overland flow is either impractical or unacceptable.
Major Road - A road to which is assigned a permanent priority for traffic movement over that
of other roads.
Major Storm - The design storm with an average recurrence interval selected on the basis of
satisfying requirements for flood immunity and safety. Design may vary in accordance with
local authority guidelines. For most development in Queensland, the Major Storm has an ARI
of 100 years.
Mannings Roughness Coefficient - A measure of the surface roughness of a conduit or
channel to be applied in the Mannings equation.
Minor Drainage System - The minor drainage system includes kerbs and channels, roadside
channels, inlets, underground drainage, junction pits or access chambers and outlets designed to
fully contain and convey a design minor stormwater flow of specified Average Recurrence
Interval. This arrangement may also include: (a) field gully inlet pits, installed to collect surface
runoff from within allotments, as well as the roofwater drainage provisions for buildings; (b)
Cross drainage under minor roads where delay or inconvenience during major flows is
acceptable. This also includes low flow pipes or box culverts installed under floodways; and (c)
Low flow pipes installed under drainage reserves or park areas.
Minor Road - A road whose primary function is to provide access to individual properties.
Minor Storm - The design storm with an average recurrence interval selected on the basis of
satisfying requirements for convenience and safety of pedestrians and vehicles. Design may
vary in accordance with local authority guidelines. For most development in Queensland, the
Minor Storm has an ARI of between 2 and 10 years.
Mulch Anchoring - A method used to increase the effectiveness of mulch against surface
erosion by water and wind. Binding agents referred to as tackifiers are mixed with the mulch in
a water slurry prior to application.
Mulching - The application of plant residues or other suitable material to the land surface to
conserve moisture, hold the soil in place, aid in establishing plant cover, increase infiltration and
minimize temperature fluctuations.

March 2010

xi

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Normal Flow Conditions - A condition in open channel flow where the depth and velocity of
flow achieved is consistent with the prevailing channel shape, slope and roughness.
Obvert - The highest portion of the internal surface of a culvert or arch.
Open Graded Asphalt - A bituminous mix using aggregate containing only small amounts of
fine material, and providing a high percentage of air voids.
Outlet - The point at which water discharges from a stream, river, lake, tidewater, or artificial
drain.
Outlet Control - The situation where factors downstream of the culvert entry such as high
water level at the outlet govern the discharge characteristics.
Outlet Protection - scour protection placed downstream of a pipe or culvert outlet to complete
the transition between pipe flow and open channel flow. Pipe outlet protection may be provided
by energy dissipaters, channel protection (non-structural and structural methods), or a
combination of the two. See also Apron.
Overland Flow Path - Open space floodway channels, road reserves, pavement expanses and
other flow paths that convey flows typically in excess of the capacity of the Minor Drainage
System.
Pavement Immunity - An immunity defined to ensure part or all of a pavement is not
submerged in a flood with a given ARI.
Permeability - The property of a material by virtue of which a fluid such as water can pass
through it.
Pervious Surface (Pervious Area) - A surface or area within a drainage catchment where some
of the rainfall will infiltrate thus resulting in a reduced volume and rate of runoff e.g. grassed
playing fields, lawns etc.
Piping - Refer to Tunnel Erosion.
Pressure Change Coefficient - Refers to Head Loss Coefficient (Pressure).
Probable Maximum Flood - The theoretically greatest runoff event from a catchment.
Probable Maximum Precipitation - The greatest theoretical depth of precipitation for a given
duration that is physically possible over a catchment.
Rainfall Intensity - The rate of rainfall in millimetres per hour.
Reach - A section or length of stream/channel/river over which the physical characteristics are
similar.
Retention Basin - See Detention Basin.
Revetment - A facing of stone or other material laid on a sloping face of earth to maintain the
slope in position or to protect it from erosion.
Review of Environmental Factors REF (Concept) - Broadly identifies, describes and assesses
environmental advantages, disadvantages and constraints associated with particular broadly
defined routes or corridors during the Concept Phase of the Roads Delivery Program.

March 2010

xii

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Review of Environmental Factors REF (Planning) - Identifies, describes and assesses the
environmental advantages, disadvantages and constraints associated with chosen route options
during the Planning and Preliminary Design Phase.
Riprap - Medium to large size rock protection applied (usually by dumping) to the face of an
embankment, in a waterway or as an outlet protection from a storage.
Road - A route trafficable by motor vehicles; in law, the public right-of-way between
boundaries of adjoining property.
Rock Sediment Trap - Barriers consisting of rock placed in a channel to trap sediment through
the temporary detention of water.
Runoff - That portion of the water precipitated onto a catchment area which flows as surface
discharge from the catchment area past a specified point.
Run-On - Water that accumulates at a site (compared with runoff water that exits a site).
Scour - A term commonly used to mean localised erosion of a bank or channel which typically
occurs due to excessive slope, turbulence or flow velocity.
Sediment Basin - A basin or tank in which stormwater containing settleable solids is retained to
remove by gravity or filtration a part of the suspended matter.
Sediment Curtain - A piece of material, typically geotextile, attached to floats and weights and
extending from the floor of a water body (e.g. sea, lake, river) to the surface, and used to trap
sediments. Also referred to as Silt Curtain.
Sediment Fence - A barrier typically consisting of permeable material stretched between and
attached to supporting posts and entrenched in the earth.
Sediment Trap - Generally, sediment traps are smaller versions of sediment basins.
Sedimentation - Deposition of material of varying size, both mineral and organic, away from
its site of origin by the action of water, wind, gravity or ice.
Seeding - Refers to the establishment of perennial warm-season grasses for the stabilisation of
disturbed soils.
Sheet Flow - Water, usually storm runoff, flowing in a thin layer over the ground surface. Also
referred to as overland flow.
Silt - An alluvial material intermediate in particle size between sand and clay (0.002-0.02mm).
It is usually non-plastic.
Sodicity - A measure of the exchangeable sodium percentage (ESP) of soil material. Soil
material with an ESP of < 6 is referred to as non sodic. Soil material with an ESP of 6-15 is
referred to as sodic. Soil material with an ESP > 15 is referred to as strongly sodic.
Soil Dispersion - The process by which soil aggregates disperse into individual particles (clay,
silt and sand) in water.

March 2010

xiii

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Specific Energy - Is the measure of the total energy of water flow at a particular location. It is
the combination of static, velocity and pressure heads and is measured as a height with the
channel bottom as the datum and expressed in metres of water.
For any cross sectional shape, the specific energy E at a particular section is defined as:

E y V

2g

Where V is the mean velocity, P is the pressure, is the density of water, y is the flow
depth, g is acceleration due to gravity and is the kinetic energy correction factor
which accounts for velocity variations across the section.

Spillway - An open or closed outlet used to convey water from a reservoir or basin. Usually
used to convey a given design runoff.
Sprayed Seal (Flush Seal) - A thin layer of binder sprayed onto a pavement surface with a
layer of aggregate incorporated and which is impervious to water.
Stabilisation (Soil) - The provision of adequate measures (vegetation, mulches, geotextiles, riprap, and other structural measures) to prevent erosion from occurring.
Stabilised Construction Exit - Stabilisation of exposed soil at construction exits to reduce or
eliminate sediment from leaving the construction site. Common materials or controls include
gravel, aggregate cover and timber, and cattle grids.
Stabilised Material (Soil) - A natural material which has been modified to improve or maintain
its load carrying capacity or reduce erosion. Modification may be by the addition of other
natural materials such as sand, loam or clay or of manufactured materials such as bitumen, lime
and cement.
Stream Bank Protection - Measures used to protect existing stream banks from eroding.
Measures may include loose or anchored materials such as large boulders, brush mats,
geotextiles, logs, or concrete.
Structure - Soil structure refers to the size, shape and arrangement of particles and aggregates,
and the size, shape and arrangement of voids or spaces separating the particles and aggregates.
Subcritical Flow - Flow in a channel or conduit which has a Froude number less than 1, a depth
greater than the critical depth and a velocity less than the critical velocity. In practice,
subcritical flows are controlled by the downstream flow conditions.
Subsoil Drain - A drain below the ground surface, which collects subsurface water throughout
its length.
Supercritical Flow - Flow in a channel or conduit which has a Froude number greater than 1, a
depth less than the critical depth and a velocity greater than the critical velocity. In practice,
supercritical flows are controlled from upstream.
Superelevation - (1) The continuous transverse slope normally given to the carriageway at
horizontal curves; (2) The phenomenon where flow around a horizontal curve in an open
channel is at a higher level at the outer edge than at the inner edge of the curve.

March 2010

xiv

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Surcharge Overflow - That portion of the flow which is forced out of a piped system at a gully
inlet, manhole or surcharge structure when the downstream pipe system capacity is exceeded.
Surface Condition - Surface condition refers to the characteristic appearance of the surface soil
when dry. Conditions including cracking, firm, loose and soft.
Swale - A shallow constructed channel, often grass-lined, which is used as an alternative to kerb
and channel, or as a pre-treatment to other measures. Swales are generally characterised by a
broad top width to depth ratio and gentle grades.
Table Drain - The side drain of a road adjacent to the shoulders, having its invert lower than
the subgrade level and being part of the formation.
Temporary Seeding - Refers to the use of soil stabilisation with grasses that will establish
quickly and have longevity of one year or less.
Terrace - A former flood plain on which alluvial deposition and erosion are barely active or
inactive.
Terracing - Grading technique which reduces slope length through the creation of benches.
Tidal Definitions:
(a) Highest Astronomical Tide (HAT) - Highest tide level which can be predicted to occur
under average meteorological conditions and under any combination of astronomical
conditions.
(b) Lowest Astronomical Tide (LAT) - Lowest tide level which can be predicted to occur
under average meteorological conditions and under any combination of astronomical
conditions.
(c) Mean High Water Springs (MHWS) - The long term average of the heights of two
successive high tides when the range of tide is greatest, at full moon and new moon.
(d) Mean Low Water Springs (MLWS) - The long term average of the heights of two
successive low tides when the range of tide is greatest, at full moon and new moon.
(e) Mean High Water Neaps (MHWN) - The long term average of the heights of two
successive high tides when the range of tide is the least, at the time of the first and last
quarter of the moon.
(f) Mean Low Water Neaps (MLWN) - The long term average of the heights of two
successive low tides when the range of tide is the least, at the time of the first and last
quarter of the moon.
(g) Mean Sea Level (MSL) - The average level of the sea over a long period.
(h) Storm Surge -The increase in sea level occurring during a cyclone resulting from the
combines effect of reduced atmospheric pressure and the build up of water against the
shore caused by onshore wind (wind stress).
(i) Wave Setup - The raising of sea level inside the surf zone resulting from the momentum
flux of broken waves.

March 2010

xv

Department of Transport and Main Roads


Road Drainage Manual

Glossary

Time of Concentration - The shortest time necessary for all points on a catchment to contribute
simultaneously to runoff past a specified point.
Trafficability - A road is defined as trafficable when the total head (static plus velocity) across
a carriageway is less than or equal to 300 mm.
Transition Loss Coefficient - Coefficient associated with head losses at open channel
transitions.
Tunnel Erosion - The removal of subsoil by water while the surface remains relatively intact;
also referred to as piping.
Vegetative Protection - Stabilisation of erodible areas through covering with vegetation.
Velocity Head - A measure of the kinetic energy of flow in a pipe or channel and equal to V2/2g
where V is the average velocity of flow.
Water Surface Elevation - The elevation of the water surface reached in a gully inlet, manhole,
junction structure or open channel.
Water Table - (1) In an aquifer, the upper limit of the portion of ground saturated with water;
(2) The natural level at which water stands in a bore-hole or well under conditions of
equilibrium.
Watercourse - A river, creek or stream in which water flows permanently or intermittently in a
natural or artificial channel. For legal definition, refer Water Act 2000.
Waterway - (1) A channel or stream; (2) The area available for water to pass through or under a
structure.

March 2010

xvi

Department of Transport and Main Roads


Road Drainage Manual

References

References

March 2010

Department of Transport and Main Roads


Road Drainage Manual

References

References Amendments Mar 2010


Revision Register
Issue/
Rev
No.

Reference
Section

March 2010

ii

Description of Revision

Initial Release of 2nd Ed of manual.

Authorised
by

Date

Steering
Committee

Mar
2010

Department of Transport and Main Roads


Road Drainage Manual

References

References
The following lists the references used (either cited or consulted) in compiling this manual.
American Association of State Highway and Transportation Officials (AASHTO). 1991. Model Drainage
Manual. Washington: unknown.
American Society of Civil Engineers (ASCE). 1992. Design and Construction of Urban Stormwater
Management Systems. Engineering Practice No 77. New York: American Society of Civil Engineers.
Australian National Committee On Large Dams (ANCOLD). 2000. Guidelines on Selection of Acceptable
Flood Capacity for Dams.
Argue, J. R. 1960. New Structure for Roadway Pipe Culverts. The Journal of the Institution of Engineers,
Australia. Sydney: Institution of Engineers, Australia.
Argue, J. R. 1986. Special Report No 34. Storm Drainage Design in Small Urban Catchments. Australia:
Australian Road Research Board.
Australian and New Zealand Environment and Conservation Council (ANZECC). 2000. Australian and
New Zealand Guidelines for fresh and marine water quality. Australia: unknown.
Australian Road Research Board (ARRB). 1987. Special Report No 35. Subsurface Drainage of Road
Structures. Australia: Australian Road Research Board.
Australian Road Research Board (ARRB). 1998. Part I Biodiversity and Road Drainage Issues, Part II
Practical Rules to Reduce Impacts of Drainage on Ecosystems and Part III Applications of Improved
Drainage Management. Draft. Australia: Australian Road Research Board.
Australian Standard (AS) 1348. 2002. Glossary of terms - Roads and traffic engineering. Australia:
Standards Australia.
Australian Standard (AS) 1726. 1993. Geotechnical site investigations. Australia: Standards Australia.
Australian Standard (AS) 1761. 1985. Helical lock-seam corrugated steel pipes. Australia: Standards
Australia.
Australian Standard (AS) 1762. 1984. Helical lock-seam corrugated steel pipes - Design and installation.
Australia: Standards Australia.
Australian Standard (AS) 2041. 1998. Buried corrugated metal structures. Australia: Standards Australia.
Australian Standard (AS) 3725. 1989. Loads on Buried Concrete Pipes. Australia: Standards Australia.
Australian Standard (AS) 3996. 2006. Access covers and grates. Australia: Standards Australia.
Australian Standard (AS) 4058. 2007. Precast concrete pipes (pressure and non-pressure). Australia:
Standards Australia.
Australian Standard (AS) 4685.1. 2004. Playground equipment. Australia: Standards Australia.
Australian Water and Coastal Studies (AWACS). 1990. Smeaton Grange Industrial Development Hydraulic Model Testing of Drop Structure. Report 90/15. Sydney: unknown.
Austroads. 1994. Waterway Design - A guide to the hydraulic design of bridges, culverts and floodways.
Sydney: Austroads.
Austroads. 2005. Guidelines for the management of road surface skid resistance. Sydney: Austroads Inc.
Austroads. 2008a. Guide to Road Design. Part 5: Drainage Design. Sydney: Austroads Inc.
Austroads. 2008b. Guide to Road Design, Part 7: Geotechnical Investigation and Design. Sydney:
Austroads Inc.

March 2010

iii

Department of Transport and Main Roads


Road Drainage Manual

References

Ball, J.E., Jenks, R. and Aubourg, D. 1998. An assessment of the availability of pollutant constituents on
road surfaces. The Science of the Total Environment, 209: 243-254.
Barnes, D., P. J. Bliss, B. W. Gould, and H. R. Vallentine. 1981. Water and Wastewater Engineering
Systems. Great Britain: Pitman.
Beecham, S. C. and S. J. Sablatnig. 1994. Hydraulic modelling of stormwater trash racks. Proceeding of
the International Conference on Hydraulics in Civil Engineering, NCP 94/1, 97-104.
Blaisdell, F. W. 1948. Development and Hydraulic Design, The SAF Stilling Basin. Report SCS-TP-79.
United States of America: US Soil Conservation Service.
Blaisdell, F. W. 1948. Development and Hydraulic Design, Saint Antony Falls Stilling Basin.
Transactions of the American Society of Civil Engineers, Volume 113. United States of America:
ASCE.
Blaisdell, F. W. 1949. The SAF Stilling Basin. Technical Bulletin No.79. United States of America: US
Soil Conservation Service.
BMT WBM 2007. State-wide Review of the Use of Road Runoff Treatment Devices and Water
Monitoring Programs. Brisbane: Queensland Department of Main Roads & BMT WBM Pty Ltd.
Bohan, J. P. 1970. Erosion and Riprap Requirements at Culvert and Storm-Drain Outlets. US Army
Engineer Waterways Experiment Station, Research Report H-70-2. Vicksburg, Mississippi: US Army.
Brater, E. F., H. W. King, J. E. Lindell and C. Y. Wei. 1996. Handbook of Hydraulics (Seventh Edition).
USA: McGraw-Hill
Brisbane City Council (BCC). 1996. Environmental Best Management Practice (Volume 1) for Erosion
and Sediment Control and for Waterways and Wetlands. Brisbane: Brisbane City Council.
Brisbane City Council (BCC). 2001. Sediment Basin Design, Construction and Maintenance Guidelines.
Brisbane: Brisbane City Council.
Bureau of Public Roads. 1965. Hydraulic Charts for the Selection of Highway Culverts. Washington DC:
US Department of Commerce.
Bureau of Reclamation. 1960. Design of Small Dams. Washington DC: US Department of the Interior.
Bureau of Reclamation. 1964. Hydraulic Design of Stilling Basins and Energy Dissipaters, Engineering
Monograph No.25. Washington DC: US Department of the Interior.
Cameron and McNamara. 1966. Report on Model Investigation of Causeway Design for Commonwealth
Department of Works, Darwin and Queensland Main Roads Department. Unknown:
Camp, Dresser and McKee. 1993, California Storm Water Best Management Practice Handbooks:
Municipal. (prepared for California Stormwater Quality Task Force). Unknown:
Chiew, F. H. S., L. B. Mudgway, H. P. Duncan, and T. A. McMahon. 1997. Urban stormwater pollution.
Industry Report 97/5. Australia: CRC for Catchment Hydrology.
Chiu, A and W. R. Rahmann. 1980. Drainage Design and Outlet Protection. Brisbane: Department of
Main Roads.
Chow, V. T. 1959. Open Channel Hydraulics. New York: McGraw-Hill Book Company Inc.
Cotterell, E. 1998. Fish Passage in Streams: Fisheries guidelines for design of stream crossings.
Brisbane: Department of Primary Industries.
Cottman, N. H., and G. R. Mackay. 1990. Bridges and Culverts Reduced in Size and Cost by Use of
Critical Flow Transitions. Proceedings Institution Civil Engineers (London), Part 1, Volume 88.
Commonwealth Scientific and Industrial Research Organisation CSIRO. 1989. Greenhouse: Planning for
Climate Change. Melbourne: CSIRO Publishing.
Dash, D. M. 1977. Life Cycle Study F3 Sydney to Newcastle Highway, Surface Characteristics. Australia:
Roads and Transport Technology.
Dash, D. 2006. Aquaplaning literature review. Sydney: Roads and Traffic Authority (NSW).

March 2010

iv

Department of Transport and Main Roads


Road Drainage Manual

References

Department of Agricultural and Biological Engineering. 1996. Reducing Sediment Discharge from
Sedimentation Basins with Barriers and a Skimmer. Pennsylvania: Pennsylvania State University.
Department of Agricultural and Biological Engineering. 1996. Sedimentation Basins Evaluation and
Design Improvements. Pennsylvania: Pennsylvania State University.
Department of Agriculture, Soil Conservation Service (USA). 1975. Standards and Specifications for Soil
Erosion and Sediment Control in Developing Areas. Maryland:
Department of Conservation and Land Management (NSW). 1992. Urban Erosion and Sediment Control.
Sydney:
Department of Environment (QLD). 1997. A User's Guide to Queensland Environmental Protection
(Water) Policy. Brisbane:
Department of Employment, Economic Development and Innovation (QLD). 2009. Fish Habitat
Management Operational Policy (FHMOP 008) Waterway barrier works development approvals.
Brisbane: Queensland Fisheries
Department of Employment, Economic Development and Innovation (QLD). 2009. Fish Habitat
Management Operational Policy (FHMOP 008) Waterway barrier works development approvals.
Brisbane: Queensland Fisheries
Department of Housing (NSW). 1998. Managing Urban Stormwater - Soils and Construction (Third
Edition). Sydney:
Department of Land and Water Conservation (NSW). 1998. The constructed wetlands manual, Volume 1
and 2 Department of Land and Water Conservation. Sydney:
Department of Main Roads (QLD) (DMR). 1997a. Erosion and Sediment Control Manual. Brisbane:
Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 1997b. Roads in the Wet Tropics - Planning, Design,
Construction, Maintenance and Operation Best Practice Manual. Brisbane: Queensland Department
of Main Roads.
Department of Main Roads (QLD) (DMR). 1997c. Road Landscape Manual. Brisbane: Queensland
Department of Main Roads.
Department of Main Roads (QLD) (DMR). 1998. Cultural Heritage Manual. Brisbane: Queensland
Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2000. Fauna Sensitive Road Design, Volume 1 - Past and
Existing Practices. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2001a. Queensland Environmental Legislation Register, 2nd
edition. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2001b. Road Maintenance Contracts, Volume 2: Open
Competition. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2002a. Asset Maintenance Guidelines. Brisbane: Queensland
Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2002b Road Drainage and Design Manual. Brisbane:
Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2003. Manual of Uniform Traffic Control Devices. Brisbane:
Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2004a. Road Maintenance Contracts, Volume 3: Guidelines
for undertaking maintenance. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2004b. Road Project Environmental Processes Manual.
Brisbane: Queensland Department of Main Roads.

March 2010

Department of Transport and Main Roads


Road Drainage Manual

References

Department of Main Roads (QLD) (DMR). 2005a. Preconstruction Processes Manual. Brisbane:
Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2005b. Road construction erosion & sediment control
workshop. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2006. Road Planning & Design Manual (RPDM) Bridges and
Retaining Walls (Chapter 22). Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2007a. Road Maintenance Contracts, Volume 1: Sole Invitee.
Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2007b. Road Traffic Noise Management: Code of Practice.
Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2008a. Guidelines for Strategic Road Network Planning.
Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2008b. Macrotexture and Microtexture for Road Surfacings Development of Interim Initial Standards. Brisbane: Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2009a. Project Cost Estimating Manual. Brisbane:
Queensland Department of Main Roads.
Department of Main Roads (QLD) (DMR). 2009b. Standard Drawings (Roads) Manual. Brisbane:
Queensland Department of Main Roads.
Department of Natural Resources & Water (QLD) (NR&W). 2007. Guidelines on Acceptable Flood
Capacity for Dams. Brisbane: Department of Natural Resources & Water
Department of Natural Resources & Water (QLD) (NR&W). 2008. Queensland Urban Drainage Manual
(QUDM). Brisbane: Department of Natural Resources & Water
Department of Primary Industries (DPI) (QLD). 1978. Soil Conservation Handbook. Brisbane:
Department of Primary Industries (DPI) (QLD). 1992. Soil Conservation Handbook Part 9 - Design of
Structures. Brisbane:
Department of Primary Industries (DPI) (QLD). 1998. Fisheries, Fish Passage in Streams - Fisheries
Guidelines for Design of Stream Crossings. Brisbane:
Department of Transport. 1978. Charts for the Hydraulic Design of Channels and Pipes. London, UK:
HMSO.
Department of Transport (QLD). 1975. Urban Road Design Manual - Volume 2. Brisbane:
Department of Transport and Main Roads (QLD) (TMR). 2010a. Road Planning & Design Manual - A
guide to QLD practice. Brisbane: Queensland Department of Transport and Main Roads.
Department of Transport and Main Roads (QLD) (TMR). 2010b. Soils Manual. Brisbane: Queensland
Department of Transport and Main Roads.
Department of Transport and Main Roads (QLD) (TMR). 2010c. Standard Specification Roads.
Brisbane: Queensland Department of Transport and Main Roads.
Department of Transport (USA). 1972. Hydraulic Design of Improved Inlets for Culverts, Hydraulic
Engineering Circular No.13. Washington DC:
Department of Transport (USA). 1983. Hydraulic Design of Energy Dissipaters for Culverts and
Channels. Hydraulic Engineering Circular No.14. Washington DC:
Department of Transportation (USA). 1995. Best Management Practices of Erosion and Sediment
Control. Washington DC:

March 2010

vi

Department of Transport and Main Roads


Road Drainage Manual

References

Department of Urban Services (ACT). 1994. Urban Stormwater: Standard Engineering Practices.
Canberra:
Drapper, D., A. Bell, and R. Tomlinson. 1997. Analysing Runoff from Road Surfaces; Formulating a
Baseline for the Queensland Department of Main Roads. Storm Water and Soil Erosion 97
Conference, Brisbane.
Earley, P. C. 1979. Gully Inlet Spacing Design. Australia: University of Western Australia.
Emerson, W. W. 1967. A classification of soil aggregates based on their coherence in water. Australian
Journal of Soil Research No.5, 47-57.
Emerson, W. W. 1977. Determination of the contents of clay sized particles in soils. Journal of Soil
Science No.22, 50-59.
Engineers Australia. 2006. The Australian Runoff Quality A guide to Water Sensitive Urban Design.
Australia: Engineers Australia.
Environmental Protection Agency (NSW). 1997. Treatment techniques, Managing urban stormwater
series. Sydney:
Environmental Protection Agency (QLD). 2006. Queensland Water Quality Guidelines (QWQG).
Brisbane: EPA
Environmental Protection Agency (QLD). 2008. Best Practice Urban Stormwater Management - Erosion
and Sediment Control Guideline. Brisbane: EPA
eWater (2008). Catchment Modelling Toolkit, (http://www.toolkit.net.au/music). Canberra: eWater
Cooperative Research Centre (CRC).
Frankcombe, M. and R. Coventry. 2001. Townsville City Council, Revegetation Contractors, Natural
Resource Assessment, Catchments & Creeks. Erosion and Sediment Control Planning for North
Queensland. Unknown:
Franzini, J. B. and E. J. Finnemore. 1997. Fluid Mechanics with Engineering Applications (Nineth
Edition). USA: WCB McGraw-Hill
Freeman, G and L. Howells. 1995. Saline Flocculation in Stormwater Quality Management. Proceedings
of the 3rd Annual Conference on Soil and Water Management for Urban Development. Sydney:
International Erosion Control Association of Australasia.
Foucard, J. 2005. Role of tires in the skid resistance phenomenon. France:
Gallaway, B. M, R. E. Schiller Jr and J. G. Rose. 1971. The Effects of Rainfall Intensity, Pavement Cross
Slope, Surface Texture and Drainage Length on Pavement Water Depths. Texas: Texas
Transportation Institute
Gallaway, B. M., G. G. Hayes, D. L. Ivey, W. B. Ledbetter, R. M. Olsen, H. E. Ross Jr, R. E. Schiller Jr,
and D. L. Woods. 1979. Pavement and Geometric Design Criteria for Minimising Hydroplaning.
Washington, DC: Federal Highway Administration.
Gillespie, T.D. 1992. Fundamentals of Vehicle Dynamics. Warrendale, PA: Society of Automotive
Engineers, Inc.
Goldrick, D. A., I. Yassini, and B. G. Jones. 1996. Flocculation Behaviour of West Dapto and Albion
Park Soils Types in Urban Runoff Treatments. Proceedings of the 4th Annual Conference on Soil and
Water Management for Urban Development. Sydney: International Erosion Control Association of
Australasia.
Gothi, M. 2005. Water Influence on Skid Resistance, Surface Friction: Roads & Runways Conference, 14 May 2005, New Zealand.
Griffith University and RUST PPK. 1997. Erosion Treatment for Urban Creeks, Guidelines for Selecting
Remedial Works Guidelines for Brisbane City. Brisbane:
Hare, C. M. 1983. Magnitude of Hydraulic Losses at Junctions in Piped Drainage Systems. Australian
Transactions, Volume CE25. Australia: Institution of Engineers.

March 2010

vii

Department of Transport and Main Roads


Road Drainage Manual

References

Hee, M. 1969. Hydraulics of Culvert Design Including the Constant Energy Concept. Queensland Local
Government Engineers' Conference, Brisbane.
Hee, M. 1978a. Hydraulic Calculations for Bridges and Culverts. 2nd Conference, Road Engineering
Association of Asia and Australasia. Manila, Philippines.
Hee, M. 1978b. Selected Case Histories. Workshop on Minimum Energy Design of Culvert and Bridge
Waterways. Australian Road Research Board.
Hee, M. 1993. Practical and Legal Aspects of Floodway Design. First Annual Conference, the Institution
of Municipal Engineers, Queensland Division.
Henderson, F. M. 1970. Open Channel Flow. United States of America: Prentice Hall
Higgins, R. G. and R. J. Campbell. 1997. Pacific Motorway (Logan to Nerang) - Case Study:
Management of Water Quality Impacts. Brisbane: Department of Main Roads.
Highway Department. 1962. Hydraulic Manual. Texas:
Highway Research Board. 1971. Design of Culverts, Energy Dissipaters and Filter Systems, Record
No.373. Washington DC:
Horner, R. R., J. J. Skupien, E. H. Livingston, and H. E. Shaver. 1994. Fundamentals of urban runoff
management. Washington DC: Terrene Institute.
Horne, W. B. 1968. Tire Hydroplaning and its Effects on Tire Traction. HRR. No. 214.
Hunt, J. S. 1992. Urban Erosion and Sediment Control. (revised edition). Sydney: Department of
Conservation and Land Management.
Ibrahim, A and F. Hall. 1994. Effect of Adverse Weather Conditions on Speed-Flow-Occupancy
Relationships. TRR No. 1457
IECA (Australasia) 2008. Best practice erosion and sediment control. International Erosion Control
Association (Australasia).
Institution of Engineers, Australia (IEAust). 1977. Australian Rainfall and Runoff Flood Analysis and
Design. Barton, ACT: Institution of Engineers, Australia
Institution of Engineers, Australia (IEAust). 1985. Guidelines for the Design of Detention Basins and
Grassed Waterways for Urban Drainage Systems. Sydney:
Institution of Engineers, Australia (IEAust). 1987a. Australian Rainfall and Runoff - Guide to Flood
Estimation, Vol 1. Barton, ACT: Institution of Engineers, Australia
Institution of Engineers, Australia (IEAust). 1987b. Australian Rainfall and Runoff - Guide to Flood
Estimation, Vol 2. Barton, ACT: Institution of Engineers, Australia
Institution of Engineers, Australia (IEAust). 1996. Soil Erosion and Sediment Control Engineering
Guidelines for Queensland Construction Sites. (prepared by G. Witheridge and R. Walker). Brisbane:
Institution of Engineers, Australia (IEAust). 2001. Australian Rainfall and Runoff - Guide to Flood
Estimation, Vol 1. Barton, ACT: Institution of Engineers, Australia
Isbell, R.F. 2002. The Australian soil classification. Revised edition, CSIRO Publishing.
James Cook University of North Queensland (JCU). 1977. Numerical Simulation of Tropical Cyclone
Storm Surge Along the Queensland Coast, Parts I to X. (prepared for Beach Protection Authority,
Department of Environment and Heritage). Townsville:
Keim, S. R. 1962. The Contra Costa Energy Dissipater. Journal of the Hydraulic Division, Proceedings
of the American Society of Civil Engineers, Vo188, No.HY2.
Korom, S. F., S. Sarikelle and A. L. Simon. 1990. Design of Hydraulic Jump Chamber. Journal of the
Hydraulic Division, Proceedings of the American Society of Civil Engineers, Vol 116, No.HY2.
Lawrence, I. and Breen, P. 1998. Design Guidelines: Stormwater Pollution Control Ponds and Wetlands.
Canberra: Cooperative Research Centre (CRC) for Freshwater Ecology.

March 2010

viii

Department of Transport and Main Roads


Road Drainage Manual

References

Landcom 2004. Managing urban stormwater: Soils and construction. (Fourth edition). Landcom NSW.
Local Government and Shires Association (NSW). 1982. Residential Standards Manual. Sydney:
Loch, R. J., B. K. Slater, and C. Devoil. 1998. Soil erodibility (Km) values for some Australian soils.
Australian Journal of Soil Research.
Maccaferri Gabions (London). 1973. River and Sea Gabions, London:
MacDonald, T. C. 1969. Energy Dissipaters for Large Culverts. Journal of the Hydraulic Division,
Proceedings of the American Society of Civil Engineers, Vo195, No.HY6.
Matthews, M. 2007. Manual for Erosion and Sediment Control Version 1.1. Maroochy Shire Council,
Qld.
McDonald, R. C., R. F. Isbell, J. G. Speight, J. Walker, and M. S. Hopkins. 1990. Australian soil and
land survey field handbook. (Second edition). Australia: Inkata Press.
McDonald, R. C., R. F. Isbell, J. G. Speight, J. Walker, M. S. Hopkins, L. J. Gregory, R. J. Hnatiuk and
R. Thackway. 2009. Australian Soil and Land Survey Field Handbook. (Third edition). Australia:
CSIRO Publishing.
Meredith, D.D. 1975. Model Study of Culvert Energy Dissipater. Journal of the Hydraulic Division,
Proceedings of the American Society of Civil Engineers, Vol 101, No.HY3.
Ministry of Works and Development (NZ). 1978. Culvert Manual. Christchurch:
Mockmore, C. E. 1944. Flow Round Bends in Stable Channels. Transactions of the American Society of
Civil Engineers, Volume 109. United States of America: ASCE.
Moore, D. F. 1975. The friction of pneumatic tyres. Amsterdam: Elsevier Scientific Publishing Company.
Morse, McVey and Associates. 1993. Soil and Water Management for Urban Development. (prepared for
New South Wales Department of Housing). Sydney:
Moreton Bay Waterways and Catchments Partnership (MBWCP). 2006. Water Sensitive Urban Design
Technical Design Guidelines for South East Queensland. Brisbane: Moreton Bay Waterways and
Catchments Partnership
Mudgway, L. B., H. P. Duncan, T. A. McMahon and F. H. S. Chiew. 1997. Best Practice Environment
Management Guidelines for Urban Stormwater. Australia: CRC for Catchment Hydrology.
National Association of Australian State Road Authorities (NAASRA). 1983. Guide to the Control of
Moisture in Roads. Sydney: NAASRA
National Association of Australian State Road Authorities (NAASRA). 1986. Guide to the design of Road
Surface Drainage. Sydney: NAASRA
National Association of Australian State Road Authorities (NAASRA). 1989. Bridge Waterways
Hydrology and Design. Sydney: NAASRA
National Capital Development Commission. 1988. Design Manual for Urban Erosion and Sediment
Control. Unknown:
Neill, C. R. 1973. Guide to Bridge Hydraulics. Canada: Roads and Transportation Association of Canada.
New South Wales Soil Conservation Service. 1998. Construction Site Erosion and Sediment Control
Course Notes. Sydney: (authors: A. King, D. Greentree and D. Height)
North Carolina Department of Transportation. 1992. Erosion and Sediment Control Guidelines for
Division Maintenance Operations. Raleigh, North Carolina:
North Carolina Department of Transportation. 1995. Erosion and Sediment Control Guidelines for
Contract Construction. Raleigh, North Carolina:
North Carolina Sediment Control Commission. 1992. Erosion and Sediment Control - Inspector's Guide.
Raleigh, North Carolina:

March 2010

ix

Department of Transport and Main Roads


Road Drainage Manual

References

North Carolina Sediment Control Commission. 1993. Erosion and Sediment Control Planning and
Design Manual, 1993. Raleigh, North Carolina:
O'Loughlin, E. M. 1960. Culvert Investigations by Hydraulic Models, Harbours and Marines Branch,
Hydraulic Laboratory. Sydney: Department of Public Works.
Oliver, J. W. H. 1979. Surface Drainage of Road Pavements, the Problem. Australia: ARRB.
Orange County Planning Department. 1989. Soil Erosion and Sediment Control Manual. Hillsborough,
North Carolina:
Peterka, A.J. 1984. Hydraulic Design of Stilling Basins and Energy Dissipaters. USBR Engineering
Monograph 25. Denver: US Bureau of Reclamation.
Premier's Department (QLD). 1989. The Greenhouse Effect: Implications for Queensland A Discussion
Paper. Brisbane:
Queensland Fisheries (QF). 2009. Fish Habitat Management Operational Policy (FHMOP 008)
Waterway barrier works development approvals. Brisbane: Queensland Fisheries (part of Department
of Employment, Economic Development and Innovation)
Rand, W. 1955. Flow Geometry at Straight Drop Spillways, Transactions of the American Society of
Civil Engineers. Volume 81, No 791. United States of America: ASCE.
Rice, C. E. and K. C Kadavy. 1991. HGL Elevation at Pipe Exit of USBR Type VI Impact Basin. Journal
of the Hydraulic Division, Proceedings of the American Society of Civil Engineers, Volume 117,
No.HY7.
Ritchie, J.A. 1963. Earthwork Tunnelling and the Application of Soil Testing Procedures. Journal of the
Soil Conservation Service of New South Wales, No.19, 111-129.
Roads and Traffic Authority (NSW). 1993. Road Design Guide - Section 8 "Erosion and Sedimentation".
Sydney:
Roads and Traffic Authority (NSW). 1994. Draft Guidelines for Limiting the Risk of Aquaplaning on
Roads. Sydney:
Roads and Traffic Authority (NSW). 1995. Water Quality Management. Sydney:
Roads and Traffic Authority (NSW). 1996. Road Design Guidelines. Sydney:
Roads and Transportation Association of Canada. 1982. Drainage Manual, Volume 1. Canada:
Rosenthal, K. M. and B. J. White. 1980. Distribution of a rainfall erosion index in Queensland. Land
Utilisation Report 80/8. Australia: Queensland Department of Primary Industries.
Rosewell, C. J. and J. B. Turner. 1992. Rainfall Erosivity in New South Wales. CaLM Technical Report
No.20, NSW. Department of Soil Conservation and Land Management. Sydney:
Rosewell, C. J. and R. J. Loch. 2002. Estimation of the RUSLE soil erodibility factor, in soil physical
measurement and interpretation for land evaluation. (McKenzie, N., Coughlan, K. and Cresswell,
H.) of CSIRO Publishing.
Ross, N. F. and K. Russam. 1968. The Depth of Rain Water on Road Surfaces. Berkshire: Road Research
Laboratory.
Rouven, N., R-M. Li and D. B. Simons. 1981. Flow Resistance in Vegetated Waterways. United States of
America: American Society of Agricultural Engineers.
Sangster, W. M., H. W. Wood, E. T. Smerdon and H. G. S. Bossy. 1958. Pressure Changes at Storm
Drain Junctions, Engineering Series Bulletin No 41, Engineering Experimental Station. United States
of America: University of Missouri.
Schueler, T. R., P. A. Kumble and M. A. Heraty. 1992. A Current Assessment of Urban Best Management
Practices: Techniques for Reducing Non-point Source Pollution in the Coastal Zone. United States of
America: US Environment Protection Authority.

March 2010

Department of Transport and Main Roads


Road Drainage Manual

References

SMEC 2008. Literature Review: Road Runoff Quality and Treatment Technologies. Queensland
Department of Main Roads & SMEC.
Smith, C. D. and J. Yu. 1966. Use of Baffles in Open Channel Expansions. Journal of the Hydraulic
Division, Proceedings of the American Society of Civil Engineers, Vo192, No.HY2, plus various
discussions: Skogerboe, G. V., M. Leon Hyatt and M. M. Soliman. 1966. ASCE No.HY5, 255-261;
Rao, B. V., V. J. Galay and W. H. R. Nimmo. 1966. ASCE No.HY6, 212-215; Dake, J. M. K., S. Kar,
S. Raghunathan, V. C. Kulandaiswamy and M. Narayanan. 1967. ASCE No.HYl, 78-85; Smith, C. D.
and J. Yu. 1967. ASCE No.HY4, 273-275.
Speight, J. G. S. 1990. Landform, in Australian soil and land survey field handbook (second edition).
Australia: Inkata Press.
Vallentine, H. R., R. T. Hattersley and B. A.Cornish. 1961. Low Cost Scour Control at Culvert Outlets
Reports 48 and 62. Sydney: University of New South Wales.
VicRoads. 1995. Road Design Guidelines, Part 7, Drainage. Melbourne: VicRoads.
Water Resources Commission. 1998. A Manual for Managing Urban Stormwater Quality in Western
Australia. Perth: Water Resources Commission.
Willing and Partners Pty Ltd. 1992. Design guidelines for gross pollutant traps. (prepared for ACT
Department of Environment, Land and Planning and the Department of Urban Services). Canberra:
Wischmeier, W. H. and D. D. Smith. 1978. Predicting rainfall erosion losses, a guide to conservation
planning, USDA Agriculture Handbook Number 537. United States of America: unknown.
Witheridge, G. M. and R. B. Tomlinson. 1995. Water Under the Bridge, Engineering Update, Institution
of Engineers, Australia, Queensland Division, Technical Papers. Brisbane: IEAust.
Witheridge, G. M. 1997a. Design of Energy Dissipators: Engineering Update Volume 5 No.2. Brisbane:
IEAust.
Witheridge, G. M. 1997b. Design of Energy Dissipators. Technical Paper 3, Engineering Update,
Institution of Engineers, Australia, Queensland Division, pp.3-10. Brisbane: IEAust.
Witheridge, G. S. 1997c. Erosion and Sediment Control in Road Design and Construction.
Wong, T. H. F. 2000. Improving Urban Stormwater Quality From Theory to Implementation. Water
Journal of the Australian Water Association, Vol. 27 No. 6, November/December, 2000, pp. 28-31.
Wong, T., T. Fletcher, H. Duncan, J. Coleman, and G. Jenkins. 2002. A Model for Urban Stormwater
Improvement Conceptualisation (MUSIC) (Version 1.00). Melbourne: CRC for Catchment Hydrology.
Wright-McLaughlin Engineers. 1969. Urban Storm Drainage - Criteria Manual, Volume 2. Denver:

Yeager, R. W. 1974. Tire Hydroplaning: Testing, Analysis and Design. Plenum Press.

March 2010

xi

You might also like