You are on page 1of 242

Flood Risk Management (Scotland) Act 2009

Methods to Screen and Quantify Natural


Flood Management Effects
Scottish Environment Protection Agency & Forestry
Commission Scotland, May 2012

Version history
Issue

Date

Author

Draft issue for comment

16/03/12

Neil Nutt CEng MICE Halcrow Group Ltd


16 Abercromby Place, Edinburgh, EH3 6LB

Finalised report

11/05/12

Neil Nutt CEng MICE Halcrow Group Ltd


16 Abercromby Place, Edinburgh, EH3 6LB

Acknowledgements
Section 20 Technical Advisory Group
Andrea Johnstonova (Chair)

SEPA

Drew Aitken

SEPA

Heather Forbes

SEPA

Julia Garritt

Forestry Commission Scotland

Lorna Harris

SEPA

Kirsty Jack

SEPA

Richard Jefferies

SEPA

Roy Richardson

SEPA

David Scott

SEPA

Nadeem Shah

Forest Research

Mark Williams

SEPA

Dominic Habron

SEPA Corresponding Member

Mark McLaughlin

SEPA Corresponding Member

External Review by CREW


CREW (Centre of Expertise for Waters) is a hub which ensures that water research
and expertise is available and accessible to the Scottish Government and its
agencies. This is designed to ensure that existing and new research and expertise
can feed into the development of water related policy in Scotland in a timely and
effective manner.
The Author and Technical Advisory Group would like to thank Prof Alan Werritty
(Dundee University), Dr Scott Arthur (Heriot-Watt University) and Dr Tom Ball
(Dundee University) for their detailed review and subsequent guidance within a tight
timescale during the development of this report.

Executive summary
The Flood Risk Management (Scotland) Act 20091 (the FRM Act) introduces a new
sustainable approach to managing flood risk and places new duties on SEPA and
Responsible Authorities. One responsibility placed on SEPA is to assess and
consider the role which Natural Flood Management can play in reducing flood risk.
Natural Flood Management was defined by SAIFF (2011)2 as:
Natural Flood Management can be defined as those techniques that aim to work
with natural hydrological and morphological processes, features and characteristics
to manage the sources and pathways of flood waters. These techniques include the
restoration, enhancement and alteration of natural features and characteristics, but
exclude traditional flood defence engineering that works against or disrupts these
natural processes.
The aim of this study is to advance development of the methodology for assessing
the contribution that Natural Flood Management can make to managing flood risk.
This report builds on previous work undertaken by SEPA, Jacobs and others in the
field of Natural Flood Management. The main tasks have included:

Review the previous work by Jacobs and the proposed GIS method for
assessing Natural Flood Management;

Collate and review additional literature on Natural Flood Management,


focusing on non-fluvial Natural Flood Management techniques;

Collate and review existing methods of assessing Natural Flood Management


techniques; and

Propose a way forward for development of methods to identify opportunities


and appraise the contribution of range of Natural Flood Management
techniques.

As part of the delivery of Section 20 of the FRM Act SEPA appointed Jacobs to
develop a methodology to assess Natural Flood Management measures (Jacobs,
2011)3. The Jacobs method was founded on a comprehensive and up-to-date
literature review (Werritty et al., 2012)4. Unfortunately this methodology is associated
with licensing issues, technical difficulties and a catchment based approach which
did not allow national identification of opportunities led to this tool not being adopted
by SEPA.
In this study SEPA has sought to build on the Jacobs literature review by extending it
to include all flood mechanisms (pluvial, fluvial, groundwater, coastal and urban) and
by expanding it to include a wider literature base. The extended literature review has
formed the foundation for the development of more robust methods for screening and
assessing Natural Flood Management measures.
Section 4 of the report proposes methods for identifying and appraising Natural Flood
Management opportunities. The methods will provide an overview of the effect of
Natural Flood Management measures that will inform the appraisal process. It is
crucial that these predictions are kept in context. At no point should these estimates
be taken as the correct answer, rather they should be seen as an approximate guide
to what potentially might happen within the bounds of the reported uncertainty.

-i-

It is proposed that the identification and appraisal of Natural Flood Management


measures will be undertaken using screening and assessment tools as summarised
in Figure 1.

Section 20 Identification of potential


Running of Screening Tools
Output: Maps showing the potential for Natural
Flood Management in catchments with
Potentially Vulnerable Area

2012
SEPA

Output: Section 20 maps

Section 28 Appraisals
Using output of Screening in combination with
catchment characteristics and further
information to develop a long list of measures
and refinement to short list. More detailed
appraisal producing a list of preffered
measures.

2013 - 2014
SEPA / Responsible
Authorities

Output: A list of preferred FRM measures for


inclusion in FRM Strategies

Section 34 Local FRM Plans


Development of the preferred measure using
appropriate modelling and assessment tools
(not exclusively the proposed hydrological
assessment tool)

2015 16
The relevant
Responsible Authority

Output: Local FRM Plans


Figure 1: Outline process for the delivery of Natural Flood Management requirements
under the FRM Act

It is proposed that the identification and appraisal of Natural Flood Management will
include three phases of work:
1.
Identification phase: A national screening process to identify opportunities for
Natural Flood Management measures. This screening process will identify areas
within catchments with Potentially Vulnerable Areas that have natural flood
management opportunities but will make no consideration of constraints or other
benefits. The screening process will not directly recommend which specific measure
should be implemented where, nor will the screening facilitate the quantification of
the flood risk management benefits of undertaking a specific natural flood
management activity. However the process will facilitate the identification of areas
that are worth further investigations at a later stage. The main output of the
screening process will be six maps showing:

-ii-

Areas of high runoff generation;

Areas of floodplain storage potential;

Opportunities to remove hydraulic constrictions;

Areas of heightened hydromorphological activity;

Areas of estuarine surge attenuation potential; and

Wave energy attenuation potential.

It is proposed that these maps, will meet the requirements of Section 20 of the FRM
Act and they will be made publically available along with appropriate supporting
descriptions and guidance.
2.
Appraisal phase: The second phase will seek to bring together the outputs
from the screening process with further information about catchment characteristics
in order to develop better understanding of what may be achievable where within the
catchment. Using the findings of the screening process in combination with other
catchment information the measures will be presented as options to manage the
sources and pathways of floodwaters, as summarised in the Natural Flood
Management Summary Table (Appendix B). This information together will be used to
appraise natural flood management measures alongside other flood risk
management measures. Over the course of this stage the long list will be reduced to
a short list and the short list appraised to identify a basket of the most sustainable
measures. The process will give due consideration to environmental, economic and
social costs and benefits. The end result of this phase is a short list of options that
will be agreed by all responsible authorities.
3.
Further development through local FRMPs: The third phase will include more
detailed assessment of the agreed, measures using existing and new assessment
approaches. This more detailed assessment will be carried out by the relevant
responsible authority as part of a local flood risk management plan.
The literature review has identified that some measures can be assessed using
existing tools and methods. It is therefore proposed that where possible, this further
assessment is carried out using existing assessment tools. However, it is also
proposed that a new hydrological assessment method is developed to facilitate the
quantification of fluvial and pluvial Natural Flood Management measures. This study
has investigated how this assessment could be conducted and the report presents a
proposed method for how this could be accomplished.
The proposed hydrological assessment methodology is a single event spatially
distributed model which builds on the uniformly distributed PDM (Probability
Distributed Model) runoff generation model. In the absence of viable alternatives
many elements of the PDM configuration are based on the ReFH (Revitalised Flood
Hydrograph). The effect of land use change is included within the PDM runoff
generation model through the inclusion of a canopy interception moisture store and
variation of the antecedent moisture condition based on the vegetations moisture
demand. There is also provision to assess the effect of soil degradation on the soil
moisture capacity as a consequence of intensive land management practices.

-iii-

Flow routing within the hydrological model is based on a time to outlet grid which is
generated for a representative bankfull flood event. This grid summarises the travel
time for runoff to travel from any location to the catchment outlet. It facilitates the
generation of a hydrograph at the catchment outlet by the summation of the runoff
generated for the array of PDMs covering the catchment. This approach allows the
effect of any number of diffuse land management changes to be accumulated over a
wide range of catchment sizes. The 1D kinematic wave hydraulic routine within the
flow routing engine permits the quantification of land cover roughness changes prior
to flood water reaching the watercourses in addition to in channel modifications such
as the increase in riparian vegetation or restoration of a watercourses planform.
Although not included within the core of the proposed flow routing method, there is
sufficiently flexible to allow the time to outlet grid to be calculated using an alternative
method where appropriate. This flexibility permits the use of more detailed 1D-2D
hydraulic models to test the time lag effects of floodplain interventions such as the
planting of trees on floodplains to increase roughness.
There is currently no software available to assist in the application of the fluvial
hydrological assessment methodology proposed by this study. Testing undertaken
as part of this study has identified that it would be feasible to construct it as a
Toolbox in ArcGIS or similar. If or when a tool is developed it will be necessary to
undertake a comprehensive calibration and verification process using a large
selection of catchments across Britain. It is essential that the calibration and
verification process considers catchments with a wide range of land use, soils,
rainfall, potential evaporation, size, topography and network shapes. The calibration
process should not solely focus on recreating the reported effects at a small number
of Natural Flood Management test catchments which are likely to only represent the
conditions found within a small proportion of catchments which the methodology
might ultimately be applied to.

-iv-

Table of Contents
Executive summary.................................................................................................. i
1

Introduction ..................................................................................................... 1

Review of Jacobs Methodology..................................................................... 5


2.1

Introduction ................................................................................................ 5

2.2

Review of Jacobs Section 20 Review (CREW, 2012)................................. 5

2.2.1

Comments relating to literature review ............................................... 6

2.2.2

Comments relating to Section 20 assessment tool ............................. 6

2.3
Natural Flood Management Implementation Learning from Practice
Workshop Report (SNIFFER & SEPA, 2011) ........................................................ 7

2.4

Additional comments.................................................................................. 7

2.5

Summary ................................................................................................. 12

Literature review ............................................................................................ 13


3.1

Introduction .............................................................................................. 13

3.2

Review of previous literature reviews on Natural Flood Management ...... 13

3.3

Science relating to a range of Natural Flood Management types ............. 29

3.3.2

Summary.......................................................................................... 72

3.4

Review of Natural Flood Management screening and assessment tools .. 74

3.5

Literature review summary ....................................................................... 86

4
Proposed methodology for assessing and considering the contribution of
Natural Flood Management .................................................................................. 87
4.1

Introduction .............................................................................................. 87

Identification of Natural Flood Management measures ........................................ 91


4.2

Screening for Natural Flood Management opportunities .......................... 92

4.2.1

Introduction ...................................................................................... 92

4.2.2

Screening methods .......................................................................... 92

4.2.3

Summary........................................................................................ 108

4.3
Quantifying the flood risk management benefits of Natural Flood
Management measures ..................................................................................... 110
4.3.1

Introduction .................................................................................... 110

4.3.2

Specification of proposed S20 hydrological assessment tool.......... 113

4.3.3

Data requirements.......................................................................... 137

4.3.4

The capabilities of the assessment tool .......................................... 139

4.3.5

The inclusion of elements of the Refh ............................................ 140

4.3.6

Calibration, verification and improvement ....................................... 140

4.3.7

Flooding and the role of forestry ..................................................... 142

4.3.8

Summary........................................................................................ 143

-v-

Summary of key recommendations ........................................................... 144

References ........................................................................................................... 146

-vi-

Table of Figures
Figure 1: Outline process for the delivery of Natural Flood Management requirements
under the FRM Act ..................................................................................................... ii
Figure 2: Natural Flood Management techniques (SAIFF, 2011) .............................. 2
Figure 3: Example of the differences between the catchment delineations used within
the Jacobs S20 tool ................................................................................................... 8
Figure 4: Example of a riparian intervention that generated a 0.64% reduction in the 1
in 200yr flood event without any of the proposed riparian woodland being positioned
in the riparian zone. (Proposed riparian woodland shown in pink). ......................... 8
Figure 5: Example of the output from the proposed Jacobs S20 assessment tool
showing the hydrological benefits of restoring the floodplain area (purple) to facilitate
an increase in flood depths of 300mm with a 50% efficiency factor. ........................ 10
Figure 6: Derivation of the risk of compaction classes (Holman et al, 2001) (GLU
Grazing Livestock Unit) ........................................................................................... 33
Figure 7: Mass infiltration rates (Holtan and Kirkpatrick, 1950) ................................ 34
Figure 8: Monitored soil moisture at two adjacent sites in the Allt AMharcaidh
catchment in the Cairngorms: a) mature natural Scots Pine woodland, b) wet heath
moorland. (Dots = averaged field measurement, Unbroken line = HYLUC model
simulation)............................................................................................................... 42
Figure 9: Simulated average seasonal soil moisture deficit for upland grass, heather
and coniferous woodland for the River Calder catchment, Renfrewshire. (Period
simulated: 1983 1993) .......................................................................................... 43
Figure 10: Predicted attenuation of: a) winter, b) summer flood peaks due to a
hypothetical 100% conversion of the Calder catchment (Renfrewshire) from upland
grass to various types of woodland\forest or lowland grass (Price et al, 2000). Given
as a percentage of the peak flows predicted for upland grass. (Prediction uncertainty
band reflects the range of values attributed to the model parameters coming from
different field experiments). ..................................................................................... 43
Figure 11: Modelled potential reduction of flood peaks on the Kamp following
afforestation (increase in pine forest from 47% to 86%), plus the predicted increase
of flood peaks following deforestation of the existing forest (from 47% to 0% forest
cover). [Source: Frances et al (2008)] ..................................................................... 45
Figure 12: Modelled potential reduction of flood peaks on the Iller following a 29%
increase in pine forest cover. [Source: Frances et al (2008)] Note the 25 50 year
event is estimated to be approximately 800 cumecs................................................ 46
Figure 13: The changes to the speed of catchment response parameter (TP) and the
percentage of storm event rainfall forming quick runoff (PR) observed at the Coalburn
experimental peat catchment following intensive upland drainage. (Values calculated
from data presented in Robinson et al, 1998. Speculated canopy cover estimated by
Jacobs based on experience and the limited information provided by Robinson et al.)
................................................................................................................................ 52
Figure 14: Unit costs of implemented realignment plotted against size and year.
Scott et al. (2011). ................................................................................................... 71
Figure 15: Sample output, River Ouse combined sensitivity (Environment Agency,
2008) ....................................................................................................................... 75
Figure 16: Sample output, National assessment combined sensitivity (Environment
Agency, 2008) ......................................................................................................... 75

-vii-

Figure 17: User interface of the Land Management CFMP Tool (Environment
Agency, 2008) ......................................................................................................... 76
Figure 18: High priority areas with the greatest potential for woodland planting to
reduce downstream flooding (Broadmeadows & Nisbet, 2009) ............................... 78
Figure 19: Sample output from the POLYSCAPE tool for the Pontbren catchment
(Pagnell, 2009). ....................................................................................................... 81
Figure 20: Example of current UK farms and likely impacts of land use management
scenarios and how they can be mapped onto the decision support matrix (Hewett et
al., 2006) ................................................................................................................. 82
Figure 21: Four hillslope runoff risk scenarios for the same common land unit
(Hewett et al, 2006). ................................................................................................ 82
Figure 22: Example of a time map generated by OVERFLOW for the Pickering Beck
catchment, in this instance a 14mm per day gross rainfall even, assuming a
percentage runoff of ~55% after adjusting for evaporation and groundwater losses.
The timescale is in hours to the catchment outlet. ................................................... 84
Figure 23: Outline process for the delivery of Natural Flood Management
requirements under the FRM Act............................................................................. 88
Figure 24: Flow chart showing the inclusion of Natural Flood Management within the
FRM Act .................................................................................................................. 89
Figure 25: Variation in percentage runoff with a changing leaf area index using
typical Scottish catchment descriptors ..................................................................... 95
Figure 26: Variation in PROPWET across northern Britain (Bayliss, 1999)............. 96
Figure 27: Variation of PROPWET with SAAR (Data for HiFlows catchments) ....... 97
Figure 28: Potential means of displaying sub-catchment desynchronisation
information .............................................................................................................101
Figure 29: Example output from ST:REAM for the Taff catchment in Wales (Parker, in
press) .....................................................................................................................103
Figure 30: Sample output for the Surge Attenuation Potential ...............................104
Figure 31: Summary of the European Environment Agency wave climate data . Error!
Bookmark not defined.
Figure 32: Pictorial overview of the proposed Natural Flood Management
assessment model .................................................................................................114
Figure 33: Schematic representation of the proposed Natural Flood Management
assessment model .................................................................................................115
Figure 34: Design rainfall profiles for summer and winter, drawn as normalised
hyetographs (Kjeldsen, 2007) ................................................................................116
Figure 35: A comparison of the variation in canopy interception capacity with leaf
area index (LAI) using MORECS (Hough and Jones, 1997) and Hoyningen-Huene
(1981) ....................................................................................................................118
Figure 36: Equal water content ( C ) across stores of different capacity (Kjeldsen,
2007) ......................................................................................................................119
Figure 37: Plot comparing the regression based estimate of PROPWET with the
FEH CD PROPWET ...............................................................................................124
Figure 38: Plot comparing the regression based estimate of PROPWET with the
FEH CD PROPWET ...............................................................................................125
-viii-

Figure 39: An example of a time to outlet grid generated for a sample Scottish
catchment as part of this study ...............................................................................128
Figure 40: Roughness values recommended by USDA (2010) for shallow surface
runoff for flow depths of less than 30mm. ...............................................................129
Figure 41: Example of how the time lag effects of reservoirs, lochs and floodplains
could be included within the assessment method ...................................................132
Figure 42: Summary of the proposed flow routing methodology ............................133
Figure 43: Example of a five band time to outlet grid for a sample Scottish catchment
...............................................................................................................................135
Figure 44: Example hydrograph using five runoff bands for a generalised catchment
...............................................................................................................................135

-ix-

Table of Tables
Table 1: Priority research questions ...................................................................... 20
Table 2: Questions requiring further research......................................................... 20
Table 3: Summary of catchment scale estimates (from monitoring or modelling) of the
effect of afforestation on peak flood flows (Jacobs, 2011). ...................................... 25
Table 4: Summary of riparian woodland studies (Jacobs, 2011) ............................. 26
Table 5: Summary of floodplain woodland studies (Jacobs, 2011) ......................... 26
Table 6: Summary of annual transpiration and interception data for a range of land
cover types (Nisbet, 2005)....................................................................................... 37
Table 7: Wave run-up friction factors (CIRIA, 2007) ..............................................106
Table 8: Summary of the preferred Natural Flood Management screening processes
and the data requirements......................................................................................108
Table 9: Summary of the screening processes which are not recommended without
further scientific support or data .............................................................................109
Table 10: Summary of assessment methods for Natural Flood Management
measures and the identification of areas where there are currently gaps in
assessment capability ............................................................................................112
Table 11: Leaf Area Index values as used in MORECS 2.0 (Hough and Jones, 1997)
...............................................................................................................................117
Table 12: Effect of soil degradation via the use of analogue HOST soil types
(Packman et al, 2004) ............................................................................................121
Table 13: The range and quartiles for variables used in the PROPWET descriptor
regression analysis ................................................................................................123
Table 14: Summary of flow routing used within existing hydrological models and
methods .................................................................................................................127
Table 15: Examples of land cover types and the roughness bands which could be
employed ...............................................................................................................130
Table 17: Example of memory requirements for a range of catchment sizes .........136
Table 18: Summary of the data required for the proposed methodology................137
Table 19: Summary of the data which could improve the quality of the tool ............138
Table 20: Summary of how a range of Natural Flood Management measures could
be assessed using the proposed methodology .......................................................139

-x-

Appendices
Appendix A - Review of Jacobs Section 20 Review (CREW, 2012)
Appendix B - Summary Table of Natural Flood Management Measures
Appendix C - Other Natural Flood Management considerations
Appendix D - Requirements on SEPA in relation to Natural Flood Management under
the FRM Act
Appendix E Runoff generation model examples and initial calibration of flow routing
model

-xi-

Notation and terms


(The) Act
BFIHOST

- The Flood Risk Management (Scotland) Act 2009


- Baseflow index (BFI) based on HOST data

Cini

- Initial soil moisture capacity (mm)

C max

- Maximum soil moisture capacity (mm)

D
DDF
DPLBAR
DPSBAR
DTM
FARL
FEH
FSR
G2G
GLU
HOST

- Duration (hours)
- Depth duration frequency model
- Mean drainage path length (km)
- Mean drainage path slope (m/km)
- Digital terrain model
- Flood attenuation from reservoirs and lakes, FEH descriptor
- Flood Estimation Handbook
- Flood Studies Report
- Grid to Grid Hydrological Model
- Grazing livestock unit
- Hydrology of Soil Types (soil classification)

LAI summer / w int er - Leaf area index, half the leaf area per unit area of ground (sqm/sqm)
LCM
- Land cover map (1990, 2000 or 2007)
Lumped model - A model that treats the whole of a catchment as a single accounting
unit and predicts only values of variables averaged over the
catchment area
MORECS
- Met Office Rainfall and Evaporation Calculation System
PDM
- Probability Distributed Model
PE
- Potential evapotranspiration (reference land cover) (mm/yr)
PR
- Percentage runoff (%)
PROPWET - FEH catchment descriptor, proportion of time when the estimated
SMD was less than or equal to 6mm (1961-1990)
PVA
- Potentially Vulnerable Area
q
- Direct runoff (cumecs)
Q
- Flow (cumecs)
QMed
- Median annual maximum flood

rsc

- Vegetation (canopy) resistance (s/m)

ReFH
- Revitalised Flood Hydrograph
SAAR
- Standard average annual rainfall (mm/yr) (1961-1990)
SAIFF
- Scottish Advisory and Implementation Forum for Flooding
Semi-physical model - a hybrid model which represents some hydrological
processes in a physical manner and others via mathematical functions
that have no direct physical basis.
SEPA
- Scottish Environment Protection Agency
SM
- Soil moisture depth (mm)
SMD
- Soil moisture deficit (mm)
SNH
- Scottish Natural Heritage
SPR
- Standard percentage runoff (%)

-xii-

SuDS
T

- Sustainable (urban) drainage systems


- Return period (years)

Tp

- Time to peak (hours)

URBEXT
WFD

- Extent of urban and suburban land cover (1990 or 2000)


- Water Framework Directive

-xiii-

-xiv-

1 Introduction
The Flood Risk Management (Scotland) Act 20091 (the 'FRM Act') was enacted in June
2009. The FRM Act places new responsibilities on the Scottish Environment Protection
Agency (SEPA) and other Responsible Authorities in relation to the sustainable
management of flood risk. Part of the sustainable approach is the consideration of the
contribution that Natural Flood Management can make to manage flood risk. This report
concerns the delivery of Section 20 of the FRM Act which stipulates that SEPA must
undertake an assessment of the role that the restoration, enhancement and alteration of
natural features and characteristics can have in managing flood risk in Scotland and
subsequent use of this information through the appraisal process and in local flood risk
management plans. This restoration, enhancement or alteration of natural features and
characteristics is commonly referred to as Natural Flood Management (NFM).
What is Natural Flood Management?
In reflection of the broad-spectrum of measures that have been termed Natural Flood
Management there has been some ambiguity over what it entails despite there being
apparent consensus on what it was trying to achieve. The Scottish Advisory and
Implementation Forum for Flooding (SAIFF, 2011)2 provides the following definition of
Natural Flood Management which has been used by this study:
Natural Flood Management can be defined as those techniques that aim to work with
natural hydrological and morphological processes, features and characteristics to
manage the sources and pathways of flood waters. These techniques include the
restoration, enhancement and alteration of natural features and characteristics, but
exclude traditional flood defence engineering that works against or disrupts these natural
processes.
It should be recognised that alteration has been included within the definition so that
techniques associated with changes in land management and partial restoration are not
excluded from Natural Flood Management. In some instances it will be difficult to draw a
line between what is considered Natural Flood Management and traditional flood
defences. Figure 2 provides additional insight to SAIFF interpretation of what is included
within Natural Flood Management.

-1-

Figure 2: Natural Flood Management techniques (SAIFF, 2011)

Natural Flood Management under the FRM Act: legislative requirements


Section 20 of the FRM Act requires SEPA, by 22 December 2013, to assess whether
alteration (including enhancement) or restoration of natural features and characteristics of
any river basin or coastal area in a flood risk management district could contribute to the
management of flood risk for that district. Such assessment must be reviewed and
updated at least every 6 years or as directed by Ministers.
Section 20 specifies that natural features and characteristics include such features and
characteristics that can assist in the retention of flood water, whether on a permanent on
temporary basis, (such as floodplains, woodlands and wetlands) or in slowing the flow of
such water (such as woodlands or other vegetation), those which contribute to the
transporting and depositing of sediments, and the shape of rivers and coastal areas.
The assessment must take into account the flood risk assessment, any flood hazard
maps and flood risk maps and any flood risk management plan for the time applicable,
and refer to a map showing where natural features and characteristics could contribute to
management of flood risk.
The map must be prepared at a scale that would assist SEPA in the identification of
measures and assist local authorities in preparing local flood risk management plans.
Assessment under Section 20 and maps produced as part of the assessment must be
made publically available.
Section 28 further requires SEPA to consider Section 20 assessment when setting
objectives and identifying sustainable measures to manage flood risk. Where Section 20
identifies the potential for a natural feature or characteristic to contribute to managing
flood risk, and such feature is not selected to be included in a flood risk management
plan, a justification must be provided in the relevant part of a flood risk management plan
as to the reasons for not selecting such measure.
Finally, section 34 (3) requires that a local flood risk management plan (the
supplementary part) also includes supplemental information on about how implementing
the measures may alter (including enhance) or restore natural features and
characteristics.

-2-

The SFM guidance published by the Scottish Government encourages that Section 20
assessment also creates information that can be used in the appraisals of FRM options.
This means developing information about the potential benefits and adverse impacts of
Natural Flood Management measures that can feed into the appraisal process.
Introduction to this report
Section 2 of this document provides an overview of the review process for a recent report
commissioned by SEPA on Natural Flood Management (Jacobs, 2011)3. As part of the
quality management process the report was peer reviewed by leading Scottish
academics in the field of flood risk management via the Centre of Expertise for Waters
(CREW) (Werritty et al., 2012)4. This peer review process identified that the Jacobs
literature review was exhaustive, authoritative and up-to-date, leading to only minor
suggestions of additional sources with no notable changes to the outcome of the studys
findings. A proposed methodology for assessing Natural Flood Management measures
was also presented within the report (Jacobs, 2011)3. Unfortunately this methodology is
associated with licensing issues, technical difficulties and a catchment based approach
which did not lend itself to national identification of opportunities led to this tool not being
adopted by SEPA.
The literature base presented by Jacobs has been revisited in Section 3 of this report with
efforts made to expand the depth of the review by bringing in additional literature
recommended by the peer review process. The range of flood mechanisms was also
expanded in an attempt to give due consideration to the role Natural Flood Management
can play in all modes of flood risk, including:

Pluvial;

Fluvial;

Groundwater;

Coastal: and

Urban drainage.

The updated literature review has been divided into three sub-sections so that a specific
focus can be directed at each area in turn:

Section 3.2 Review of previous literature reviews on Natural Flood Management

Section 3.3 Science relating to a range of Natural Flood Management types

Section 3.4 Review of Natural Flood Management screening and assessment


tools

Having updated the literature review a summary table detailing the key information for a
broad range of Natural Flood Management measures (29 in total) has been prepared.
This table is presented in Appendix B of this report. The table summarises the most
relevant scientific literature for each measure, how the flood risk reduction can be
quantified and the proposed methodology for inclusion within the methodology for the
delivery Section 20 of the Act.
Section 4 details the proposed methodology for the delivery of Section 20 of the Act. This
section details how the delivery will be divided into three phases; The identification (or
screening) phase using national GIS screening; Appraisal phase to assess NFM
measures against and in combination with other measures to manage flood risk, and
finally further development of appraised measures through Local Flood Risk Management
Plans.

-3-

The Screening, process, will be undertaken at a national level to identify opportunities for
Natural Flood Management. The output will be the production of six national maps
detailing the potential within the landscape to use Natural Flood Management measures
to manage flood risk. Each map will detail the opportunities for a specific type of
measure, such that localised opportunities for the source control of runoff or the
attenuation of estuarine surge can be rapidly identified following consideration of further
catchment information as part of the Appraisal process. The scoring system used within
the screening maps will not include non-flood risk constraints or opportunities, nor will the
screening process directly specify which measure would be most suited to address
flooding at a particular location. It is intended that the output of the screening process will
assist in generating ideas at the earliest stages of measures development ahead of more
focused assessment during the Appraisal.
The last stage in the process - quantification of the effects of agreed measures - will be
taken forward as part of local flood risk management plans. The literature in Section 3
identifies that there are currently gaps in the methodologies that can be used to quantify
the effect of Natural Flood Management measures, in particular for fluvial and pluvial
driven flooding. Where there are existing tools and methodologies which can be used to
for assessment then these should be used, however in the case of fluvial and pluvial the
report recommends that a new hydrological methodology is developed. Section 4.4
presents a specification for a proposed semi-physical rainfall-runoff modelling
methodology which could be used to assess fluvial and pluvial measures. It should be
noted that the method is currently only a proposal, and its successful implementation
would require a comprehensive calibration and verification phase prior to it becoming a
valuable hydrological assessment tool.

-4-

2 Review of Jacobs Methodology


2.1 Introduction
As part of SEPAs programme to meet the requirements of the Flood Risk Management
Act 20091 Jacobs was commissioned to undertake a study to develop a methodology for
Section 20 which states:

SEPA must assess whether alteration (including enhancement) or restoration of


natural features and characteristics of any river basin or coastal area in a flood risk
management district could contribute to the management of flood risk for the district.
A literature review of the science relating to Natural Flood Management measures was
undertaken as part of this study to inform the configuration of the Jacobs tool. The
findings of the study are reported in Development of Methodology for Assessment
Required under Section 20 (Jacobs, October 2011)3. On completion the report was
reviewed by leading Scottish academics and it was presented to a select group of flood
risk management practitioners via a seminar breakout session. As a result two
documents provide comment on the Jacobs report.

Werritty A., Arthur S., Ball T. and Wallerstein N., Review of Jacobs Section 20
Review, CREW-The James Hutton Institute, January 20124.

SNIFFER & SEPA Natural Flood Management Implementation Learning from


Practice Workshop Report, November 20115

Informal comments have also been provided by Forest Research via email to the author
of this document, these have been incorporated within the authors comments in Section
2.4.

2.2 Review of Jacobs Section 20 Review (CREW, 2012)4


A review was undertaken by Scottish academics at CREW (CentRe of Expertise for
Waters) base on 11 questions set by SEPA:
Q1) The literature review was limited to those studies that provided quantitative evidence
of the effects of Natural Flood Management techniques on catchment scale flooding. In
your view, is this the right approach or have any important studies been missed?
Q2) In your view, are the studies identified in the literature review representative of
Scottish catchments?
Q3) Do you agree with conclusions and recommendations in sections 2.2.4, 2.3.4 and
2.4.4[of the Jacobs Report]?
Q4) Do you agree with the outlined approach described in section 3.1[of the Jacobs
Report]?
Q5) Do you have any observations in relation to the testing of the model described in
section 3.3[Of the Jacobs Report]?
Q6) In your opinion, is the preliminary rule set described in section 3.4 [of the Jacobs
Report] acceptable and are there sufficient grounds for the rule set? Suggest, if possible,
alternative ways to improve the rule set.
Q7) In your opinion is the outline method for coastal floodplains correct?

-5-

Q8) Do you agree with the limitations specified in section 3.7 [of the Jacobs Report] and
recommendations in section 8.8[of the Jacobs Report]?
Q9) Do you have any other observations from Section 3 [of the Jacobs Report] that you
would like to highlight?
Q10) Overall, would you recommend the use of the proposed GIS tool and the rule for
use as the screening tool to provide an indication of the effect of Natural Flood
Management techniques?
Q11) Do you have any recommendations for future work that would be beneficial to
further develop the rule set and the modelling tool?
To avoid duplication of the CREW review it has been appended to this document in
Appendix A.

2.2.1 Comments relating to literature review


Both the CREW reviews are complementary of the literature undertaken by Jacobs with
one reviewer describing it as exhaustive, authoritative and up to date. The reviewers
pick up on the omission of land management activities such as reducing stock densities,
land use change and field drainage as areas where additional information could have
been provided. However it is accepted that these omissions could be addressed in later
development of the S20 GIS tool as proposed by Jacobs. Both reviewers suggest a
handful of additional literature sources to be given consideration in later literature
reviews.
The findings relating to reduction in flow introduced by woodland cover are broadly
accepted however one reviewer suggests that the reduction in flows are probably over
optimistic. The proposed reduction in flows generated by introducing riparian woodland
and floodplain woodland are highlighted by the reviewers as being subject to a serious
degree of uncertainty due to the small number of studies which can be used to support
the proposed reductions.

2.2.2 Comments relating to Section 20 assessment tool


It should be noted that only limited access to the GIS tool was provided to reviewers due
to data licensing issues. As a result, reviewers were only able to comment on the
principles behind the tool rather than tool functionality. Concerns are raised about the
division of the target catchment into thirds (upper, middle and lower by catchment
descriptor estimated time to peak) by both reviewers, with one reviewer referring to it as a
straight jacket which needs to be loosened.
Both reviewers have issue with how floodplain enhancement is included within the tool,
with one reviewer highlighting that it is very dependent on its calibration to the Enrick
Water and that it may not be possible to assess floodplain enhancement without site
specific hydraulic modelling. The other reviewer considers the underlying science related
to floodplain enhancement as too tenuous at present and strongly advises that
floodplain enhancement is removed until evidence is available to substantiate the values.
The assumption that the effects of different Natural Flood Management measures within
a catchment will be accumulative also raises concern, with one reviewer highlighting it as
extremely simplistic, and unlikely to be the case in realitythis issue should be flagged
as a serious issue.
The issue of the absence of ground truthing is raised by both reviewers, one reviewer
recommends that the intensively gauged Eddleston Water could be used to test the tool
against real data.

-6-

2.3 Natural Flood Management Implementation Learning from


Practice Workshop Report (SNIFFER & SEPA, 2011)5
A breakout session was held as part of the above seminar to give delegates the
opportunity to comment on the recently released Section 20 methodology proposed by
Jacobs. It is reported that in overview the audience considered the tool to be very
promising in what is a difficult area. Some delegates did raise concern that the tools
should only be used by practitioners who are aware of its limitations. The breakout group
delegates made the following recommendations:

The tool should be validated by running it on a range of Scottish catchments

The model should be peer-reviewed

It was suggested that the model incorporate land capability data

Concerns were raised over the validity of the upper/middle/lower division hard
coded within the tool. It was suggested that a more subtle subdivision of
catchments is needed in catchment headwater areas.

It was suggested that the uncertainty may be so large that decision making will
still remain based on local expert judgement.

Many of the encoded Natural Flood Management measures are very generalised
and will need to be improved in the future.

It would be helpful to know what factors are important to understand within


different types of catchments.

2.4 Additional comments


In undertaking this study the author of this document was able to test the functioning of
the proposed Jacobs S20 assessment tool, it should be noted that none of the other
reviewers were able to gain access to the tool due to data and software licensing issues.
In undertaking the testing of the tool only the prepared data for the Allan Water catchment
was used and no other catchments were investigated.
In addition to the comments raised by the CREW (2012)4 review and at the SNIFFERSEPA Seminar (2011)5 the author would like to raise the follow points:

The S20 tool is fast and relatively stable (within the normal expectations ESRIs
ArcGIS platform). It has an intuitive easy to use interface.

The prescriptive structure of the model does retain a good level of control over the
models outputs, this reduces the risk of an inexperienced user applying the tool
incorrectly.

The method is reliant on CEH Flow grid coverage for the entire catchment and the
tool is hard coded on the assumption that the CEH Flow grid is correct. In
essence the tool calculates the change in flows and not the actual flows, therefore
the predictions made by the tool are limited by the underlying CEH Flow grid on
which it has been built. The tool should not be seen as a hydrological model,
rather a method for scaling existing hydrological predictions.

It is unclear what catchment delineation the tool functions with. The sample data
provided by Jacobs for the Allan Water shows two catchment delineations for the
Allan Water which disagree with each other. It is understood that parts of the tool
work with the assumption that the CEH Flow Grid is correct while other elements
work with a user supplied catchment delineation. The ambiguity regarding what
catchment area is being considered is not helpful. An example of the Forglen
-7-

Burn (a tributary of the Allan Water), which drains the loch at the University of
Stirling, is shown in Figure 3.

Figure 3: Example of the differences between the catchment delineations used within the
Jacobs S20 tool

The tool allows benefits of riparian planting to be accrued in non-riparian areas.


For example using the riparian tool to cover the entire upper two thirds of the Allan
Water catchment less all riparian corridors indicates a central estimate of flow
reductions for the 1 in 200yr flood event of approximately 0.64% despite no
woodland being positioned within the riparian margin. Refer to Figure 4 for more
information. The same area covered by forest cover achieves a 1.7% central
estimate reduction for the 1 in 200yr event.

Figure 4: Example of a riparian intervention that generated a 0.64% reduction in the 1 in


200yr flood event without any of the proposed riparian woodland being positioned in the
riparian zone. (Proposed riparian woodland shown in pink).

The current configuration of the tool does not permit predefined shapefiles to be
uploaded into the tool nor can manually drawn Natural Flood Management
-8-

outlines be exported from the tool. It is likely that it would be feasible to include
improved import/export functionality in an updated version of the tool.

It is up to the user to classify the function of placed woodland. For example if the
woodland tool is used across a broad area which incorporates floodplain and
riparian areas the tool will only assess the woodland for its runoff reducing
capabilities and not for the riparian or floodplain benefits.

The woodland cover within the LCM2000 dataset is hard coded as the actual
woodland cover. While it is accepted that it is a good representation of woodland
cover at the time the data was acquired, in many instances the LCM2000 data
does not correctly classify the present day local land cover correctly.

The tool does not allow a way for the results to be exported, for example the
revised hydrograph cannot be extracted to allow the flows to be applied to a
hydraulic model.

Currently the tool only allows the assessment of the hydrological benefits for the 1
in 200yr and 1 in 5yr flood events. It is acknowledged that the report indicates
other return periods could be added.

The drain blocking element of the tool does not take account of local soil
conditions, drain geometry or drain direction which are likely to play a significant
role in the both the magnitude and effect (reducing or increasing) flood risk.
(Refer to Section 3.3(D)

The effect of catchment scale on the effectiveness of Natural Flood Management


measures is not accounted for within the tool. There is a general consensus
within the literature that the effectiveness of runoff reducing measures is inversely
related to catchment size.

The adopted hydrological method does not conserve mass. That is, when the
flood peak is reduced this is achieved by cropping the top off the hydrograph.
This lost flow volume is not reintroduced later in the flood hydrograph. The
method of cropping the flood also introduces the curiosity that floodplain
restoration might have a notable impact at reducing flows around the peak of the
1 in 5yr event, but for a 1 in 200yr event there is no modification to the hydrograph
at that same flow rate, (refer to Figure 5).

-9-

Figure 5: Example of the output from the proposed Jacobs S20 assessment tool showing
the hydrological benefits of restoring the floodplain area (purple) to facilitate an increase in
flood depths of 300mm with a 50% efficiency factor.

The assessment tool is catchment outlet specific, therefore the tool is not
appropriate for national screening of Natural Flood Management opportunities.

In preparing the literature review, Jacobs did not undertake a review of existing
hydrological methods which have been used to represent the hydrological benefits
for Natural Flood Management measures. Nor does the review take account of
the recommendations for future Natural Flood Management assessment tools
such as those by OConnell et al. (2004):
The modelling should be distributed and be capable of running continuous
simulations. It should also be partially or wholly physically based so that the
physical properties of local landscapes, soils and vegetation can be represented,
and it should include detailed modelling of surface water flow so that the effects of
changes can be tracked downstream.
It is unfortunate that the Jacobs S20 is not aligned with these recommendations,
namely it is not a distributed model that is capable of continuous simulation, nor
does it have a physical basis or a detailed flow routing engine to permit the effects
to be tracked downstream.

The hydrological benefit of floodplain restoration does not include any component
to account for the hydrograph volume. The ratio of hydrograph volume to
floodplain volume is a key variable (Section 3.3(L)).

The increase in flood depth as a consequence of floodplain planting is weak. It


would be beneficial for a stronger physical basis to the selection of increased
floodplain depth so that land cover (roughness), discharge intensity (flow/unit
width) and hydraulic gradient (or valley slope) are used to estimate the likely
increase in floodplain depth.

It is not clear how the Jacobs S20 tool can quantify Natural Flood Management
benefits in catchments that have significant lochs or reservoirs, particularly if these
waterbodies are artificially managed or are close to the catchment outlet.

-10-

It is unclear how hydrometric data, such as a real data estimation of QMed or the
catchments observed time to peak can be incorporated within the tool. It would
appear that the tool is hardcoded to remain a no data methodology.

Jacobs recommends that the S20 tool is only used to provide an initial broadscale
identification of possible benefits of various Natural Flood Management measures;
going on to say that measures should be more fully appraised before adoption.
This implies that a different method must be put in place to allow for the detailed
assessment and that the tool will not allow Natural Flood Management measures
to be directly compared against hard engineered measures. A robust hydrological
assessment framework which can be refined as the project develops, and which is
consistent with the underlying principles of current British hydrological practice,
would be more progressive.

The Jacobs S20 tool only provides a means of assessing the hydrological benefits
of Natural Flood Management measures in fluvial systems. Limited discussion of
a potential means to configure a saltflats restoration tool for the estimation of
reduction in estuarine surge is provided however there are notable gaps in the
methodology which is discussed only in broad concept. There is no discussion of
using saltflats to dissipate wave energy and the areas of groundwater and pluvial
flooding are not considered.

In some cases there has been criticism that the Jacobs S20 tool does not give
enough consideration to other benefits. While it would be ideal to have a tool
which would consider all benefits and disbenefits of Natural Flood Management,
to do so would be immensely challenging. For the time being, the focus of energy
should be on the development of a hydrological assessment methodology for
Natural Flood Management techniques thus facilitating the fulfilment of Section 20
of the Act.

By direct correspondence with the author, Nadeem Shah at Forest Research


provided the following comments:
o

Riparian woodland should not be considered a subset of floodplain


woodland. He recommended that consideration is given to the definition
provided within the UK Forestry Standard Water Guidelines (Forestry
Commission, 2011)76

There is a need for the tool to consider other forest mechanisms for flood
alleviation, namely infiltration and water use.

Would be keen to see more mention of multiple benefits

The OConnell et al. (2004)83 study identified three fundamental


unresolved issues of rainfall-runoff models yet it is unclear that the Jacobs
S20 methodology takes appropriate steps to address these issues.

It would be beneficial for the model output to incorporate confidence bands


to assist in decision making.

The assumptions underlying the simplified and generic models should be


commented on in the report or an appropriate paper.

-11-

2.5 Summary
It can be summarised that the literature review undertaken by Jacobs is generally well
accepted. Subject to inclusion of a handful of papers, including some recently published,
the literature review should be considered authoritative and up to date.
The Jacobs S20 Tool is based on modifications to the CEH Flow Grid. In developing the
methodology for modelling the effect of Natural Flood Management measures Jacobs
have not drawn on the recommendations of OConnell et al. (2004)83 in regards to how
the model should be configured. Instead the method is centred on pro-rata cropping of
the hydrograph peak via best estimates made from the literature rather than adopting a
more physical based approach to modify parameters contained within the hydrological
model.
There are some concerns over the schematisation of the model, with the most notable
being related to the absence of a hydraulic based flow routing element to the model
(division of catchment into thirds). This formulation prevents accurate accumulation of
hydrological benefits moving down through the catchment towards receptors and the
assessment of delay/attenuation measures (floodplain enhancement and riparian
intervention). There are other areas of concern relating to how the tool accumulates the
benefits of multiple types of interventions, the absence of ground truthing and the
suggested benefits are questioned for some measures (floodplain enhancement).
The requirement to specify a single location as the catchment outlet will lead to
challenges in using the tool to undertake a national screening of Natural Flood
Management opportunities. The tool also lacks functionality to identify flood risk
management benefits for coastal, groundwater and pluvial flood mechanisms.
It is not recommended that the current formulation of the Jacobs S20 assessment tool is
used for the identification or appraisal of Natural Flood Management measures in
Scotland.

-12-

3 Literature review
3.1 Introduction
Natural Flood Management is a broad area covering many of the aspects of hydrology,
coastal science and their related subjects. This literature review will be divided into the
following sections:

A summary of previous Natural Flood Management literature reviews


concentrating on reviews relevant to the catchments and coastal conditions typical
to Scotland (Section 3.2)

A summary of the most comprehensive science relating to each of the main


Natural Flood Management measures within a Scottish context (Section 3.3)

A summary of the existing tools for screening and assessing Natural Flood
Management measures (Section 3.4)

3.2 Review of previous literature reviews on Natural Flood


Management
The understanding of Natural Flood Management techniques and their effectiveness in
reducing flood risk is a growing area of science. In order to identify key research
questions, the major British Natural Flood Management literature reviews were reviewed
and outcomes summarised in this part of the report. These reviews included:
(A) FD2114 Review of Impacts of Rural Land Use and Management on Flood Generation
(OConnell et al., 2004)83
(B) R&D update review of impact of rural land use and management on flooding
(Environment Agency - Atkins, 2007)
(C) FD2120: Analysis of historical data sets to look for impacts of land use and
management change on flood generation, R&D Technical Report (Defra / EA, 2008)
(D) Delivery of Making Space for Water, The role of land use and land management in
delivering flood risk management (Environment Agency Halcrow, 2008)
(E) Land Use Jigsaw: Prioritising scientific research on land management and FCRM
(Environment Agency, 2008)
(F) Evidence Based Review: Does Land Management Attenuate Runoff? (Haycock
Associates National Trust, 2008)
(G) The way forward for Natural Flood Management in Scotland - a report to Scottish
Environment LINK (MNV, 2008)
(H) Technical review of NFRM techniques, their effectiveness and wider benefits (RSPB,
2009)169
(I) Natural approaches to flood management under the Flood Risk Management
(Scotland) Act 2009: Development of Methodology under Section 20 - a report to SEPA
(Jacobs, 2011)
(J) Greater working with natural processes in flood and coastal erosion risk management
(Environment Agency, 2012)
It is clear that literature reviews vary in their quality and effectiveness. Some of the
studies had not been peer reviewed, lessening their accuracy and validity of their
conclusions. The most robust technical reviews of Natural Flood Management
techniques were commissioned by Defra and the Environment Agency, but these studies
-13-

reflect the issues in England and Wales. There is a tendency for Scottish Natural Flood
Management literature reviews to not be as scientifically robust as they have not been
subjected to a peer review process, but never the less they do form useful sources.

(A)

FD2114 Review of Impacts of Rural Land Use and Management on


Flood Generation (OConnell et al., 2004)83

This was the first major review of the impacts of land use and management on flood
generation undertaken by Defra and the Environment Agency. The purpose of the study
was to review the factors contributing to runoff and flooding in the rural (managed, not
natural) environment, and to scope the research needed to improve the identification of
the management policies and interventions to reduce the impact of flooding. The study
focused on inland flooding mechanisms such as pluvial and fluvial flooding.
The study was undertaken by a team drawn from the disciplines of agriculture, soil
science, hydrology, hydrogeology and socio-economic science and provides much of our
current understanding of the science with respect to land use management and its impact
on flood generation.
It is worth noting that this review focused on land use and land management techniques,
therefore focusing on runoff generation and thus excluding other Natural Flood
Management techniques such as floodplain restoration, river restoration, washlands
creation and coastal realignment.
The main conclusions of FD2114 are as follows (taken from the Summary of the
document):

Significant changes in land use and management practices in the last fifty years
have resulted in a general intensification of agricultural land use;

There is substantial evidence that changes in land use and management


practices affect surface runoff generation at the local scale, but the effects are
complex;

There is only limited evidence that local changes in runoff are transferred to the
surface water network and propagate downstream;

Analyses of peak runoff records provide very little firm evidence of catchment
scale impacts of land use management;

There are many measures that can be taken to mitigate local flooding by delaying
runoff;

There is considerable uncertainty about how effectively land managers will


respond to promotions or policies related to particular flood prevention or
mitigation measures;

Rainfall-runoff modelling to predict the effects of changes in rural land use and
management on flood generation is in its infancy: there is no generally accepted
theoretical basis for the design of a model suitable to predict impacts; and

The uncertainty in the response of land managers noted above needs to be


accounted for when modelling the overall outcomes when flood prevention and
mitigation practices are promoted.

FD2114 also proposed a number of research priorities centring on collating existing


knowledge, data gathering and the development of assessment tools.

-14-

(B)

R&D update review of impact of rural land use and management on


flooding (Environment Agency - Atkins, 2007)6

Subsequently in 2007, Defra and the EA commissioned Atkins to undertake a further


review of the evidence base, updating the findings of the 2004 Defra Research project
FD2114 with the latest research. This project was commissioned as part of delivery of
Making Space for Water (Defra activities HA6 and HA7) and included a review of
current research relevant to the role of land use in flood risk management in conjunction
with the identification of future research and development needs. This research project
was mainly concerned with the processes governing runoff response of catchments and
the findings of this review are of relevance to Natural Flood Management in Scotland.
This report is one of the most valuable and authoritative sources of information on Natural
Flood Management measures.
The objectives of the review were to:

Update evidence from Defra Research project FD2114 on the impacts of rural
land use changes and management on flood generation in rural areas, at both the
local and catchment scale within the UK; and

To review and prioritise future research needs in this area.

The review was developed from published literature, ongoing studies and additional
information obtained through consultation with a number of key individuals between
January and March 2007. The review reflected the most up-to-date evidence and
provided an indication of where current research is being targeted.
The R&D update focused upon the most important land use management practices, as
identified within FD2114 and subsequent consultation with the Environment Agency.
Unlike FD2114, the update was limited to UK sources and evidence. The findings were
presented according to five main land use or management practices:

Afforestation and deforestation;

Agricultural drainage;

Peat drainage and moorland gripping;

Pasture; and

Cultivation techniques.

The key findings of the study were:

There is continued uncertainty in how local scale effects propagate downstream.

For large catchments, existing modelling studies suggest that a large extent of
land-use or land management change is required to produce a relatively modest
reduction or delay in downstream flood peaks.

The location of local changes or interventions within a catchment is critically


important. Changes will have different effects depending on the relative phasing
within the catchment. Catchment scale models that couple hillslope responses to
downstream channel networks are vital to enable prediction of these effects.

For particular soil types and geologies (and therefore locations within the UK),
groundwater becomes important for flooding. The effects of land use on
groundwater flooding remains an under-researched area.

Multiple interventions have and continue to hamper our ability to predict the
impacts from specific land-use changes.

-15-

Climate variability is the dominant factor in influencing the frequency and


magnitude of flood events within any given catchment.

Afforestation and deforestation:

There is strong evidence that upland conifer forests have a negligible effect on the
reduction of flood flows. (This point is subsequently contradicted implying that it
relates to out of date forestry practices.)

There is very limited potential for afforestation to reduce flood risk at catchment
scale. While upland plantations may reduce peak flows for smaller events, they
are thought to have a negligible impact on extreme events. Equally there is scant
evidence that deforestation increases flood risk although there is evidence that
poor logging practices may have temporary negative impacts.

Land management practices associated with conifer afforestation such as plough


drainage and access road construction can exacerbate flooding. However these
effects are limited by current best practice.

There is evidence that at local scale there is some potential for strategic tree
planting to increase infiltration in compacted soils.

There is limited knowledge on the hydrology of natural unmanaged woodland in


the UK. It is not thought that these woodlands would be any more likely to reduce
extreme floods than conifer plantations, except perhaps for short duration summer
and autumn events. In some circumstances the increased water use may lower
antecedent water thus reducing flood risk.

There is a growing body of evidence that suggests that riparian and floodplain
woodland can attenuate flood propagation.

Agricultural practices:

Agricultural policy has been a major driver of land management practices. As yet
there is little evidence of the impact of new agri-environment schemes.

The evidence shows that there is a need to consider the land management history
of a catchment down to the field scale when assessing its flows.

It is possible to predict with a significant degree of confidence the effect of


agricultural drainage on runoff and river flow from a given site by an examination
and understanding of the individual site properties. Areas could be mapped at
local scale and high risk areas identified.

Peat and moorland drainage

Tools have been and are being developed that help to predict the effect of artificial
drainage or drain blocking on the timing and magnitude of runoff. These should
help target interventions. Catchment scale models also need to be used to take
account of the location of the land-use change within the catchment and its
influence on the timing and magnitude at receptors.

There is uncertainty to whether long-term changes to peat can be reversed.

Pasture

There is evidence that land management in grassland environments influences


runoff at a local scale and small catchment scale. There is also evidence for yearto-year variability in soil structure, soil vulnerability to degradation, and in runoff
processes.

The Pontbren study will be important at bridging the gaps between hillslopes,
small catchment and downstream impacts on flood frequency.
-16-

(C)

Existing modelling studies suggest that a large extent of grassland land-use or


land management change is required to produce a relatively modest modification
to downstream floods.

FD2120: Analysis of historical data sets to look for impacts of land


use and management change on flood generation, R&D Technical
Report (Defra / EA, 2008)7

This study was commissioned as part of joint Defra/EA Flood and Coastal Erosion Risk
Management R&D Programme. Its purpose was to provide an analysis of historical data
on land use and land management to ascertain whether any impacts of land use and
management change on flood generation could be identified. The change identification
methodologies used were Dynamic Harmonic Regression to examine seasonal scale
trends, and Data Based Mechanistic models were used to examine changes in storm
responses.
The study found that, in general the variability between years and inconsistencies in the
rainfall and flow data appear to dominate any tendency of other changes. Only seasonal
flows could be shown to exhibit any significant temporal trend and then only for two of the
nine catchments studied. The assessment of hydrograph responses showed no clear
changes over time. Where tendencies in hydrograph characteristics with time were
evident they were obscured by year to year variability.
The project concluded with a set of policy recommendations:

Both climate variability (particularly rainfall variability) and land use affect flood
runoff. Changes in discharge should not be analysed without consideration of
changes in catchment rainfall inputs.

The preliminary study of catchment responses within different event classifications


(antecedent conditions) was the most promising form of analysis developed during
this project. Different classification schemes should be investigated to check the
nature of changes, including a more complete uncertainty analysis. Careful quality
control of existing datasets is necessary in carrying out such analyses.

Adequate information about past land management changes and soil conditions is
not readily available but will need to be collected and made available in future for
different land use categories if improved understanding of the links between runoff
and land management is to be gained and used at catchment scales.

The results of this project show that there will be a real difficulty of estimating the
benefits of land use change measures in respect of any reduction of flood risk.
Further monitoring of studies aimed at reducing runoff should be carried out to
evaluate the effectiveness of different types of measure at the local level in the
context of agri-environment schemes.

The difficulty in identifying consistent changes given the limitations of the available
data means that land management measures cannot be relied on as alternatives
to more proven flood risk management options.

The difficulty in identifying consistent change given the limitations of the available data
does not mean that change is not happening and should not be taken to imply a policy of
doing nothing. Contextually relevant management practice guidelines (linked to land use,
soil type, antecedent condition) should be developed and monitored to deliver multiple
benefits including reduced runoff generation and local flood risk.

-17-

(D)

Delivery of Making Space for Water, The role of land use and land
management in delivering flood risk management (Environment
Agency Halcrow, 2008)8

This study was commissioned by the Environment Agency with the objectives to:

Summarise the strength of underlying science and knowledge gaps with respect
to land-use/management impacts on flood risk;

Illustrate how land-use/management changes relevant to flood risk are being


delivered through a review of case studies; and

Identify policy mechanisms for delivering changes in land-use management


relevant to flood risk.

The study provides a summary of recent catchment scale modelling studies:

The lumped sub-catchment hydrological model with 1-D flow routing model
approach used in the Ripon study potentially provides a simple framework for
testing future land use scenarios which could be applied generically to different
catchments. However, this approach is as yet untested on other catchments:
o

Scale issues -Land use change modelling is undertaken at a subcatchment scale and does not take account of localised changes;

Underlying science issues Land use changes are simulated by making


changes to parameters within the Flood Estimation Handbook (FEH)
rainfall-runoff method, which provides only crude assessments of surface
runoff impacts; and

Limitations in model input data quality and setup.

The physically based distributed hydrological model used in the Parrett catchment
is considered to be more robust because surface, subsurface and groundwater
are better represented throughout the catchment using a grid approach. This
approach can potentially be more reliably applied to catchments with very different
characteristics and as the Natural Flood Management interventions are modelled
using physically based modifications to the model there is better scope to
accurately represent the accumulative impact of Natural Flood Management
measures at a catchment scale. The study identified:
o

The potential to identify sub-catchments where land management


changes would offer the greatest benefit;

Potential to appraise land management measures such as buffer strips


and contour planting;

Models are data hungry and computationally expensive; and

Models can be costly to setup and require an experienced user.

Simple modelling tools (such as the lumped models applied to the Ripon) should
be used wherever only basic level information is available. It should be
recognised that simpler tools will provide simple and more approximate
predictions.

Research undertaken in Upper Wharfedale has shown that rapid aggradation is a


principle driver of flooding and that the sedimentation risk could be controlled via
Natural Flood Management measures.

-18-

(E)

Land Use Jigsaw: Prioritising scientific research on land


management and FCRM (Environment Agency, 2008)9

The Land Use Jigsaw is an ongoing (as of time of writing) research prioritisation tool
maintained by the Environment Agency that seeks to guide the development of policy
relevant research for Flood and Coastal Risk Management (FCRM) in England and
Wales. The tool, which takes the form of a community maintained register, seeks to
promote current research and determine gaps and overlaps in evidence. It is worth
noting that the Land Use Jigsaw has been focused on the evidence base for individual
measures and has not considered the combined effects of measures on reducing flood
risk. In addition to this it should be noted that it is an Environment Agency tool and
therefore focused on issues faced by the Environment Agency in England and Wales.
While there are marked similarities between hydrological processes across the British
Isles care should be taken in applying the findings in Scotland.
The starting point for the Land Use Jigsaw was the current understanding as presented in
previous studies as discussed above. In summary:

Effects of land use management on flood risk at a large catchment scale are
difficult to determine as they require aggregating many local-scale effects and are
dependent on catchment characteristics;

The lack of understanding does not necessarily mean that there is no catchmentscale effect, but rather that the nature of the effect differs between different
catchments and is difficult to detect;

There is strong evidence that small scale catchment management approaches


deliver flood risk management benefits. Farmland can store floodwater to reduce
downstream flood risk. Large washland (storage) or managed realignment areas
can be designed to store excess water and slow down flood peaks;

Although there can be benefits for local flood mitigation, the evidence to date has
not demonstrated the reduction of extreme flood events at the catchment scale
when using most land management interventions. The project will continue to
support research into large catchment scale impacts of land management
changes on flood risk, water quality and resource protection and will actively seek
partnership projects; and

Policy needs to be based on sound scientific evidence. In order to deliver the


evidence base there needs to be a coordinated effort between all those
undertaking research in this area to focus their research efforts in the correct
areas.

The key priority research questions identified in the 2008 summary report are
summarised in Table 1.

-19-

No.

Evidence

Need

Category

Question

Q020

Evidence
gap

Needed
now

Under
sowing

2a Are there quantifiable flood risk benefits


at the local scale?

Q031

Evidence
gap

Needed
now

Reduce soil 3a Where in the catchment will this be


compaction most effective?

Q038

Evidence
gap

Needed
now

Crop choice 2a Are there quantifiable flood risk benefits


at the local scale?

Q039

Evidence
gap

Needed
now

Crop choice 2b What are the costs for implementation


and maintenance?

Q070

Evidence
gap

Needed
now

Wetland
creation

4a Is there a quantifiable flood risk benefit


at the catchment scale?

Q088

Evidence
gap

Needed
now

On farm
storage

4a Is there a quantifiable flood risk benefit


at the catchment scale?

Table 1: Priority research questions

The most significant gaps in scientific knowledge identified in the 2008 summary report
are summarised in Table 2.
No.

Evidence

Need

Category

Question

Q112

Ongoing
work

Needed
now

Grip
blocking

3a Where in the catchment will this be


most effective?

Q113

Ongoing
work

Needed
now

Grip
blocking

3b How long does the intervention last?

Q121

Ongoing
work

Needed
now

Peat
3a Where in the catchment will this be
restoration most effective?

Q122

Ongoing
work

Needed
now

Peat
3b How long does the intervention last?
restoration

Q139

Ongoing
work

Needed
now

Floodplain
woodland

3a Where in the catchment will this be


most effective?

Table 2: Questions requiring further research

(F)

Evidence Based Review: Does Land Management Attenuate Runoff?


(Haycock Associates National Trust, 2008)10

This report was commissioned by the National Trust to provide an overview of the current
literature in order that future operational policy is relevant and informed. It focuses on
changing land use impacts upon runoff and the impact of land management upon
catchment level studies (>100sqkm).
The study used a systematic approach to identify published and unpublished research
relating to land use/management and runoff with a strict quality control procedure. In
total 29,772 articles (including duplicates) were indentified by the literature search. Of
these articles, 108 were identified as sufficiently relevant for full text viewing and
-20-

subsequently 31 were found to be sufficiently relevant and robust. These 31 articles


covered a total of 34 study sites.
The review found that there is significant evidence that land management is an effective
method for impacting runoff at an experimental scale, however there is not significant
evidence at representative catchment scales. Despite this there is increasing evidence of
local scale impacts propagating downstream.
There is consistent evidence to support that at the representative catchment and
experience scale afforestation will lead to an overall reduction in runoff. Similar strength
trends are observed in relation to agricultural intensification and deforestation which
suggest increases in discharge, peak flows, baseflow and surface flow are frequent to
observe in studies at the representative catchment and experimental level.
Studies into the impact of moorland drainage on runoff follow similar trends as found in
deforested and intensified areas, although surface flow is actually found to be lower in
these areas. In the small number of studies that looked into the impact of soil degradation
on runoff, it was demonstrated that discharge and peak flows were found to be higher in
degraded soils.
There were no trends to be observed in the impact of agricultural drainage upon overall
runoff. However, the lack of a trend may be due to the fact that the exact impact that
agricultural drainage has upon the hydrology of a site is largely dependent upon the soil
type as well as the type, size and extent of the installed drainage. Further complications
associated with the life expectancy of a drainage system need to be understood. It has
been identified that the impacts that a newly installed drainage system has upon runoff,
cannot be assumed to be applicable indefinitely. For example mole drainage channels
(traditionally used in clayey soil types) are typically effective for 2-10 years, again
dependent upon the soil type. Pipe drainage can last anywhere between 10-50 years,
dependent upon the gradients used and the proportion of fine silts in the soil.
The following observations were made within the studys findings:

In an experimental scale there is substantial quantitative evidence than land


management attenuates runoff;

The number of representative catchment studies is relatively low;

There is not substantial quantitative evidence at a representative catchment scale


that land management attenuates runoff, however there is an increasing number
of computational studies and a large range of qualitative evidence. There is
emphasis that a lack of evidence does not necessarily equate to a lack of effect;

Need to adopt the precautionary principle when assessing the influence of land
management upon runoff;

At a large catchment scale there is some evidence that land use changes impact
runoff;

Need to appraise current methodologies for investigating impact of land


management, in response to the lack of statistics and unit hydrograph analysis in
representative catchment studies;

From the papers reviewed it is not clear what impact land management has on
different return periods; and

It is important to note that following permanent changes in vegetation it takes


more than 5 years for a catchment to reach new equilibrium.

-21-

(G)

The way forward for Natural Flood Management in Scotland - a


report to Scottish Environment LINK (MNV, 2008)11

This report was commissioned by Scottish Environment LINK to inform the development
of flood risk management policies in Scotland. The non-peer reviewed report focuses on
how Natural Flood Management is incorporated within sustainable flood risk
management. The key messages from the report summary include:

Most hydrological studies of upland forestry have concentrated on the effects on


the catchment water yield and not the effects of flood generation processes;

Some previous river restoration projects have indicated that there are benefits for
flood management but very often the claim is not substantiated by modelling or
data collection;

Restoration of wetlands and the reconnection of floodplains have been shown to


be effective as flood management techniques and some data exist to quantify how
effective they can be. There is a need to sub-divide wetlands, at least in terms of
upland and lowland, to better understand their role in Natural Flood Management;

There is clear evidence that riparian vegetation will slow down flows in channels
and increase local floodwater storage. There is a need to understand the role of
vegetation type over a floodplain and the ability to temporarily hold back water
(the leaky barrier principle); and

Several studies have been carried out which demonstrate the benefits of some
agricultural practices in flood management. These are mostly related to
increasing infiltration rates into the soils and reducing the amount of artificial
drainage.

(H)

Technical review of NFRM techniques, their effectiveness and wider


benefits (RSPB, 2009)169

This is a comprehensive review of Natural Flood Management techniques, discussing


Natural Flood Management measures from the catchment headwaters to the coast,
covering:

Upland soils;

Upland woodlands;

Upland streams;

Farm management;

Washlands;

River (and floodplain) restoration;

Restoring urban waterbodies; and

Regulated tidal exchange and managed re-alignment

The review provides valuable information about the selected Natural Flood Management
measures over a broad front, addressing the following areas:

Hydrological benefits;

Potential biodiversity benefits; and

Other benefits

The review is less constrained by the hard science approach which forces a more
conservative stance to be reported by government funded studies such as those
-22-

undertaken by DEFRA. This less constrained approach allows an optimistic or more


pragmatic stance to be reported in some circumstances, this is not to say the study is
unscientific.
Key conclusions taken from the report include:

Blocking drainage has complex hydrological response, often having both positive
and negative impact on flood peaks. The restoration of peat is thought to reduce
the 'flashiness' of the catchment, but once saturated, this effect may be reduced.

Enhancement of hydraulic roughness by natural vegetation slows overland flow.


Reducing stocking densities in uplands can aid vegetation recovery and benefit
flood management. Control of burning in sensitive areas can also bring benefits.

Restoring upland woodlands increases interception, with deciduous woodland


having beneficial effects during summer months, benefiting the management of
summer storms.

Planting of coniferous species can have all year round benefits; however, during
and after felling of commercial plantations, peak flows can increase for several
years.

Restoring rivers in uplands can delay response to rainfall by increasing channel


roughness and slowing the conveyance of water and increasing the storage
capacity.

Large woody debris can play an important role in storing and attenuating flood
flows.

Land use and management has significant effect on the soil's capacity to
attenuate water. For example, some arable practices increase the risk of local
muddy floods, and changing the direction of tillage can delay or reduce run-off.

The effect of drainage in lowlands has a complex hydrological effect, and the
response depends upon geology, soil and drainage network. Reducing drainage
capacity on permeable soils can reduce catchment flood risk by slowing and
reducing water export from farmland to the river channel.

Restoration of floodplains and river channels can have positive effects on


catchments ability to store water. Restoring wet woodlands, rough grasslands,
and hedgerows can increase floodplain roughness, resulting in attenuation of the
hydrograph.

Intertidal habitats play an important role in absorbing wave energy and reducing
wave height.

Coastal realignment increases the area of the tidal floodplain, making space for
tidal flooding and thus reduces estuarine surge. The benefits of managed
realignment and tidal exchange have been well demonstrated.

-23-

(I)

Natural approaches to flood management under the Flood Risk


Management (Scotland) Act 2009: Development of Methodology
under Section 20 - a report to SEPA (Jacobs, 2011)3

A review of published literature was undertaken as part of a SEPA commissioned project,


delivered by Jacobs. The literature review aimed to inform the development of a
modelling tool to undertake an assessment of Natural Flood Management potential in
Scotland. The purpose of the literature review was to provide an informed understanding
of the likely impact and effectiveness of the various natural features and characteristics
being considered as candidate Natural Flood Management measures. The aim was to
focus on those studies that have provided quantitative evidence of the effects with an
emphasis on catchment scale flooding. A detailed commentary on the report, including
comments by leading Scottish academics, is provided in Section 2. The review reported
on the following Natural Flood Management techniques:

Upland forest cover;

Riparian woodland;

Floodplain woodland;

Upland drain blocking; and

Wetlands and floodplains.

Upland Forest Cover


The report found that the current literature suggests that the effectiveness of forest cover
is influenced by a number of factors:

Plantation and natural woodlands have different hydrological characteristics, with


commercial forests being subject to more research than native woodlands;

The forestry management procedures have changed with time;

The types of trees planted is important, with annual evaporative losses being
greater from conifer forests than from deciduous;

Caution needs to be applied when considering studies from different climatic


zones;

Dependence on soil type and soil condition;

Seasonality of leaf cover, which may introduce variability in the efficiency of flood
attenuation; and

Time to establish, with mature forests providing the desired effect.

Other issues that need to be considered relate to physical constraints upon commercial
forest cultivation, land use change and possible effects on sediment deposition and
erosion rates.
The study identified that the 10% attenuation of flood flows from a fully forested
catchment are most likely during a high order flood event and that a 25% reduction is
likely for a low order event.
The literature review considered investigations via monitoring of catchments and by the
use of hydrological modelling. Summary of the reviewed literature is provided in Table 3.

-24-

Table 3: Summary of catchment scale estimates (from monitoring or modelling) of the


effect of afforestation on peak flood flows (Jacobs, 2011).

Riparian Woodland
In terms of flood management benefits, riparian woodlands function in a similar way to
floodplain woodlands by increasing roughness, slowing the flow and increasing storage but on a different, smaller scale. In addition, the presence of large woody debris within
the stream channel can delay flood flow, promote out of bank flow and increase storage.
These riparian 'buffer strips' can also have wider benefits associated with sediment
control and erosion, protection of water quality and maintenance of habitat diversity.
The study reports that the literature indicates that 60 minute delays in flood propagation
are achievable for both upper and middle catchment locations. These benefits were
found to hold true for both low and high order flood events. A summary of reviewed
studies is provided in Table 4.

-25-

Table 4: Summary of riparian woodland studies (Jacobs, 2011)

Floodplain Woodland
Floodplain woodlands can provide flood risk management benefit by providing a physical
obstacle to the movement of water. In addition, woodlands also enhance soils storage
potential, by improving infiltration and increasing water uptake. The increase in hydraulic
roughness and diverse network of backwaters and ponds characterised in natural
floodplain woodlands increases water storage, retains water on the floodplain and could
therefore have significant potential in flood management.
The two main studies reviewed (the River Carey and the River Laver) suggest that for an
increase in floodplain roughness alone, due to woodland planting, the increase in flood
water level could be up to around 300mm, possibly higher. These studies consider high
order events and though there is a suggestion in some studies that floodplain woodland
may be less effective during extreme events this range is still considered reasonable (on
the basis of the Carey and Laver studies). A summary of these studies is provided in
Table 5.

Table 5: Summary of floodplain woodland studies (Jacobs, 2011)

-26-

Upland drain blocking


The scientific literature was reviewed for upland drain blocking, including both monitoring
and modelling studies. Investigations undertaken for the Coalburn and Blacklaw Moss
studies were considered of particular relevance and with their findings being consistent
with other studies. It was determined that introduction of ditching reduced time-to-peak
by 20% in Blacklaw Moss Law although the effect would diminish over a 20 year period.
In the experimental peat catchment of Coalburn, intensive upland drainage decreased
time-to-peak by 20% initially and over a 20 year period it was assessed this would reduce
to about 5%. Benefits are likely to focus on upland parts of the catchment and diminish
downstream.
Wetlands and floodplains
Wetlands cover a range of different types. It was considered that literature of wider
catchment wetlands is less extensive than that for other Natural Flood Management
techniques and there is therefore less certainty over the effects of wetland creation within
the wider catchment. The two main hydraulic processes with a bearing on flood
management that might be influenced by wetland restoration and enhancement are likely
to be conveyance and water storage capacity. These in turn are affected by area of
wetland / floodplain available, soils condition, surface roughness and topography and
slope.
These aspects are therefore considered along with any other influences that might be
specific to the particular type of wetland. The hydrological role of wetlands is widely
accepted as being significant, either influencing floods, recharging groundwater or
augmenting low flows. Consequently they comprise an important consideration for water
management.
The literature reports that wetlands perform varying functions and their effects are not
always easy to discern. However, the vast majority of studies into floodplain wetlands
reviewed in their paper concluded that these wetlands did reduce or delay floods.
However the roles of other wetlands on slopes and in the headwaters were not as
conclusive.
Based on the studies reviewed, the consensus is there is a general benefit from
floodplain restoration and enhancement whether by a direct increase in the floodplain
storage capacity or through floodplain reconnection. In some cases the attenuation effect
has been identified to be approximately 10 to 15%.

(J)

Greater working with natural processes in flood and coastal erosion


risk management (Environment Agency, 2012)12

This report was prepared in response to Sir Michael Pitts recommendation for Defra, the
Environment Agency and Natural England to work with partners to establish Catchment
Flood Management Plans. The report provides additional explanation on what natural
processes are in flood management and to assist practitioners in achieving greater
working with natural processes. While the report does provide some commentary on the
evidence base for Natural Flood Management measures, this is considerably lighter than
the commentary provided in other literature reviews. There is some beneficial guidance
on how to maximise project and programme success, the guidance is aligned with the
system in England and Wales and is not applicable beyond general principles in
Scotland. Finally the report provides an overview of current initiatives to support the
increase in flood risk management via Natural Flood Management methods:

The Land Use Jigsaw discussed in item (E)

The Land Management CFMP Tool The EAs CFMP tool allows the quantitative
assessment of land use and land management changes on the flood hydrograph.

-27-

The tool helps target land management approaches that reduce flooding and
potentially opens up the comparison of Natural Flood Management measures
against traditional engineering.

Defra Multi-objective Projects three flagship demonstration projects to


demonstrate Natural Flood Management projects. (Source to Sea, Moors for the
Future & Slowing the Flow at Pickering)

A practical framework for integrating biodiversity A Natural England project


which suggests a framework to identify projects which have benefits for both
biodiversity and flooding. A toolkit was developed as part of the project to assist
the practitioner.

Appraisal Guidance the Environment Agencys Flood and Coastal Erosion Risk
Management Appraisal Guidance includes greater working with natural processes
for future flood and coastal erosion projects. It encourages multi-criteria analysis
through the assessment of environment, social and economic benefits.

Working with natural processes support guidance additional guidance showing


how flood risk management projects can work with natural processes.

Land management flooding resource pack the Environment Agency and Natural
England are currently working together to develop a training pack intended for
catchment sensitive farm officers, environment officers and agri-environment
advisors.

Flood storage area project to allow for greater use of flood storage areas, and
for greater biodiversity benefits when flood storage areas are constructed, a tool
has been developed to assist in planning and design. It includes a GIS database
of previous flood storage schemes, a cost estimating tool and a design decision
flow chart.

-28-

3.3 Science relating to a range of Natural Flood Management


types
Following the definition of Natural Flood Management provided in Section 1, a literature
review of the science relating to a wide range of Natural Flood Management measures
has been undertaken. This review has been compiled based on published material
identified within previous Natural Flood Management literature reviews as presented in
Section 3.2, internet searches, the recommendations of CREW (Werritty et al., 2012)4,
and through consultation with members of the projects Technical Advisor Group. The
measures are presented in order of their general occurrence within the hydrological cycle,
starting with processes found in catchment headwaters and finishing with coastal
flooding:
(A) Upland land use issues
(B) Pasture land use issues
(C) Arable land use issues
(D) Field drainage
(E) Afforestation of uplands
(F) Upland drain blocking
0 The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of data
sets and wetlands are subdivided into classes based on wetland type. This data set
represents a useful resource for a national overview of wetlands in addition to more site
specific information.
Constructed farm wetlands
(H) Wetland creation (non-floodplain)
(I) Gully (cleugh) planting
(J) Floodplain reconnection
(K) Using vegetation to slow channel flows and stabilise banks (riparian planting)
(L) Using vegetation to slow out of bank flows (floodplain planting)
(M) Washland creation
(N) Artificial large woody debris dams and boulder placements
(O) River restorationError! Reference source not found.
(P) Beach recharge
(Q) Sand dune management
(R) Wave attenuation over saltmarsh and mudflats
(S) Managed coastal realignment (reduction of estuarine surge)
(T) Reefs and breakwaters
(U) Strandline and beach management techniques

-29-

(A)

Upland land use issues

This section discusses land use issues including grazing and land management
practices. Given that there is a notable volume of literature relating to upland drainage
this area is discussed separately in Section 3.3(F).
Upland soils, which are typically peat based in Scotland, are sensitive and can be easily
degraded by land use practices such as overgrazing, drainage or poorly conducted
muirburn. In many areas this has led to vegetation change and retreat, often exposing
peat causing it to desiccate in dry weather, leaving it prone to erosion from both wind and
water. Over the decades this has lead to measurable impact upon the timing and
magnitude of floods (Longfield et al, 1999; Conway and Millar, 1960; Robinson,
1960)131415.
Reducing stocking density encourages regeneration of upland vegetation communities,
including ericaceous plants that tend to have higher rainfall interception values compared
to grasses and bare ground. For example heather provides a consistent year round
interception loss (16 to 19%) although this high interception rate is somewhat offset by a
relatively low transpiration rate (200 to 420mm/yr) (Nisbet, 2005)64. Note that the effects
upon vegetation will be seasonal with the magnitude of these fluctuations being
dependent on the vegetation type.
Lower livestock density can also reduce trampling and subsequent compaction of the soil,
reducing the antecedent water content and increasing the infiltration capacity (Godwin
and Dresser, 2003)16.
The shooting of grouse is one of the main forms of income for the uplands and despite a
recent decline in grouse numbers grouse shooting has continued to remain a popular and
profitable sport in the UK (Davies, 2005)17. The decline has been linked to several
reasons which include the loss of heather moorland and subsequent closure of several
shooting estates, reduction in gamekeepers managing the moor, disease and an increase
in predator species (Cox, 2003; Hudson, 1992)1819.
Prescribed burning of moorland heather (Calluna vulgaris), often referred to as muirburn,
is a traditional management tool to improve grouse densities for sport shooting. Burning
helps prevent establishment of woody species, reduce plant litter and release nutrients
(Gimingham, 1975)20. Such management then stimulates earlier growth of vegetation
and temporarily increases the accessibility, palatability, and nutrient content of forage for
grouse and livestock, thus optimising the populations for maximum profit (Lawton, 1990;
Tucker, 2003)2122. Although most of the burning is conducted on shallow peat and
podzolic soils, it also encroaches onto adjacent areas of deeper peat found on blanket
bog (Yallop et al, 2006; OBrien et al., 2006)2324.
The mosaic of different heather ages created by muirburn is viewed by Scottish Natural
Heritage as offering benefits to native wildlife such as birds, insects and reptiles (SNH,
2012)25. The Single Farm Payment require moorland to be maintained in Good
Agricultural and Environmental Condition, the Muirburn Code (Scottish Government,
2011)26 is used to define the standards expected of land managers.
The replacement of a thick blanket of mature heather and associated understory of litter
with a patchwork of immature bare ground, heather and grasses will reduce the canopy
interception from that reported above for heathers to much lower rates. It is also reported
that the infiltration capacities of cotton grass moorland were as low as 2mm/hr while
cowberry areas have an infiltration rate just less than 20mm/hr (OBrien et al, 2007)27. A
mature ground cover will have a higher hydraulic roughness and will serve to slow
overland flow (Lane et al., 2003)28.
A study by English Nature on the effect of heather burning on soils, hydrology and
ecology summarises the key literature relating to muirburn. Work by Mallik et al (1984)29
-30-

found that infiltration rates on freely drained brown podzolic soils declined by 74% on
burnt heather plots compared to unburnt plots, probably as a result of ash particles
clogging the soil pores in the upper surface layers. Their results concur with similar
studies in other countries, but it is worth noting Kinakos (1975)30 Phd thesis reported that
burning in a heathland in Scotland increased infiltration by 30%. Mallik et al suggest that
Kinakos method of measuring infiltration may have been unsatisfactory, as it was
unsuccessful when tried in their study. Mallik et al (1984)29 also reported that the water
retaining capacity of the burnt podzolic soils was significantly higher than the unburnt
soils, particularly in the upper 2 cm of the soil profile but the difference was detectable to
10 cm. This would have had the joint effect of reducing the rate of water percolation
through the soil and increasing its water retention capacity.
Severely burnt soils would be expected to generate an increased proportion of rapid
runoff, through increased water retention in the soils and the absence of
evapotranspiration and rainfall interception by vegetation (Mallik et al., 1984)29. An
experimental plot scale study on Dartmoor, observed no increase in runoff from an area
of managed burn (Meyles, 2002)31. Further to this no differences in soil moisture levels
were detected by Meyles under wet conditions between burnt and unburnt areas, for up
to two months after burning.
One of the few catchment studies on blanket peat was carried out in the 1950s at Moor
House in the Pennines, the results of which suggested that severe burning of the peat
surface reduced the water storage capacity of the soil and lowered dry weather flows, but
increased peak flows in drainage ditches (Conway and Millar 1960)32. By analysing a
longer sequence of Conway and Millars data, Robinson (1985)33 confirmed their
conclusions. Although moderate burning, without total destruction of the mire community
or the heather, tended to produce a flashier runoff hydrograph, the effect was very much
less than that of drainage by gripping.
A study in the North York Moors, reported via a Phd thesis (Dunn, 1986)34, identified that
land-use change from open heather moorland to burnt ground promoted reductions both
in evapotranspiration levels, especially potential demand and in moisture deficits. The
study identified that complete catchment muirburn could lead to a doubling of the
hydrograph peak predominately because of the increased antecedent moisture condition.
Ramchunder et al (2009)35 highlights the absence of detailed knowledge relating to the
hydrological impact of rotational burning and provides some hypotheses on the effects
that rotational burning might have at a catchment scale. These predictions are
judgement based.
The EMBER Project (Effects of Moorland Burning on the Ecohydrology of River systems)
at the University of Leeds includes increasing the understanding of moorland burning on
hydrology (EMBER, 2012)36. The project is using a paired catchments approach
consisting of five control catchments and a further five were muirburn is undertaken. The
catchments in the Pennines are all less than 3sqkm. Unfortunately to date the project
has not published any material.

(B)

Pasture land use issues

In relation to the grazing of pasture DEFRA research project FD2114 (OConnell et al.,
2004)83 concludes that:

There is quantifiable evidence from managed grassland in south west England


that overgrazing and trampling by stock can decrease surface infiltration by up to
80%(Heathwaite et al., 1989)37 and can double surface runoff at the field and hill
slope scale (Heathwaite et al., 1990)38

There is evidence from field sites of reduced infiltration and increased field-scale
surface runoff, when some livestock management practices are undertaken in less
-31-

than ideal soil and weather conditions (Holman et al., 2001; Goodwin and
Dresser, 2003; Hollis et al., 2003)394041.
DEFRA research project (BD2304) (Cranfield University, 2007)42 reported that there is a
clearly demonstrable link within the literature between soil compaction and livestock
densities. This finding echoes the concluding message of an earlier Environment Agency
R&D project (Johns, 1998)43 which was predominately based on anecdotal and prima
facie evidence. Compaction damage due to trampling may also be cumulative with time
but this could be dependent on soil texture. Therefore the 2007 DEFRA project adopted
the following approach:

Mapping the extent of soil compaction, by;


o

Estimating the vulnerability of soils to compaction

Estimating the stress which soils are subjected to

The hydrological impacts of compaction

The study identified that the vulnerability of soils to compaction can be estimated using
the Soil Suitability for Grassland method proposed by Harrod (1979)44 with the method
being based on:

Topsoil retained water capacity

Depth to the impermeable horizon

A combined soil and climate factor (soil wetness class)

Unfortunately it is thought that such a dataset of the Harrod (1979)44 soil compaction
vulnerability does not exist for Scotland nor is it clear if it is technically feasible for an
identical Scottish version to be produced. Further to this, it is not clear if an alternative
approach could be developed. Work by Palmer (2004)45 using a technique developed
during work on flood-affected catchments in 2000 provides the most significant body of
evidence of the extent of compaction but only covers twelve catchments mostly in the
south and west of England (one is from Wales) and the sites cannot be claimed to be
representative of Britain as a whole. By analysis of Palmers published field data the
2007 DEFRA research project identifies a correlation between vulnerability and actual
degradation in so far as low vulnerability soils were found to have predominantly low
levels of actual degradation and, conversely, high vulnerability soils were linked to the
highest proportion of high degradation sites. It also suggest that soils under ley
grassland (younger than 5 years) are more susceptible to compaction that those under
permanent grassland.
Modelling using MIKE SHE of the River Camel catchment (Williams, 2004)46 suggested
that minor changes in the moisture characteristics of soils beneath pastures (12% by
catchment area) produced an increase of approximately 18.75% increase in flood peaks.
These changes are presented within the paper as being representative of degradation
due to intensive grazing. A site based study of a small catchment in Dartmoor identified
that the soil water field capacity is lower for heavily grazed pastures (Meyles et al.,
2006)47. The study also proposes that intense grazing causes improved rapid surface
flow routes due to the presence of trampled livestock tracks. The dominance of the
saturated excess overland flow pathway within grazed grassland is highlighted following
observations within the Pontbren catchment (Wheater et al, 2008)104. Observations at
Pontbren also indicated that, for the soils present in the study area, surface runoff as a
consequence of exceeding the soil infiltration capacity was not detected.
The 2007 DEFRA research project (Cranfield University, 2007)42 assessed the soil stress
class based on a simplified Grazing Livestock Unit (GLU) density (per unit area)
calculation where all cows were assigned a value of 1GLU and sheep a value of 0.1GLU.
Parish livestock statistics were used to gain an estimate of local livestock numbers. It is
-32-

then assumed that soil stress is directly proportional to the GLU density allowing the GLU
density map to be divided into three classes (high, medium and low). The report accepts
that practices such as winter grazing and trafficking are omitted from this approach
however in the absence of viable data it is currently the only means of gaining at least a
slight insight into the compaction caused by livestock. The GLU density was calculated
on a 1km grid for England and Wales for all grid cells with more than 40 percent
grassland. It should be noted that the density was not based on the area of grassland
within the 1km grid cell as this reportedly caused the inappropriate shifting of high stress
areas to the lowlands. Instead the density was based on the number within each 1km
grid cell regardless of grassland proportion.
The spatial estimates of vulnerability to compaction and the stress which soils are under
was then combined to create a measure of the severity of compaction using Figure 6.

Figure 6: Derivation of the risk of compaction classes


Grazing Livestock Unit)

42

(Holman et al, 2001) (GLU

Following the identification of a method to estimate the spatial distribution and severity of
compaction the 2007 DEFRA research project (Cranfield, 2007)42 identifies the work
undertaken by Packman et al. (2004)48 to quantify the hydrological impacts of
compactions. This methodology is based on the assignment of an appropriate analogue
HOST (Hydrology of Soil Types, Boorman et al., 1995137) class to represent the degraded
soil. The analogue HOST class can then be used to estimate the Standard Percentage
Runoff (SPR) and Baseflow Index (BFI). The rationale for the proposed changes is that
soil structural degradation, in the form of topsoil and upper subsoil compaction or
seasonal capping and sealing of soil surfaces, causes a reduction in the effective soil
storage, which in turn results in increased surface runoff. The method recommended that
the modified HOST is adopted for all areas within the catchment which are cereal or
lowland grass, with all other areas outside this being based on the published HOST class
by Boorman et al. (1995)49.
The DEFRA research project (Cranfield, 2007)42 reported using a modified version of
Packman et al. (2004)48. The study identified an indicative absolute increase in SPR as
an average of less than 6% for the England and Wales study area. The largest increase
-33-

in SPR was identified for catchments with a Boorman et al. (1995)49 SPR of 25 to 45%
(undamaged). Catchments with high SPR showed little impact from soil compaction due
to the lower density of managed grasslands and wetter antecedent conditions. The small
absolute increases in SPR represents relative increases of less than 13% in most
catchments, although a number (mostly in the south west of England) showed indicative
increases of 13 to 41%. It should be noted that the relative increase in SPR are typically
much lower than those observed for plot studies (Heathewaite et al., 1989; Heathewaite
et al., 1990)3738. The DEFRA research project (Cranfield, 2007)42 suggests these
differences are due to the moderating effects of landscape connectivity.
Although modification of HOST class proposed by Packman et al. (2004)48 was initially
developed to be used with the Flood Estimation Handbooks Rainfall-Runoff Method via
the revision of SPR, it is proposed that the same approach could be used to modify BFI
values which are used within the Revitalised Flood Hydrograph Method (Kjeldsen,
2007)50.

(C)

Arable land use issues

There is a significant body of information reporting that arable land has an impact on
runoff generation (Cranfield, 2007)42. The most relevant literature regarding arable crops
appears to have been completed by Holtan and Kirkpatrick (1950)51. This American
observation based research indicates that long established pasture has a higher
infiltration rate than arable crops, especially during periods of the year when cropped
ground is bare or crushed. It should be noted that the published rates do not account for
soil type or for variation in land management practices. Importantly, the size of
agricultural machinery has changed dramatically since 1950, thus present day soils are
likely to be more trafficked and consequently more compacted (Cranfield, 2007)42.

Figure 7: Mass infiltration rates (Holtan and Kirkpatrick, 1950)

-34-

51

Packman et al. (2004)52 proposes a modification of HOST class to assess the


hydrological impact of arable land use to that proposed for the impact of grazing. The
method is based on the application of the FARM tool (refer to Section 3.4) but the recent
Land Management CFMP Tool (JBA, 2008)53 developed for the Environment Agency is
also based on the methods developed by Packman et al. (2004)52. The FARM tool takes
account of the following factors for arable land:

Soil exposure

Land slope

Soil degradation

Crop Management Practices

Crop cover

Land drains and ditches

Hillslope form

Tramlines, tyre tracks, roads and trafficking

Hedgerows and buffer zones

Wetlands, water-logged zones, ponds and flood storage ponds

Through assessing the degradation of the land the tool assists in the selection of an
alternate HOST class using the guidance provided by Packman et al. (2004)54. It should
be noted that the guidance reminds the user that the FARM tool functions at a local field
or farm scale while the FEH Rainfall-Runoff Model (and ReFH Rainfall-Runoff Model) is a
catchment scale lumped hydrological model. In applying the methodology it is therefore
necessary to appropriately weight the baseflow index (BFI) or standard percentage runoff
(SPR) value based on the prevalence of the degraded land within the modelled
catchment.

(D)

Field drainage

This section discusses the literature relating to the blockage of agricultural field drainage,
upland drain blocking is discussed in Section (F).
Field-drainage and associated subsoil treatments can increase or decrease peak drain
flows and the time to peak flow by as much as two to three times either way with the
behaviour apparently dependent on soil type and wetness (OConnell et al., 2007)55.
These very variable impacts of field-drainage systems are likely to render flood
management gains from changes to field drainage systems a complex, data intensive
process which retain a high level of uncertainty. It was noted that drainage was found to
alter runoff timing, but it had little impact on runoff volume.
Institute of Hydrology Report 113 (Robinson, 1990)56 reported a comprehensive series of
field studies, catchment studies and regression analysis to investigate the impact of field
drainage on flood flows. The study found that general statements that field drainage
causes or reduces downstream flooding are over-simplifications of a complex process.
At a field scale it was concluded that agricultural drains reduce peak flows originating
from clay soils while having the opposite effect for permeable soils. The type of drainage
was also shown to be significant with moling in combination with pipe drains increasing
peak rates, while open ditches give the highest discharge rates. At a catchment scale it
was identified that it was necessary to differentiate between open channel drainage and
piped drainage. Arterial open drains were found to increase flows downstream due to
accelerated flows and reduced overbank flooding, with the effect increasing with channel
size and flow.

-35-

(E)

Afforestation of uplands

This section discusses the literature relating to the change in runoff rate and timing which
afforesting uplands can introduce. For details of literature relating to the attenuation and
other effects which can be introduce by placing vegetation within flow pathways (riparian,
floodplain and gully woodland) refer to items (K), (L) and (I).
Much of Britains native upland woodland has historically been felled, as well as being
lost to a combination of overgrazing and burning over time. That which remains is often
degraded because natural regeneration is prevented by grazing and browsing pressure
from sheep and/or deer, which have increased in population over the last 50-60 years
and which often use the woods for shelter (Hardiman et al., 2003)57.
DEFRA commissioned a major study (FD2114) (OConnell et al., 2004)83 which reviewed
the impacts of a range of rural land uses including forest on flood generation. In the
reports critical assessment of the monitored evidence for afforestation affecting river
flood flows the following was concluded:

There is quantifiable evidence for the effect of conifer afforestation, but it is


difficult to interpret. Most catchment monitoring studies in the UK have focussed
on upland catchments dominated by conifer forest or rough grassland. These
have shown that there is a tendency for the water yield to be less from forest than
pasture. There is evidence that afforestation affects peak flows and times to peak.
However this evidence shows that the impact of forests on flood generation
cannot be predicted simply by using the above-mentioned water yield data. In
their general review of the history of forest hydrology, McCulloch and Robinson
(1993) conclude that forests should reduce flood peaks, except for the effects of
drainage and forest roads. In the Coalburn experiment, peak flows actually
increased by 20% in the first 5 years after forest planting (decreasing to 5% after
20 years) and times to peak decreased. This is thought to be the result of plough
drainage and ditching;

A review of results from 28 monitoring sites throughout Europe (Robinson et al


2003) concluded that the potential for forests to reduce peak flows is much less
than has often been claimed, and that forestry appears to ... probably have a
relatively small role to play in managing regional or large scale flood risk.

There is consensus within the literature that the introduction of woodland cover can
influence a number of key hydrological processes, namely:

Rainfall interception (and subsequent evaporation)

Moisture transpiration

In turn the following hydrological processes might reasonably be expected to be


influenced, although the extent to which they are affected is subject to uncertainty within
the literature:

Antecedent moisture conditions

Rainfall infiltration rates and volumes

Surface runoff rates and volumes

Groundwater flow volumes

Rainfall interception is the capture and subsequent evaporation of the water before it
reaches the ground surface, it is the difference between the gross rainfall above the
vegetation canopy and the net rainfall at the soils surface. Broadleaved woodland in the
UK typically intercepts 10 to 25% of annual rainfall, compared to 25 to 45% for conifer
stands (Calder et al, 2003; Hall; Nisbet, 2002)585960. Fitting with what might be reasonably

-36-

expected, it has been observed that the greatest interception for broadleaved trees is
when they are in leaf, averaging 40% or more (Roberts and Rosier, 2005)61, in contrast
with 3 to 12% when trees are leafless (Roberts and Rosier, 2005; Harding et al, 1992)6261.
Hardiman et al (2003)57 proposes that this high summer interception rate may become
increasingly important in attenuating local run-off given climate change predictions citing
an increase in the frequency and intensity of summer storms. However the seasonality of
Scottish flooding is reported as being predominately autumn/winter (Reed et al., 1999)63
and the assumption that interception losses within the tree canopy will be present during
the design event peak is possibly overoptimistic. Trees with lighter canopies, such as
ash have been observed to intercept only 10 to 15% per annum compared to 15 to 25%
by oak and beech (Roberts and Rosier, 2005; Harding et al, 1992)6162. It is reported that
the annual interception rates (by percentage) remain remarkably constant over a wide
range of total rainfalls (Nisbet, 2005)64. A summary of the reported interception rates for
conifers and broadleaf species is provided in Table 6.
It is reported that interception loss as a proportion of rainfall reduces with increasing
storm size, reaching a maximum of 6-7mm/day (Nisbet et al, 2011)65. However,
interception varies greatly throughout the year and in particular, declines with the size
and intensity of a given rainstorm. This reflects the relatively small water holding capacity
of forest canopies, equivalent to only a few mm of rainwater. As a result, interception
losses are likely to be <10% for individual major storm events (Nisbet, 2005)64.
It has been reported that transpiration rates vary little between the two forest types, with
annual losses mainly falling within a relatively narrow range of 300350mm (Roberts,
1983)66. Higher values of 390 to 410mm have been derived for native broadleaved
woodland in southern England, with a range of 360mm to 490mm varying by study
(Roberts and Rosier, 2005)61. It can be inferred that if both interception and transpiration
are considered together, and assuming an annual rainfall of 1000 mm, conifers could be
expected to use some 550 to 800mm of water compared with 400 to 640mm for
broadleaves (Nisbet, 2005)64. Other factors influencing use of water are woodland
structure and age, however given the already complex issue between transpiration and
flooding these will not be discussed. A summary of the reported annual transpiration
rates for conifers and broadleaf species is provided in Table 6.
Land
cover
Conifer
Broadleaf
Grass
Heather
Bracken
Arable

Transpiration
(mm/yr)
300 350
300 390
400 600
200 420
400 600
370 - 430

Annual
interception
(%)
25 45%
10 25%
16 19%
20%
-

Winter
interception
(%)
25 45%
3 12%
16 19%
?

Summer
interception
(%)
25 45%
40%
16 19%
50%
?

Table 6: Summary of annual transpiration and interception data for a range of land cover
64
types (Nisbet, 2005)

There is a general consensus in the published literature that, within a British context,
woodlands reduce annual water yield (Calder, 2009)67. The reduction in water yields as a
consequence of the increased evapotranspiration by trees might be of concern for
catchments which are already water stressed.
It should be noted that soil type and condition are significant in determining the pathways
and consequent timing of water draining from the land. Shallow, poorly draining soils are
prone to rapid surface runoff responses, while deep, freely draining soils facilitate much
slower, more attenuated, groundwater pathways. Native woodland rarely involves the
formation of continuous cultivation channels or drainage treatments to aid establishment

-37-

and the lack of disturbance helps to increase soil organic matter and improves soil
structure, resulting in an increase in ability infiltrate and store water (Forestry
Commission, 1999)68. Observations in Pontbren in Wales have found infiltration rates to
be up to 60 times higher under young native woodland than heavily grazed pasture, with
rapid improvement over just two years after stock removal and planting (Bird et al.,
2003)69. More recent modelling work undertaken by the same team suggests that whole
catchment forestation would reduce peak flows by 10 to 54% while being reduced by 2 to
11% via the use of optimally placed shelter belts (Jackson et al, 2008)70. Infiltration and
permeability will be predominately characterised by the soil type however there does
seem to be consensus that forest soils are more permeable than grassland (Aaloui,
2011)71. Therefore, by the conservation of mass they are a reasoned choice when
deciding land cover for flood management purposes.
Forest Research undertook a review of Short Rotation Forestry (SRF), incorporating
Short Rotation Coppice (SRC) (Nisbet et al., 2011)72. This review identifies that the fast
growing species often used in SRF and SRC have a high water demand. This increased
water demand is attributed as a consequence of high transpiration rates with interception
being comparable to conventional forestry. The review focuses on the issue of reduced
water yields particularly in relation to already stressed waterbodies, however the potential
to reduce groundwater levels, and by extension groundwater flood risk, is discussed
using the example of Eucalyptus plantations (Calder, 1992)73. Calder observed an Indian
plantation to have evapotranspiration rates exceeding the available rainfall (3400mm/yr
and 2100mm/yr respectively), it is anticipated that similar effects may be observed using
high water demand native species.
The development of the Flood Studies Report (NERC, 1975) established a link between
catchment wetness (and in particular the presence of any soil moisture deficit) and the
resulting design flood magnitude. It could be hypothesised based on this evidence that as
afforestation leads to an appreciable alteration of the typical seasonal soil moisture
deficits experienced in a catchment the reduction may be sufficient to influence the likely
magnitude of floods. However it should be noted that the Flood Studies Report
concluded that the extent of woodland in the UK is not statistically significant in affecting
peak flows of large flood events and hence the absence of a forest cover catchment
descriptor within the Flood Studies Report and subsequent Flood Estimation Handbook
hydrological methods.
A study at Leadburn, Scotland, implicated pre-planting plough drainage of deep peat as a
major source of local runoff (David and Ledger, 1988)74, whilst other research has
focussed upon the influence of forestry tracks and localised soil compaction. Historically
speaking intensive drainage and plough based planting techniques have been employed
in commercial conifer plantations (Rutter, 1963)75, such techniques are now out-dated
(Forestry Commission, 2011)76. That said, 30 years of research at Coalburn, England,
indicates that even with best practice, felling induces an increase in peak flows for
several years after felling, until new trees are well established (Robinson, 1986; Robinson
et al., 1988; Archer and Newson, 2002)777879. Because of the clear-fell cultivation cycles,
commercial forestry may be expected to have a much more variable medium-term effect
upon runoff and flood generation than native woodland of any type which is likely to be
more consistent over this timescale but may alternatively experience seasonal
fluctuations. This is supported by a broad review of upland afforestation across Europe,
which concluded that complete clear felling of a catchment can increase annual peak
flows by 10 to 20% with complete afforestation having a similar but opposite effect
(Calder et al., 2008)80. Although the cycle varies by location and species among other
variables, typically commercially grown conifers in the upland basins of North West
Europe have a short management cycle of 30 to 60years (Robinson et al., 2003)81.
The influence of native woodland restoration on upland hydrology will thus depend in part
upon what it replaces, and its situation within the catchment. For example, a literature
-38-

review based study for the Woodland Trust (the study itself being peer reviewed)
concluded that restoring native broadleaves from upland conifers could lead to a 10%
increase in localised flooding of minor floods (up to 1 in 5yr) due to lower water use and
greater potential runoff in winter and spring (Calder et al., 2009)67. However, replacing
rough or unimproved grassland with native woodland should have a roughly equal and
opposite effect, with further attenuation in summer from species with a high water use
such as willow (Calder et al., 2008)80. Conversely, research in the UK to date does
demonstrate that changes to hill slope woodland is unlikely to have a measurable impact
upon runoff or flood hydrographs in extreme rainfall events (McCulloch and Robinson,
1993)82.
There is an absence of evidence for afforestation in catchments greater than around
10sqkm (OConnell et al., 2004)83. This is in part due to the complexity of measuring
these effects when compared to other activities occurring within the catchment and
climatic variability. Modelling in Plynlimon in Wales found that it was frequently too
challenging to discriminate the effects of woodlands against, for example, climatic
variation or farming pressures further downstream (OConnell et al., 2004; Robinson and
Dupevrat, 2005; Kirby et al., 1991; Neal, 1997)84838586. This is likely to be a factor of the
very low woodland coverage in the UK. The Woodland Trust review (Calder et al., 2009)67
indicated that unless the coverage is more than 20%, the effects of woodland planting
would be unlikely to have any significance on the flood hydrograph. That said, it should
be noted that a likely mechanism for the achieving the forestation of catchments could be
via a large number of small schemes leading to a more notable net effect rather than a
small number of extensive whole catchment interventions, so it would be unwise to
consider in isolation that a small intervention is too minor. Conversely, decisions to
undertake extensive planting raises implications for other considerations such as habitat
priorities in a given upland area, such as open moor and semi-natural grassland
biodiversity, farming and other land uses. Nevertheless, the relative scarcity of seminatural woodland in the UK, especially in upland areas, may provide a weight for its
expansion.
Furthermore, much of the evidence of woodland effects upon flooding has been focussed
beyond the UK (particularly in the USA), where landscape, forestry practices and
woodland extent and type can be significantly different, making comparisons difficult. The
Centre for International Forestry Research cautions against referring to woodlands as
sponges and suggests that ill-informed policies can have detrimental effects on local
communities (CIFOR, 2005)87. Based on a recent report (CIFOR, 2005)87 the Centre for
International Forestry Research agrees with the UK literature consensus that
predominantly forested catchments often show little evidence of actually preventing large
scale floods. The European Basins (Robinson et al., 2003)88 report concluded that
forestry appears to have a relatively small part to play in managing regional or large scale
flood risk.
Jacobs (2011)3 provides the following concise summary of published catchment
monitoring and catchment modelling studies relating to the impact of forests on flood
flows:
a) Catchment monitoring findings
Catchment monitoring studies in the UK are few. The Institute of Hydrology (now
known as the CEH Wallingford) ran a number of long-term research catchments
studying the effects of commercial coniferous forest on hydrology (viz Plynlimon in
central Wales, Balquhidder in the central southern Highlands of Scotland, and
Coalburn in northern England). The Balquhidder work did not include an assessment
of the effects of forestry on flood flows, whilst the Coalburn study has been more
focused on quantifying the effects of drainage ditches, though Robinson (1986)15

-39-

does speculate that the negative effects of drainage ditches are in part balanced out
by the establishment of the trees.
The Institute of Hydrology used Plynlimon to assess the forest influence on floods
(Kirby et al., 1991)85. The findings of this were that an identifiable effect could be
detected for small floods (<2-year events) but that no effect was discernable for
larger ones. These findings could however be partially affected by the presence of
the old style drainage ditches used in the preparation of the land at the planting
stage.
The findings of other catchment studies were sought in the review of Price et al
(2000)99. Only relevant studies from the USA were identified at the time. These all
considered small (<1.5 km2) upland catchments that were instrumented to assess
the hydrological effects of clearfelling. These studies were relatively short term and
tended to sample events of limited magnitude. Bates and Henry (1928)89 observed a
55% increase in flood peak flows and a 23% increase in flood volumes immediately
after the clearfelling of a natural mixed conifer and deciduous forest cover in
Colorado. Hornbeck et al (1970)90 observed an average 13% and 21% increase in
flood peaks and volumes respectively after a mixed conifer and deciduous vegetation
cover was clearfelled in New England. Rothacher (1970)91 however detected no
significant difference for the largest floods when a Douglas Fir forest was clearfelled
from a catchment in Oregon. Swank and Crossley (1988)92 observed peak flow
increases of 15% and flood volume increases of 10% when a mixed forest was
clearfelled with minimal ground disturbance in North Carolina. Higher increases of
30% (peak flows) and 17% (flood volumes) were observed in an adjacent catchment
when it had 48% of its tree cover clearfelled together with 29% thinned and 23%
undercut. In all, apart from the Swank and Crossley study, the possible compaction
of soil during the felling process cannot be ignored.
Other approaches investigating whether the signature of land-use and land-use
change can be detected within the flood flows of the UK river gauge network have
been tried. Beven et al (2008)7 attempted to identify catchment-scale effects of land
use and land management change using available UK datasets, but failed to identify
a clear relationship between these and flows.
The DEFRA FD2114 study was updated in 20076. This summarised the conclusions
of the original FD2114 project with respect to afforestation/deforestation in the
following additional manner:

Climate variability is the dominant factor in influencing the frequency and


magnitude of flood events within a given catchment
There is no clear evidence that local scale impacts can be identified at the
catchment scale. This is probably due to local impacts being smoothed out
during the longer duration flood generating storms and also from the
combined influence of the patchwork of different land uses often found in a
typical catchment.
Upland conifer plantations tend to significantly increase annual evaporation
rates when mature forested catchments are compared to grassland. As a
result, annual water yields from forested catchments are lower than for
comparable grass or moorland catchments.
Conifer forests may reduce low flows through a combination of high
interception and evaporation losses during wet periods and increased
transpiration losses in dry periods. These processes act to increase soil
moisture deficits in dry periods when compared to other vegetation-cover
types and thereby increase the moisture deficit of the soil, delaying the onset
of saturation-excess runoff.

-40-

There is only a limited potential for forests to reduce peak flows for larger
flood events, with a much greater effect on small to medium magnitude flood
events.

Their updated (circa 2007) interpretation of the overall role of forest in terms of the
hypothesis that forests reduce flooding was:
In summary, this hypothesis cannot be accepted for the generation of extreme flood
events at a catchment scale on the basis of available evidence. Nevertheless, well
managed forests can help reduce local flooding and the peak flows of smaller, more
frequent events and potentially for infiltration-excess driven extreme summer
events.
b) Modelling studies
An alternative means of investigating the likely influence of forest is to use modelling
techniques that capture the key land-use effects.
Trent, Severn and Thames modelling study:
Naden et al (1996)93 developed a semi-distributed daily rainfall-runoff model
incorporating a soil water balance model, a runoff model, and a flood routing model
to estimate the impacts of both climatic and land-use change on flood response in
the three large (approximately 10,000km2) English basins: the Trent, Severn and
Thames. Projections suggested that completely covering the basins in coniferous
forest would reduce the 50-year flood peaks by an average of 20%, whilst total
conversion to deciduous forest was predicted to reduce peaks by an average of
14%. The authors acknowledged that the although the model had accounted for the
vegetation cover effects via evapotranspiration and water balance of the soil no
account of possible shifts in timing nor the partitioning between fast and slow
components of the runoff were made.
River Calder (Renfrewshire) modelling study:
Price et al (2000)99 investigated the likely influence of a range of Scottish upland
vegetation covers to flood flows in the relatively small (33 km2) River Calder
catchment in Renfrewshire. Vegetation covers considered were: natural coniferous
woodland, broadleaf woodland, conifer plantation, heather moorland, upland and
lowland grass. The Institute of Hydrologys HYLUC model (Calder, 1990)94,
calibrated on Scottish field data (Hall, 198795; Hall & Harding, 199396; Wright &
Harding97, 1993; Haria & Price, 200098), was used to predict the average seasonal
soil moisture status under the varying vegetation covers. An example of the
significant differences that can occur is given in Figure 8. This shows the monitored
difference in soil moisture deficits at adjacent mature Scots Pine woodland and wet
heathland sites at in the Cairngorms, together with the simulated deficits using the
HYLUC model. The simulations for the River Calder catchment were averaged over
an 11-year period (Figure 9) and used to derive the average Flood Studies Report
(FSR) catchment wetness index for different seasons. This was used together with
design rainfall (minus vegetation dependent interception losses) to run the FSR
rainfall-runoff model to test the resulting peak flow sensitivity to varying land covers
during typical winter and the summer periods (Figure 10).

-41-

(a)

Figure 8: Monitored soil moisture at two adjacent sites in the Allt AMharcaidh
catchment in the Cairngorms: a) mature natural Scots Pine woodland, b) wet heath
moorland. (Dots = averaged field measurement, Unbroken line = HYLUC model
simulation)

Figure 8 and Figure 9 indicate that summer season soil moisture deficits below forest
and particularly coniferous trees tend to be greater than under short stature
vegetation such as grasses or heather. Also the replenishment of the deficit during
the late autumn/early winter occurs later by several weeks under the tree cover than
the other vegetation types. Both these aspects are predicted to exert positive benefit
in terms of flood runoff. However during the winter period (December to February)
the deficits have been replenished under all of the vegetation types and by this
measure the catchment flood response is likely to be similar regardless of vegetation
type. This is reflected in the seasonal flood peak reductions which show a likely
greater effect during the summer than the winter (Figure 10).

-42-

20
0

Average soil moisture deficit (mm)

31

61

91

121

151

181

211

241

271

301

331

361

-20
-40
-60
-80
-100
-120

Conifer
Heather

-140

Upland grass

-160

Day number in year (1 = 1st Jan)

Figure 9: Simulated average seasonal soil moisture deficit for upland grass, heather
and coniferous woodland for the River Calder catchment, Renfrewshire. (Period
simulated: 1983 1993)

a)

b)

Figure 10: Predicted attenuation of: a) winter, b) summer flood peaks due to a
hypothetical 100% conversion of the Calder catchment (Renfrewshire) from upland
grass to various types of woodland\forest or lowland grass (Price et al, 2000). Given as

-43-

a percentage of the peak flows predicted for upland grass. (Prediction uncertainty
band reflects the range of values attributed to the model parameters coming from
different field experiments).

The study by Price et al (2000)99 only attempts to account for likely antecedent
conditions and interception losses but that the effects related to changes to: i) the
soil characteristics (permeability/infiltration rates) and ii) snow accumulation and melt
processes have not been considered due to a lack of sufficient understanding, and
these may also exert an appreciable influence.
The DEFRA FD211483 investigated the modelling approach and concluded the
following in general:
there are serious shortcomings in the rainfall-runoff models and methods
available for the use in the operational assessment of the impacts of land use
change and land management practices. There are three fundamental unresolved
issues: there is no generally-accepted theoretical basis for the design of a model
suitable to predict impacts, it is not known which data have the most value when
predicting impacts, and there are limitations in the methods available for estimating
the uncertainty in predictions. Some general recommendations can, however, be
made for a way forward in rainfall-runoff modelling for predicting impacts. The
modelling should be distributed and be capable of running continuous simulations. It
should also be partially or wholly physically based so that the physical properties of
local landscapes, soils and vegetation can be represented, and it should include
detailed modelling of surface water flow so that the effects of changes can be
tracked downstream. A considerable amount of high-quality field data on impacts will
be needed to support the development of robust methods for predicting impacts.
Since the publication of the DEFRA FD2114 study several detailed modelling studies
that begin to tackle the issue in line with the above modelling recommendations have
been published.
Frances et al (2008)100 report the findings of modelling studies in three catchments
located in Spain (Poyo, 380 km2), Austria (Kamp, 1500 km) and Germany (Iller, 954
km2), with three different climates (semi-arid, sub-alpine and alpine). The models
(different in each case) were all spatially distributed process based models coupled
to hydraulic models of the rivers.
Poyo modelling study:
For the Spanish site a 55% afforestation by pine was predicted to lead to 15% - 38%
peak flow reductions; with the lower reductions more representative of the events
starting with wetter initial conditions and the higher reductions for the drier initial
condition events. Much predicted variability was thought to be due to the
dependence on the combination of the temporal and spatial characteristics of the
storms and the spatial physical catchment characteristics. Peak flows of
approximately the 1 in 200 annual probability (200-year) event were less influenced
by the afforestation with reductions between 3 14%. The largest events (>> 1 in 200
annual probability) had the lowest predicted reductions 0 - 10%. No predicted
change in the timing of the flood peak was reported.
Kamp modelling study:
For the Austrian study pine forest cover was increased from the existing 47% to 86%
of the catchment. A range of reductions to peak flows were reported varying from
very small to 30%. The smaller reductions were predicted for the largest of floods (>
50 year return period), whilst the larger reductions were for the smaller events (<5
year return period). Deforestation of the existing forest from 47% to 0% of the
catchment was also modelled. Very small to 75% peak flow increases were predicted
with again the smallest increases for the largest of floods and the largest increases
-44-

for the small flood events. The asymmetry of afforestation and deforestation
responses was related to where within the catchment the changes are manifest (i.e.
flat deeper soils or shallow steep soils). The modelled changes to peak flows as a
function of both the magnitude of the event and the initial wetness of the catchment
are given in Figure 11. No change in the timing of the peaks was discernable.

Figure 11: Modelled potential reduction of flood peaks on the Kamp following
afforestation (increase in pine forest from 47% to 86%), plus the predicted increase of
flood peaks following deforestation of the existing forest (from 47% to 0% forest
cover). [Source: Frances et al (2008)]

Iller modelling study:


For the German study pine forest was increased from 32% to 61% of the catchment.
Predictions for the large flood events (25 50 year return period) suggested that
both peak flows and flood volumes would only reduce by 0.5 1.0%, whilst for
smaller events (<2 year return period) would experience peak flow and flood volume
reduction of between 2-8%. Figure 12 presents the predicted reductions in flood
peak flows as a function of flood magnitude. The reduction of effect at higher flows
was interpreted as the storage being exhausted at the early stage of medium and
large events, resulting in very small flood peak reductions. No change in the timing of
the peaks was discernable.

-45-

Figure 12: Modelled potential reduction of flood peaks on the Iller following a 29%
increase in pine forest cover. [Source: Frances et al (2008)] Note the 25 50 year event
is estimated to be approximately 800 cumecs.

The studies at each of the three sites suggest that antecedent soil moisture
conditions are important in determining the effectiveness of the forest to reduce
downstream flood peaks. Each study also suggests that the relative size of the
reduction is a function of the magnitude of the event with very large floods little
affected by the presence of forest cover. The authors link this to the concept of a
retention volume provided by the forest. Their view is that although the retention
volume offered by afforestation is likely to be small, it may be useful to combine with
other flood management techniques to bring about a more substantial reduction in
downstream flows.
The authors of the report indicate that appreciable uncertainty exists in the
description and parameterisation of afforestation hydrological processes and that it is
necessary to view the presented predictions in this context.
In the UK several modelling studies have also been undertaken or are on going:
Parrett modelling study:
Park et al (2006)101 report the findings of an initial investigation in to the likely flood
management potential of afforestation in the Parrett catchment (south west England).
The investigation used a high resolution, fully distributed model based on the MIKE
SHE model which encompasses both the hydrological and hydraulic routing models.
Its operation requires a large number of parameters to be used.
The scenarios modelled were: i) arable land converted to coniferous forest (61% of
total area changed), and ii) grassland to coniferous forest (22% of total area
changed). Only autumn and winter rainfall events were investigated. Overall, there
was no clear evidence to indicate that the introduction of woodland either reduced or
increased flooding in the Parrett catchment to a significant extent, but that these
predictions should be viewed only as initial provisional findings.
Pont Bren modelling study:
A detailed multi-scale investigation has, and continues to be, undertaken at
Pontbren, mid Wales (Wheater et al, 2010)103. The study has investigated both the
effects of soil degradation due to compaction by livestock and the effects of tree
cover in the form of both shelter belts and complete catchment cover on flood runoff.

-46-

The experiments focus on soil properties and runoff processes, based on plot-scale
and hillslope-scale measurements nested within gauged first and second-order
catchments (Marshall et al, 2006)102. Plot studies were undertaken on the varying
land uses from which it was identified that infiltration rates and pore space in the A
horizon soil of the woodland were greater than that from the same soil for grazed
grassland. In addition these properties, at this particular site, had a seasonal
dependence probably related to seasonal wetting and drying and increases in
biological behaviour.
The modelling strategy was to: i) first establish a detailed, physically based local
scale model, ii) use this to train a simpler conceptual model, iii) develop a catchment
scale semi-distributed model such that the effects can be simulated downstream
through the river network. Wheater et al (2010)103 present example output from the
above approach for a 4km2 Pontbren sub-catchment for the following scenarios
compared to the baseline condition of the current-day land use:
a)

b)

c)

Removal of recent tree planting (shelterbelts planted [tree type unspecified] in


the last decade, extent not quantified but likely to be <20% of the catchment)
leaving the catchment dominated by intensively grazed grass.
Addition of shelter belts to all grazed grassland sites (again type of tree and
extent not specified but estimated to be likely to be between 10 20% of the
catchment).
Entire catchment is woodland.

The simulated changes in flood peak for these scenarios are: removal of all trees
causes a 20% increase in flood peaks; adding tree shelter belts to all grazed
grassland sites causes a 20% decrease in flood peaks from baseline condition; and
full afforestation causes a 60% decrease in flood peaks from the baseline condition.
The floods simulated are described as relatively frequent events and are
interpreted here as the size of event that is likely to be experienced several times a
year. In Wheater et al (2008)104, the authors also report a speculative simulation for
Pontbren using the extreme rainfall that generated the Carlisle flood of January 2005
(2 day rainfall = 140mm with a rainfall return period of 180 years). The simulated
changes in this extreme case were predicted to be: removal of all trees causes a 37% increase in flood peaks; adding tree shelter belts to all grazed grassland sites
causes a 2-11% decrease in flood peaks from baseline condition; and full
afforestation causes a 10-54% decrease in flood peaks from the baseline condition.
It was also noted that the full afforestation scenario increased the time-to-peak by 30
minutes (though it is unclear whether this may, in part, be attributable to an assumed
deterioration of the field drain system).
The authors emphasise that quite large uncertainty exists for the predictions due to
parameter uncertainty. Additionally attention is drawn in general to the site-specific
nature of these findings. The authors suggested that much work may remain to
provide a more general methodology for national application.
There appear to be an array of local variables (soil, climate, event magnitude, forest
species, forest age, topography, forest extent, catchment shape, other changes in the
catchment etc) which mask the easy development of a simple rule to accurately predict
how the change in forest cover can affect flood flows. While there is consensus in the
literature that the forestation of small catchments can reduce flood flows, particularly for
smaller flood events, based on the available science it is not felt appropriate to directly
relate changes in percentage forest area to a percentage reduction in flow as to do so
would overly simplify what is an immensely complex process. A potential means of
achieving this may be to adopt a more physically based hydrological modelling approach
to represent the role trees play within existing hydrological models. The Leaf Area Index
-47-

(LAI) based method (Equation 1) developed by von Hoyningen-Huene (1981)105 is


recognised as an accurate, non-crop specific means of estimating the maximum canopy
storage capacity (Kozak, 2007)106. This method could be used to modify the storage
capacity of models such as the Revitalised Flood Hydrograph. The LAI is defined as one
half of the total green leaf area per unit ground surface area (Law et al., 2008)107, it
ranges from 0 (bare ground) to over 10 (dense conifer forests).

S Max 0.935 0.498( LAI ) 0.00575( LAI ) 2


SMax = maximum canopy interception storage (mm)
LAI = Leaf area index
Equation 1: Estimation of maximum canopy storage based on the LAI (Hoyningen-Huene,
105
1981)

There may be potential to use a 1km global data set published on an 8 day basis by
MODIS108 to setup of database of LAI values for Scottish land cover and there may be
scope to make account of cover maturity and season. Alternatively, the methodology
adopted by MORECS (Hough & Jones, 1997)216 for estimating LAI variation with land
cover and the effect of LAI could be applied.
As previously discussed, forests can also serve to reduce antecedent moisture
conditions. A range of potential methods of quantifying this effect are discussed in
Section 4.4.2(D).

(F)

Upland drain blocking

This section discusses the literature relating to upland drain blocking in headwater areas
of catchment which are typically moorland or peat bogs. For details of the literature
relating to agricultural field drains refer to Section 3.3(D).
The majority of catchment monitoring studies took place prior to 1990 and may consider
the impact of now outdated drainage practices. There are a number of monitoring studies
which have investigated the impact of introducing drainage while qualitative monitoring
information is sparse when it comes to blockage information. Adopting the simplistic
approach of assuming no long term damage has taken place to the soil and that simply
blocking the drainage channel will return the soil to an identical condition to its predrained state may be overly optimistic.
A literature review undertaken during the work on peatland hydrology by Armstrong et al
(2006)123 noted that grips are typically 50cm deep and 50-70cm wide, and have been
installed in a large proportion of UK peatlands. Their work was on upland blanket peats,
and they note that drainage of lowland raised bogs has included drains both significantly
smaller and larger. Drainage has been shown to influence both the properties of the peat
and the runoff characteristics of outflowing streams. Ramchunder et al (2009)110 noted
impacts on peat shrinkage and consolidation, microbial activity and decomposition, all
influence hydraulic conductivity and water storage capacity as well as flow rates and
processes and susceptibility of the peat to erosion.
Price et al (2003)109 identified that the efficiency of drainage depends on the depth of the
ditch, the ditch spacing and the hydraulic conductivity of the peat. Quoting several studies
they proposed that water tables might be drawn down up to 50m from the ditch in fibrous
peat, but hardly affected at all in decomposed fen peat. On gentle slopes, closer drain
spacing increases the likelihood that water tables will be depressed over large areas of
the landscape. In addition to drawing down the groundwater by their effect on subsurface
flows, ditches can have a significant effect on peatland by interrupting water which would
naturally flow over or near the surface of the peat. Ramchunder et al. (2009)110 noted that

-48-

many blanket peatland drains were excavated so that they run across the slope, typically
following contours and so interrupting natural flow routes.
In other cases, however, grips are often channelled directly into an adjacent watercourse
and in some cases herringbone patterns either side of watercourses have been created.
Ditches running up and down slope are typically thought to produce rapid flow velocities
and have a very different effect on water tables from ditches following the contours, which
have been shown to produce the asymmetric patterns of water table upslope and
downslope of the drain. Holden et al (2004)121 also noted the importance of the position of
grips in the catchment, because ditches serve to lower the watertable (and so may
increase storage and reduce flow peaks in the stream) but they also increase the velocity
of flow once it reaches the drain, and may thus increase the peak flow by speeding water
from one point to another.
As reported in Labadz et al (2010)111 although large amounts of water are held within the
peat of a blanket bog, it is important to understand that little of the water received as
rainfall is retained. Peat has extremely high water content and, in an intact peatland, most
of the water storage capacity is already full. Even intact blanket peat is highly productive
of rapid (rainfall event) runoff and, by contrast, generates little long-term baseflow during
dry periods. For example, in the Feshie catchment in the Cairngorms, Soulsby et al
(2006)112 showed streams draining blanket bog are often ephemeral in their flow regime,
ceasing to flow in driest periods. Many bogs do not act to delay flow into streams (Rydin
and Jeglum, 2006)113. The oftenquoted idea of a peatland as a sponge that soaks up
rainfall and then releases it slowly into rivers is therefore viewed as erroneous.
Ramchunder et al (2009)11041 noted that although drain blocking prevents efficient delivery
of water through the artificial network, and alters hydrological routing to give noncontinuous flow, there is little evidence as yet of larger scale impacts. Grayson et al
(2010)114 have also noted that although carbon storage and flood mitigation are
increasingly used to justify the expenditure on peatland restoration, there is a lack of
reliable evidence of impacts on the flood peak downstream of grip blocking and
revegetation of bare/eroded peat.
Ballard et al (2010)124 applied a physically-based modelling approach to predicting the
effect of drain blocking on peak flows and found that it may depend on site specific
conditions, sometimes increasing and sometimes decreasing peak flows. The study
suggested that the greatest benefits for flood management would be achieved by
blocking drains on steep slopes that are poorly vegetated, but acknowledged that there
are many uncertainties and suggested that better characterisation of overland flow and
drain roughness is required. Subsequent study using the same small scale physically
based hydrological model identified that the drainage of peatlands will increase peak
flows but that drainage blockage will not necessarily reduce flood peaks (Ballard et al,
2011)115. The research suggested that drainage blockage can be prioritised to steeper
and smoother channels but it concludes that field studies are needed to provide data for
future modelling.
A concise summary of the literature relating to upland drain blocking catchment
monitoring and modelling studies has been provided by Jacobs (2011)3 as presented
within this breakout box:
a) Catchment monitoring studies
Perhaps the earliest study in Britain was by Lewis (1957)116 of the Alwen and Brenig
catchments in north Wales. Although principally concerned with the water yield from
upland catchments he noted that forestry drainage operations resulted in a more
flashy runoff response, but gave no further details.
Conway and Millar (1960)14 compared the flows from adjacent upland drained and

-49-

undrained small upland peat catchments in the northern Pennines and found that a
higher flood response came from the drained catchments. Their results and further
data subsequently collected were examined by Robinson (1985)117 which confirmed
their main conclusions regarding flood flows.
The following three experiment sites\catchments offer a more detailed and quantified
understanding of the effects of upland drainage:
Blacklaw Moss (Lanark, Scotland):
The 7 ha experimental site was instrumented for 5 years from 1959-1964 by the
Macaulay Institute and the Hill Farming Research Organisation (Robertson et al,
1968)118. The site is described as a raised basin mire used for rough grazing with
deep peat (average depth 5.8 m) soils that are almost permanently waterlogged. The
vegetation was dominated by heather (Calluna vulgaris) and cotton grass
(Eriophorum species), with a discontinuous carpet of mosses including Sphagnum
species. After a 3-year calibration period the land was drained by cutting open
ditches about 40cm wide and 36cm deep at 9 metre spacings.
There was little difference between the two periods in the characteristics of the
storms, but there was a large increase in the observed peaks (Robinson, 1990)56. In
part this was due to an increase in the percentage of rainfall that formed runoff, but
mainly due to an increase in the flashiness of the site. These findings are numerically
reported in terms of how the standard unit response of the site changed using the
parameters of the Flood Studies Report136 Unit Hydrograph rainfall-runoff model. The
time-to-peak decreased to just 0.33 of it pre-drainage value; the percentage runoff
increased from 46% to 58%, and the peak of the unit hydrograph increased by a
factor of 2.6.
The increased in responsiveness of the site was interpreted as the result of the large
increase in the channel network speeding up flows by shortening the slower flow
paths through the soil to the channels, and (bearing in mind the high water retention
of the peat, together with capillary rise above the water table, the soil remained wet
for long periods after drainage) there was very little compensating increase in the
available storage capacity of the soil (giving limited additional runoff attenuation).
Llanbrynmair (central Wales):
An instrumented catchment study investigating the effects of forestry on hydrology
and water quality (Leeks and Roberts, 1987)119 comprised both a 3 km2 and a nested
0.34 km2 peat moorland catchment that were planned to be afforested. The
catchment was progressively drained over a 4-year period until 70% of the area was
affected (the residual area was either too steep or not to be planted). Robinson
(1990)56 undertook similar analysis on the smaller of the two catchments as
undertaken for Blacklaw Moss. The comparison of unit hydrographs before and after
the drainage indicated similar hydrological effects to those at Blacklaw, namely open
drainage resulted in a much peakier storm flow response.
Drainage of the main catchment started on the higher ground most distant from the
outlet gauge, and then moved to the valley bottom near to the gauge. Drainage of the
higher land (affecting 45% of the catchment area) gave rise to a much peakier runoff
response recorded at the outlet gauge, but when the valley bottom was subsequently
drained (15% of catchment) there was no further increase in peaks, although the
response time of the catchment was shortened. This was interpreted as the speeding
up of flow from the areas near the gauge, building up the early part of the runoff
hydrograph; any increase in the peak response was offset due to this water leaving
the catchment before the arrival of flows from the more distant parts of the
catchment. Subsequent drainage of a further 10% of the area near the centre of the
catchment produces no further increase in peaks. In fact the maximum flows slightly
-50-

reduced and this was thought potentially connected to the earlier drains becoming
hydraulically less efficient.
The effect of location of drainage was also reported by Acreman (1985)120 for the
extensive pre-planting upland drainage that occurred in the Ettrick catchment in
southern Scotland. Initially most of the ploughing occurred near to the catchment
outlet and then later in the headwaters. The result was a small decrease in peak
flows when the lower areas were ploughed, but as the drainage extended into the
headwaters the peak flows increased. As for Llanbrynmair this increase reduced
within a few years as the drainage was speculated to have become less effective.
Coalburn (northern England)
Robinson et al (1998) report the hydrological findings of the Coalburn study from its
inception in 1967 to 1996.The catchment comprises a 1.5 km2 upland peat
catchment (altitude range 270 330 m AOD) that has a mean annual rainfall of about
1350mm. 75% of the catchment is covered by peat bog, and the remaining 25% by
peaty gley soils. The peat depth is generally within the range of 0.5 3 m, though the
maximum recorded depth is 10 m. Prior land use was for rough grazing (vegetation
comprised: Molinia grassland and bog species including Eriophorum, Sphagnum,
Juncus and Plantago). Some old relic grip drainage existed from the 1940s and
1950s, but these were considered to have become largely inoperative. Hydrological
data was collected for 5 years before the whole catchment was subject to the
ploughing of open drains in 1972. Spacing of drains was about 5 m apart and they
were aligned with the ground slope. The long continuous channels were 80 90 cm
deep and over 100cm wide. Water from these drains was either intercepted by
deeper drains or allowed to connect directly to the natural water course. This is an
example of the old style environmentally insensitive upland drainage and can be
considered to be a relatively aggressive form from that era.
As for Blacklaw Moss the characteristic response of the catchment was investigating
using the parameters of the Flood Studies Report Unit Hydrograph rainfall-runoff
model. The key parameters for the model were established before drainage and at
subsequent periods after. In the 5-year period after the drainage the time-to-peak
parameter reduced on average by 22%. The effect diminished over the following 20
years (Figure 13), though the apparent effectiveness of the drainage may be affected
by the establishment of the forest cover. It is therefore not possible to reliably
assume that the effectiveness of the drainage would have been the same had the
trees not been planted. The authors suggest the increase in catchment flashiness to
be a result of a greater density of drainage channels which speeds up the removal of
surface waters. The diminution of the effect over time is suggested to be a result of a
reduction in the hydraulic efficiency of the drains due to the furrows becoming
colonised by vegetation and later by litter from needle fall.

-51-

100%

-25%

80%

-20%
60%
-15%
40%
-10%
-5%

20%

0%

0%

1970

1975

1980

1985

1990

Percentage of catchment covered


by canopy

Percentage change in Tp and PR

-30%

1995

Decrease in Tp parameter (increase in flashiness)


Reductiion in percentage of rainfall forming quick runoff (PR)
When ploughing of drainage occurred
When conifers planted
Speculated conifer canopy cover

Figure 13: The changes to the speed of catchment response parameter (TP) and the
percentage of storm event rainfall forming quick runoff (PR) observed at the Coalburn
experimental peat catchment following intensive upland drainage. (Values calculated
from data presented in Robinson et al, 1998. Speculated canopy cover estimated by
Jacobs based on experience and the limited information provided by Robinson et al.)

The reduction in the proportion of rainfall forming fast flow is small over the first ten
years and is not statistically significant. Later estimates maybe suggest a reduction,
but again this may be related to the establishing tree cover.
Based only on the change in flashiness (TP) design flood flows (using the FSR Unit
Hydrograph method) were predicted to have been increased by 15 20% in the first
5 years after the ploughing, after which the effect reduced. Importantly the authors
also note:
This is not, however, to say that the increases indicated here would be the same for
extreme events: the unit hydrograph analysis was based on events that were
generally smaller than the mean annual peak. In extreme (and rare) events, system
threshold may be surpassed. For instance, if a storm exceeded the interception
capacity of the vegetation and the available depression storage of the many partly
infilled and blocked drains, then the extensive artificial channel network could
potentially produce much higher peak outflows than the original moorland That is the
direct consequence of the much greater velocities of open channel flow than
subsurface or overland flow.
Holden et al (2004)121 present an overview of studies that have examined the
hydrological response to artificial drainage in peat. The majority of the studies noted
an increase in annual flood peaks following drainage, although some showed a
decrease. They deduced that flood responses decrease in cases where the peat,
topography and ditch characteristics cause water tables to fall. This may affect both
the saturation excess overland flow and the occurrence of soil pipe/macropore flow.

-52-

(However they also note that a fall in watertable may lead to peat decomposition and
subsidence, so that the temporary increase in storage is lost and runoff becomes
more flashy again). In most peats, however, they indicate that the soils have low
hydraulic conductivity, so in most cases the effect on watertable is limited to a very
localised zone along the ditches. They therefore see that the primary effect of ditches
is to speed up the rate that water leaves the catchment, which increases flood peaks.
They also recognise that changes in the timing of runoff may alter downstream flood
peaks in different ways, depending on the phasing with flow contributions from other
sub-catchments.
Holden et al (2006)122 review the long-term hydrological impacts of artificial drainage
of peat at two north Pennine catchments drained in the 1950s compared to two
control catchments at Moor House. These include the catchments considered by
Conway and Millar (1960)14. Their findings were:

Both drained and undrained catchments have a flashy response to rainfall.


However the drained catchments were flashier.
The drained catchments were less sensitive to storm rainfall in 2002-2004
than shortly after the ditching in the 1950s.
The intact control catchments were dominated by saturation-excess overland
flow.
In the drained catchments, soil matrix throughflow formed a much greater
proportion of storm runoff (around 23% of the total) than intact catchments.
Macropore flow, soil pipe and the density of soil pipes was much greater in
the drained catchments. This structural change is considered to have
developed over many years.
Watertable is not controlled by drawdown around ditches but by the
relationship between ditch position and topography.

Several upland catchment drainage blocking studies are being undertaken:

River Ashop catchment in Derbyshire grip blocking trial for Severn Trent
Water, the National Trust and Natural England involving monitoring before
and after blockage of eroded gully channels (OBrien et al, 2004)

Whitendale in Bowland study for United Utilities studying the blocking of


agricultural grips and monitoring of individual drains (Armstrong at al,
2006)123.

Allensheads in Teesdale grip blocking study for Northumbrian Water Ltd.

Sapling Clough in the River Hodder headwaters in north west England - grip
blocking study as part of United Utilities Sustainable Catchment Management
Programme (SCaMP).

The focus of these studies is primarily the impact on colouration (of particular interest
to the water companies) but the studies also have the potential to provide valuable
insight into the impact upon the hydrological behaviour during flood events. Findings
from these monitoring studies have not become apparent to this review, but SEPA
are recommended to seek the findings in the future.
b) Modelling studies
Modelling studies have largely been focused on trying to predict how effective
peatland drain blocking will be, as opposed to the catchment studies that have
focused on the impact of installing drainage.
The SCIMAP (Sensitive Catchment Integrated Modelling and Analysis Platform)
study (Lane et al, 2003)28 investigated drain blocking using a model that linked a
-53-

hydrological model to a detailed digital elevation model. From this they concluded
that grips affect flood runoff more via the increased speed of water shed from the
catchment than by the extra soil moisture storage that results from the drainage. In
terms of grip blocking they concluded that a catchment scale model that represents
the spatial arrangement of drains and their connectivity to the drainage network is
needed to a) determine the catchment scale impact of drainage or drainage blocking
on downstream runoff, and b) to identify which channels would be most effective to
block.
Ballard et al (2010)124 used a simplified physics based model to simulate the flood
response of a 200m x 200m plot of upland peat. The study assumed that no channel
structural changes and no vegetation changes occurred after drain blocking and that
no bypass flow around drain dams occurred. Drain depth was fixed at 0.6m. Storm
events from 2006 were used to drive the model using a 15-minute time step for i) no
drain blocking, and ii) drain blocking over a range of: drain spacings, surface slopes,
drain angles, and hydraulic roughness values representing different vegetation
covers and channel surfaces. The simulations suggested that flood peak response
was sensitive to the characteristics of the storm and the antecedent conditions.
Sensitivity analysis suggested that one average drain blocking leads to the greatest
reduction in flooding for sites with larger drain spacing, steeper drain angle, steeper
slope, rougher plant cover, smoother drains and a thin acrotelm (surface layer).
However there was much variability in the outcome with increases in peak flows also
predicted as well as decreases depending on the event and parameter set used. The
authors conclude that their work suggests that drain blocking may not necessarily
always reduce flow peaks, with some cases showing negligible changes in runoff and
other cases actually indicating an increase in flow peaks. But that the results from
their preliminary study could be used to prioritise works for drain blocking. Drains that
are steeper and smoother are most likely to show the greatest reduction in flood
peaks following blocking.
Geris et al. (2010)125 describe preliminary work involving numerical modelling of grip
blocking at Sapling Clough (1.7km2), a headwater of the River Hodder catchment in
north west England. However the model is in the stages of early development so no
direction regarding the effectiveness of blocking was presented.
JBA (2007)126 report upon a modelling study that considered the potential benefits of
upland drain blocking upstream of Ripon in North Yorkshire. The catchment down to
Ripon has an area of 120km2 (comprising the rivers Laver, Skell and Kex Beck), an
altitude range of 30 410m, average annual rainfall of 900mm with the headwaters
receiving about 35% more rain than the easterly lowlands). The soils in the higher
rainfall, upland headwaters (approximately 30% of the overall catchment area) are
dominated by various types of deep and surface peats overlying mineral soils. The
land cover in these upland areas comprises moorland grass and heather. During the
1970s this area was subject to partial upland grip drainage (67km in length; equating
to about 22% of the overall catchment to Ripon). In the lower portions of the
catchment the soils are a mixture of free draining and relatively impermeable mineral
soils that are used for pasture, arable crops and some forest\woodland cover. In the
model the overall catchment was split into 10 sub-catchments. The hydrological
response from each was modelled using the Probability Distributed Moisture (PDM)
rainfall-runoff model, and linked to an ISIS hydraulic routing model to simulate the
passage of the river flows through the catchment down to Ripon.
The effect of drain blocking in the headwaters was simulated by increasing the
various timing parameters in the PDM hydrological model by 1 hour which has the
effect of reducing the speed of sub-catchment response. No evidence was given to
justify this size of parameter amendment. Its selection was indicated to be reliant on

-54-

the authors judgment. No change to the parameter controlling the percentage of


runoff was made. This is to say that only the timing element of the model was
changed and the total volume of runoff was kept the same. Simulations of observed
winter and summer events together with design events were undertaken for the
baseline and for a scenario in which all the upland drains were blocked.
For the simulation of the observed winter events the drain blocking was predicted on
average to reduce the Ripon peaks by 4.4% (standard deviation of +/- 1.6%;
minimum = 1% & maximum = 8%); and the arrival of the peak was on average
delayed by 19 minutes (+/- 9 minutes). Only a few summer events were considered
suggesting similar results.
For the design events the predictions at Ripon suggested a 5% reduction in a 1%
annual probability event (100-year flood) peak flow plus a 15 minute delay in arrival.
In the meantime the reduction in peak flow was only 1% with no delay in arrival for
the 10% annual probability event (10-year flood).
Common to all simulations was that the percentage change in peak flow decreased
the further downstream one considered.
The authors considered that appreciable uncertainty existed in the land management
specification (which has been interpreted to relate to both the areal extent and the
choice of parameter amendments). The authors also caution that the scale of the
predicted changes may not be appropriate for catchments with different
characteristics and that the value should be viewed as indicative.
Johnson (2007)127 provides predicted impacts to floods due to the blocking of artificial
drains at Glendey. The catchment is located within the headwaters of the River
Devon in the Ochills, Scotland. The catchment has an area of 2km2 within which a
demonstration site of 0.0175km2 was considered for an array of Natural Flood
Management techniques. A hydraulic model was used to estimate how effective the
blocking of artificial drains in the wetland portion of the site would be. This suggested
that a 4% to 6% reduction may result, with the lower value being estimated for the
biggest event that was believed to represent a 0.04 annual probability event (25-year
event). This study however differs to the preceding ones in that the drained area had
an artificial watercourse running through it that was re-aligned into a meandering
channel. The results are therefore interpreted as reflecting both the effects of drain
blocking and the effect of re-arranging the watercourse flow path and propensity to
cause out-of-bank inundation.
The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of data
sets and wetlands are subdivided into classes based on wetland type. This data set
represents a useful resource for a national overview of wetlands in addition to more site
specific information.

(G)

Constructed farm wetlands

Constructed farm wetlands can be used to treat lightly contaminated surface water runoff
from farms and they are a means of managing diffuse pollution (Scottish Government,
2009)128. The Constructed Farm Wetland Design Manual for Scotland and Northern
Ireland (Carty et al., 2008)129 provides a best practice manual for the construction of farm
wetlands; however the manual focuses on designing wetlands to manage diffuse pollution
and only mentions in passing that the constructed ponds may also provide flood
attenuation benefits. There is no literature specific to the flood attenuation benefits of
constructed farm wetlands which have been designed for the treatment of diffuse
pollution. It is speculated that the wetlands perform in a similar manner to wetlands
which are used in SUDS. The SUDS Manual (CIRIA, 2007)130 identifies that it is
-55-

necessary to provide additional storage volume over the permanent retained volume and
an outlet control which will ensure that this attenuation volume is utilised effectively during
flood events. The design of flood attenuation features is a complex undertaking and it is
considered unlikely that farm wetlands will provide any significant attenuation due to both
the requirement to provide additional storage over the permanent treatment volume
advocated within current best practice and the need to provide a carefully tuned flow
control at the outlet to ensure the volume is utilised at the flood peak, not before or after
it.
The use of an array of offline and online flood storage areas, which are referred to as
Runoff Attenuation Features (RAFs), is presented for the town of Belford by Wilkinson et
al (2010)131. The paper reports that a recent flood alleviation scheme feasibility study
identified that it was not viable to protect the 31 properties in the town which are
estimated to be at risk of flooding during the 1 in 200yr event. However it is noted that
the Environment Agency has recently undertaken minor works to improve the situation for
a handful of worst affected properties affected in the town by increasing the channel
capacity and constructing a short flood wall (Environment Agency, 2011)132. Wilkinson et
al (2010) presents the use of four RAFs providing a total of 2,800cu.m of storage within
the 6sqkm catchment, these are online and offline flood storage areas which are intended
to primarily attenuate floods rather than address diffuse pollution problems. By a
simplified comparison of the hydrograph volume and the available storage it is proposed
that the current flood storage reduces the 1 in 5yr flood peak by approximately 8%,
however this assessment takes no account of any inherent inefficiencies in utilising the
storage (Ackers and Barlet, 2010)133. The paper presents the protection of the town from
flooding during a 1 in 48yr rainfall event in 2008 as evidence that the measures are
effective, however the question of whether the town would have experienced flooding
prior to the installation of the RAFs for this 36 hour 96mm event is not answered
conclusively. It is noted, that in the paper the duration of flooding is approximately two
hours, therefore the 1 in 48yr, 36 hour rainfall event is not representative of the flow
return period. Elsewhere the paper estimates the attenuated 2008 peak flow as 2.1+/0.4cumec, putting the event as substantially less than the 4.56cumec (1 in 5yr) threshold
of flooding once the optimistic 0.4cumec reduction of flood is accounted for. Finally it is
reported that during the 2008 event all the flood storage cells were full around the time of
the flood peak, implying that during a larger event the storage would be utilised ahead of
the flood peak and thus not attenuating the hydrograph peak.
It is accepted that small scale flood storage areas, such as those in use in Belford, can be
an effective flood risk management method, however there remain a number of issues
which need to be addressed:

For the RAFs to be effective they all need to be designed to throttle the flow
effectively so that the flood storage they offer is efficiently used to reduce the peak
of larger floods and not the leading part of the hydrograph (Hall et al., 1993)134.

The large number of sites, often located on private land in remote parts of the
catchment, introduces maintenance challenges. It is unclear if RAFs can be relied
upon to function in the desired way without frequent routine maintenance,
especially considering the relatively short design life of the structures, the
blockage risk to small openings used to throttle flows and the multi-year periods
between the sites being required to perform their function.

The large number of sites leads to a large number of in-channel structures, each
forming a potential barrier to wildlife movement and a morphological pressure on
the watercourse.

As with all flood storage, care must be taken to ensure that flood peaks on flashier
sub-catchments are not delayed so that they coincide with other sub-catchments

-56-

although the same effect can be applied to desynchronise peaks (Hall et al.,
1993)134.

(H)

Wetland creation (non-floodplain)

This section relates to the creation or restoration of non-floodplain wetlands. Detailed


discussion on upland drainage is provided in item (F) and on floodplain reconnection in
item (J).
A comprehensive review of 439 published statements by Bullock and Acreman (2003)135
identifies that there are many examples where wetlands reduce floods, recharge
groundwater or augment flows. However the same paper also states that there are also
many examples where wetlands increase floods, act as barriers to recharge and reduce
low flows. The paper surmises the variation that wetlands can have within a hydrological
system concludesing that:

Wetlands can alter the hydrological response of catchments

Wetlands in catchment headwaters have an uncertain effect approximately half


(30 of 66) decrease flood peaks, while the other half (27 of 66) have been found
to increase flood peaks

In approximately half of cases (11 of 20) headwater catchments increase the


immediate response of rivers to rainfall and generate higher volume floods.

Two thirds of studies have concluded that wetlands increase average annual
evaporation or reduce average annual river flow. It should be noted that a
minority conclude the opposite.

Two thirds of studies conclude that wetlands reduce dry weather flows.

Many bogs do not act to delay flow into streams (Rydin and Jeglum, 2006)113. The Winter
Rainfall Acceptance Potential (WRAP) was developed as part of the Flood Studies
Report (NERC, 1975)136 uses five WRAP classes to inform the spatial percentage runoff.
Soil and land cover types normally indicative of wetlands are typically in WRAP class 5,
this class has the highest percentage runoff within the Flood Studies Reports RainfallRunoff Method (Boorman et al., 1995)137. The underlying messages are that each
wetland is unique and that many wetlands are saturated at the start of flood events
therefore any incident rainfall is rapidly converted to runoff.
The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of data
sets and wetlands are subdivided into classes based on wetland type. This data set
represents a useful resource for a national overview of wetlands in addition to more site
specific information.

(I)

Gully (cleugh) planting

There is limited published literature on the specific benefits of gully planting within a
Scottish flood risk management context. There are a few papers on the subject of the
formation of gullies within loess soils (alluvial silts) in other parts of the world. These
gullies are notably smaller than the gullies or cleughs that are typically found in
Scotland, with depths, widths and lengths on the order of a few meters rather than tens of
meters often encounter in places such as the Scottish Borders.
It has been reported in a number of locations across Scotland that high rates of
deposition may be a major contributor to flooding due to a loss of channel capacity as
observed at Selkirk (Long Philip Burn), Hawick (River Teviot) and Marykirk (Balmaleedy
Burn).

-57-

Poesen et al. (2003)138 reports that gully erosion represents an important sediment
source in a range of environments and gullies are effective links for transferring runoff
and sediment from the uplands to valley bottoms. Poesen et al identified that most
research effort in recent decades has concentrated on sheet (interrill) and rill erosion
processes operating at the plot scale with few studies investigating gully erosion.
A study of the Upper Wharfedale by Lane (2006)139 identified that the forested
headwaters of the catchment produced comparatively little coarse gravels while tributary
gills (terminology used for gully in some areas of England) were a major source of the
problematic gravels which were causing the channel aggradation and subsequent
flooding further downstream. It was identified that the gills were cutting down into old
glacial deposits. Lane comments that in this case it was most appropriate to address the
sediments at source rather than seeking to convey the sediment through the problematic
reach (which was in an essentially natural state). Through modelling he proposed that
very localised planting of gully slopes could reduce the delivery of coarse sediments by
85%.
A historic mapping/aerial photography/digital terrain model study of a 25sqkm catchment
in Spain identified a sediment production rate in gullies of 1322+/-142 ton/Ha/yr
140
(Martn
ez-Casasnovas, 2003) . While it would be unwise to adopt anything more than a
sense of possible magnitude from this data it does give an indication of the scale of the
sediment source gullies represent. A catchment study in Yangjiagou identified that
manmade forests have reduced sediment yields by as much as 92.5% (Zhuo, 1992)141.
Raven et al (2010)142 provides a detailed overview of the issues associated with the
management of gravels in upland rivers. The paper concludes that there is a need for a
shift in how rivers are managed based around catchment scale approaches. The
management of sediment sources on the hillslopes may provide long term mitigation, but
it will be necessary to manage sediments which have already entered river systems. In
many cases the use of bank protection prevents the process of safely deposing coarse
sediments at the inside of bends in exchange for the mobilisation of fine sediments. The
uncertainties imply that the best long term management strategy may be for people to
learn live with bank erosion and occasional flooding.
In preparing this section it is noted that:

Gully planting could be considered a variety of riparian planting

The steep topography of a gully is likely to make any intensive land use difficult if
not impossible

Erosion, gully formation and deposition are natural processes

Based on specific streampower, reducing flow rate, increased the wetted


perimeter or increased roughness could reduce the erosive power of a
watercourse flowing through a gully or potential gully.

The steep sides of a gully would infer very rapid hydraulic connectivity for any
runoff that is generated within a gully.

There is an absence of transferable data relating specifically to the hydrological benefits


of gully planting. For this reason at this stage it is not thought possible to treat gully
planting any differently to ordinary riparian planting when assessing hydrological benefits.
Where sediment deposition has been identified to be a contributor to flooding then
targeting the forestation of actively eroding gullies would seem a logical decision although
recognition of the potentially decade scale travel time for sediments already in the system
should be made.

-58-

(J)

Floodplain reconnection

This section refers to the literature relating to the physical re-introduction of flow to the
floodplain during flood events. Refer to item (L) for literature relating to using vegetation
to slow out of bank flow.
A modelling study of the Cherwell catchment was undertaken to assess the hydrological
impact of restoring the connection of floodplains within the catchment (Acreman et al.,
2002)143. The study employed a 1D hydraulic model and a semi-distributed hydrological
model to assess the impact of restoring entrenched river reaches to a natural state. The
natural conditions were based on the findings of a geomorphological study. The study
found that the complexity of floodplain hydraulics presented as considerable challenge to
the modelling work. The results showed that where embankments separated the river
from the floodplain the peak flow rate could be increased by 50 to 150% (represented by
adding embankments to reaches presently without embankments) however it reports that
restoring the channel through the floodplain to its pre-engineered dimensions could
reduce flood flows by 10 to 15%. Within the body of the report it is presented that some
floodplain reconnections have a negligible impact, hence it may be more representative
to report a 0% to 15% reduction in flows where the watercourse is entrenched and for
cases where floodplains are completely disconnected the benefits may be greater. The
report does discuss the variation on time of peak for some of the test reaches, in one
case it is reported that the peak could be brought forward by 40 hours should
embankments be introduced and delayed by 17 hours if the channel was restored to a
natural condition. Unfortunately the study does not provide a time to peak or provide an
estimation of the change in flood wave velocity. It should be noted that the 1D modelling
approach did not allow for the conveyance of water on the floodplain (extended crosssections within ISIS using a Mannings value of 0.0), while this approach is a commonly
adopted method for modelling floodplains it is likely to result in the overestimation of flood
attenuation. The increasing usage of 2D hydraulic modelling techniques which would
avoid the need for extended 1D river sections render this study somewhat outdated.
While this study adds to the case that floodplain reconnection can be used to attenuate
flood flows the quality of this study is insufficient to allow it to be applied elsewhere.
The benefit of floodplain restoration is discussed qualitatively by Blackwell et al (2006)144
and includes the presentation of a number of projects where floodplain reconnections
have been implemented across Europe. Unfortunately the report does not provide a
summary of the hydrological benefits. It does report that reduction in flood risk is
dependent on local conditions.
A modelling study on the River Suir in south-east Ireland (Ahilian, 2009)145 reported that
storm duration is a significant factor in the attenuation of the flood wave. To this end it
was identified that the ratio of the storage volume available on the floodplain to the
hydrograph volume is significant. The study also found that the floodplain width and
floodplain slope were also significant.
A comprehensive literature review of the hydrological role of wetlands was undertaken by
Bullock and Acreman (2003)146. The review collated 439 published statements on the
role of wetlands within the hydrological cycle. Part of the review focused on the function
of floodplain wetlands in reducing flood peaks. The review found that most (23 of 28)
show that floodplain wetlands reduce or delay floods. The study concluded that there is
no simple relationship between wetland types and the hydrological function they perform
and that apparently similar wetlands can be driven by different hydrological processes.
For this reason the papers authors recommend the development of an improved wetland
classification system which makes better account of function.
Jacobs (2011)3 proposes the application of a tool which functions by scaling the
generated floodplain area by model results generated for a donor catchment. The
suggested methodology includes an allowance for slope (1:250 to 1:1000), the location
-59-

within the catchment (upper, middle and lower) and an efficiency factor. The method
implies that a reduction in flood peak of approximately 0.0% to 4.6% per 10,000cu.m of
created storage is achievable. The method does not take account of the ratio of
floodplain volume to hydrograph volume which has been found by others to be of key
significance Ahilan (2009)145 and Potter (2006)147. It is accepted within the Jacobs (2011)
study that this is an attempt at simplifying a complex process which varies markedly
between locations. As this proposed methodology is heavily based on the findings of one
donor catchment and the judgement of its author it is difficult to recommend its
application.
Within the scientific literature there is evidence that floodplain reconnection can in some
circumstances increase the downstream flood risk (Bullock and Acreman, 2003)146. This
is because improving the connection to a floodplain could increase the conveyance
capacity offsetting the increase in storage in a similar fashion to the effect of removing
constrictive bridges and culverts which is well documented. Particular care and attention
should be made when removing flood embankments, especially where they form a
constriction at the downstream end of an existing functioning floodplain.
It should be noted that some means of reconnecting a floodplain (e.g. reducing channel
entrenchment) can result in an increase in water level. The backwater effect should be
investigated as discussed in Appendix C.
As presented by Jacobs (2011)3, it is difficult to draw general conclusions on the
magnitude of the benefits as these are invariably case specific and depend on the nature
of the enhancement (compared to the existing condition), the specific flood event and the
proximity to the flood receptor. Going a step further, it is probably impossible to make a
generalised statement about the effect which floodplain reconnection can have and that
the only means of making a meaningful assessment is to use a flood routing element
within a hydrological model. As presented in Ahilan (2009) the ratio of floodplain storage
generated to hydrograph volume is a useful indicator of the potential impact. In the
absence of sufficient data to develop a regression based analysis of the link between the
floodplain volume to hydrograph volume ratio in addition to other key variables the only
viable means of assessing the flow attenuation capabilities of floodplain reconnection is
to construct a hydraulic model.

(K)

Using vegetation to slow channel flows and stabilise banks (riparian


planting)

The term riparian is relatively loose and there are a number of different definitions. The
UK Forestry Standard Water Guidelines (Forestry Commission, 2011)76 defines the
riparian zone as:
the area of land adjoining a river channel (including the river bank but not the wider
floodplain) capable of directly influencing the condition of the aquatic ecosystem (e.g. by
shading and leaf litter input).
Elsewhere, the riparian zone is defined by the Environment Agency (1998)148 as:
the land either side of the river channel, extending to the limits of associated floodplain
wetland, or 50m, whichever is greater.
In reviewing the literature it is necessary to take due consideration of the varying
definition of riparian.
The main functions of the riparian buffer are considered by Broadmeadow and Nisbet
(2004)149 to be:

Sediment removal and erosion protection;

Protection of water quality;

-60-

Moderation of shade and water temperature;

Maintenance of habitat structural diversity and integrity; and

Improvement of landscape quality

From a flood risk management perspective riparian woodland can alter the following
processes:

Increase river bank roughness, thus slowing in channel flows;

Protect river banks from erosion;

Slow surface runoff as it enters the watercourse;

Provide a supply of large woody debris therefore facilitating the natural formation
of large woody debris dams leading to the slow of flows; and

Provide a supply of organic material which can cause the blockage of downstream
structures. The elevated risk of blockage to structures is discussed in Appendix
C.

Nisbet et al. (2011)65 presents a review of the potential benefits of riparian woodland
referencing earlier work which demonstrated that woody debris dams can significantly
delay flood peak travel time and lengthen catchment response times. It is suggested that
to be effective at a larger catchment scale would require extended reaches of riparian
woodland and the subsequent formation of large woody debris dams along tributary
streams. It is also noted that questions remain over the ability of riparian woodland (and
floodplain woodland) to capture or release large woody debris. The risk of woody debris
blocking pinch points at downstream flood receptors and increasing flood risk at such
locations also needs to be considered.
The Robinwood study (Forest Research, 2008)150 into the effect of large woody debris in
watercourses indicated that large woody debris can slow in-channel flows and thus help
to attenuate floods. Modelling (1D hydraulic) undertaken as part of the project indentified
that during a 1% AEP event each large woody debris dam would delay the flood peak on
the River Y Fenni by an average of 2 to 3 minutes. It was found that the effect of large
woody debris on flow attenuation diminished with event magnitude as the features would
become drowned out during high flows. The report indicates that larger benefits can be
achieved when large woody debris dams improve the floodplain connection by raising inchannel levels and thus facilitating flood water to enter the floodplain earlier in the
hydrograph. It is suggested that large woody debris dams in headwater streams are
expected to be most effective primarily because of the increased chance of an effective
blockage forming in small constrictive channels. For this reason large woody debris
interventions are best suited for middle and upper reaches of a catchment.
Odoni and Lane (2010)198 undertook a modelling study of the Pickering Beck catchment
using OVERFLOW to assess the effect of catchment riparian intervention measures
(CRIMS) and woody debris dams. The CRIMS are described as 30m wide riparian
woodland corridors on either side of the channel. The woody debris dams were assumed
to form along the channel at a natural spacing of 7 to 10 times the channel width, based
on work by Linstead and Gurnell (1998)151. The woody debris dams and riparian
woodland were represented within the distributed hydrological model by localised
increases to the Mannings roughness values. The study represented large woody debris
dams via a Mannings value of 0.18 and riparian woodland via a value of 0.2. The
authors acknowledge that the selected Mannings values are at the upper limit of what
can be supported within the literature. As discussed in item (L) there is extensive
published guidance relating to the selection of Mannings numbers. The modelling
identified that the installation of 100 debris dams could reduce a flood event from
29.5cumecs to 27.3cumecs (7.5% reduction). It was identified that the magnitude of flow
-61-

reductions increased with event magnitude, thus suggesting that the effectiveness of
riparian woodland and large woody debris dams improves for the larger more extreme
events which cause flooding. This increase in effect goes against the findings of the
Robinwood study (Forest Research, 2008)150 and Gregory et al (1985)152. As with other
flood measures which desynchronise flows there is no hard and fast rule of thumb to
dictate where slowing flood flows will reduce downstream flood risk, indeed, many
potential locations for placing large woody debris were found to increase rather than
decrease downstream flood risk. Like the Robinwood study it was identified that the best
locations for riparian interventions tend to be in the upper catchment.
Gregory et al. (1985)152 studied vegetation dams in a small watercourse in the New
Forest over a 12 month period. The paper reports that 36% of the debris dams changed
location or were destroyed during this period, however the overall number remained
relatively constant, with a small increase being attributed to stormy weather. It was
estimated that the flood wave travel time at bank full condition was delayed by 10 minutes
over the 4028m study reach and that the delay in propagation was inversely proportional
to flow. It was reported that it was necessary to reduce the Mannings roughness value of
the reaches containing the debris dams to replicate the reduction in flood wave travel
times with a roughness of 1.02 being representing of lows flows and 0.31 being
representative of high (bank full) flows. This paper introduces terminology which
subsequent British literature has adopted:

Active Debris Dams dams that introduce a step in water elevation during low
flows

Complete Debris Dams dams which span the entire channel but do not generate
a step in water level during low flows

Partial Debris Dams dams which do not span the full channel width

Modelling of the River Laver by Nisbet and Thomas (2008)155 to assess the effect of
floodplain woodland on reducing downstream flood risk also included riparian woodland.
Unfortunately it is not possible to distinguish the effect of riparian woodland as all
presented model results include both riparian and floodplain woodland. Like the
Pickering Beck study the in-channel Mannings roughness values were increased (by
comparable amounts) in areas of riparian woodland to represent the presence of bank
vegetation and large woody debris within the channel. The findings of the River Laver
study were broadly inline with the other riparian woodland studies.
Modelling of proposed large woody debris dams in the Great Triley Wood was
undertaken by Forest Research to assess the impact of the placed structures on flood
propagation (Thomas et al., 2012)153. This study represented the large woody debris
dams within a calibrated InfoWorks RS model by using the Channel Blockage function
(application of 70% channel blockage) rather than increasing roughness. The approach
of applying a channel blockage allows the diminishing effect of large woody debris dams
with flow to be represented, with the study citing Gregory et al (1985)152 as evidence for
this reducing effect. Large woody dams were placed at an interval of 7 to 10 times the
channel width, consistent with Linstead and Gurnell (1999)151. The study identified the
large woody debris dam on average increased the reach time of travel by 2-3 minutes per
dam but that there was only a very minor change to flow. The suggestion that the large
woody dams facilitated the reconnection of the floodplain was seen as important. The
reconnection was partly influenced by the deposition of sediments upstream of the debris
dams which raised the bed elevation and proved a desirable location for natural
vegetation to colonise thus further increasing channel roughness.
The general consensus is that the creation of riparian woodland serves to slow in-channel
flow through the presence of bank vegetation and in-channel vegetation and that the
decrease in flow velocity can be represented via an increase in the Mannings roughness

-62-

value but recent work has been undertaken by partial channel blockages. It is agreed
that riparian woodland has the largest role to play in the upper and middle catchment
where watercourses are narrower and more prone to effective blockage by large woody
debris. While reductions of up to 10% in flood peaks are reported in some instances
these reductions appear to be subject to local conditions and the modelling technique
used. It is felt that the most appropriate means of accounting for riparian woodland is to
specifically apply a localised blockage as undertaken by Thomas et al., (2012)153 within a
hydrological model that includes a physically based flow routing engine. The alternative
is to use elevated Mannings roughnesses, however this approach will suggest that the
effect of large woody debris increases with flow and hence may require the Mannings
roughness to be varied with flow.

(L)

Using vegetation to slow out of bank flows (floodplain planting)

A detailed modelling based assessment of the hydrological impact of floodplain


roughening via tree planting was undertaken by Thomas and Nisbet (2006)154. Using the
River Cary sub-catchment of the River Parret the study assessed the impact of applying
whole floodplain (2.2km reach, 133Ha) and a 50Ha (~700m reach by ~700m width)
forestation via the use of a 1D and 2D models. By extension a 2D hydraulic model is an
accurate (but computationally expensive) means of estimating flood routing. The
modelling used a donor 1 in 100yr flood hydrograph to estimate the impact foresting
floodplain which was open pasture in the baseline case. The paper concludes:

The complete forestation of the floodplain increased the floodplain storage volume
by 71% and delayed the flood peak progression by 140 minutes (baseline case
~3hrs).

The use of a 50Ha woodland block in the middle of the test reach increased
floodplain storage volume by 15% and delayed the travel time by 30 minutes
(baseline ~3hrs).

Peak water depths within the forested areas were found to be 50 270mm
deeper. This was based on increasing the Mannings roughness values for the
out of channel areas from 0.04 to 0.15.

The backwater effect of the 50Ha forest block (raised water levels by 120 to
180mm) was identified to extend nearly 400m upstream.

There is considerable scope for using floodplain woodland as an aid to flood


control. More extensive woodland than tested in the study could exert a much
greater downstream impact.

A detailed analysis of the flood hydrograph would identify where the restoration of
floodplain woodland would have the greatest benefits in terms of desynchronising
sub-catchments. Desynchronising, however, could extend the flood hydrograph,
with possible implications for longer duration or consecutive events.

Although it is very unlikely that floodplain woodland on its own would be able to
provide complete protection, it could make a valuable contribution along side
existing defences. Similarly, it could help manage smaller-scale flooding
problems where the high cost of constructing hard defences cannot be justified.

The paper does not directly report the change in velocity of the flood wave, however
using the graphs provided it can be estimated to be reduced from 0.73km/hr for the
baseline to 0.42km/hr and 0.63km/hr for the complete floodplain and 50Ha block
respectively. Unfortunately the paper doesnt specify the gradient of the watercourse,
however using the published hydraulic long section the gradient can be estimated as
0.00045m/m.

-63-

The paper does not report the reduction in flood flows as a result of the introduction of
floodplain forestation, however it is questioned quite how a percentage reduction for a
single hydrograph could be used since the increase in travel time for the reach would
likely increase the time to peak of the catchment. The longer storm duration required to
generate the critical storm would result in a marginally higher volume flood hydrograph,
thus any reported percentage reduction for a single storm hydrograph may be misleading.
Nisbet and Thomas (2008)155 undertook a subsequent study of the River Laver. The
modelling based study was undertaken to inform the placement of 40Ha of proposed
floodplain woodland. The study used 1D hydraulic model (InfoWorks RS) and a donor 1
in 100yr hydrograph. The modelling study identified that:

Floodplain planting increased flood depths by approximately 440mm to 800mm


with the average over the four sites considered being 610mm.

At each site the planting of woodland raised water levels over a backwater
ranging from 120m to 330m long. It was identified that the extent of the
backwater was a function of the river gradient and the increase in flood depth
caused by the planting.

There was a negligible impact on the peak flow rate at the downstream end of the
modelled reach with a maximum reduction of only 0.3% being identified.

The lag for each of the four sites considered ranged between 15 minutes and 20
minutes over the baseline case. The lag for the scenario where all four sites was
considered was estimated to be approximately 55 minutes.

The lag per hectare of floodplain planting was estimated to range between
1min/Ha to 3.5min/Ha, with the greatest lag per unit area being generated by the
smallest interventions. This reduction in effect suggests a law of diminishing
returns.

It was proposed that flooding in the town of Ripon could be reduced by the
strategic planting of the River Laver which joins the similarly sized River Skell just
upstream of the town. By delaying the peak on the River Laver and thus
desynchronising flows within the River Skell it was hypothesised that the flood
peak in Ripon could be reduced by 1 to 2% with the possibility of much greater
benefits if the extent of the planting was increased.

Roughness values of 0.03 and 0.05 were assigned to the channel and floodplain,
respectively, to represent the bed roughness associated with the nature of the existing
river channel and the baseline grassland or arable land cover. The establishment of a
cover of native floodplain woodland was represented by increasing the channel
roughness to a value of 0.10 and the floodplain roughness to 0.3. The report authors
accept that the adopted value for floodplain woodland is considered to be at the upper
limit of possible values for floodplain woodland and to attain such a high roughness the
woodland would need to be particularly dense with plenty of undergrowth, low branches
and fallen woody debris. The high channel roughness was justified as being a
consequence of an increased supply of large woody debris leading to an increase in large
woody dam formation.
Based on the provided river long section within the report body, the hydraulic gradient of
the watercourse in the vicinity of the floodplain roughening was identified to be typically
0.006m/m.
Initially 40Ha of woodland was considered feasible and a subsequent baseline monitoring
programme was put in place to collect valuable real data. Unfortunately landowner
interests resulted in the extent of the woodland planting being limited to an 8Ha plot at a
location prescribed by the landowner. This resulted in the baseline monitoring
programme being dropped.
-64-

Sensitivity testing of partial afforestation of the floodplain sites suggested that the most
effective placement for delaying flood propagation was within the lowest lying parts of the
floodplain, suggesting that corridors of planting along the sides of the river may exert
greater than average effect i.e. riparian woodland. However in preparing this review it is
considered possible that this increase may be a side effect of the 1-D modelling approach
and it may be more appropriate to conclude that the effectiveness of increasing
roughness is dependent on flow velocity (which will be highest in deepest areas of a 1-D
river cross-section). The areas of high flow velocity will be typically found in the lowest
areas of the floodplain so there is no issue with the overriding conclusion.
Floodplain woodland is presented within Woodland for Water: Woodland measures for
meeting Water Framework Objectives (Nisbet et al., 2011)65 via reference to the Cary and
Laver studies (also completed by Forest Research). The report proposes that floodplain
woodland might have a decreasing effect with more extreme floods.
Staffordshire Wildlife Trust has led a project in the River Sow and Penk (Jones, 2010)156.
A review of the project report indicates that the study focused on ecological service and
that while the implemented measures will have hydrological benefits no modelling was
undertaken to assist in the design of the works. It is understood that minimal floodplain
woodland has been introduced however a number of study sites involve the creation of
wet meadows with higher roughness introduced by ground scrapes and longer grasses.
A hydrological monitoring program is in place however it is apparent that this focuses on
the functioning of the ecological features and together with an absence of baseline data
may not be appropriate for assessing the hydrological impact of the measures during
extreme events.
Verbal communication with Tom Ball of University of Dundee has identified that as part of
the Eddleston Water project it is understood that floodplain woodland is proposed on the
Longcote Burn. The primary function of this floodplain woodland is to attenuate flood
flows. It is understood that a 1D modelling study will be used to inform the placement of
woodland and to assess its effectiveness. There is an extensive hydrological monitoring
programme in place which will provide much needed real data.
A study has been undertaken at Trier University in Germany to investigate the function of
forest management on the flood retention characteristics of floodplain forests (Sartor,
unknown year)157. The unpublished study presents the 2001 reforestation of 2.2% (23Ha)
of the River Blies in Germany floodplain, a catchment of approximately 1800sqkm. Using
hydraulic modelling it was estimated that extreme events (1 in 75 year) would be reduced
by 1%. The paper reports that another investigation in 2005 confirms the authors own
findings, unfortunately this supporting paper is in German so has not been accessed
(Aatz & Musong, 2005)158. It is reported that the complete forestation of the floodplain
was assessed to have the potential to reduce the peak flow rate by up to 15%.
Forest Research undertook a review of Short Rotation Forestry (SRF), incorporating
Short Rotation Coppice (SRC)(Nisbet et al., 2011)72. This review proposes that the
greater planting density and faster growth of SRF could be expect to provide a rapid
means of roughening floodplains in addition to achieving roughnesses greater than
convention woodland. As a further benefit the report highlights that the lack of large
woody debris generated by SRF crops [not mentioned: the ability to entrap floating debris
via the close plant spacing] would reduce downstream blockage risks (discussed in
Appendix C). Importantly SRF may be more attractive to landowners than conventional
forestry. Care would be needed to avoid SRF in water stressed catchments as discussed
in item (E).
There are a number of sources which can provide guidance on the selecting of Mannings
roughness values and this area can be considered well documented. It is beyond the
scope of this literature review to provide guidance on how to select appropriate Mannings
values for channel and floodplain flow, however further reading sources include:
-65-

Chow V.T., Open-channel hydraulics159 - A series of published tables for a range


of channel forms and vegetative cover. Despite being over 50 years old this
source remains a key source for many hydraulic modellers today.

Barnes H., Roughness Characteristics of Natural Channels160, A series of


photographed river sections with estimates of Mannings roughness values which
are based on observed flow and level data. This is a key source for American
hydraulic modellers.

Conveyance Estimation System161 - The CES is the culmination of a detailed


research and development programme (McGahey and Samuels, 2003)162 which
was lead by the Environment Agency and part funded by DEFRA, the Scottish
Executive and DARD NI. It provides an easy to use interface for practicing
hydraulic modellers to select scientifically appropriate Mannings roughness
values for most river and floodplain systems found in the United Kingdom. The
CES is available via current free versions of ISIS (Halcrow, 2011)163.

Acrement and Schneider, Guide for Selecting Mannings Roughness Coefficients


for Natural Channels and Flood Plains164 - This American source provides a
selection of methods to permit the estimation of Mannings numbers for a range of
channel and floodplain types.

Lane S. N., Roughness time for a re-evaluation165 - this published paper


provides a valuable commentary on how best to optimise roughness values for
open channel hydraulics.

Whatmore S. and Landstrom C., Mannings n: putting roughness to work, in How


well do facts travel?166 provides a commentary on the poor adoption of current
scientific knowledge relating to Mannings roughness.

In a consultation response to the Scottish Government the Forestry Commission (Nisbet


& Thomas, 2008)167 highlighted that opportunities for creating a continuous area of
woodland across the floodplain is usually greatest in the middle and upper reaches of
river systems, where housing and other forms of habitation pose less of a constraint. The
narrow nature of headwater floodplains and high water velocities would require extended
lengths of floodplain/riparian woodland to achieve a significant reduction in peak flows.
The Understanding Environmental Knowledge Controversies: The Case for Flood Risk
Management research programme project report output has at its very core the fact that
increasing floodplain storage decreases downstream flooding by the simple requirement
of a hydrological system to conserve mass (Whatmore, 2010)168. The research
programme identified that to assess changes to floodplains it is necessary to use an
appropriate model, hence the development of the OVERFLOW model for assessment of
the hydrological impact to the River Uck as part of the protection of Uckfield using
CRIMS.
In summary while there appears to be a weight of academic agreement that that
floodplain roughening is a beneficial flood risk management measure, there are relatively
few quality peer reviewed published papers on the subject. There is a general consensus
that the assessment of floodplain roughening requires hydraulic modelling or high quality
flow routing within the hydrological model. It should be noted that all literature identifies
that local (at site and upstream) flood levels will be increased as a result of introducing
vegetation to the floodplain. A review of the extent of the backwater generated by this
increase is presented in Appendix C.

(M)

Washland creation

Washlands are areas of managed floodplain that are allowed to flood, or are deliberately
flooded, this can be for a range of reasons including flood attenuation or habitat creation
-66-

(Hardiman et al., 2011)169. Typically, manual interventions are used to control how water
enters or exits the washland storage area so that the available storage is reserved for the
flood peak or flood waters are artificially held back to be discharged after the flood peak
has passed. It has been demonstrated that washlands generally provide a higher level of
flood attenuation than more natural floodplain systems (Shultz and Leitch, 2003)170. As
washlands are artificially altered floodplains much of the literature relating to floodplain
connections discussed in item (N) are applicable to washlands however given the
increased variability due to manual intervention the need for specific modelling of the
system is likely to be even greater than floodplain systems.
A review of washlands is provided within Morris et al. (2004)171. The English Nature
commissioned report defines washlands as an area of the floodplain that is allowed to
flood or is deliberately flooded by a river or stream for flood management purposes, with
potential to form a wetland habitat. The report identified that most washlands have
either been used mainly for flood storage or for wetland habitat and few attempts have
been made to integrate the two objectives. There are some conflicts between ecological
and flood risk management considerations including the need to keep washlands wet in
the interflood period to sustain ecology, consequently reducing the volume of storage
which is available to attenuate floods. Washlands vary between the duration of flooding,
the inlet and outlet mechanisms, and the habitat types found within them. Some
washlands are configured so as to reserve the washland flood storage for the flood peak
thus ensuring the maximum benefit in attenuation can be drawn from the available
storage volume. The outlet at some washlands can be manually controlled, thus there is
the potential to delay the release of flood water until the flow in the receiving watercourse
has receded.

(N)

Artificial large woody debris dams and boulder placements

There is limited literature relating to the use of placed large woody debris dams and
boulder placements. Odoni and Lane (2010)198 present the use of woody debris dams in
the Pickering Beck catchment to reduce downstream flood risk by slowing in channel
flows and thus creating additional flood storage volume by raising in channel water levels
and increasing floodplain connectivity. It should be noted that this approach requires a
large number of discrete interventions which has the potential to be expensive. There are
no hard and fast rules of thumb to dictate where the placement of large woody debris
would have the desired impact of reducing flood flows, indeed, many potential locations
for placing large woody debris were found to increase rather than decrease downstream
flood risk. Modelling using OVERFLOW identified that the installation of 100 debris dams
could reduce a flood event from 29.5cumecs to 27.3cumecs (7.5% reduction). The study
found that the large flood events experienced a greater percentage reduction than
smaller events however it should be noted that the placed woody debris was modelling by
localised elevation of Mannings roughness values which cannot take account of the
woody debris becoming drowned.
There is significant cross over between artificial large woody debris and riparian
woodland hence section (K) should be read for further information.
No literature relating to the Natural Flood Management benefits of in channel boulder
placements could be identified. It is assumed that these would have a similar, but notably
smaller impact than placed large woody debris.

(O)

River restoration

Many river restoration projects cite flood risk management benefits in support of the
restoration project, however it has been highlighted that there is an absence of data
detailing the nature of the flood risk management benefits these projects can have (MNV,
2008)11. In many cases it is evident that river restoration is undertaken to address

-67-

specific local issues. River restoration can have a role to play in delivering Natural Flood
Management by three means:
1.

Attenuating flood water as it moves through the catchment by improving


channel-floodplain connectivity in locations where the channel has become
incised.

2.

Extending the flow path length through the reintroduction of meanders thus
reducing the channel slope, reducing in-channel flow velocity and increasing
travel time.

3.

Removing artificial structures such as constrictive culverts and bridges which


do not have adequate flow capacity to discharge flood flows.

In both instances using river restoration to manage flooding is likely to be heavily


dependent on the local conditions.
A modelling study of the Cherwell (Acreman et al., 2002)172 catchment was undertaken to
assess the hydrological impact of restoring the connection of floodplains within the
catchment. The study employed a 1D hydraulic model and a semi-distributed
hydrological model to assess the impact of restoring entrenched river reaches to a natural
state. The natural conditions were based on the findings of a geomorphological study.
The study found that the complexity of floodplain hydraulics presented as considerable
challenge to the modelling work. The results showed that were embankments separated
the river from the floodplain the peak flow rate could be increased by 50 to 150%
(represented by adding embankments to reach presently without embankments) however
it reports that restoring the channel through the floodplain to its pre-engineered
dimensions could reduce flood flows by 10 to 15%.
Riley (1998)173 provides a comprehensive discussion on the use of river restoration to
increase channel capacity in North America. The source indicates that flood risk
management benefits of removing constrictive structures such as bridges or culverts can
be readily assessed using normal hydraulic modelling techniques.

(P)

Beach recharge

The Beach Recharge Manual (CIRIA, 2011)174 provides an authoritative overview of


beach recharge (commonly referred to as beach nourishment). The manual reports that
beaches are frequently the sole barriers to coastal flooding and erosion, while in other
locations they play a key role in supplementing engineered defences. The manual
advocates that beach recharge is a cyclic process of monitoring performance, gathering
information, assessing it and taking any necessary actions followed by the process being
restarted. Beach recharge projects are predominately undertaken to manage coastal
erosion but the practice can also be used to dissipate wave energy and reduce coastal
flooding risk via the overtopping of waves into defended areas.
Beach recharge was used in Aberdeen in 2006 to supplement a 600m section of eroding
beach which forms a component of the coastal protection to the city (University of
Dundee and Aberdeen City Council, 2010)175. The beach was nourished with
70,000cu.m of dredged sand. Artificial reefs were placed offshore to increase the wave
energy dissipation. It is understood that the beach nourishment at Aberdeen was
necessary to prevent the failure of the existing hard shore revetment which if it had failed
would have resulted in the loss of the esplanade, the release of contaminated material
and the eventual loss of built property.

(Q)

Sand dune management

Defra FD1302 (Pye et al, 2007)176 provides an authoritative overview of the use of sand
dunes for flood risk management. Part 2 of the report indicates that dune systems which

-68-

are both wide and high provide the greatest flood defence value. It is noted that narrower
systems do provide a reduced benefit however it is common for these systems to be
receding due to erosion. It is reported that dune systems in the Netherlands have been
assessed to offer a 1 in 10,000yr standard of protection to large areas. Part 3 of the
report describes dune restoration activities which predominately include fencing, control
of visitors, construction of board walks and revegetation. A guide to managing coastal
erosion in beach/dune systems (SNH, 2000)188 provides detailed guidance for sand dune
restoration projects.

(R)

Wave attenuation over saltmarsh and mudflats

Defras Saltmarsh Management Manual (Environment Agency and Royal Haskoning,


2005)177 reports it has long been assumed that saltmarsh provide a beneficial effect for
attenuating wave energy and thus contribute to flood and erosion protection. The manual
provides a concise literature review relating to the wave attenuation that can be provided
by coastal flats. It is reported by Mller that saltmarsh is approximately four times more
effective at attenuating wave energy than unvegetated sand and mud flats (Mller et al.,
1999; Mller and Spencer 2002) 178179. In the Wash the reduction in wave energy over
existing saltmarsh has been measured as between 72% and 97% (Cooper, 2001)180.
There are numerous factors which affect the efficiency of wave attenuation over
saltmarsh which include water depth, bed roughness, marsh edge characteristics and
vegetation type (EA and Haskoning, 2005)177, the most significant being depth. Mller
(2001)181 reports that in water depths of less than 1.1m on saltmarsh, the observed
reduction in wave energy is typically around 87%, dropping to around 72% for depths
over 1.1m. In sand flat areas the energy dissipation was reported as 37% and 27%
respectively.
The wave energy dissipation does suggest that saltmarshes do have a role to play in
coastal defence (EA and Haskoning, 2005)177, however it should be noted that the energy
dissipation efficiency of saltmarsh over plain sand decreases with depth. At
approximately 3.7m depth there is minimal difference between the efficiency of saltmarsh
and open sand (Mller et al., 2001)181. Thus in extreme events, i.e. those that might
cause flooding, the water depth would exceed this threshold thus implying that the
saltmarshes do not offer an increase in protection over open sand or mud. However it
should be noted that saltmarsh does offer benefits during more normal conditions and the
protection offered to the toe of defence structures should not be overlooked (Mller et al.,
2001)181.
It has been shown at Dengie Peninsula that the rate of energy dissipation is not constant
over the width of the flat and that a significant (40%) of incident energy is dissipated
within the first 10m width of the flat, of the remaining energy 60% on average is
dissipated over the next 28m (EA and Haskoning, 2005)177.
A study by Mller and Spencer (2002)179 showed that wave energy dissipation is greater
at steeper eroding saltmarsh boundaries (8%/m) than accreting (4%/m), which are
typically gradually inclined. It should be noted that an accreting system will gradually be
becoming wider, thus it should not be assumed that an eroding system is a more
sustainable long term approach.
Mller et al. (1999)178 estimated that if the saltmarsh at Stiffkey was lost completely, it
would cause a doubling of wave heights at the shore, due to the reduction in the surface
friction factor. This effect would vary depending on the site, in particular the elevation of
the saltmarsh and the nature of the vegetation. These results indicate that if saltmarsh
areas were lost and replaced by sand or mud flat, average wave heights would increase
at the shore or defence line.

-69-

The area of saltmarsh immediately landward of the permanently vegetated edge is most
significant in terms of attenuating wave energy, so vegetation changes in this area are
likely to have the greatest influence. Mller and Spencer (2002)179 observed a correlation
between lowered wave attenuation and seasonal decreases in vegetation density over a
gradually sloping saltmarsh edge. The same study observed no such relationship over a
steep eroding saltmarsh margin, suggesting that the dominant effect on wave dissipation
at this type of site is the eroding topography rather than the vegetation.
The vegetation type can also influence wave attenuation rates, due to the different
physical structure of the plants. Across Dengie, Mller et al. (1999)178 identified a possible
link between vegetation density, wave attenuation and the type of vegetation, with flexible
grasses in the upper saltmarsh having less effect on wave attenuation than woody plants.

(S)

Managed coastal realignment (reduction of estuarine surge)

Reed (2009)182 provides a concise overview of the function of estuarine tidal flats in
relation to flood risk by highlighting their value in damping storm surges from the ocean or
by providing storage of riverine floodwaters.
A review of coastal managed coastal alignment practices in England and Germany was
undertaken in 2002 (Rupp and Nicholls, 2002)183. The comparative study involved the
interview of six key managed realignment experts, three in each country. The general
consensus was that coastal realignment is correctly viewed as a cheap and
environmentally friendly option, provided the practical conditions on the ground are not an
insurmountable barrier. In both countries it was identified that at that time (2002) there
was a shortage of well documented managed retreats, with three pilot sites in the
Blackwater estuary in Essex being the best documented English site. All tidal retreat
schemes in the UK to that date amounted to little more than 100Ha. In most cases
managed retreats have been undertaken primarily for habitat creation purposes. It is
reported that there is agreement between the experts that the additional flats created by
retreat will assist in attenuating wave energy during storms. It should be noted that it was
identified by the probably more reliable work by Mller (2001)181 that the effect of wave
attenuation decreases with water depth and thus attenuation might be minimal during
extreme tidal surge events when it is most needed.
A published modelling study was undertaken to assess the feasibility of adopting
managed realignment as an adaptive management to sea-level rise (French, 2008)184.
The study uses the Blyth estuary in Suffolk as a case-study, reporting the estuary to be a
morphologically complex and heavily engineered. The reported model results indicate
that a sea-level rise of 0.3m (central estimate for 2050) would necessitate extensive
upgrades to existing tidal floodplain defences. Realignment (or managed retreat) of
these defences can reduce local flood defence costs by eliminating unsustainable
seawall but needs to be evaluated in the light of wider impacts. Modelling of hypothetical
realignment scenarios shows that restoration of tidal exchange to the largest flood
compartments could have an immediate effect on outer estuary hydrodynamics that is
larger than worst case scenarios for half a century of accelerated sea-level rise. More
generally, incompatibilities are apparent between flood defence and habitat restoration
objectives, such that the appropriateness and feasibility of large-scale flood defence
realignment should not be assumed in estuarine contexts.
As part of flood risk management within the Humber Estuary the Environment Agency is
utilising coastal realignment as a tool within a sustainable flood risk management
strategy. To date four coastal retreat sites have been created (OMReg, 2012)185:

Paull Holme Strays (Managed breach) 2003 80Ha

Chowder Ness (Defence removal) 2006 15Ha

Alkborough (Managed breach) 2006 370Ha


-70-

Welwick (Defence removal) 2006 54Ha

It is understood that the Paull Holme Strays, Chowder Ness and Welwick sites where
undertaken for habitat compensation purposes. It is reported that the Alkborough site
increases the level of flood protection to an area stretching from the Humber Bridge to
Goole up the tidal River Ouse and as far as Keadby Bridge on the tidal River Trent. The
scheme features include a 20 metre-wide breach of the existing flood bank, a 1,500
metre length of lowered embankment or spillway and a new section of floodbank to
protect assets at the edge of the site. For a 1 in 200yr event the scheme at Alkborough
will reduce extreme water levels by more than 150 mm (EA, 2006)186.
A directory of managed coastal realignment projects is hosted by ABP Marine
Environmental Research Ltd (OMReg, 2012)185. The database includes detailed
information on 51 managed breaches, regulated exchanges and defence removal
projects across the UK since 2001. A review of the data held within the database was
presented to the ICE by Scott et al (2011). In total 1,300Ha of coastal retreat has been
created since 1990, and as a result there is an increasing knowledge base relating to the
subject. The paper identifies that the scale and number of coastal retreat projects have
both increased in recent years as shown Figure 14. The paper also provides a useful
insight into the cost of implementing coastal realignment projects and finishes with the
closing proposal that it may be time to consider very large scale retreat projects
(substantially greater than 1000Ha) to improve the prospect of reaching national
ecological targets. CIRIAs C628 (Legget et al., 2004)187 provides a useful overview of
the issues associated with managed realignment projects.

Figure 14: Unit costs of implemented realignment plotted against size and year. Scott et al.
(2011).

(T)

Reefs and breakwaters

A guide to managing coastal erosion in beach/dune systems (SNH, 2011)188 describes


artificial reefs as shore parallel rock mound structures set part way down the beach face.
They can be long continuous structures or a series of short reefs and they are
distinguished from near shore breakwaters by being submerged for at least part of the
tidal cycle. The use of both artificial reefs and near shore breakwaters to reduce coastal
erosion is well documented and examples include:

Aberdeen artificial reefs

-71-

Sea Palling, Norfolk parallel offshore breakwaters

Newbiggin Bay, Northumberland isolated parallel offshore breakwater

Elmer, West Sussex parallel offshore breakwater

The design of artificial reefs and breakwaters is well documented including:

Environment Agency (2008) Estuary Edges: Ecological Design Guidance.


http://www.environment-agency.gov.uk/business/sectors/100745.aspx

Chasten, M.A., Rosati, J.D. Mc Cormick, J.W., Randall, R.E. (1993) Engineering
design guidance for detached breakwaters and shore stabilisation structures.
CERC-93-19, US Army Corps of Engineers Waterways Experiment Station,
Vicksburg, Mississippi.

CIRIA, (2007) The Rock Manual, Second Edition, CIRIA C683. ISBN 0-86017683-5.

Nottage A. S. and Robertson P. A. (2005) The saltmarsh creation handbook: a


project managers guide to the creation of saltmarsh and intertidal mudflats. The
RSPB, Sandy and CIWEM, London.

Pilarczyk, A. and Zeidler, R. (1996) Offshore Breakwaters and Shore Evolution


Control, Rotterdam, A.A. Balkema Publishers.

Simm, J. D., Brampton, A. H., Beech, N. W., Brooke, J. S. (1996) Beach


Management Manual, CIRIA Report 153, London.

The application of biogenic reefs to act as breakwaters in the Mobile Bay, Gulf of Mexico,
indicates that oyster beds can be used to protect an eroding shoreline (Cyphers, et al.,
2011)189. The study sites used loose shells held temporarily in place by mesh until
adequate oysters are recruited to cement the loose shell substrate in place. This
naturally cemented substrate then can provide an effective protection to medium energy
eroding sites. It was found that the biogenic reef slowed the rate of erosion by 40% when
compared against a control site. The study accepts that this technique is not viable for
high energy sites where due consideration may need to be given to more traditional
measures.

(U)

Strandline and beach management techniques

It is understood that West Sands Partnership in Fife is trialling techniques which will
improve the erosive resistance offered by the strandline and that St Andrews and Abertay
Universities are undertaking research to assess its effectiveness (pers comm). The trial
includes the hand picking of litter rather than using mechanical collectors with the
intention of leaving natural material to form a natural strandline. With time it is hoped that
the approach may improve the natural colonisation of the standline by vegetation allowing
the seaward extension of the dune system. Early indications are that the modified
maintenance regime is proving very effective at stabilising the seaward edge of sand
dunes.

3.3.2 Summary
The literature relating to a broad range of Natural Flood Management measures or
techniques have been presented in this Section of the report. The quality and extent of
literature relating to the 21 techniques is variable, however it is apparent that all
measures have a role to play in the future sustaibable management of flood risk. Some
measures have been shown to be effective in most cicumstances while evidence
suggests that other measures can be counterproductive if implemented in the wrong

-72-

circumstances. A full summary of the Natural Flood Management measures considered


is presented in the Natural Flood Management Summary Table in Appendix B.

-73-

3.4 Review of Natural Flood Management screening and


assessment tools
The DEFRA FD2114 study (OConnell et al., 2004)190 provided an authoritative position
statement on the status of current hydrological modelling techniques that were available
at that time to investigate the impacts of Natural Flood Management type activities:
there are serious shortcomings in the rainfall-runoff models and methods available for
the use in the operational assessment of the impacts of land use change and land
management practices. There are three fundamental unresolved issues: there is no
generally-accepted theoretical basis for the design of a model suitable to predict impacts,
it is not known which data have the most value when predicting impacts, and there are
limitations in the methods available for estimating the uncertainty in predictions. Some
general recommendations can, however, be made for a way forward in rainfall-runoff
modelling for predicting impacts. The modelling should be distributed and be capable
of running continuous simulations. It should also be partially or wholly physically
based so that the physical properties of local landscapes, soils and vegetation can
be represented, and it should include detailed modelling of surface water flow so
that the effects of changes can be tracked downstream. A considerable amount of
high-quality field data on impacts will be needed to support the development of robust
methods for predicting impacts.
Since 2004 a significant research effort has been applied to this area as presented in the
following items.

Environment Agency Methodology for the identification of


catchments sensitive to land use change (JBA, 2008)191

(A)

This is a national screening tool which has been developed for the Environment Agency.
The tool is GIS based and employs commercially available national data to calculate
Combine Runoff Sensitivity on a 1km grid for the whole of England and Wales. The tool
considers the following aspects:

Land cover class, based on LCM2000, scored 1 to 4

Soil class, based on NATMAP1000 (there may be scope to use HOST data),
scored 1 to 4

Slope class, based on NEXTMAP DTM, scored 1 to 4

Rainfall class, based on Standard Average Annual Rainfall, scored 1 to 4

These scores are then combined via the following equation to produce a national grid of
sensitivity to land use change:

Sample output from the tool are shown in Figure 15 and Figure 16,
The tool is configured for English and Welsh type catchments and would require some
minor modifications before it could be applied in Scotland, i.e. it the rainfall banding runs
from <600mm a year (Very low) to >900mm a year (High).
The tool does not include a flow routing element for flood flows and only looks at runoff
generation potential. The incorporation of a flow routing element together with
incorporation of known downstream flood receptors is a recommendation for a future
extension of the tool.

-74-

Figure 15: Sample output, River Ouse combined sensitivity (Environment Agency, 2008)

Figure 16: Sample output, National assessment combined sensitivity (Environment


Agency, 2008)

-75-

Environment Agency Land management CFMP Tool (JBA, 2008)192

(B)

The purpose of this bespoke tool was to assist the Environment Agency in the
implementation of a delivery plan for the Catchment Flood Management Plans (CFMP) by
assessing the effect of rural land use and land management on the flood hydrograph.
The tool employs the WaSim daily lumped soil water balance model to simulate the
effects of climate, soil type, field condition and land cover on daily runoff depth and
antecedent moisture conditions. The changes in maximum daily runoff depth have then
been used in the Revitalised Flood Hydrograph (ReFH) model to assess the change in
peak flow.
The Land Management CFMP Tool allows users with local knowledge to explore how
realistic changes in land use can quantitatively affect the downstream hydrograph. Within
the Excel interface the user has the option to change the proportion of the following broad
land cover classes:

Managed grassland;

Cereals;

Horticulture/non-cereal crops;

Semi-natural; and

Woodland.

The effects of land management can be explored through changing the distribution of
field condition which encapsulates both the condition of the soil and the presence of land
management practices which alter runoff transmission.
The report stresses that, despite the inevitable assumptions and simplifications
necessary to develop the tool for national applicability, the scenarios suggested changes
in peak runoff which are reasonable within the context of the uncertainty of current
scientific understanding. The methodology is reported to be repeatable, academically
sound and supported by the Environment Agency.

Figure 17: User interface of the Land Management CFMP Tool (Environment Agency, 2008)

-76-

(C)

Forest Research screening method (Broadmeadows & Nisbet,


2009)193

A range of GIS data sets were used to locate land vulnerable to flooding and unaffected
by constraints to woodland planting. The approach built on previous work in the
catchments of the River Parrett in Somerset and Bassenthwaite Lake in Cumbria. The
main output is a series of maps showing opportunities for planting floodplain and riparian
woodland for flood mitigation within the region. Areas were prioritised according to the
scope for planting to generate added value for nature conservation and water quality.
Floodplain sites were ranked by their potential to create an extended forest habitat
network, while riparian zones took into account the risk of the adjacent land delivering
sediment to watercourses.
The project also assessed opportunities for woodland planting to assist flood and water
pollution management within the wider catchment. This used data sets that classified the
catchment soils by their vulnerability to generate rapid surface runoff, degrade structurally
and/or deliver sediment to watercourses. The result is a map ranking areas as low,
medium or high priority for woodland planting.
Significant opportunities exist within the region to restore floodplain woodland for
sustainable flood management. A total of 35,328 ha of the fluvial floodplain is potentially
available for woodland planting, including 168 major sites with an individual area of >50
ha. The majority of these lie within the catchments of the River Derwent, River Swale,
River Ure and central section of the River Ouse, where there is potential to help reduce
the flooding of small towns and villages, as well as major towns such as York.
There are also many opportunities across the region for using riparian woodland to help
reduce flood flows. A total of 18,730 ha are available for planting in the upper reaches of
most rivers, including those catchments in the Southern Pennine Fringe at high flood risk
but with limited potential for planting floodplain woodland. Around 2,562 ha is adjoined by
land at high risk of sediment delivery, where woodland creation could potentially benefit
both flood risk management and diffuse pollution control. Some 997ha lay within ECSFDI
priority catchments, although most of this is located in the catchment of the River
Derwent.
Finally, 65% of the land in the wider region is potentially available for woodland creation
for multiple benefits, including flood reduction. Some 1,538sqkm or 16% of the region is
mapped as high priority for woodland planting on the basis of soil propensity to generate
rapid runoff, soil sensitivity to structural degradation, and/or high risk of sediment delivery
to watercourses. This land tends to be concentrated in the upper parts of river
catchments and is most extensive within the River Ure, Swale, Derwent and Nidd
catchments.
It is hoped that the opportunity maps will be used by regional stakeholders to promote the
use of woodland in sustainable flood management and in so doing, help to meet the
following objectives:

Delivery of a sustainable flood risk management strategy for the region

Delivery of reduced flood risk through effective and better integrated Catchment
Flood Management Plans

Delivery of flood alleviation for smaller communities where traditional methods of


flood defence are not cost effective

Delivery of the regional Biodiversity Habitat Action Plan targets for wet woodland
and an enhanced forest habitat network

Delivery of Regional and National Forestry Strategies, including climate change


adaptation
-77-

Contribute to a reduction in diffuse water pollution and an improvement in


hydromorphology, thereby helping to meet EU Water Framework Directive targets
for all water bodies to reach good water status by 2015

Contribute to the England Catchment Sensitive Farming Delivery Initiative

Develop partnerships to establish floodplain, riparian and wider woodland creation


demonstration projects within the region as a way of developing a local evidence
base and communicating the expected success of this option for flood and water
pollution management

Figure 18: High priority areas with the greatest potential for woodland planting to reduce
193
downstream flooding (Broadmeadows & Nisbet, 2009)

Limitations:

(D)

The tool does not make account of spatial variations in precipitation

The tool provides an overview of opportunities within the catchment, it does not
quantify the benefits which might be achieve from implementing that measure nor
does it provide an indication of which measure is most beneficial

There is not assessment of time to outlet hence the tool does not allow
assessment of the contribution that the identified opportunity might have to the
flood peak

SEPA, Allan Water (Halcrow and CRESS, 2011)194

A single event distributed hydrological model which estimates runoff rates using the
American Soil Conservation Services Runoff Curve Number methodology. The ArcGIS
model includes overland and channel flow routing. Channel flow routing is via a modified
normal depth routine (kinematic wave). The method uses a grid smoothing technique to

-78-

provide a means of inferring the backwater profile, while being less accurate than a
diffuse wave method the approach is numerically robust and computationally efficient.
The model predicts channel dimensions and roughness values based on estimates of
median annual flood and channel slope. The channel routing routine includes
functionality to represent floodplain flow however floodplain width is based on typical
floodplain widths (based on valley slope) found within the catchment. Floodplain
roughness values are based on a user supplied grid and the threshold for spill of water
from the channel to the floodplain can be user defined (site reconnaissance or review of
the morphological pressures database). The transition from overland runoff to channel
flow can be user defined, i.e. using a mapped river network or once a catchment area
threshold is exceeded. Overland flow routing is based on a simple normal depth routine
which assumes a constant depth over the cell width, the roughness used for this routine
is via user supplied grid which can be based on landuse.
Vegetation interception and initial ground infiltration is based on the simplified initial
abstraction provided within the SCS methodology. It should be noted that this has been
found to be dependent on local antecedent moisture conditions and should be reviewed
to maintain a Scottish context.
The method does not simulate the whole storm hydrograph, instead the hydrograph peak
is estimated by estimating the time that the flood peak would occur at and then
accumulating the runoff generation from across the catchment that would arrive at the
catchment outlet at that time.
The model does not include groundwater flow and it is assumed that all infiltrated water is
lost from the system. It would also be desirable to improve the channel routing routine to
be based on the more accurate diffusive wave.
The model has not been formally tested however tests on the Allan Water indicate time to
peaks similar to estimates made using the FEH time to peak by catchment descriptors.
The model is encoded as a ToolBox within ArcGIS in an incomplete state; that is no clear
user interface, user prompts or help files. The incomplete state also means that applying
the model is user intensive requiring output from one routine to be manually passed to
the next routine.

SEPA, S20 GIS Tool (Jacobs, 2011)3

(E)

The Natural Flood Management tool proposed by Jacobs as part of a SEPA funded
project to develop a methodology to apply Section 20 of the Flood Risk Management
(Scotland) Act is presented in detail within Section 2 of this report. In summary the
ArcGIS tool allows the user to scale hydrograph peaks based on land use changes within
the catchment. The percentage changes included within the tool are based on published
literature and uncertainty is managed by the tool via the use of qualitative uncertainty
bands pessimistic, central and optimistic. The tool incorporates functionality to
estimate the percentage change in flows for the 1 in 5yr and 1 in 200yr flood events for
the following land use change types:

Forestation;

Riparian woodland;

Floodplain woodland; and

Drain blocking.

The S20 GIS tool uses the Automated Flow Grid produced by CEH as the baseline
hydrology, this gives the tool good national coverage but introduces a challenge for
distributing the tool due to licensing. As presented in Section 2 of this report the tool
suffers from a number of fundamental technical issues including how the hydrological

-79-

benefits of the measures, particularly floodplain enhancement, are assessed, and the
assumptions based around the division of the catchment into thirds.

University of Bangor, POLYSCAPE (Pagella et al., 2009)195

(F)

POLYSCAPE is a multiple criteria GIS toolbox designed to assist decision making of


where to place Natural Flood Management and landscape scale ecosystem service. The
tool evolved from the research undertaken within the Pont Bren catchment (Wales) where
it was observed that strategically planted shelter belts could significantly reduce runoff,
sediment mobilisation and water quality. The tool was intended to balance the interests
of water, biodiversity, agriculture and forestry which all have their own conflicting or
complementary interests.
It is reported that the tool can be applied to any landscape with readily available data and
that although it has been developed for a small (10sqkm) catchment it could be
applicable for medium (1000sqkm) catchments. The information on how the tool actually
functions is sparse however it is understood that it comprises of the following modules or
components:

Farm impact:
o

Habitat (forest) connectivity:


o

Impact based on slope, current land use and water regime


Impact assessed using habitat networks via the BEETLE196 algorithm

Flood mitigation:
o

Via an undefined flow accumulation algorithm which is modified by soil


type and land cover

It is evident that the tool is centred around the principle of optimally placed
shelter belts based on the research conducted by H. Wheater at Imperial
College London, but there is no published information on how this is
achieved.

Tradeoffs:
o

Tool explores trade offs, multiple benefits and opportunities for change

The tool operates on a traffic light system, with:

Green areas highlighting high opportunities for change

Red areas highlighting existing utility and hence should be wary of


change

Orange areas highlighting either marginal opportunity for service


provision or areas of trade off

Sample output from the tool is shown in Figure 19.

-80-

Figure 19:
Sample output
from the
POLYSCAPE tool
for the Pontbren
catchment
(Pagnell, 2009).

Previous consultation in 2010 by Scottish Borders Council with the author of the toolbox
indicated it was unclear if he was going to continue the development of the tool which
was at the time in an unfinished state. There is no evidence that the development of the
tool has continued since 2010, hence this tool does not represent a viable option for the
screening or assessment of Natural Flood Management measures in Scotland.

(G)

University of Durham - FARM (Floods and Agricultural Risk Matrix)


(Hewett et al., 2006)197

FARM is a decision support matrix designed to allow farmers and land use planners to
investigate land use interventions at a field scale. The tool is written in Excel and is
currently available for download.
The tool is aimed at farmers and local flood risk managers who have a high level of local
knowledge but less awareness of what constitutes a high or low runoff risk. As the
FD2114 Impact Study Report has stated, the factors affecting catchment scale runoff are
difficult to quantify and are still open to some debate, hence, the FARM tool has been
deliberately aimed at the farm scale where runoff processes can at least be observed and
are backed up by the ISR findings. Aggregation to the catchment scale can be performed
via a FEH toolkit (Packman et al., 2004)48.
The tool functions by asking the user a series of questions to identify the soil infiltration,
storage and tillage regime in addition to the flow connectivity. Throughout the process
the user can see the impact of their answers through the animated highlight of the matrix
(Figure 20). Examples are provided to assist the user in answering the questions (Figure
21).

-81-

Soil and
management

Connectivity
Figure 20: Example of current UK farms and likely impacts of land use management
scenarios and how they can be mapped onto the decision support matrix (Hewett et al.,
2006)

Figure 21: Four hillslope runoff risk scenarios for the same common land unit (Hewett et al,
2006).

-82-

Given the simplicity and field scale of the tool, it is considered that FARM is most
appropriate for educating land managers. Other tools such as the Land Management
CFMP Tool (Section (B)) will be a more valuable resource for professional practitioner
use.

University of Durham OVERFLOW (Odoni and Lane, 2010)198

(H)

OVERFLOW is a distributed hydrological model which incorporates a simplified runoff


generation algorithm with an integrated surface flow routing model. The model has been
developed for the purpose of modelling the runoff response of very wet catchments and
therefore includes a number of simplifications. These simplifications offer a significant
reduction in computational time when compared to conventional models such as CRUM3
(Lane et al., 2003)199. Assessment of the impact of upstream land management
measures on flood flows in Pickering Beck using OVERFLOW (Odoni & Lane, 2010)198
does not provide a detailed description of OVERFLOW, however it does state that it is
intended to write up in due course with a view to publication in mainstream scientific
literature. An MSc Thesis (Porter, 2011)200 provides additional information on the model
and indicates that two papers are in preparation:

Odoni, N. A. and Lane, S. N. (Forthcoming) "Development of the OVERLOW


distributed model for testing catchment-wide land management interventions to
reduce flood risk: (1) model conception and testing." Hydrological Processes.

Odoni, N. A. and Lane, S. N. (In preparation) "Knowledge-Theoretic Models in


Hydrology."

It is understood that the model incorporates the following assumptions (Odoni & Lane,
2010)198:

Groundwater flow is represented via a catchment wide approximation of the


baseflow which can be made using flow records from an existing river gauge
network.

Rivers are assumed to form when the flow exceeds 0.06cumecs, with watercourse
dimensions being estimated using generic channel width and depth equations
such as those published by Leopold and Maddock (1953)201.

In channel and out of bank flow resistances are estimated using representative
catchment wide Mannings roughness values. It is noted that there is potential to
vary Mannings values in any way that the user chooses provided there is
adequate data and justification to do this.

With assumptions made concerning input rainfall, baseflow, runoff percentage, channel
geometry and Mannings, the model has all the components needed in order to run. A
flow accumulation and routing algorithm calculates how rain falling on the catchment is
made to flow through the landscape, with the assumption that the catchment is saturated
and the water that contributes to the flood wave moves primarily by overland flow. Flow
paths are allowed to vary as a function of rainfall depth, as well as in situations where
some of the flow is out of bank and routed across floodplains. The flows are converted to
inferred flow depths and velocities via Mannings thus allowing an estimation of travel
time. By accumulating these along flow paths, starting at the outlet, a flow time map is
generated for the rain rate in question (Figure 22).

-83-

Figure 22: Example of a time map


generated by OVERFLOW for the
Pickering Beck catchment, in this
instance a 14mm per day gross
rainfall even, assuming a percentage
runoff of ~55% after adjusting for
evaporation and groundwater losses.
The timescale is in hours to the
catchment outlet.

It is understood that the tool does not include a means of representing spatially varying
rainfall or percentage runoff across the catchment although in the available material it is
stated this is something that the author is working towards. There is no discussion on
how the attenuation effect of lochs and reservoirs might be incorporated within the model.
OVERFLOW has to date been only applied to assess CRIMS (Catchment Riparian
Intervention Measures) such as placed large woody debris and riparian/floodplain
planting. Should the model be improved to include modules to assess the spatial
variation in runoff generation (based on land use and soil types) and to allow the variation
in rainfall across the catchment then the modelling method could be of use for assessing
Natural Flood Management measures in Scotland.
It is unclear if the tool would be commercially available for the Section 20 assessment as
previous approaches by Scottish Borders Council have found that the developer is not
keen to release the model to industry at this time requiring any assessment to be
undertaken by staff at Durham University.

(I)

SEPA, Diffuse Pollution (ADAS and MLURI, 2006)202

The Diffuse Pollution Screening Tool presented within report WFD19 (ADAS & MLURI,
2006)202 includes a water balance model. The model functions on a 1km grid and
produces monthly estimates of the generation of overland flow, sub-surface flow and
drainage to groundwater. The model makes allowance of the spatial variation in
precipitation, evapotranspiration, land use and soils and has been used to produce
monthly estimates of overland flow, sub-surface flow and ground water for the 10 year
period from 1989 to 1998. It should be noted that the tool was not developed for flood
risk management purposes but it may present a means of quantifying likely antecedent
moisture conditions following land use changes in addition to useful data sources.

-84-

(J)

Physically based distributed hydrological model

There are nearly as many hydrological models as there are hydrologists (Beven, 2012)203,
with many hydrologists keen to develop and promote their own hydrological model. Many
of these hydrological models are physically based distributed models such as:

IHDM Institute of Hydrology Distributed Model

SHE System Hydrologique Europeen

CRUM3

HYLUC Hydrological Landuse Change model

PRMS Precipitation Runoff Modelling System

CLASSIC Climate and Land-use Scenario Simulation In Catchments.

It is argued by Beven (1988)204 that there are


fundamental problems in the application of
physically-based models for practical
hydrological predictions and that attempts to
represent increasingly fine detail within
hydrological processes at ever reduced
resolutions leads to the replacement of one
uncertainty with others. Ultimately complex
physically based models are heavily dependent
on data which cannot be collected or verified.

Physical model
- a model that
seeks to represent hydrological
processes in a physically realistic
manner.
Semi-physical model
-a
hydrbid model which represents
some hydrological processes in a
physical manner and others via
functions that have no direct physical
basis.
Lumped model - A model that treats
the whole of a catchment as a single
accounting unit and predicts only
values of variables averaged over the
catchment area.

Beven (2012)203 advocates the use of semiphysical models which lie in a grey area
between physical and conceptual lumped
models, two such models are GRID-TO-GRID
(G2G) and TOPMOD. These semi-physical
models have a much reduced parameter set
when compared to physical models, therefore
the process of supplying data and calibration is made much more straightforward.
Normally the parameters sets have an underlying physical basis to key hydrological
processes but there is no attempt to represent every process that occurs within the
catchment. He notes that even the current state of the art distributed physical models do
not present a significant improvement in accuracy over simple lumped models and he
warns that an increase in complexity does not necessarily lead to an increase in
accuracy.
There are currently a number of research initiatives seeking to improve the conditioning
of rainfall-runoff hydrological models; two recent publications include Bulygina et al
(2009)205 and Moore et al (2007)206.

(K)

Louisianas 2012 Coastal Management Plan - Planning Tool207

It is understood that a planning tool has been developed as part of the Louisianas
Coastal Management Plan development which is being developed in the wake of
hurricane Katrina. The tool functions around an index which is based on the location and
amount of land in proximity to flood receptors, vegetation and elevation. The index also
includes the effect that changes might have on the attenuation of waves and surges.
The approach has allowed 400 project options, considering of terabytes of data, to be
assessed in a user friendly manner covering areas such as costs, duration, landscape
and stakeholder preferences. The planning tool is not designed to make decisions but
instead it provides the information to assist in the decision making process of grouping
projects together to achieve coast wide goals.
-85-

3.5 Literature review summary


There is a wealth of scientific literature relating to the broad area of Natural Flood
Management, which at a broadscale covers much of the scientific literature relating to
hydrology, hydrogeology, hydromorphology, hydraulics and coastal science. As
presented by previous literature reviews (Section 3.2 of this document) the scientific basis
for Natural Flood Management is relatively weak, particularly for diffuse fluvial and pluvial
flood mechanisms.
The Natural Flood Management Summary Table in Appendix B provides a
comprehensive overview of the broad range of measures which have been identified by
this study. Initially this table was developed with the intention of summarising the
pertinent issues relating to all identified measures so that ineffective measures could be
eliminated from Section 20 and subsequent stages of the Flood Risk Management Plan
development process. However, on completion of the literature review process it was
realised that all Natural Flood Management techniques have their place, and it would be
inappropriate to eliminate specific measures.
The literature presented in this section of the document (Section 3) is built upon in
Section 4 by bringing together the current understanding of Natural Flood Management
measures and existing assessment techniques to develop a means of identifying and
appraising Natural Flood Management measures.

-86-

4 Proposed methodology for assessing and


considering the contribution of Natural Flood
Management
4.1 Introduction
The summary of requirements on SEPA in relation to Natural Flood Management
under the FRM Act is provided in Appendix D. These requirements essentially fall
into two broad categories:

To assess whether Natural Flood Management can contribute to the


management of flood risk (FRM Act Section 20); and

To take into account the identified Natural Flood Management potential when
setting objectives and identifying measures in flood risk management
strategies and plans (FRM Sections 27, 28 and 34).

The methodologies proposed in this report aim to deliver requirement under Section
20 to produce maps of Natural Flood Management potential, whilst also proposing
how this information can be used in the objectives setting and identification of
measures in flood risk management strategies. Figure 23 summarises the overall
approach to delivering Natural Flood Management requirements.

-87-

Section 20 Identification of potential


Running of Screening Tools
Output: Maps showing the potential for Natural
Flood Management in catchments with
Potentially Vulnerable Area

2012
SEPA

Output: Section 20 maps

Section 28 Appraisals

2013 - 2014

Using output of Screening in combination with


catchment characteristics and further
information to develop a long list of measures
and refinement to short list. More detailed
appraisal producing a list of preffered
measures.

SEPA / Responsible
Authorities

Output: A list of preferred FRM measures

Section 34 Local FRM Plans

2015 16

Development of the preferred measure using


appropriate modelling and assessment tools
(not exclusively the proposed hydrological
assessment tool)

The relevant
Responsible Authority

Output: Local FRM Plans

Figure 23: Outline process for the delivery of Natural Flood Management requirements
under the FRM Act

-88-

National screening for NFM


potential

NFM assessment tools


A range of tools which will allow
the flood risk management
benefits of NFM measures to be
quantified.

GIS rule based assessment to


provide information on what NFM
measures might be effective where

National maps of NFM


Opportunities (Section 20)
Section 20: Identification of potential

Appraisal of NFM against as


Agreed
measures for
inclusion in
LFRMP

well as in combination with


other measures
(Sections 27 and 28)

Baseline
national
pluvial, fluvial
and coastal
flood maps

Appraisal
methodology
and additional
guidance

Catchment
characteristics
and local
information

NFM Fluvial Assessment


Tool and other existing
methodologies

Implementation of
measures

Section 28: Appraisal

Section 34: Local FRM Plans

Figure 24: Flow chart showing the inclusion of Natural Flood Management within the
FRM Act

It is proposed that the identification and appraisal will be divided into three main
phases:
1.
A national screening process to identify opportunities for Natural Flood
Management measures. This screening process will identify areas within catchments
with Potentially Vulnerable Areas that have natural flood management opportunities
but will make no consideration of constraints or other benefits. The screening
process will not directly recommend which specific measure should be implemented
where, nor will the screening facilitate the quantification of the flood risk management
benefits of undertaking a specific natural flood management activity. However the

-89-

process will facilitate the identification of areas that are worth further investigations at
a later stage. The main output of the screening process will be six maps showing:

Areas of high runoff generation;

Areas of floodplain storage potential;

Opportunities to remove hydraulic constrictions;

Areas of heightened hydromorphological activity;

Areas of estuarine surge attenuation potential; and

Wave energy attenuation potential.

The proposed methods for constructing the GIS based screening tools which will be
used for this assessment are presented in Section 4.3.2. Please note that subcatchment desynchronisation will not be undertaken due to unacceptable uncertainty
in the underlying science and data constraints.
It is proposed that these maps, will meet the requirements of Section 20 of the FRM
Act. These maps will be made publically available along with appropriate supporting
descriptions and guidance.
2.
The second phase will seek to bring together the outputs from the screening
process with further information about catchment characteristics in order to develop
better understanding of what may be achievable where within the catchment. Using
the findings of the screening process in combination with other catchment information
the measures will be presented as options to manage the sources and pathways of
floodwaters, as summarised in the Natural Flood Management Summary Table
(Appendix B). This information together will be used to appraise natural flood
management measures alongside other flood risk management measures. Over the
course of this stage the long list will be reduced to a short list and the short list
appraised to identify a basket of the most sustainable measures. The process will
give due consideration to environmental, economic and social costs and benefits.
The end result of this stage is a short list of options that will be agreed by all
responsible authorities.
3.
The third phase will include more detailed assessment of the agreed,
measures using existing and new assessment approaches. This more detailed
assessment will be carried out by the relevant responsible authority as part of local
flood risk management plans.
The Section 3 literature review has identified that some measures can be assessed
using existing tools and methods, while the assessment of other Natural Flood
Management techniques require the development of new tools. It is proposed that
the assessment is undertaken using existing tools when available. In the case of
fluvial and pluvial flood risk there is a general absence of existing tools and methods
which can be used to quantify the effects of Natural Flood Management measures. It
is therefore proposed that a new hydrological method is developed to facilitate the
quantification of fluvial and pluvial Natural Flood Management measures. Section
4.4 presents a proposed specification for how this could be accomplished using a
single event spatially distributed model based on the uniformly distributed PDM
(Probability Distributed Model) runoff generation model.

-90-

4.2 Identification of Natural Flood Management measures


Following the completion of the literature review an attempt has been made to
concisely summarise all Natural Flood Management measures and the key
information relating to these measures within the Natural Flood Management
Summary Table (Appendix B). The 29 identified measures have been divided into
five broad categories based on the flooding mechanisms:
A) Fluvial - 19
B) Pluvial 8 (all also fluvial measures)
C) Coastal - 8
D) Groundwater - 1
E) Urban 1 (wide umbrella of SuDS)
These measures have then been sub-classified based on how the measures
function:
A) Runoff reduction reducing flooding by controlling at source
B) Attenuation reducing the magnitude of a flood wave before it reaches the
receptor
C) Delay/desync delaying the progression of a flood wave thus potentially
increasing the warning time or allowing flood peaks to be desynchronised
D) Protect reducing the flood risk by raising the threshold for flooding at the
flood receptor
In addition to classifying Natural Flood Management measures, the summary table
also provides a simplistic overview of how the Natural Flood Management measures
function, where the measures might typically be found, pertinent literature, modelling
methods and knowledge gaps. This classification system is comparable to the
source, pathway and receptor terminology which has been previously presented
however the presented terminology provides a much more immediate description of
how the measures function. The information contained within the table has then
been used to identify the course of action for enabling each Natural Flood
Management measure to be incorporated within Section 20 of the Act and
subsequent Flood Risk Management Plans. There are in general, three different
mechanisms for including the identified measures:
A) Screening This will be undertaken where it has been assessed that the
Natural Flood Management measure is likely to offer significant potential at a
national scale and where there is a technically feasible means of undertaking
a meaningful screening process. Not all Natural Flood Management
measures will be screened for. Section 4.3 details the proposed methodology
for the screening processes.
B) Assessment using existing methods It was identified that a notable number
of Natural Flood Management measures can be assessed using existing tools
and methods as detailed in Table 10. It is proposed that these measures are
assessed on a site specific basis using these existing tools.
C) Assessment using the proposed fluvial assessment tool In the case of
fluvial and pluvial measures there is currently an absence of appropriate
assessment methods for evaluating the flood risk reductions. Section 4.4.2
details the proposed fluvial assessment methodology.

-91-

4.3 Screening for Natural Flood Management opportunities


4.3.1 Introduction
A Natural Flood Management screening process will be undertaken to assist in the
identification of areas where Natural Flood Management can contribute to the
sustainable management of flood risk. To complete this process it will be necessary
to divide the screening into a number of sub-processes each targeting different types
of Natural Flood Management measures on the basis of how the Natural Flood
Management measure function:
A) Areas of high runoff generation
B) Areas of floodplain storage potential
C) Opportunities for sub-catchment desynchronisation
D) Opportunities to remove hydraulic constrictions
E) Areas of heightened hydromorphological activity
F) Areas of estuarine surge attenuation potential
G) Areas of wave energy attenuation potential
It is essential that the Natural Flood Management screening is undertaken
consistently at a national scale, it is therefore essential that all data required to
complete it is available nationally to a consistent standard. Further information on
the proposed data sources is presented in Section 4.3.3.
For screening it will not be possible to accumulate the effects of Natural Flood
Management measures on reducing flood risk a receptor level as to do so would
require each flood receptor to be considered individually. The accumulated effects of
multiple Natural Flood Management measures will not be accumulated until the
appraisal phase.
There will be no national screening for potential Natural Flood Management
measures in regards to urban and groundwater flooding as there is inadequate
consistent national data to allow such a screening process to be undertaken. This is
not to say that Natural Flood Management does not have a valid role to play in
managing these mechanisms. It will therefore be necessary to use local knowledge
and subsequent site specific studies to screen for Natural Flood Management
opportunities. In urban areas, NFM measures should be considered as part of
surface water management planning.
It is envisaged that the maps produced at the end of the screening exercise will be
useful resources in generating ideas for further validation through appraisals process
and local flood risk management plans. However the screening tools will not make
any allowance for local constraints such as the presence of infrastructure,
environmental designations with conflicting interests or areas of cultural heritage.
These constraints will be considered in the appraisals to prevent the incorrect early
exclusion of valid measures.

4.3.2 Screening methods


(A)

Areas of high runoff generation

This screening process aims to produce a national dataset of runoff potential with the
intention of providing an insight into which areas contribute most to the generation of
fluvial and pluvial flows. With this information, it will be possible to review the existing
-92-

landuse, then in conjunction with the information contained in the Natural Flood
Management Summary Table in Appendix B and the local constraints or
opportunities an alternate reduced runoff landuse can be identified. In this way runoff
reducing activities can be targeted to the areas where they will be most effective
before a quantitative assessment of the hydrological effect is undertaken. A number
of means of completing the screening for high runoff areas have been identified:
1. Using the Environment Agencys Method for the identification of catchments
sensitive to land use change (JBA, 2008)191
This previously applied method creates a Combined Sensitivity Score via a method
described in Section 3.4(A). It takes into account land cover, soil, slope and rainfall,
but does not apply a weighting to these factors. There is also a data availability issue
regarding a degraded soil dataset used by the method. The method is built around
qualitative judgement based scores which are representative of conditions found in
England and Wales, thus application in Scotland would require additional study.
Equation 2 summarises how the land cover, soil, slope and rainfall classes are
combined to generate a combined score.

Equation 2: Combined runoff sensitivity score used within the Environment Agencys
identification of catchments which are sensitive to land use change.

The supporting report for the method indicates that it could be improved by weighting
the effect of the model variables but it does not recommend a means of achieving
this.
2. Using output from the Diffuse Pollution Screening Tool (ADAS and MLURI,
2006)202
The existing monthly overland and sub-surface flow data produced as part of this
study could be used to identify areas of high runoff generation. It should be noted
that the tool was not developed with the intention of assessing flood risk and
therefore it may be inappropriate to use the output in this way. The model makes
allowance of the spatial variation in precipitation, evapotranspiration, land use and
soils. It is also not clear if only monthly averages were produced or whether the
totals for each month in the 10 year period are available thus allowing an analysis
based on annual maximums.
As this method was not developed for the purpose of assessing flood events it is
considered inappropriate to use it for this purpose.
3. Using a modified version of the QMED by catchment descriptor equation
(Kjeldsen et al., 2008)208
The median annual maximum flood ( QMED ) for a catchment can be calculated
using Equation 3.

QMED 8.3062 AREA

0.8510

0.1536

1000
SAAR

FARL3.44510.0460 BFIHOST

Equation 3: Estimation of QMed via catchment descriptors (Kjeldsen et al., 2008)

208

The catchment area ( AREA ) and catchment storage ( FARL ) elements could be
dropped out of the equation to produce a measure of the median annual flood
generated per unit area ( q ) as detailed in Equation 4.

-93-

q 8.3062 0.1536

1000
SAAR

0.0460 BFIHOST

Equation 4: Estimation of runoff generation per unit area

Some land use pressures could be represented using the method of degraded soils
via analogue HOST classes proposed by Packman et al (2004)48. However the
method would not permit the inclusion of topography or the representation of existing
land use activities.
The alignment of this approach with current British flood hydrological methodology
and its simplicity make it a good alternative should the preferred (next) method not be
feasible.
4. Combination of the Environment Agencys Method for the identification of
catchments sensitive to land use change (JBA, 2008)191 and the modified QMED
by catchment descriptor equation (Kjeldsen et al., 2008)209
There is potential to produce a hybrid method based on options 1 and 3. This
approach would facilitate the quantitative of method 3 to be modified by factors
relating to land cover and slope. It would be necessary to undertake further study to
develop what weightings might be appropriate for the various components. A
potential relationship could take the form of Equation 5, where local topographic
gradient ( g ) and leaf area index ( LAI ) are modified by yet to be defined coefficients
( , , , , ).

q 8.3062 0.1536

1000
SAAR

0.0460 BFIHOST slope _ weight landuse _ weight

slope _ weight 1 g

landuse _ weight LAI

Equation 5: Potential hybrid method for screening for runoff generation

Initial suggestions for the slope coefficients could be in the region of 0.1
and 1 , these values would limit the effects of slopes to increasing the runoff
generation score ( q ) by 15 to 25%. It should be noted that these factors are
suggested based on speculation only. The factors must be selected and robustly
justified as part of the screening process. The variation of soil moisture capacity with
slope, as discussed in Section 4.4.2(C), could provide a scientific basis to this
weighting factor.
In regards to land use effects, there may be scope to track through the effect of LAI
on percentage runoff (Equation 26) by applying a range of LAI to Equation 31
together with typical Scottish SAAR, Potential Evapotranspiration (PE) and bulk
(canopy) resistance ( rsc ). Figure 25 shows one example of tracking the effect of LAI
through using some typical Scottish values, this suggests the relationship for the land
use weighting as detailed in Equation 6. Again it is essential that these factors are
reviewed and robustly justified as part of the screening process.

-94-

Link between LAI and runoff generation


0.7
BFI = 0.5
PE = 500mm/yr
SAAR = 1500mm/yr
r = 70s/m

0.69

Percentage Runoff

0.68
Median LAI for Scotland
is ~4 (4^4.7 = 675)

0.67

Generated using a
50mm rainfall depth

0.66
0.65

PR = -0.00002680(LAI^4.7) + 0.68854927
0.64
0.63
0

200

400

600

800

1000

1200

1400

1600

1800

2000

LAI^4.7

Figure 25: Variation in percentage runoff with a changing leaf area index using typical
Scottish catchment descriptors

landuse _ weight

0.0000268LAI

4.7
summer

0.672

0.6885 1.0244 0.0000399LAI

4.7
summer

Equation 6: Potential method for the inclusion of land use effects within the runoff
generation screening process

Equation 6 does not include any effect of soil degradation or the effect of canopy
interception. There is the potential to include the effect of soil degradation via the
analogue degraded HOST class (Packman et al., 2004)48 discussed in Section
4.4.2(C). The effect of canopy interception could be included through reducing the
rainfall as discussed in Section 4.4.2(B), however the effect of canopy interception is
anticipated to be negligible on anything but the smallest rainfall events as canopy
interception is likely to be completely utilised ahead of the hyetograph peak.
This is currently the preferred method due to the speed which it could be
implemented in combination with the useful insights the data could provide. It is
accepted that further work is required to develop a more robust means of including
land use effects and topography.
5. Using a distributed PDM model (modified version of the ReFH)
This approach could estimate the total runoff depth at field scale using the modified
ReFH losses model as presented in Section 4.4.2(C). This model would facilitate the
inclusion of soil type, land cover, topography and rainfall depth within the
assessment. It would be necessary to undertake the assessment for a standardised
rainfall return period and duration, this could be based on a carefully selected return
period and duration (i.e. 3hr 1 in 50yr rainfall) requiring the calculation of a national
grid of values.
Importantly the method would permit the existing runoff generation scenario to be
compared to an idealised land cover scenario thus facilitating an insight into the
value of altering land use within that area. This approach could also used to assess
the effect of degradation linking to the requirements of Section 19.

-95-

While BFIHOST can be calculated from vector HOST data to approaching field scale
resolutions via the method laid out in IH126 (Boorman et al., 1995)137 generating a
grid of PROPWET values to a similar resolution would be more problematic. Section
4.4.2(D) presents a number of potential means of improving the spatial resolution for
estimates of PROPWET . It is noted that the small power applied to PROPWET (0.24) has the effect of largely negating its influence on CMax as the range of

PROPWET across Scotland typically varies from 0.5 to 0.8 as shown in Figure 26.
As demonstrated within Figure 27 SAAR can be used as a reasonable predictor
of PROPWET , hence Equation 28 could be replaced with Equation 7 without a
detrimental loss of accuracy.

CMax 596.7 BFIHOST 0.95 0.7706Log10 SAAR 1.8539

0.24

Equation 7: Alternative calculation of

CMax which is not reliant on prior knowledge of

PROPWET
Figure 26: Variation
in PROPWET across
northern Britain
223
(Bayliss, 1999)

-96-

Figure 27: Variation of PROPWET with SAAR (Data for HiFlows catchments)

The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of
data sets and wetlands are subdivided into classes based on wetland type. This data
set represents a useful resource for a national overview of wetlands in addition to
more site specific information. It is recommended that consideration is given to
whether an element of this database could be presented alongside this screening
process. In presenting this data it is important to acknowledge that the analogy of
wetlands as sponges is technically incorrect as presented in Sections 3.3(F) and
3.3(H).

(B)

Areas of floodplain storage potential

This screening process aims to produce a national dataset of areas of additional


floodplain storage potential using floodplain roughening. Due to the site specific
information and detailed modelling required to identify currently effective floodplain
storage areas this assessment will not be able to discriminate between currently
connected and currently disconnected floodplain.
Three methods have been identified for identifying areas of floodplain storage
potential. All are based on the use of the next generation of fluvial flood maps which
are currently in production by Halcrow using ISIS2D for a broad range of flood return
periods and scenarios. All the methods will include hydraulic gradient based on the
analysis of the derived water surface. As a last resort, should the next generation
flood maps not be available it will be necessary to replace the hydraulic gradient with
the valley gradient; this approach would result in a substantial reduction in quality.
Based on the literature findings presented in Section 3.3(L) and the discussion on
backwaters in Appendix C floodplain roughening activities will be more effective in
more gently sloping river reaches hence hydraulic gradient is an important
consideration. All three methods could be prepared on a grid (~25-100m) covering
the entire extent of the mapped flood envelope (for 1 in 200yr event). It may be
desirable to remove urbanised areas and open waterbodies within the identified
floodplain extent prior to undertaking the assessment.

-97-

1. Storage Potential based on valley slope


By rearranging Mannings Equation to the form of a wide channel (where channel
width equals wetted perimeter) and constraining flow to 1cumecs/m (as it is
unknown) and Mannings Roughness to 1 (assuming universally constant), the
Storage Potential can be defined as detailed in Equation 8 based on valley slope ( S ,
m/m).

Qn
y

5
3

1 3
1
1
SP
5
S
S
S6
Equation 8: Proposed equation for generating a measure of indicative Storage
Potential

This is not the preferred method, as it overlooks existing landuse and therefore would
require an alternate method for Section 19.
2. Increased Storage Potential based on hydraulic slope and the current landuse
roughness
By rearranging Mannings Equation to the form of a wide channel (where channel
width equals wetted perimeter) and constraining flow to 0.1cumecs/m (as it is
unknown), setting the maximum floodplain roughness (possible range 0.075-0.2) and
estimating the existing land roughness based on land coverage, the Increased
Storage Potential can be defined as detailed in Equation 9 based on the estimated
hydraulic gradient ( S , m/m).
5
3

Qnexisting 3

Qn
y max
S

n
ISP max
S

0.1nmax 0.1nexisting
nexisting


5
S
S6
5
3

5
3

5
3

Equation 9: Proposed equation for generating a measure of indicative Increased


Storage Potential

This method will enable a consistent screening method to be used in Section 20 and
Section 19. Equation 10 has been presented in case it is desirable to reduce the
complexity of Equation 9.

ISP

nexisting

max

Equation 10: Potential simplification of Equation 9 should it be necessary to reduce


processing time

-98-

Land capability maps (forestry and agriculture) could be used to inform the selection
of maximum roughness values to improve the quality of the estimation of the
capability to store water at a given location, however the existing land cover and land
capability must be of comparable qualities to avoid the introduction of uncertainty.
3. Increased Storage Potential based on hydraulic slope, flow velocity and the
current landuse roughness
This method builds on the previous screening method but introduces the use of flow
velocity which can be based on the data provided by the national flood mapping
study. It is necessary to use the estimated velocity and slope for both the existing
and tested scenario. Equation 11 presents the proposed method.

3
2

3
2
max
3
4

v n
S

3
2

v n

3
2
existing
3

S4

3
2

3
3

v 2
2

ISP 3 nmax nexisting

4
S

Equation 11: Proposed equation for generating a measure of indicative Increased


Storage Potential

This is the preferred method as it will enable a consistent screening method to be


used in Section 20 and Section 19 and it is anticipated to provide a good quality
output.
4. Relating the Storage Potential or Increased Storage Potential to the hydrograph
volume
As presented by Ahilan et al (2009)145 and Potter (2006)147 the ratio of floodplain
storage to flood hydrograph volume is an important factor, it would be beneficial to
capture a measure of this within the screening process. An indicative hydrograph
volume ( HV ) for the nearest watercourse can be estimated by assuming a triangular
hydrograph with a base width two times the estimated time to peak times ( T p )
multiplied by the estimated median annual flood ( QMed ) using Equation 12. Both
these values are nationally available to SEPA within the CEH Flow Grid.

HV T p QMed
Equation 12: Proposed equation for generating a measure of indicative Hydrograph
Volume

Hence the Storage Potential (or Increased Storage Potential) score could be divided
by the indicative Hydrograph Volume to generate a Storage Value ( SV ) score as
detailed in Equation 13.

SV

SP
1

HV ST p QMed

Equation 13: Proposed equation for generating a measure of indicative Storage


Potential

-99-

The use of QMed and T p from the nearest watercourse may invalidate this method
as it would be more applicable to use the QMed and T p values for the location that
experiences the flood, therefore the method is not preferred however if time permits
there may be merit in undertaking additional study in this area.
It would be beneficial to capture knowledge of the morphological condition of the river
channel as an enlarged channel is likely to have a reduced connection to its
floodplain. While it is attractive to attempt to somehow weight the above Storage
Potential score using the morphological condition of the watercourse, this would
place an overemphasis on the quality of the morphological data. It is proposed that
selected elements of the Morphological Pressures Database (channel realignments
and embankments) are laid over the above score so that the user can quickly link the
potential to store water in that area with the morphological condition of the
watercourse.
The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of
data sets and wetlands are subdivided into classes based on wetland type. This data
set represents a useful resource for a national overview of wetlands in addition to
more site specific information. It is recommended that consideration is given to
whether an element of this database could be presented alongside this screening
process to assist in the identification of areas which have previously been classified
as having habitat types which are indicative of good channel-floodplain connectivity.
Consideration should be given to the likely backwater extent generated by floodplain
roughening activities (and other activities which might raise water levels) so that
upstream receptors do not experience an increase in flood risk due to elevated water
levels. It may be feasible to relate the likely extent of a backwater based on the
valley slope however this will require numerical analysis of the backwater extent
equation (Appendix C).

(C)

Opportunities for sub-catchment desynchronisation

Managing flood risk through the desynchronisation of sub-catchments is entirely


speculative and the literature review presented earlier in this document identified no
hard evidence that flood risk management benefits can be achieved. Ultimately,
even if desynchronisation of sub-catchments can achieve tangible flood risk
management benefits the approach will be sensitive to the direction of storm
progression over the wider area.
If included, the aim of this screening process would be to produce a national dataset
of notable confluences. Using CEH Time to Peak data it will be possible to identify
the Time to Peak ( T p ) of the joining watercourses. A score of Time to Peak
Similarity ( TS ) could be generated based on the Equation 14. This scoring was
developed by summarising the main effects of adding standardised hydrographs
(based on gamma distributions). It should be viewed as a judgement based
subjective score rather than a true quantification of potential.

T p ,min
TS
T
p ,max

Equation 14: Proposed method for identifying the Time to Peak Similarity of
catchments

-100-

A similar means could be used to quantify a second score to represent the value of
desynchronisation by replacing T p with the estimated Median Annual Maximum
Flood ( QMed ).
The largest challenge will be displaying the data in a concise and coherent manner.
One potential means may be to present the watercourse network as polylines and to
adjust the colour of the lines upstream of confluences depending on whether it has
the longer of shorter time to peak. At the junction a point of variable colour could be
used to visually present the Time to peak Similarity Score ( TS ) with the size of the
point being based on the Median Annual Maximum Flood Score ( QS ).

TS=0.96

Figure 28: Potential means of displaying sub-catchment desynchronisation


information

By identifying potential desynchronisation opportunities it may assist scheduling flood


risk management works by prioritising Natural Flood Management features which
delay flood progression in one sub-catchment over another. This screening process
may be of interest where engineered flood storage areas are being considered within
a catchment as it can be more efficient to attenuate slower responding catchments.
It may be more appropriate to only use the output of the analysis to avoid flood risk
management interventions that increase the synchronisation of sub-catchments.
There are licensing constraints on the CEH Time to Peak data which may prevent the
output for this screening process being released to bodies external to SEPA. If the
screening process was to be prepared great care would need to be taken to ensuring
that adequate health warnings were provided regarding the risks of attempting to rely
on catchment desynchronisation, even if the data is not intended to be made
generally available.
Given the uncertainties of this method it is recommended that this screening process
is not undertaken until there is additional scientific support in place to demonstrate
that the overall approach of catchment desynchronisation is viable.
It should be noted that the time to outlet grid produced as part of the catchment
specific hydrological assessment (Section 4.4.2(E)) will provide a useful insight into
the timing of runoff generation. However as the generation of the time to outlet grid
requires the catchment outlet to be defined it cannot be generated for a non-site
specific national assessment.

-101-

(D)

Opportunities to remove hydraulic constrictions

The aim of this screening process is to identify potential locations where the removal
of unnatural hydraulic constrictions might achieve a local reduction in flood risk. A
constriction might be caused by a confined river corridor, culvert, bridge or weir. The
restoration of the river reach or the replacement with a higher capacity structure
might offer a significant reduction in flood risk to neighbouring flood receptors.
One means of identifying conveyance improvements could be by overlaying the
latest fluvial flood maps with the locations of hydraulic structures identified within
SEPAs Morphological Pressures Database. However this approach will provide no
further information beyond simply overlaying an Ordnance Survey map with the
mapped flood extent as the Morphological Pressures Database was generated using
OS Master Map data.
A more relevant means of undertaking this screening process would require the
national fluvial flood mapping to be undertaken using an additional alternate scenario
representing a naturalised catchment case. The naturalised scenario would be
schematised by removing all hydraulic structures. This would facilitate a comparison
of the with and without hydraulic structures flood extents or flood depths to identify
areas that would benefit from the removal of hydraulic structures. It is understood
that the Section 19 assessment of the role existing artificial structures play in
reducing flood risk will already be undertaking a very similar assessment, therefore
this screening process should be aligned with the method adopted for Section 19.

(E)

Areas of heightened hydromorphological activity

The aim of this screening process is to identify areas of heightened


geomorphological activity, such as actively eroding gullies, reaches which might
experience future aggradation or areas of potential channel avulsion. This
information can then inform the decision making process to address how already
identified morphology driven flood mechanisms might best managed. It may also be
useful for identifying locations where future morphological issues might occur.
SEPAs Hydromorphology Team is currently undertaking a national sediment budget
modelling study using ST:REAM (Parker, in press)210. The model will classify
reaches which are expected to be eroding, transporting or depositing sediment at a
broadscale. Based on sample data for the Taff catchment in Wales (Figure 29) the
output from this study could form the basis of the morphological activity screening
process.

-102-

Figure 29: Example output


from ST:REAM for the Taff
catchment in Wales (Parker, in
210
press)

The first issue of model results is programmed for summer 2012. It is recommended
that once data is available it is reviewed to ensure that it does meet the aim of this
screening process.

(F)

Areas of estuarine surge attenuation potential

The aim of this screening process is to identify areas where coastal realignment
could be undertaken for the purpose of reducing estuarine surge.
It is proposed that this screening process would be based on identifying land which
could be allowed to flood during a tidal surge event. It will be assumed that the
benefit offered will be directly proportional to the depth of water during a coastal flood
event. That is, the Surge Attenuation Potential ( SAP ) will be the estimated 200yr
water level ( y 200 ) minus the local ground level ( Z ) and the mean sea level ( y ) as
shown in Equation 15.

SAP y 200 y Z
Equation 15: Estimation of the Surge Attenuation Potential

-103-

Figure 30: Sample


output for the Surge
Attenuation Potential

(Data used to generate


figure is not
representative of final
screening process)

The location of known direct coastal defences, as identified by Section 19, will be
presented to facilitate the visual identification of areas which could potentially be
allowed to flood in the future following a managed realignment. There is no
consistent national data set detailing the condition and height of coastal defences
therefore it will not be possible to identify land which is currently protected within the
national screening process.
The SEPA Wetlands Inventory (SEPA, 2012) presents a useful spatial database of all
known wetland habitats in Scotland. It is based on the accumulation of a range of
data sets and wetlands are subdivided into classes based on wetland type. This data
set represents a useful resource for a national overview of wetlands in addition to
more site specific information. It is recommended that consideration is given to
whether an element of this database could be presented alongside this screening
process to assist in the identification of salt habitats which are already likely to
experience coastal inundation.
The location of artificial drainage ditches or channels which may also serve to
accelerate water getting onto tidal marshes will also be presented as part of this
screening. The assessment of the effect these flow conduits might have on the
movement of surge water is anticipated to be very complex requiring the use of a
hydraulic model. For this reason the effect of flow conduits will not be included in the
screening output, but the presence of conduits will be presented.

(G)

Areas of wave energy attenuation potential

The aim of this screening process is to identify areas where saltmarshes or other
measures such as artificial reefs could be used to reduce the wave energy arriving at
the shore. The attenuation of wave energy could be used to supplement existing
direct defences to either extend their lifespan or to increase the standard of
protection that they provide.
It is proposed that this screening process would be based on estimating the Incident
Wave Power ( IWP ) using Equation 16 from knowledge of the significant wave height
( H s ) and the representative wave period ( T ). The best source of wave climate data
identified to date is the Atlas of UK Marine Renewable Energy Resources (DTI,
2008)211 which provides higher resolution than Hydrodynamics and Sea Level Rise
(European Environment Agency, 2005). This data is freely available and details the
mean annual wave power for non-estuarine coastal waters.

-104-

IWP

g 2
0.5H S2T 0.5H S2T
64

Equation 16: Incident Wave Power equation

The Space for Attenuation ( SA ) of wave energy could be estimated based on the
intertidal slope which, to avoid the problematic calculation of local tide levels, could
be taken as the horizontal distance between Mean High Water Springs and Mean
High Water Springs ( x ).

SA x MHWS MLWS
Equation 17: Space for Wave Attenuation

Subsequently the Capability of Wave Attenuation ( CWA ) could be estimated for the
entire coastline by multiplying the Incident Wave Power by the available Space for
Attenuation as shown in Equation 18.

CWA IWP SA
Equation 18: Capability of Wave Attenuation

This is the preferred means of undertaking the wave attenuation screening process
due to its simplicity.
Figure 31: Annual mean
significant wave power
(kW/m of wave crest). Data
from the Atlas of UK Marine
Renewable Energy
Resources, Commissioned
by DTI and produced by
ABPmer, The Met Office and
Proudman Oceanographc
Laboratory (2008,
www.renewables-atlas.info/)

-105-

A more detailed means of calculating the wave run-up or defence overtopping could
be undertaken using the guidance contained within the CIRIA Rock Manual (CIRIA,
2007)212. While potentially offering more accurate output the methods also requires
additional data. It is felt that an attempt to quantify overtopping would be misguided
and would give the impression of a greater level of knowledge than is actually
available. Therefore it is proposed that the estimation of wave run-up would be a
more valuable descriptor to be used for the screening process. According to the
C683 the wave run-up can be estimated to within 2% by Equation 19 with knowledge
of the significant wave height ( H m 0 ), Iribarren Number using the energy wave period
( m 1.0 ) and a friction factor ( f ).

Ru 2%

2.5H m 0 f

m 1,0

Equation 19: Estimation of wave run-up

The Iribarren Number can be estimated using the shore slope ( ) and the fictitious
wave steepness ( S 0 ) as shown in Equation 20.

m 1,0

tan
S0

Equation 20: Estimation of the Iribarren Number

The fictitious wave steepness can be calculated using Equation 21 based on the
significant wave height ( H m 0 ) and the energy wave period ( Tm,1, 0 ).

S0

2H m 0
gTm21, 0

Equation 21: Estimation of the fictitious wave steepness

Shore type

Friction factor ( f )

Concrete, grass or asphalt

1.0

Pitched stone

0.80-0.95

Single layer of armour stone on an


impermeable base

0.7

Double layer of armour stone on an


impermeable base

0.55

Table 7: Wave run-up friction factors (CIRIA, 2007)

The calculation of wave run-up is a data intensive and complex process. The
additional complexity associated with the method will not bring a notable
improvement in the quality of the screening process and therefore it is not preferred
over the Capability of Wave Attenuation score presented in Equation 18.
It should be acknowledged that the Flood Risk Management (Scotland) Act makes no
reference to erosion therefore the inclusion of a measure of potential coastal erosion
-106-

is not appropriate unless it is considered within the context of causing damage to


flood defence assets. The calculation of erosion is a very complex process requiring
detailed site specific information. It is considered that the Incident Wave Power in
combination with site reconnaissance, local knowledge, geological maps and
Ordnance Survey mapping will provide an appropriate overview of the erosion
potential.

-107-

4.3.3 Summary
The methods outlined in Table 8 are recommended as the most valid for undertaking
the Natural Flood Management screening process.
Preferred Method

Data requirements

Areas of high runoff generation:

BFIHOST vector 250k and HOST key

Using the hybrid of the EA screening method and


the modified QMed by descriptors

SAAR 1km grid

q 8.3062 0.1536

1000
SAAR

0.0460

BFIHOST 2

slope _ score landuse _ score

LAI via Land Cover Map


DTM NextMap adequate
SEPA Wetland Inventory

Areas of floodplain storage potential:


Using an estimate of the Increased Storage
Potential

S hydraulic gradient from national flood


maps
v - flow velocity form national flood maps
nexisting - via Land Cover Map

3
2

3
3

v 2
2

ISP 3 nmax nexisting

S4

n max - via land capability maps


SEPA Wetland Inventory

Opportunities to remove hydraulic constrictions:

Data requirements to be confirmed

Using S19 methodology


Areas of heightened hydromorphological activity:
Using the results of the sediment budget
modelling currently being undertaken by SEPAs
hydromorphology team
Areas of estuarine surge attenuation potential:
Using the estimated flooded depth during the
200yr flood event to generate a score of
attenuation value.

Output from the ST:REAM sediment


budget model

DTM LiDAR much preferable, NextMap


would bring significant uncertainty
Coastal defence locations output of
S19
Estimated 200yr coastal water levels
latest coastal flood maps
SEPA Wetland Inventory

Areas of wave energy attenuation potential: (to be Mean high and low water springs OS
confirmed)
Master Map
Wave Power and the Space for Wave Attenuation

Wave climate data Atlas of UK Marine


Renewable Energy Resources (DTI,
ABPmer, The Met Office & POL, 2008)

Table 8: Summary of the preferred Natural Flood Management screening processes


and the data requirements

Table 9 outlines the screening method that will not be undertaken due to low
scientific uncertainty in the underlying Natural Flood Management method.

-108-

Method

Data requirements

Opportunities for sub-catchment


desynchronisation:

CEH time to peak points for the entire


river network

Using a Time to Peak Similarity score

T p ,min
TS
T
p ,max

Table 9: Summary of the screening processes which are not recommended without
further scientific support or data

-109-

4.4 Quantifying the flood risk management benefits of Natural


Flood Management measures
4.4.1 Introduction
In order to quantify the effects of Natural Flood Management a step change in the
hydrological assessment techniques which are employed is needed. It should be
noted that NFM measures have a level of uncertainty associated and the effect with
diffuse soft interventions will never be as certain as that for localised hard engineered
interventions. However, without knowledge of the anticipated effect quoted in the
context of its uncertainty, it will not impossible to make an informed decision.
A summary of the capability of existing tools to quantify the benefits of Natural Flood
Management measures is summarised in Table 10. The table is based on the
findings of the literature review presented in Section 3 and it recommends how each
identified measure can be included within the quantitative assessment of Natural
Flood Management which is to be undertaken as part of the appraisal process.
Natural Flood
Management measure

Capability of existing
assessment techniques

Recommendation actions to facilitate


quantification of impact on flood risk

Conifer woodland

No existing tools
Weak literature base in place
No existing tools
Very weak literature base in place
No existing tools
Very weak literature base in place

Requires new hydrological assessment


methodology
Requires new hydrological assessment
methodology
Requires new hydrological assessment
methodology

Methods for small catchments


Weak literature base in place
No existing tools
Very weak literature base in place
Only research hydrological and
groundwater models
Methods for small catchments
Weak literature base in place

Requires new hydrological assessment


methodology
Method unclear, local assessment team to
develop method if required
Local assessment team to select
appropriate existing method
Requires new hydrological assessment
methodology

No existing tools
Weak literature base in place
No existing tools
Weak literature base in place
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques

Method unclear, local assessment team to


develop method if required
Method unclear, local assessment team to
develop method if required
Local assessment team to select
appropriate existing method

Broadleaf woodland
Good muirburn practice
(compliance with muirburn
code)
Reducing grazing pressure
on pasture
Creation of cross slope tree
shelter belts
Creation / restoration of nonfloodplain wetlands
Reducing soil compaction in
arable areas, improving soil
texture and reducing bare
earth in wetter seasons
Changing agricultural field
drainage
Upland drain blocking
Floodplain reconnection

Creation of washlands

Creation of constructed farm


wetlands or ponds

-110-

Local assessment team to select


appropriate existing method

Local assessment team to select


appropriate existing method

Natural Flood
Management measure

Capability of existing
assessment techniques

Recommendation actions to facilitate


quantification of impact on flood risk

Afforestation of floodplains

Localised effects can be estimated


using existing hydraulic models
when used in conjunction with
existing hydrological techniques
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques
No existing tools specific for gullies
Very weak literature base
Localised effects can be estimated
using existing hydraulic models
when used in conjunction with
existing hydrological techniques
Effects can be estimated using
existing hydraulic models when
used in conjunction with existing
hydrological techniques
Current approaches rely on the
experience of a geomorphologist.
Some modelling approaches but
these are subject to large
uncertainty.

Local assessment team to select


appropriate existing method

Use of SUDS

Existing assessment techniques are


appropriate

Local assessment team to select


appropriate existing method

Using high water demand


vegetation to reduce
groundwater levels

Existing groundwater modelling


techniques are appropriate

Local assessment team to select


appropriate existing method

Beach management (beach


recharge schemes and
shingle management)
Artificial/ biogenic reefs and
detached breakwaters
Sand dune restoration (e.g.
dune fencing and thatching,
marram grass planting).
Standline and beach
management techniques
Creation/ restoration of
intertidal area including
mudflats and saltmarsh
(Wave energy dissipation)
Creation/ restoration of
intertidal area including
mudflats and saltmarsh
(Surge attenuation)
Managed realignment of
assets or no active
intervention along shores
subject to frequent flooding
and coastal erosion

Existing techniques are appropriate

Local assessment team to select


appropriate existing method

Effects can be assessed using


existing techniques
Existing assessment techniques are
appropriate

Local assessment team to select


appropriate existing method
Local assessment team to select
appropriate existing method

Current approaches rely on the


experience of a geomorphologist.
Site specific wave run-up modelling
using currently available software
Models available to provide insight
into hydromorphology
Attenuation effects can be modelled
using existing hydrodynamic
models

Local assessment team to select


appropriate existing method
Local assessment team to select
appropriate existing method

Existing techniques are appropriate

Local assessment team to select


appropriate existing method

Reach restoration (planform


restoration)

Creation of riparian woodland

Gully woodland planting


Placed large woody debris
and boulders
Reach restoration removal
of hydraulic constrictions

Managing channel
instabilities

-111-

Requires new hydrological assessment


methodology
Supplemented by local assessment using
existing tools
Requires new hydrological assessment
methodology
Supplemented by
Local assessment using existing tools
Method unclear, local assessment team to
develop method if required
Requires new hydrological assessment
methodology
Supplemented by
Local assessment using existing tools
Local assessment team to select
appropriate existing method

Local assessment team to select


appropriate existing method

Local assessment team to select


appropriate existing method

Natural Flood
Management measure

Capability of existing
assessment techniques

Recommendation actions to facilitate


quantification of impact on flood risk

Regulated tidal exchange

Effects can be modelled using


existing hydrodynamic models
Effects can be modelled using
existing hydrodynamic models

Local assessment team to select


appropriate existing method
Local assessment team to select
appropriate existing method

Removal of artificial channels

Table 10: Summary of assessment methods for Natural Flood Management measures
and the identification of areas where there are currently gaps in assessment capability

In summary there are adequate assessment techniques in place for the quantification
of coastal, groundwater and urban Natural Flood Management measures. However
there are the following caveats to this conclusion:

It is necessary to give due consideration to the location of coastal Natural


Flood Management interventions. Current national coastal flood mapping
techniques rely on projecting estimated extreme waterlevels to map flood
extents. The quantification of the effect of Natural Flood Management
interventions requires full hydrodynamic modelling of the estuarine system
which may raise computation barriers.

Where coastal flooding is heavily interrelated with sediment movement it is


necessary to give due consideration to how modifications will impact on
sediment budgets. This sediment modelling may push the existing modelling
and it is not apparent if adequate data is available.

Where groundwater flooding is influenced by complex sub-surface drift and


solid geology it may result in there being inadequate data to assess
groundwater flooding mechanisms. Assessing complex groundwater systems
will remain out of reach for the foreseeable future. As identified in the
Eddleston Catchment (Werritty et al, 2010)213 groundwater can play a
significant role in contributing to fluvial flood flows. It is not clear whether this
is an isolated case or a common occurrence. The suitability of existing
ground water hazard maps is also unclear, however this is a different issue to
the suitability of tools to assess the impact of Natural Flood Management
measures on identified groundwater flooding.

Unfortunately there is a large gap in the assessment methodologies that can be used
for quantifying the hydrological benefits of fluvial Natural Flood Management. To
allow fluvial measures to be compared against more traditional engineered it will be
necessary to develop a methodology which can close this gap. A proposed
hydrological assessment methodology to do this is presented in Section 4.4.2.

-112-

4.4.2 Specification of proposed S20 hydrological assessment


tool
Following the guidance of DEFRA FD2114 study (OConnell et al., 2004)190 to assess
the hydrological benefits of diffuse land use changes it is necessary to adopt a
modelling approach which is:
1. Distributed;
2. Capable of running continuous simulation;
3. Partly or wholly physically based so that landscape, soil and vegetation can
be represented; and
4. Detailed modelling of surface flow so that effects can be tracked downstream.
No existing models meet these specific requirements with the exception of research
level wholly physically based hydrological models. The Land Management CFMP
Tool (JBA et al, 2008)192 goes part way to addressing these points but only by
enabling an assessment of runoff generation for a limited set of predefined
catchments in England and Wales and by the reports own admission the method
does not quantify changes in runoff timing. The research physical models are data
intensive, computationally expensive and until recently normally developed to assess
annual yields, consequently they do not lend themselves well to being used by the
flood risk management industry to assess flood risk. It is therefore proposed that a
semi-physical distributed approach is adopted based on principles taken from the
Grid-to-Grid, OVERFLOW, Allan Water Natural Flood Management method and the
Revitalised Flood Hydrograph Losses Model. An overview of the proposed model is
presented pictorially in Figure 32 and schematically in Figure 33. By configuring the
model in this way it will be consistent with current hydrological assessment
procedures within Britain while retaining a model architecture that can meet all of the
requirements set out by DEFRA FD2114 (OConnell et al., 2004)190.

-113-

Rain storms
assumed to be
standard FSR storm
profile and
instantaneously
cover the entire
catchment

No return of water to the


floodplain, all flow in bank
at normal depth

All rainfall directed at


canopy store until
exhausted, no
evaporation during
event

Requirement for continuous


simulation removed using ReFH
method to calculate antecedent
moisture

Runoff velocity (shallow


and surface) is constant
with time and flood
return period. Flow
velocity based on slope,
land cover, soil type and
peak flow

Constant baseflow throughout


event

Figure 32: Pictorial overview of the proposed Natural Flood Management assessment
model

-114-

Rainfall
Evapotranspiration

FEH DDF

MORECS
Interception
Land use

Antecedent
moisture
Based on modified
ReFH estimation of
Cini by modifying
PROPWET with
land use.

Based on MORECS
leaf area index

Drainage network

Flow routing model


Runoff generation
Based on the ReFH
losses model,
similarities to that
used in G2G

Based on land flow


and channel flow
similar to Allan Water,
OVERFLOW and G2G
to estimate time to
outlet for each runoff
generation cell

Baseflow model

Soil type

Based on ReFH
estimate of the initial
baseflow (no variation
during event)

HOST (and
degraded HOST)

OUTFLOW

Figure 33: Schematic representation of the proposed Natural Flood Management


assessment model

It is proposed that the distributed model is initially developed to simulate a single


event, however in a similar fashion to the Revitalised Flood Hydrograph method a
continuous simulation model could be developed within the confines of the proposed
configuration.

(A)

Rainfall estimation

The design rainfall for storms will be derived in a similar manner to the ReFH and
FSR/FEH rainfall-runoff models using the Depth Duration Frequency (DDF) model as
presented by Faulkner (1999)214. The Area Reduction Factor (ARF) used to transfer
a design point rainfall depth to a catchment representative depth would also be a
logical inclusion. It is understood that SEPA already has access to a GIS toolbox
and parameter set to facilitate point rainfall estimates for varying return periods and
storm durations. It is anticipated that this toolbox can at least be included in the build
of the hydrological model to be used internally by SEPA. Outside SEPA, other
organisations may need to secure appropriate licences or revert to a manual query of
the FEH CD-Rom.
The ReFH introduced a Seasonal Correction Factor (SCF) to improve the models
calibration to the statistical analysis of flow records. It should be noted that the SCF
is a calibration coefficient which is specific to the model, following testing of the

-115-

model it may become apparent that a similar calibration coefficient may be required.
It is not recommended that a SCF is included within the initial models construction.
It should be noted that the DDF, ARF and Design Storm Profiles are generalisations
of complex systems. Kjeldsen (2007)50 provides an overview of some of the
limitations of these generalisations, and it should be remembered throughout that
these methods provide a means of making informed estimates, not a quantification of
the actual amount, and that these estimates are commonly accepted in the absence
of alternatives.
Following the derivation of the design storm rainfall depth grid (and after subtraction
of the canopy interception as presented in (B)) it is proposed to use the same 75%
winter and 50% summer rainfall profiles originally used in the FSR/FEH rainfall-runoff
method. Both are symmetric single peak profiles, and their shape does not vary
between storms (spatially, event magnitude or duration). As the tool will be applied
to predominately rural catchments (URBEXT <0.125), then the inclusion of the winter
profile will be most important within the tool (Institute of Hydrology, 1979)215.
Figure 34: Design rainfall
profiles for summer and
winter, drawn as normalised
hyetographs (Kjeldsen,
50
2007)

It is not proposed to represent the movement of a storm over the catchment as to do


so would make the analysis sensitive to the assumptions around the direction and
speed of travel of the storm. However the spatial weighting of rainfall could be
achieved using weights based on SAAR or similar so that the increased precipitation
in upland areas is taken account of.
The FEH provides a regression based equation to provide an initial estimate of
critical storm duration. This equation could be applied to give an initial estimation of
the catchments critical storm duration (based on SAAR and Tp ) however this
estimate of critical storm duration should not be relied upon when testing one Natural
Flood Management measure against another. It is essential that the critical storm
duration for the catchment in question is optimised via testing storms of different
duration so that the storm length that produces the highest flow can be identified. If
this optimisation process is not undertaken then it may be the case that an identified
change in peak flow is incorrect. This incorrect reporting is most likely to occur in
cases where the time to outlet grid is modified to quantify the effect of the measure.

-116-

SAAR
D T p 1

1000

Equation 22: Initial estimate of the critical storm duration

(B)

Interception model

As presented in Section 3.3, the literature indicates that the interception of rainfall by
the vegetation canopy is a process that reduces the runoff generation and
subsequent magnitude of flood flows. It is proposed that a basic canopy interception
model is included within the model. A simple model based on the Leaf Area Index
( LAI ) is used within the MORECS model (Hough and Jones, 1997)216. The
MORECS system places an upper threshold on daily interception ( I ) as 0.2mm per
unit of leaf area index as defined by Equation 23.

I 0.2LAI
Equation 23: Maximum daily canopy interception

A more complex means of estimating the canopy interception capacity is to use the
method proposed by von Hoyningen-Huene (1981)105 as presented in Equation 24.

I Max 0.935 0.498( LAI ) 0.00575( LAI ) 2


Equation 24: Estimation of maximum canopy storage based on the LAI (Hoyningen105
Huene, 1981)

Land cover

Summer LAI

Winter LAI

Grass (General)

Improved grassland

4*

1.5*

Semi improved grassland

5*

2*

Cereals

0 (Bare earth)
0.5 (Winter cover)

Oil-seed rape

0 (Bare earth)
0.5 (Winter cover)

Potatoes

0 (Bare earth)

Sugar beet

0 (Bare earth)

Deciduous trees

0 (Bare earth)

Conifers

Orchards

Upland (unimproved
grassland)

3.5

2*

Upland (heather)

3.5

3.5*

Bare soil, rock, water &


urban

Table 11: Leaf Area Index values as used in MORECS 2.0 (Hough and Jones, 1997)
(* Are inferred from the available data)

-117-

216

A comparison of the two methods for estimating the canopy interception capacity is
presented in Figure 35. This shows that the Hoyningen-Huene (1981)105 method
produces considerably higher estimates of the canopy interception capacity. It is
noted however that Hoyningen-Huene (1981)105 describes this as the maximum
capacity, whereas the threshold used within MORECS (Hough and Jones, 1997)217
was intended to be representative of a typical daily maximum. It is considered
probable that the true answer for the interception capacity during storm conditions
lies somewhere between the two methods. As the MORECS method was used by
the Met Office to represent typical conditions found in Britain after a robust model
development process, and because it produces a more conservative result, this
method is recommended for initial adoption within the method. In adopting this
assumption it should be noted that work by Calder (1990)94 suggests that canopy
interception by coniferous forest (LAI=6 to 10) can be up to 6 to 7mm/day. While the
variation between the two methods to the face of it appear very notable, it should be
remembered that the total interception during a flood event will be relatively small,
less than 10% of the total event rainfall (Nisbet, 2008)218 tied to the consideration that
flooding tends to occur during leafless months. In addition, this interception will occur
at the start of the event leading to the suggestion that canopy interception will have
little effect on the flood peak.
6

Interception capacity (mm)

0
0

10

Leaf area index

MORECS 2.0 (Hough and Jones, 1997)

Hoyningen-Huene (1981)

Figure 35: A comparison of the variation in canopy interception capacity with leaf area
index (LAI) using MORECS (Hough and Jones, 1997) and Hoyningen-Huene (1981)

For simplicity it is proposed that the default maximum daily canopy interception will
be taken to equal the maximum interception for a single event (as calculated using
MORECS). It is recommended that the Hoyningen-Huene (1981)105 method is
included as a user selectable alternative to permit upper bound sensitivity testing,
however the MORECS method should be the default.
The calculated interception will be subtracted from the gross rainfall at the start of the
hyetograph. Once the accumulative gross rainfall exceeds the canopy interception
capacity no further rainfall interception will occur. The intercepted rainfall will be lost
from the model with no further consideration of evaporation etc from the interception
store. It is accepted that this approach may underestimate the effect of canopy
interception at the hyetograph peak and over estimate the effect at the start of the
-118-

event. However it will be computationally efficient and proportionate to a process


which is likely to have only a very marginal impact on flood generation, with a
maximum interception of around 1.2mm per event for LAI 6. The hydrologist should
review the representation of the canopy interception model within the context of the
catchment being modelled, for low rainfall volume (short duration frequent events) it
may be appropriate to consider a more detailed canopy representation.

(C)

Runoff generation (losses model)

The role of the runoff generation or loss model is to estimate the proportion of rainfall
that will be converted to direct runoff, with the remainder being passed to soil
moisture. It is proposed that it will be based on that used in the ReFH model
(Kjeldsen, 2007)50 which is in turn based on the Probability Distributed Model (PDM)
developed by Moore (1985, 2007)219220. PDM is widely used for hydrological
applications across the UK, including in SEPAs Flood Early Warning System as well
as investigating the impact of climate change.
Conceptually, the PDM assumes the catchment consists of a number of individual
storage elements, each of a soil moisture capacity ( C ) arising from a statistical
distribution. Assuming a uniform distribution of soil moisture capacities, if the storage
elements are arranged in order from the highest ( CMax ) down to zero, the resulting
PDM distribution of soil moisture capacity is shown in Figure 36. It is further
assumed that storage elements interact so that soil moisture is constant for all
elements of capacity greater than ( Ct ), and full for those which are smaller. Incident
rainfall increases the depth of water in each storage element and when rainfall
exceeds the storage capacity, direct runoff is generated. For the short duration of
storms being considered, the effects of evapotranspiration and drainage out of the
soils will not be included.

Figure 36: Equal water content ( C ) across stores of different capacity (Kjeldsen,
50
2007)

Therefore, a pulse of rain ( P ), on the soil gives 100% runoff from the stores which
are full and increases the moisture content in all other areas. The excess amount of
rainfall converted into direct runoff ( q ) can be estimated using Equation 25.

-119-

PC
P2

CMax 2CMax

( C CMax )
( C CMax )

qP

Equation 25: Calculation of runoff using PDM

By dividing the rainfall series into timesteps the runoff at any timestep ( t ) can be
calculated using Equation 26.

Pt Ct 1
Pt 2
qt

CMax 2CMax

( Ct CMax )

qt Pt

( Ct CMax )

Equation 26: Calculation of runoff using PDM for a timestep

And from continuity the stored soil moisture can be calculated using Equation 27.

Ct Ct 1 Pt
Equation 27: Calculation of soil moisture based on previous soil moisture and incident
rainfall

It should be noted that ( Ct ) is not a measure of the catchments average soil


moisture content, rather it records the moisture content of the PDM element which
has a storage capacity equal to ( CMax ). A number of examples demonstrating how
the method works are provided in Kjeldsen (2007)50, these examples can be
recreated in a spreadsheet to ensure understanding of the fundamentals of PDM.
The next step is to consider the application of the PDM loss model on a grid of subcatchments thus the lumped model can then be distributed over a catchment.
Equation 26 and Equation 27 would then be applied locally for each grid cell so that a
local runoff specific to the conditions at that location, can be estimated.
The maximum soil moisture capacity ( CMax ) can be calculated using the estimator
equation presented by Kjeldsen (2007)50 based on the Baseflow Index ( BFIHOST )
and the proportion of time the catchment is wet ( PROPWET ) using Equation 28.

CMax 596.7 BFIHOST 0.95 PROPWET 0.24


Equation 28: Estimation of

CMax used by the ReFH

While BFIHOST can be calculated from vector HOST data to approaching field scale
resolutions via the method laid out in IH126 (Boorman et al., 1995)137 generating a
grid of PROPWET values to a similar resolution is more problematic. Section (D)
presents a number of potential means of improving the spatial resolution for
estimates of PROPWET .
It is envisaged that the runoff generation model will be constructed on a 25-100m
grid, it is important that this grid is representative of the resolution of the land cover
and soil data. It is not thought that there is a strict requirement for the runoff
generation grid to be the same as that used for the flow routing, however a variation
in grids could lead to unforeseen errors.
A method for modifying the BFIHOST to account for the effect of soil degradation
was proposed by Packman et al (2004)48. The rationale for the method is that
degraded soils suffer from structural degradation, compaction of the topsoil and

-120-

upper subsoil and seasonal capping. Together these effects reduce the effect soil
moisture capacity leading to increased runoff generation. A central theme of the
method is that the hydrogeological component remains constant while the soil
component is degraded by a class. Kjeldsen (2007)50 speculates that the method
proposed by Packman et al (2004)48 is a potential means of representing the effect of
soil degradation within the ReFH model which is the underlying basis of this loss
model. The analogue HOST classes proposed by Packman et al (2004)48 are
presented in Table 12.
HOST Class

Analogue HOST
for degraded soils

Un-degraded BFI

Degraded BFI

1.00

0.90

1.00

0.90

7?

0.90

0.79

6?

0.79

0.65

7?

0.9

0.79

0.64

0.56

7?

0.79

8 level ground

0.56

Na level ground

9 level ground

0.73

Na level ground

10

10 level ground

0.52

Na level ground

11

11 level ground

0.93

Na level ground

12

12

0.17

NA

13

3?

1.00

0.90

14

24

0.38

0.31

15

15?

0.38

16

18?

0.78

0.60

17

18

0.61

0.52

18

20

0.52

0.40

19

22

0.47

0.32

20

20?

0.52

21

23

0.34

0.22

22

27

0.32

0.26

23

25

0.22

0.17

24

25

0.31

0.21

25

25?

0.17

26

26

0.24

NA

27

27?

0.26

NA

28

28? - Upland peat

0.58

NA

29

28? - Upland peat

0.23

NA

Table 12: Effect of soil degradation via the use of analogue HOST soil types (Packman
48
et al, 2004)

There is the option to reduce the maximum soil moisture storage capacity ( CMax ) to
account for variations in topographic slope following the method presented by Bell
and Moore (1998)221 and later used in the G2G model Bell et al. (2007)222. The
reduction in maximum soil moisture capacity ( CMax ) is on the basis that steep
topographic slopes limit soil moisture to shallow levels, whereas flat areas can permit
the build up of deep soil moisture stores. It is unclear how much allowance
topographic slope is inbuilt within the method of estimating CMax based on HOST
-121-

data and therefore it is recommended that the variation in slope is not included until
after an initial modelling testing period. Equation 29 shows the potential basis of
reducing CMax based on the ratio of average grid cell topographic slope ( g ) over the
gradient at which the soil moisture storage capacity would reach zero ( g Max ). The
maximum gradient ( g Max ) varies between soil types and is related to the depth of the
soil.

g
C Max
C Max,rev 1
g
max

Equation 29: Method of reducing the soil moisture capacity based on local
221
topographic slope (Bell and Moore, 1998)

The calculation of the initial soil moisture content C ini is presented in Section 4.4.3(D)
below.

(D)

Antecedent moisture model

Following the ReFH the initial soil moisture content ( C ini ) for a winter storm can be
estimated using Equation 30.

C ini

C Max
1.20 1.70BFIHOST 0.82PROPWET
2

Equation 30: Estimation of the initial soil moisture content

To increase functionality so that degraded soils can be assessed it is proposed that


the methodology of analogue HOST soils proposed by Packman et al. (2004)48 is
adopted as presented in Section 4.4.2(C) above is used to estimate degraded
BFIHOST where appropriate.
The proportion of time soils are wet ( PROPWET ) was initially estimated by Bayliss
(1999)223 and the data made available to practicing hydrologists via the FEH CD.
This 1km grid dataset estimated the time the soil moisture deficit was less than 6mm
using MORECS data for the period 1961 to 1990. The MORECS data was produced
on a 40km grid for a restricted range of land cover types therefore it was necessary
to use weightings based on estimated land cover within the 1km grid square to
generate an estimate which is representative of local conditions. Some additional
work is required to identify the best means of including PROPWET within the final
tool:
1. The most straightforward is to use a constant PROPWET value for the entire
catchment based on a query of the FEH CD. As a minor improvement
PROPWET could be sampled for a number of subcatchments. This
approach does not provide a means of guiding the modification of
PROPWET between land use change scenarios and therefore would not
facilitate the assessment of land use on the antecedent moisture condition.
2. Develop a regression based estimate of PROPWET by investigating the link
between SAAR, BFIHOST, MORECS 40km annual potential evaporation
(PE), leaf area index (LAI) and the bulk surface (canopy) resistance ( rsc ) for
land uses in a number of known catchments. In this way the impact of land
use changes on the antecedent moisture condition could be approximated at
a higher resolution.

-122-

A preliminary scoping investigation of the potential of this regression based


method was undertaken as part of this study. The analysis used 43 randomly
selected Scottish catchments within the CEH HiFlows Database. The
LAI Summer and rsc were defined within the catchments based on the land use
(LCM2000) and the reported values in Hough and Jones (1997)216. A
relationship for PROPWET based on catchment average SAAR, PE (gauge
location taken as representative, catchment average would be better),
LAI Summer and rsc was then developed via regression. The catchments were
delineated using a coarse 500m grid resampled from the Ordnance Survey
Panorama DTM. In a number of instances the error in the delineated
catchment area was greater than 10%, these were excluded from the final
selection and only 43 of the selected catchment met the above criteria. The
identified regression based relationship is shown in Equation 31.

PROPWET 0.5365Log10 SAAR PE LAI 4.70 r 0.79 2.003E 5 0.998


R 2 0.828, n 43
Equation 31: Provisional estimator equation for PROPWET which will permit
the variation of the antecedent moisture conditions and soil moisture capacity

Value

Min

25%

Median

75%

Max

SAAR

762

1016

1334

1649

2848

PE

428.5

474.9

489.2

511.0

556.6

LAI Summer

1.917

3.656

3.943

4.330

4.911

rsc

60.699

69.521

70.769

74.607

94.006

BFIHOST

0.292

0.369

0.411

0.469

0.609

PROPWET (FEH)

0.400

0.565

0.630

0.715

0.850

Table 13: The range and quartiles for variables used in the PROPWET
descriptor regression analysis

The r-squared value for the regression based estimate of PROPWET is


0.828, this is a promising indication that a more detailed analysis might allow
the identification of a reliable method. Plots showing the performance of the
identified relationship are shown in Figure 37 and Figure 38. It is
recommended that a more detailed and comprehensive regression analysis
using a larger number of catchments, including catchments elsewhere in the
British Isles, is conducted prior to the adoption of this methodology. It should
also be noted that Equation 31 will not facilitate the estimate of PROPWET
when the Log term is less than zero, that is, when the evapotranspiration
exceeds rainfall. Similarly, the equation suggests questionable PROPWET
values (less than 0.25) catchments for situations where the
evapotranspiration approaches the rainfall, hence it is probable that an
alternate relationship or limit will need to be applied to control PROPWET in
these situations. This is probably a symptom of the absence of very high LAI
and Potential Evapotranspiration catchments being considered within the
regression analysis.

-123-

As discussed in Section 4.4.5, the analysis should be improved by extending


the catchments considered to include those with higher Potential
Evapotranspiration, lower rainfall, lower PROPWET (FEH) and a wider range
of LAI. It would also be worthwhile giving consideration to the winter LAI
within the regression analysis due to the marked seasonal evapotranspiration
variation for deciduous woodland and crops such as potatoes.
PROPWET Regression
Selected HiFlows Stations
0.90
0.80

Regression PROPWET

0.70
0.60
0.50

y = x - 5E-09
R2 = 0.8276

0.40
0.30
0.20
0.10
0.00
0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

FEH PROPWET

Figure 37: Plot comparing the regression based estimate of PROPWET with the
FEH CD PROPWET

-124-

Figure 38: Plot comparing the regression based estimate of PROPWET with the
FEH CD PROPWET

3. The most detailed means would be to gain access to the original Soil
Moisture Deficit Data generated using MORECS for the 1961 to 1990 period
for all land cover types and to reanalyse it to produce estimates of
PROPWET for each land cover type on the 40km grid. These estimates
could then be weighted based on the scenarios land cover to produce higher
resolution estimates of PROPWET as part of the hydrological modelling
process.
4. Consideration has been given to the use of soil moisture data estimated using
MOSES (Hough, 2003)224 to in turn estimate PROPWET, however it was
decided that the different output of MOSES would result in a non-comparable
PROPWET. This would lead to the need for alternative Cini and Cmax
estimator equations needing to be developed.

-125-

5. An alternative means may to be use an alternate soil moisture accounting


model such as that within the diffuse pollution screening tool, the flood
warning G2G model or a model such as HYLUC to produce more localised
land use specific estimates of the soil moisture deficit and subsequently
PROPWET . However the transfer of soil moisture levels from one model to
another is likely to introduce notable uncertainty.
Until consultation is undertaken with the holders of data it is unclear what means is
most appropriate to progress with however at this stage the regression based
approach does appear to be most promising due to the relative speed it can be
developed and the flexibility to investigate any land cover type provided SAAR, PE,
LAI and rsc are known.
As a potential later extension, an estimate of ( C ini ) for observed events could be
obtained in a similar fashion to the method used within the ReFH. The method
estimates ( C ini ) by modelling of the antecedent soil moisture content using a
separate soil moisture accounting model similar to that used within the continuous
simulation ReFH. The continuous simulation of soil moisture should be viewed as a
low priority.

(E)

Flow routing model

The ReFH model uses an instantaneous unit hydrograph approach which is founded
on the estimation of the catchments time to peak. For observed catchments the time
to peak can be estimated by inspection of observed flow and rainfall records,
however for ungauged catchments it is necessary to use a regression based
estimate of the time to peak. Unfortunately neither of these approaches allow for the
effect of changes in the catchments drainage network (i.e. increase in length by
meandering or increase in roughness by riparian planting) or the variation in time of
contribution for different regions of the catchment to the downstream hydrograph.
It is understood that G2G undertakes flow routing by estimating the velocity and
distance water takes as it passes through the catchment. From these estimates an
indication of the time of travel can be inferred. It is understood the model uses
constant channel and land flow velocities which can then be varied in order to
calibrate the timing of the runoff peak within the hydrological model where observed
data is available. The use of spatially constant velocities will prevent a meaningful
estimate of the impact of changes to the catchments drainage network (i.e. increase
in length by meandering or increase in roughness by riparian planting).
OVERFLOW and the Allan Water Natural Flood Management study method both
estimate the velocity of overland flow to the watercourses based on a hydraulic
calculation, hence the velocity is dependent on topographic slope, runoff rate and
ground cover roughness. For in channel flows OVERFLOW uses a more advanced
diffuse wave hydraulic routine which allows for flood water to be spilled onto the
floodplain, while the Allan Water Natural Flood Management method uses a steady
state kinematic wave routine. It should be noted that the simplistic kinematic wave
theory is built on the assumption of all flow being at normal depth and it therefore
prohibits the assessment of backwater effects. Both methods assume that the
channel dimensions are based on approximated dimensions based on an index
flood. The OVERFLOW model assumes that watercourses form when the flow
exceeds a threshold (0.6cumecs) whereas the Allan Water Natural Flood
Management method uses mapped watercourse network or a threshold catchment
area. The flow routing engine used by OVERFLOW is computationally expensive,
and the one employed by the Allan Water Natural Flood Management study should
be viewed as halfway between G2G and OVERFLOW.
-126-

A summary of flow routing models and methods is presented in Table 14.


Example
routing model

Land flow
velocity

Channel flow
velocity

Computational
expense

Suitability

ReFH

NA
regression
based

NA
regression
based

Very low

Gauged catchments
which are not subject
to changes

G2G

Spatially and
temporally
constant
velocity

Spatially and
temporally
constant
velocity (1D)

Low

Gauged catchments
which are not subject
to land use changes

Allan Water
Natural Flood
Management
study

Steady state
kinematic
wave (normal
depth)

Steady state
kinematic wave
(normal depth)
(1D)

Medium

Ungauged catchments
with (limited) in channel
and diffuse land use
changes

OVERFLOW

Time varying
velocity based
on kinematic
wave
hydraulics

Time varying
velocity based
on diffuse wave
hydraulics (2D)

High

Ungauged catchments
with floodplain, in
channel and diffuse
land use changes

Research
institution
distributed
models e.g. SHE

Full 2D St
Venants
Equations

Full 2D St
Venants
Equations

Very high

Ungauged catchments
with floodplain, in
channel and diffuse
land use changes

Table 14: Summary of flow routing used within existing hydrological models and
methods

It is recommended that an approach similar to that used in the Allan Water Natural
Flood Management study is included within the initial S20 assessment procedure. It
is felt that this approach offers a good balance between computational expense and
functionality by enabling the accumulation of effects from diffuse land use changes
and the limited assessment of in-channel interventions. However a flexible approach
should be adopted to allow either a decrease or increase in computational expense
by permitting the use of user supplied constant velocities (G2G) or a more detailed
computational engine.
It is proposed that a single time to outlet grid is calculated for the catchment which
will be representative for all flood return periods, in a similar fashion to how the ReFH
relies on a single time to peak for all return periods. The approach of using a single
time to outlet grid is described by Beven (2012)203 as a linearity assumption, and he
reports that the method has been found to work surprisingly well with any
inaccuracies due to the linearity assumption being generally less than those
associated with the estimation of runoff. This time to outlet grid can be recalculated
for different Natural Flood Management scenarios. It is envisaged that the time to
outlet grid would be calculated at a grid resolution of 25-200m depending on data
availability, computation constraints and catchment size. A flow chart summarising
the proposed method is shown in Figure 42. An example of a time to outlet grid for a
sample Scottish catchment is shown in Figure 39.

-127-

Figure 39: An example of a time to outlet grid generated for a sample Scottish
catchment as part of this study

The time to outlet grid would be calculated for a representative bankfull event, with
the assumption that a bankfull event is equal to the 1 in 2yr flood event. Due to the
requirement for the time to outlet grid to extend to all parts of the catchment, right
down to the scale of a single grid cell, it is proposed to use a 1 in 2yr rainfall event of
a duration equal to the FEH recommended storm duration (Equation 32) which in turn
requires knowledge of the catchments time to peak. At this stage the time to peak
can be estimated using Equation 33.

SAAR

D T p 1

1000

Equation 32: FEH estimation of design storm duration

T p 1.563PROPWET 1.09 DPLBAR 0.60 1 URBEXT1990

3.34

DPSBAR 0.28

Equation 33: FEH estimation of time to peak

On calculation of a bankfull rainfall depth which is representative of the catchments


mean using the FEH Depth Duration Frequency model, this will be factored so that it
is representative of the rainfall intensity at the time of the storm peak by multiplying
by the average rainfall intensity by 2.7 (the rainfall intensity in the middle 1/13th of a
FSR winter storm profile). This factored rainfall intensity can then be weighted
across the catchment based on SAAR .
Assuming a percentage runoff rate based on HOST data, the runoff can be
accumulated based on a rolling ball analysis of the ground surface. This rainfall
accumulation will generate an estimated flow ( Q ) per grid cell over the entire
catchment and the flow will be taken to be representative of the conditions during a
bankfull flood event.

-128-

A user supplied threshold (possibly ~0.1cumecs) will be used to define the transition
between land flow and channel flow. The threshold will be set at a value to achieve a
similar river network extent as seen on high resolution Ordnance Survey mapping
such as the OS 25k. It may be appropriate to supplement the river network by
loading in a mapped artificial drainage network. For all cells which have been
identified as watercourses, a channel width can be calculated using the relationship
identified by Robson and Reed (1999)225 (or similar i.e. Leopold and Maddock
(1953)201) which is based on QMed, as shown in Equation 34. The bankfull flow
estimated earlier will be used as surrogate for QMed.

b 0.182QMed 1.98
Equation 34: Estimation of channel width based on QMed

The next step is to calculate a grid of topographic gradient ( S ) for the entire
catchment using the updated ground surface model. It is recommended that a
smoothing process is undertaken along watercourses to overcome adverse slopes.
Using land cover, estimate roughness values ( n ) for the land flow and channel flow,
this will result in the generation of two different grids with the land flow roughness
being generally higher than channel flow. It should be noted that the land flow
roughness should be representative of the impedance that flow will experience in that
locality, be it as overland flow or shallow groundwater flow. Recognition of local
influences on flow path length should also be given consideration to (i.e. the direction
of tillage). The Mannings roughnesses presented for channels and floodplains will
not be applicable for the land flow. Although not directly applicable, some guidance
on the selection of shallow overland roughnesses is provided by USDA (2010)226, as
summarised in Figure 40. It should be noted that shallow overland flow considered
by the USDA does not include shallow groundwater flow and therefore its validity
may be questionable.

Figure 40: Roughness values recommended by USDA (2010) for shallow surface
runoff for flow depths of less than 30mm.

The roughnesses for river channel should take account of the riparian vegetation and
can theoretically be selected from published values for channels and floodplains.
-129-

Guidance on the selection of roughness values for channels and floodplains is


presented in Section 3.3(L). In assigning roughness values it is intended that the
number of roughness classes is kept small, ideally around four classes for both land
and channel flow. Some examples of the roughness bands which could be employed
are shown in Table 15, it is proposed that a single roughness value would be applied
to each band. The band roughnesses value could be varied to facilitate calibration.
Roughness band

Land flow roughness

Channel flow roughness

Low

Down slope ploughing and


open water

No riparian vegetation and


open water

Medium

Improved pasture

Limited riparian vegetation

High

Semi-improved pasture

Good riparian vegetation

Very high

Matted ground cover

Very good riparian


vegetation

Table 15: Examples of land cover types and the roughness bands which could be
employed

Now there is sufficient data to estimate the normal depth ( y n ) using Mannings
equation, and hence the velocity for the land and channel flows. For the land flow it
is assumed that the flow is spread evenly over the whole width of the grid cell so that
the wetted perimeter ( P ) is taken to equal the analysis cell size. For the channel
flow it is assumed that P is the bankfull channel width ( b ). The slope ( S ) would be
divided by a universal sinuosity factor for both the land ( SFLand ) and channel flow
( SFChannel ) so that the effect of assessing flow path lengths on a coarse grid does not
result in an overestimation of the slope. There is scope for the channel sinuosity
data set held by SEPAs Hydromorphology team to be used to inform local channel
sinuosity, this data could be overwritten should a channel planform restoration project
be considered.

5
3

1 A
n P 23

3
QnP
S

SF
v
Q

S
SF

3
5

Equation 35: Use of Mannings equation to calculate the flow velocity during normal
depth flow conditions

The calculated land and channel velocity grids can then be merged (mosaiced)
together to form a single flow velocity grid for the entire catchment. The final step is
to calculate the travel time for each grid cell ( t ) based on the calculated flow velocity
( v ) and the grid size ( l ) multiplied by the relevant sinuosity factor ( SF ). The
estimated grid cell travel times can then be accumulated working upstream, so that
the time to outlet for each grid cell can then calculated.

-130-

Calibration of the flow routing model and the application to real catchments:

Ideally the default sinuosity factors ( SF ) can be calculated by analysing flow


path lengths in a sample of catchments using a range of grid resolutions. It is
envisaged that an increase in analysis grid resolution will result in an increase
in SF . It is proposed that the user would be able to vary the SF factors to
allow the calibration of the model to observed flood events rather than by
varying roughness values.

It will be necessary to undertake an initial study using a range of test


catchments to develop guidance on the selection of roughness values,
particularly regarding the selection of land roughnesses which are likely to be
dependent on land cover, hillslope location and soil type

It is not a priority for the initial model, however the lag effect of reservoirs and
lochs could be included via a modification to the time to outlet grid should it
be required.

Initial testing of the flow routing model has been undertaken for a collection of
Scottish gauged catchment with reported observed catchment lag times in the Flood
Estimation Handbook vol 4 (1999)227. The results of this initial testing are presented
in Appendix E.
Although not implicitly within the proposed method, there is scope to create a time to
outlet grid by any means that the hydrologist feels appropriate for the catchment in
question. For an initial investigation it may be decided that it is appropriate to use
spatially constant estimates for channel and land flow velocity as assumed within
G2G. Conversely, where there may be attenuation from a reservoir or loch, it may be
seen as appropriate to calculate a lag time for the waterbody and add that lag (as a
constant) to the catchment situated upstream of the waterbody. In a similar fashion a
time lag for a floodplain could be assessed. This time lag could be calculated by
assessing the flood peak travel time using a localised 1D-2D hydraulic model of the
area where the changes will be made, or it could be estimated via a much more
simplistic means depending on what is appropriate to the circumstances. Although
the proposed method is based on a single time to outlet grid for all flood return
periods, this assumption may be questionable for the lag effect generated by open
waterbodies and could be replaced with an alternate approximation as the
hydrologist sees fit. An example of how a time lag effect could be added to the time
to outlet grid is shown in Figure 41Error! Reference source not found..

-131-

Estimation of Lag Time


0.14
0.12

Lag time

0.1

Flow

0.08
0.06
0.04
0.02

Lag time added to


time to outlet grid

Time

Figure 41: Example of how the time lag effects of reservoirs, lochs and floodplains
could be included within the assessment method

-132-

RAW DTM

SAAR GRID

DRAINAGE SHAPEFILE

FEH DDF

Create a bankfull rainfall intensity grid

Create hydrologically correct


DTM

Assuming median annual flood is bankfull using a


storm duration which is representative of the
catchment. Initial estimate of Tp made using

Update DTM by burning in


mapped watercourses,
smoothing bed of mapped
watercourses, then filling sinks

SAAR
D T p 1

1000

T p 1.56PROPWET 1.09 DPLBAR 0.60 1 URBEXT1990

3.34

DPSBAR 0.28

Grid of accumulated bankfull


rainfall (flow)

Use rainfall intensity representative of the storm peak,


2.7 times average storm intensity (intensity of mid 1/13
of winter profile)

Assuming HOST based


percentage runoff of storm peak
rainfall. Output is a grid of flows
rates per cell

Weight catchment average rainfall intensity across the


catchment using SAAR for weighting

Calculate channel widths grid

Define cells to be treated as


watercourse

Estimate channel widths using


QMed estimated bankfull flow
BFCW 0.182QMed

Cells greater than a user


specified flow rate (~0.1cumecs)

User supplied artificial drainage

Calculate grid of
topographic
gradients

Calculate grid of inchannel velocities


Using Mannings

Using the
hydrologically
corrected DTM,
calculate a grid of
topographic
gradients

CHANNEL ROUGHNESS
Based on presence of
riparian vegetation

LAND FLOW
ROUGHNESS GRID

CHANNEL
SINUOSITY

Based on land use

UNIVERSAL LAND
FLOW SINUOSITY
Calculate grid of land
flow velocities
Time to outlet grid

Using Mannings

Mosaic velocity grids


together and calculate
time from each cell to
outlet of catchment

Figure 42: Summary of the proposed flow routing methodology

-133-

(F)

Baseflow model

It is anticipated that baseflow will not form a significant component of flood flows
within typical Scottish catchments, it will be the responsibility of the hydrologist to
verify this assumption prior to the application of this method. It is proposed to adopt
a simplistic approach for baseflow by using a constant flow rate. In the absence of
observed that the estimate of initial baseflow applied within the ReFH as detailed in
Equation 36 or the FEH Rainfall Runoff Method detailed in Equation 37.

Qbaseflow 63.8Cini 120.8 5.54SAAR 10 5 AREA


Equation 36: ReFH estimation of baseflow

Qbaseflow AREA33CWI 125 3.0SAAR 5.5 10 5


Equation 37: FEH Rainfall-Runoff estimation of baseflow

As discussed in Section (A) the winter event has been adopted as it is envisaged
that the Natural Flood Management assessment tool will be predominately used for
rural catchments.

(G)

Generation of the flood hydrograph

It is proposed that the flood hydrograph will be generated by dividing the time to
outlet grid (as generated in Section 4.4.2(E)) into a number of bands of equal travel
time, or hydrochrones. Following the principles of the FEH Rainfall-Runoff methods,
it is recommended that an odd number of hydrochrones are used, and seven to
fifteen equal bands may present a good balance between accuracy and
computational expense. The design rainfall depth would be divided into the same
number of bands using the FSR storm profile as discussed in 4.4.2(A) before being
weighted across the catchment and used to estimate the runoff generation for each
grid cell via the loss model (Section 4.4.2(C)) after taking account of canopy
interception (Section 4.4.2(B)). The flow at the catchment outlet for timestep ( Qt )
would then be calculated based on the summation of the area of the hydrochrone
band ( AN ) times the average runoff rate for the band after being delayed by its runoff
band number ( q N ,t N ) using Equation 38. As discussed in Section 4.4.2(F), a
constant baseflow will be assumed throughout the event. It should be noted that
moisture accounting and the subsequent runoff generation via the losses model
would be undertaken for each grid cell and not via the use of average moisture
conditions at the hydrochrone band scale.

Qt AN q N ,t N Qbaseflow
N

N 1

Equation 38: Generation of the flow at the catchment outlet based on the summation
of the delayed runoff from each hydrochrone band

An example is presented in Figure 43 and Figure 44 using a simple five band


catchment. Ideally, this process would be refined using a large number of bands.

-134-

Figure 43: Example of a five band time to outlet grid for a sample Scottish catchment
700

2
600
4
500

400

10

300

12

Rainfall depth (mm)

Flow (cuemcs)

Band 4
Band 3
Band 2
Band 1
Band 0
Baseflow
Rainfall

200
14
100
16

18
0

10

11

12

13

14

Time step

Figure 44: Example hydrograph using five runoff bands for a generalised catchment

-135-

(H)

Model construction

It is proposed that the model is constructed within GIS as this fits well with the spatial
nature of the data that will be used. To comply with SEPAs existing software
licences the most applicable GIS product is ArcGIS 9.3. The tool could feasibly be
constructed as a series of Toolboxes, however the use of geoprocessing scripts
would bring significant computational efficiencies for the more computationally
expensive processes. The most time consuming process will be the iterative running
of the loss model and to a lesser extent the hydrograph accumulation and normal
depth hydraulics.
The file sizes of the grids associated with the method is an important consideration
as this will dictate the speed of the process and there will be significant efficiencies if
all grids can be held in memory rather than needing to written to the hard drive during
the processing. To be efficient this will require multiple grids to be held in memory at
any one time. Table 16 provides a summary of the memory requirements for a range
of analysis grid resolutions for a sample of catchments.
Catchment size
(bounding grid)

Example

Potential
grid
resolution

32bit raster
grid

5 x 10km

Minor
watercourse

25m

0.33MB

25 x 12km

River Carron

50m

0.95MB

60 x 25km

River Earn

50m

5.84MB

120 x 65km

River Tweed

100m

5.66MB

Table 16: Example of memory requirements for a range of catchment sizes

-136-

4.4.3 Data requirements


The proposed methodology for assessing the hydrological impact of fluvial Natural
Flood Management measures will require a significant volume of good quality spatial
data. The preferred and alternative means of meeting these data requirements are
presented in Table 17.
Data

Preferred

Alternative

Comments

Topographic
ground
surface

Filtered
LiDAR DTM

NextMap DTM

The quality of the DTM is not critical


as the method requires the
resampling of the DTM, even crude
DTMs are adequate to correctly
identify river valleys and
topographic ridges in all but the
flattest catchments

Rainfall

SAAR 1km

SAAR 5km

Rainfall distribution across


catchments can be significant and it
is important that good quality rainfall
distribution data is available

FEH DDF
parameters

FEH DDF
parameters
on 1km grid

Use the FEH CD


ROM to generate a
rainfall depth and
manually enter it into
the model

An estimation of the design rainfall


is essential for the method. If the
model could include the FEH DDF
model that would be the ideal,
otherwise the manual entry of
rainfall depth would be feasible.

HOST data

Vector
HOST data
250k

HOST on 1km grid

HOST data is essential for the


analysis. The 1km grid is too
coarse to allow investigations to
extend to the field scale and would
be a serious limitation on the model

Land
coverage

Vector Land
Cover Map

Grid based Land


Cover Map

Accurate land use is an important


consideration. Ideally this should
extend to the field scale.

Potential
evapotranspiration

MORECS
40km grid

It may be possible to
generate a
regression based
equation for PE if
MORECS data cant
be supplied.

The current preferred method for


modifying PROPWET relies on
knowledge of the potential
evapotranspiration.

Table 17: Summary of the data required for the proposed methodology

There would be potential to improve the fluvial assessment methodology should


additional data become available. These additional data are summarises in Table
18.

-137-

Data

Comments

Mapped river
networks

This data could be used to ensure that flow paths identified within
the routing model are correct. The model can function without it and
it would only be of particular use when channel planform restoration
is being considered.

Mapping of soil
degradation

National mapping of soil degradation would be beneficial. This


could be based on Parish livestock census data in combination with
mapped soil types but would require additional study in addition to
ground truthing.

Mapped
drainage
network

This data can only be used to extend the channel network if the
analysis grid is relatively fine. It could be used to reduce the land
flow roughness or the antecedent moisture condition following
further investigation.

Channel
sinuosity

SEPAs Hydromorphology Team has a sinuosity data set for all


main watercourses in Scotland. This could be used to inform the
selection of channel sinuosity and there would be potential for this
spatial dataset to allow the channel sinuosity to vary spatially.

Table 18: Summary of the data which could improve the quality of the tool

To improve processing time of the proposed fluvial assessment methodology it would


be beneficial to produce a number of standard pre-processed grids for the existing
conditions found across Scotland. The standardisation of these baseline grids would
also provide a means of controlling how the baseline case is modelled with any
deviations away from a standard baseline case needing to be justified by the
hydrologist completing the study. These standard grids would provide a template for
the formatting and units required for input data.
A list of potential pre-processed data sets with national coverage could comprise of
the following:

Standardised hydrologically correct 50m DTM

Existing C Max

Existing Cini,w int er

Existing LAI w int er

Existing LAI summer

Existing rsc,w int er

Existing rsc,summer

Existing nchannel

Existing nland

Existing velocity grids (channel and land combined) for a selection of RMed
durations (3hr, 6hr, 12hr & 24hr) giving potential for the estimation of
intermediate durations via interpolation or extrapolation.

-138-

4.4.4 The capabilities of the assessment tool


The proposed assessment tool can be used to assess the hydrological impact of
diffuse Natural Flood Management measures as described in Table 19.
Natural Flood
Management
measure

Means of representation

Not included

Runoff reduction
due to conifer
woodland

Canopy interception via the use of MORECS leaf area index


Antecedent moisture condition via modification of PROPWET
Reduced runoff velocity via changed ground surface roughness
within routing model
Canopy interception via the use of MORECS leaf area index (LAI)
LAI values set to representative value for the anticipated flood
season.
Antecedent moisture condition via modification of PROPWET
Reduced runoff velocity via changed ground surface roughness
within routing model
Canopy interception via the use of MORECS leaf area index
Antecedent moisture condition via modification of PROPWET
Accelerated runoff velocity via changed ground surface roughness
within routing model
Soil moisture capacity and initial soil moisture modified via the use
of analogue degraded HOST class
Accelerated runoff velocity via changed ground surface roughness
within routing model
Soil moisture capacity and initial soil moisture modified via the use
of analogue degraded HOST class
Antecedent moisture condition via modification of PROPWET
based on land cover
Accelerated runoff velocity via changed ground surface roughness
within routing model
Only via those described for deciduous woodland

The impact of drainage on the


antecedent moisture condition
or its effect on runoff velocity

Runoff reduction
from deciduous
woodland

Runoff effect of
widespread
muirburn
Reducing soil
degradation due to
overgrazing
Reducing soil
degradation due to
arable land use

Cross slope tree


shelter belts
Riparian planting
Channel planform
restoration
Floodplain related
measures

Reduction of channel velocity by increasing Mannings roughness


and thus extending the time to peak
Extending the channel length thus decreasing slope and reducing
velocity. Increased length and decreased velocity both serve to
extend the time to peak.
Not included, however could be undertaken via estimation of lag

Effects of
agricultural
drainage

Not included, outline of potential method which could be


investigated in the future proposed in Section 4.4.4

Effects of upland
drainage

Not included, outline of potential method which could be


investigated in the future proposed in Section 4.4.4

Hydraulic
constrictions

Not included

No representation of soil pore


blocking by ash particles

Infiltration of runoff arriving


from upslope
Floodplain interaction
Floodplain interaction

Model does not include


hydraulics on the floodplain
therefore the effects of
floodplain woodland or
modifying connectivity not
included
Effects on antecedent moisture
condition or soil moisture
capacity
Effects on runoff velocity
Effects on antecedent moisture
condition or soil moisture
capacity
Effects on runoff velocity
Effects of hydraulic
constrictions such as bridges
to be modelled using
appropriate hydraulic
modelling software

Table 19: Summary of how a range of Natural Flood Management measures could be
assessed using the proposed methodology

The proposed methodology does not contain a means of estimating the effect of
drainage, be that piped drainage typically found within farmland or open ditches often
-139-

found in upland areas. A discussion of the literature relating to farmland and upland
drainage is provided in Sections 3.3(D) and 3.3(F). In summary, the effects of
drainage is complex with the impact being related to soil type, land slope, drain
slope, drain capacity and an array of other considerations. It currently pushes the
capability of academic plot scale models to make meaningful predictions about small
areas. It may be feasible to develop a means of estimating the impact of drainage on
the antecedent moisture condition via regression, or another means, of the significant
contributing factors. It is likely that considerably more data will be required to
achieve this. As the drainage network is likely to be less than the relatively coarse
analysis grid size proposed by this fluvial assessment methodology the only viable
means to assessing the accelerated runoff pathway that drainage represents could
be by reducing the land covers Mannings roughness value. This modification of
roughness value would need to make consideration of drainage direction (cross or
downslope), drainage capacity, drainage condition, soil type and the impedance to
flow caused by the surrounding land cover. In combination with a modification to the
land roughness it would also be necessary to modify the initial soil moisture and the
soil moisture capacity to account for the reduced antecedent moisture condition and
the reduced capacity to store water.

4.4.5 The inclusion of elements of the ReFH


As presented in Section 4.4.2 key elements of the proposed assessment
methodology are founded on the ReFH model (Kjeldsen, 2007)50 which is currently
not recommended for application within Scotland by SEPA. These key factors are:

Estimation of Soil Moisture Capacity

Estimation of Initial Soil Moisture

The recommendation of basing estimating these model parameters on the regression


based equations utilised by the ReFH have been made in the absence of a viable
alternative. The adoption of these equations should be viewed as provisional and
should be updated should additional verification or an extended calibration exercise
is undertaken using a wider selection of catchments representative of those found in
Scotland. A number of other elements have been adopted from the Flood Estimation
Handbooks Rainfall-Runoff Method which is viewed as current standard practice.

4.4.6 Calibration, verification and improvement


Prior to the tool being released it should be put through a rigorous calibration and
verification phase to determine how the results produced compare to flood growth
curves for a collection of gauged catchments. These catchments should not be
limited to Scottish Natural Flood Management test catchments but should include a
wide range of land uses, soils, precipitation, catchment conditions, catchment sizes
and catchment shapes. In instances where land use changes have been recorded
within a gauged catchment the effects of these changes on flood generation could be
tested. However it should be noted that the driving focus should be to achieve a
good calibration to a national selection of catchments covering a wide range of
conditions rather than the calibration to a small number of Natural Flood
Management test catchments.
It is acknowledged that the tool could benefit from better calibration coefficients and
relationships in the following areas:

Default channel roughness values

Default land runoff roughness values

-140-

Default channel sinuosity factor

Default land runoff sinuosity factor

Bankfull channel width equation

Estimation of PROPWET for changing land use

Equation for providing an initial estimate of the critical storm duration

Method for estimating the role of the canopy interception

Method to estimate of baseflow

Method for including the time lag effect of lochs, reservoirs and floodplains

The Generalised Likelihood Uncertainty Estimation (GLUE) method (Beven and


Binley, 1992)228 presents a robust framework for selecting parameter values via the
use of Monte Carlo simulation. By scoring each parameter set with a Likelihood
Value based on a defined goodness of fit measure it is feasible to identify a range of
feasible outcomes by running an ensemble of model runs. This methodical approach
can provide a valuable insight into how the model operates by giving an insight into
model sensitivity to each parameter. It may be feasible to incorporate elements of
the GLUE framework within the development of the model, however it should be
noted that the GLUE methodology is specifically targeted at rainfall-runoff modelling
of a single catchment rather than the development of models which can be applied to
a wide number of catchments. The additional consideration of needing to optimise
storm duration for each parameter set implies that using the principles of GLUE
during the initial model calibration and verification phase for a large number of
catchments over a range of storm return periods may be beyond what is
computationally feasible. However the GLUE principles could be of great use when
making predictions about single catchments.
On completion of a verification process using a selection of gauged catchments it will
be feasible to develop a better understanding of how the model performs at
replicating observed flood growth curves. This information could be used to develop
; bands for the model, for example by being able to state that for a given percentage
of tested catchments the model over or under predicted the flood peak for a given
return period by a certain percentage. This analysis may identify situations where
the method performs poorly helping to direct future study and also providing a
warning to hydrologists using the tool of areas where it performs poorly.
It is proposed that on a day to day basis the hydrologist will reduce uncertainty by
first of all ensuring that the baseline model is calibrated to the local conditions. For
gauged catchments this will be a relatively straightforward process, however for
ungauged catchments this will be more challenging. It is envisaged that the most
appropriate means will be to use a nearby gauged catchment as analogue to identify
representative values for variables (Mannings roughnesses, sinuosity factors, Leaf
Area Indexes, Initial Soil Moisture etc). The importance of calibration to an analogue
gauged catchment, as identified in other hydrological methodologies (Kjeldsen,
2007)208, is stressed as being of key importance.
As with all hydrological assessments, even with the best calibration and verification
process there will still be a relatively large degree of uncertainty within the predictions
made by this methodology. It is important that this uncertainty is conveyed alongside
the headline central estimate figures. It is recommended that sensitivity testing of the
most significant variables is undertaken by testing the effect of feasible changes to
these variables on the models prediction.
As a concluding remark, it is crucial that the predictions made using the method are
kept in context. At no point should the model output be taken as being the correct
-141-

answer, rather the output should be seen as an approximate guide to what potentially
might happen.

4.4.7 Flooding and the role of forestry


There are three potential mechanisms identified by Woodland for Water (Nisbet et al,
2011)65 that trees could be used to help alleviate the effects of flooding. These
mechanisms have been included within the proposed assessment tool:
1. The role of canopy interception to reduce effective rainfall has been included
via the vegetation leaf area index
2. The ability of woodland soils to reduce runoff is considered through the
modification of the antecedent moisture condition and the soil moisture
capacity depending on the vegetation water usage and soil degradation
3. The ability for floodplain and riparian woodland to increase hydraulic
roughness therefore slowing catchment response times and reducing flood
peaks is included within the method. It should be noted that the effect of
floodplain interventions, such as increasing floodplain roughness using
woodland, requires additional location specific hydraulic modelling as
presented in Section 4.4.4.
The findings of this study are closely aligned with the estimation of impacts presented
by Nisbet et al (2011)65. Ignoring the effects of now out-dated artificial drainage
systems, all types of tree cover will serve to reduce the antecedent moisture capacity
hence increasing the probability of the soil being dry and capable of allowing the
infiltration of rainfall. As shown in Section 4.4.2(D), the higher the water usage the
larger this effect is, thus species with a high water demand are most effective at
improving the antecedent moisture condition and therefore reducing flood peaks.
The effect of leaf cover seasonality on the antecedent moisture condition has not
been directly investigated; however it is anticipated that deciduous species will
improve the antecedent moisture condition throughout the year due to high
evapotranspiration losses during the summer months.
The effect of canopy interception on flood peaks is envisaged to be marginal at most
due to the small volume of rainfall that can be stored on the canopy relative to the
volume of rainfall during flood events linked to the issue that limited capacity is
utilised at the start of the rainfall event. The relative importance of reducing
antecedent moisture over canopy interception suggests that deciduous forests are
beneficial throughout the year.
Section 4.4.2(B) of this document proposes the use of a maximum canopy
interception based on the daily threshold ( I 0.2LAI ) presented by Hough and
Jones (1997)216. It should be noted that consultation as part of the review process of
this document has identified that it is the opinion of Forest Research that this method
underestimates the canopy interception capacity and therefore will underestimate the
role of vegetation in reducing runoff. Forest Research advocates the HoyningenHuene (1981)105 methodology, however there is concern that this method introduces
the requirement to assess the initial canopy wetness at the start of the rainfall event
and to ignore this initial canopy wetness would result in an overestimation of
interception. It is recommended that the assessment tool allows the user to choose
which canopy interception model is used (MORECS, Hoyningen-Huene or none),
thus allowing sensitivity testing of findings. As experience is developed it is
recommended that the default recommendation of using the MORECS method is
revisited.

-142-

As presented in Sections 3.3(B) and 3.3(C), the effect of soil compaction has been
identified to be a cause of increased runoff generation by Packman et al (2004)48.
Along with other remedial measures, it is anticipated that the conversion of landuse
to woodland will help reduce soil compaction thus serving to bolster the flood risk
management benefits of woodland.
The ability of vegetation to impede the flow of water over the land and hence
attenuate flood water upstream of flood receptors has been partly included within the
method. The 1D kinematic wave model proposed within the core of the flow routing
method will allow the impact of channel, bank and riparian interventions to be
investigated on a broadscale through the application of Mannings roughnesses. A
method for quantifying the effects of floodplain roughening using woodland has been
presented in Section 4.4.4 using a 1D-2D hydraulic model (or alternative hydraulic
calculation) to quantify the lag time generated by the intervention. The proposed
method for assessing the effect of vegetation impeding the movement of flood water
and hence attenuating flood flows is the same as that applied by previous studies
(increased Mannings roughness). The delay effects will be related to the impedance
that the vegetation creates, with vegetation that has a high Mannings roughness
creating the largest effects. On this basis, lightly managed woodland and short
rotation forestry, might be expected to be good at maximising the attenuation effects.
As discussed in Section Error! Reference source not found., consideration should
also be given to the potential supply of debris to the watercourse which could block
constrictive structures downstream.

4.4.8 Summary
Following the recommendation of DEFRA FD2114 (OConnell et al., 2004)190 a
hydrological assessment methodology has been presented which is:
1. Distributed;
2. Capable of tracking soil moisture conditions;
3. Partly physically based; and
4. Capable of tracking effects downstream.
The method is currently not capable of continuous simulation however the PDM
principles which it is founded on do not preclude it. Although a vast simplification of a
complex set of interrelated processes, the method avoids the need for continuous
simulation by estimating the antecedent moisture conditions using the ReFH same
methodology applied within the ReFH. It is felt that this approach is in the spirit of
what was recommended by DEFRA FD2114 (OConnell et al., 2004)190.
It should be noted that this is a proposed hydrological method and that as of the time
of writing there is no software to facilitate the application of the method. Once
software has been created it will be necessary to calibrate and verify the model.

-143-

5 Summary of key recommendations


The following recommendations relating to the screening and assessment of Natural
Flood Management measures are contained within this study:
Method:

The identification and appraisal of Natural Flood Management should be


separated into three separate stages:
o

The identification of Natural Flood Management potential should be


undertaken via a national screening process. The screening does not
need to provide a quantification of the effect Natural Flood
Management can have.

Based on the output of the screening process an initial long list of


potential flood risk management options should be developed by
bringing together all known constraints and opportunities. This long
list is then refined via an initial appraisal process to produce the
preferred options.

Local Flood Risk Management Partnerships will quantitatively assess


the preferred options and refine them further before completing the
final Flood Risk Management Plan.

Identification of opportunities by screening:

It is recommended that the identification of Natural Flood Management is


undertaken via national screening;

The screen process should be divided into a number of sub-screening


processes each relevant to different types of Natural Flood Management
measures. The following screening processes have been investigated:
A) Areas of high runoff generation;
B) Areas of floodplain storage potential (using modelled hydraulic gradients);
C) Opportunities for sub-catchment desynchronisation (not to be undertaken
due to inadequate scientific support);
D) Opportunities to remove hydraulic constrictions;
E) Areas of heightened hydromorphological activity;
F) Areas of estuarine surge attenuation potential; and
G) Areas of wave energy attenuation potential.

The output of the screening process should be a range of national maps


depicting the opportunity for Natural Flood Management (six)

The screening methods presented in this document require additional


development work to be undertaken including the selection of coefficients and
the identification of appropriate assessment resolutions. It is essential that
these decisions are robust and the final methodology is recorded.

Quantification of effects using assessment tools as part of the appraisals:

Where feasible, existing tools should be used for the assessment of Natural
Flood Management measures as detailed in Table 10. Guidance should be

-144-

prepared to detail how common Natural Flood Management measures can be


assessed;

The assessment of fluvial and pluvial Natural Flood Management measures


requires the development of new hydrological assess method. This method
needs to be:
A) Distributed;
B) Capable of running continuous simulation;
C) Partly or wholly physically based so that landscape, soil and vegetation
can be represented; and
D) Detailed modelling of surface flow so that effects can be tracked
downstream.

A spatially distributed semi-physical single event hydrological model is


proposed which will facilitate the effect of diffuse land use changes to be
evaluated in a robust manner.
o

The model is founded on the ReFH runoff generation model so is


closely aligned with current British practice.

The inclusion of physically based flow routing via a time to outlet grid
permits the effects of diffuse land use changes to be accumulated for
a broad range of catchment sizes in addition to quantifying the effect
of in channel modifications

The effects of soil degradation and land cover can be investigated


scientifically by making changes to the runoff generation model

A GIS tool should be developed to allow the application of the tool;

The tool should be subjected to a rigorous calibration and verification phase


to ensure that the method is robust. This process should be extended to a
broad range of catchments and not focus solely on achieving good calibration
to a limited range of Scottish Natural Flood Management test catchments.

The calibration and verification phase will enable the assessment of model
uncertainty. When quoting results based on this method it is essential that
they are quoted alongside the evaluated uncertainty.

-145-

References
1

Scottish Government, Flood Risk Management (Scotland) Act, Scottish Parliament, June
2009. http://www.legislation.gov.uk/asp/2009/6/pdfs/asp_20090006_en.pdf (Accessed 13
February 2012)
2

SAIFF, What is meant by restoration, enhancement, and alteration under the Flood Risk
Management (Scotland) Act 2009. V2.0, SAIFF policy discussion paper. 2011.
3

Jacobs, Development of Methodology for Assessment Required under Section 20. SEPA,
2011.
4

Werritty A., Arthur S., Ball T. and Wallerstein N., Review of Jacobs Section 20 Review,
CREW-The James Hutton Institute, January 2012
5

SNIFFER & SEPA Natural Flood Management Implementation Learning from Practice
Workshop Report, November 2011
6

Atkins. R&D Update review of the impact of land use and management on flooding.
Environment Agency, 2007.
7

Beven K., Young P., Romanowicz R., FD2120: Analysis of historical data sets to look for
impacts of land use management change on flood generation. DEFRA & Environment
Agency. 2008.
8

Halcrow, Delivery of Making Space for Water; HA6 & HA7: The role of land use and land
management in delivering flood risk management. Environment Agency, 2008.
9

Environment Agency, The Land Use Jigsaw, Prioritising scientific research on land use
management and FCRM. 2008.
www.google.com/documents (insert username: land.use.jigsaw password: evidence)
10

Haycock Associates, Evidence Based Review: Does Land Management Attenuate Runoff?
National Trust, 2008.
11

MNV, The Way Forward for Natural Flood Management in Scotland A report for Scottish
Environment LINK. 2008.
12

Environment Agency, Greater working with natural processes in flood and coastal erosion
risk management, A response to Pitt Review Recommendation 27, Environment Agency,
2012
13

Longfield S.A. Macklin M.G., The influence of recent environmental change on flooding and
sediment fluxes in the Yorkshire Ouse Basin Hydrology. Hydrological Processes 13 105166,1999
14

Conway, VM. Miller, A. The hydrology of some small peat covered catchments in the
Northern Pennines Journal of the Institute of Water Engineers 14 415-24. 1960.
15

Robinson, M. Changes in catchment runoff following drainage and afforestation Journal of


Hydrology. 86 71-84. 1986.
16

Godwin R.J. and Dresser M.L. Review of soil management techniques for water retention in
the Parrett catchment Environment Agency, Almondsbury, Bristol. 2003
17

Davies R., Predation and the Profitability of Grouse Moors. British Wildlife, 16, 339-347.
2005
18

Cox R Parasites and Red Grouse. Are parasites of red grouse becoming resistant? The
Game Conservancy Trust Review, 35, 48-49. 2003
19

Hudson P., Grouse in Space and Time: The population biology of a managed gamebird.
Game Conservancy Ltd, Fordingbridge, Hampshire. 1992
20

Gimingham C.H., An Introduction to Heathland Ecology, Oliver & Boyd, Edinburgh. 1975

-146-

21

Lawton J.H., Red Grouse Populations and Moorland Management, British Ecological
Society, Shrewsbury. 1990
22

Tucker G., Review of the Impacts of Heather and Grassland Burning in the Uplands on
Soils, Hydrology and Biodiversity. English Nature Research Report Number 550. English
Nature Peterborough, 147. 2003.
23

Yallop, A R, Thacker, J, Thomas, G, Stephens, M, Clutterbuck, B, Brewer, T and Sannier,


C A D The Extent and Intensity of Management Burning in the English Uplands. Journal of
Applied Ecology, 43, 1138-1148. 2006
24

O'Brien H.E., Labadz J.C. and Butcher D.P., An Investigation of the Impact of Prescribed
Moorland Burning in the Derwent Catchment Upon Discolouration of Surface Waters. Report
to Moors for the Future. Nottingham Trent University pp.85 2006
25

Scottish Natural Heritage, Fire: Friend or foe? http://www.snh.gov.uk/land-andsea/managing-the-land/upland-and-moorland/fire-friend-or-foe/ (referenced 25 January 2012)
26

Scottish Government, The Muirburn Code. Edinburgh, 2011.

27

OBrien H., Labadz J C & Butcher D P, A review of Blanket Bog Management and
Restoration, Technical Report to DEFRA. CTE0513, Nottingham University, DEFRA, 2007.
28

Lane, SN. Brookes, CJ. Hardy, RJ. Holden, J. James, TD. Kirby, MJ. McDonald, AT. Tayefi,
V. Yu, D. Land management, flooding and environmental risk: new approaches to a very old
question Proc. Of National CIWEM Conference. 2003
29

Mallik A.U., Giminham C.H., Rahman A.A., Ecological effects of heather burning: I. Water
infiltration moisture retention and porosity of surface soil. Journal of Ecology, 72, pp. 767-776.
1984
30

Kinako, P.D.S.. Effects of heathland fires on the microhabitat and regeneration


ofvegetation. PhD Thesis. University of Aberdeen, Aberdeen. 1975
31

Meyles, E.W., Hillslope and watershed scale hydrological processes and grazing
management in a Dartmoor catchment, southwest England. PhD Thesis. University of
Plymouth, Plymouth. 2002.
32

Conway V.M. & Millar A., The hydrology of some small peat-covered catchments in the
northern Pennines. Journal of the Institute of Water Engineers, 14, pp. 415 424. 1960
33

Robinson, M., The hydrological effects of moorland gripping: a reappraisal of Moor House
research. Journal of Environmental Management, 21, pp. 205-211. 1985.
34

Dunn C.E., Hydrological responses to moorland land-use change. Phd Thesis, Department
of Geography, The University of Hull. 1086.
35

Ramchunder S.J., Brown L.E. & Holden J., Environmental effects of drainage, drain
blocking and prescribed vegetation burning in UK upland peatlands. Progress in Physical
Geography, 33, 49-79. 2009.
36

Ember, http://www.geog.leeds.ac.uk/research/projects/ember/ (accessed 16 February


2012)
37

Heathwaite A.L., Burt T.P. and Trudgill S.T. Runoff, sediment and solute delivery in
agricultural drainage basins: a scale-dependent approach. 182, International Association of
Hydrological Science Publications. 1989
38

Heathwaite A.L., Burt T.P. and Trudgill S.T., Land-use controls on sediment production in a
lowland catchment, south-west England. In: J. Boardmand, I.D.L. Foster and J.A. Dearing,
Soil Erosion on Agricultural Land. John Wiley and Sons Ltd. Chichester, UK. 1990.
39

Holman I.P., Hollis J.M. and Thompson T.R.E., Impact of agricultural soil conditions on
floods Autumn 2000. R&D Technical Report. Environment Agency. 2001.

-147-

40

Goodwin R.J. and Dresser M.L. Review of soil management techniques for water retention
in the Parrett catchment, Environment Agency, Almondsbury, Bristol. 2003.
41

Hollis J.M., Dresser M., Thompson T.R.E. and Newland T., Comparison of agricultural soil
conditions in the Uck and Bourne catchments during the winter periods of 2000/2001 and
2002/2003. Report for the Environment Agency. 2003.
42

Cranfield University, British Trust for Ornithology, The Open University, University of
Reading and Rothamstead Research. Scoping study to assess soil compaction affecting
upland and lowland grassland in England and Wales, Final Report. Project BD2304. 2007
43

Johns M., The Impact of Grazing and Upland Management on Erosion and Runoff:
Additional Information. R&D Project Record P2/035/9, Environment Agency, 1998.
44

Harrod, T.R. Soil suitability for grassland. p. 51-70. In M.G. Jarvis and D. Mackney (ed).
Soil Survey Applications. Technical Monograph No.13, Harpenden, UK 1979.
45

Palmer, R.C. Preliminary estimation of soil structural conditions within 10 soil landscapes in
South West England during 2002 and 2004. Final Report to Environment Agency for England
and Wales, Bristol. 2004
46

Williams A.G., Borthwick M.F., Mtika E., Ternan J.L. and Sullivan A., The influence of
changes in farming patterns on the runoff characteristics of the River Camel, Cornwall, UK.
st
Hydrology: Science & Practice for the 21 Century. Volume II. British Hydrological Society,
2004.
47

Meyles, E.W., Williams, A.G., Ternan, J.L., Anderson, J.M. and Dowd, J.F, The influence
of grazing on vegetation, soil properties and stream discharge in a small Dartmoor catchment,
southwest England, UK. Earth Surface Processes and Landforms, 31, 622-631. 2006
48

Packman J.C., Quinn P.F., Hollis J., OConnell P.E. Review of impacts of rural land use and
management on flood generation: Short-term improvement to the FEH Rainfall-Runoff model:
Technical Background. DEFRA R&D Project FD2114/PR3. 2004.
49

Boorman D.B., Hollis J.M. and Lilly A., Hydrology of Soil Types: A hydrologically-based
classification of soils in the United Kingdom. Institute of Hydrology Report 126 Wallingford,
UK. 1995.
50

Kjeldsen T.R., Flood Estimation Handbook, Supplementary Report No1. The revitalised
FSR/FEH rainfall-runoff method. Centre for Ecology and Hydrology, 2007.
51

Holtan, H.N. and Kirkpatrick, M.H., Rainfall, infiltration and hydraulics of flow in runoff
computation. Transactions of the American Geophysical Union, 31: 771-779. 1950.
52

Packman J.C., Quinn P.F., Hollis J., OConnell P.E. Review of impacts of rural land use and
management on flood generation: Short-term improvement to the FEH Rainfall-Runoff model:
Technical Background. DEFRA R&D Project FD2114/PR3. 2004.
53

JBA Consulting & Cranfield University, Land management CFMP Tool, Development of a
software tool to investigate the potential of changes in rural land use and land management
on flood generation, Environment Agency, 2008.
54

Packman J.C., Quinn P.F., Hollis J., OConnell P.E. Review of impacts of rural land use and
management on flood generation: Short-term improvement to the FEH Rainfall-Runoff model:
Technical Background. DEFRA R&D Project FD2114/PR3. 2004.
55

OConnell E., Ewen J., ODonnell G., Quinn P. Is there a link between agricultural land-use
management and flooding? Hydrology & Earth System Sciences, 11(1) 96-107, 2007.
56

Robinson M., Impact of improved land drainage on river flows. Report 113. Institute of
Hydrology. Wallingford, UK.1990.
57

Hardiman N. et al, Technical review of NFRM techniques, their effectiveness and wider
benefits, RSPB, 2009.

-148-

58

Calder I.R., Reid I., Nisbet T.T and Green J.C., Impact of lowland forests in England on
water resources: application of the Hydrological Land Use Change model. Water Resources
Research, 39, 2003.
59

Hall R.L, Hydrological aspects of broadleaved woodland: Implication for water supply.
Forests and water supply, proceedings of a discussion meeting. Heriot Watt University,
Edinburgh. Institute of Chartered Foresters. P44-60
60

Nisbet T.R., Implications of climate change: soil and water. Forestry Commission Bulletin
125, p53-67. Forestry Commission, 2002.
61

Roberts J. and Rosier P., The impact of broadleaved woodland on water resources in the
lowland UK: III The results from Black Wood and Bridges Farm compared with those from
other woodland and grassland sites. Hydrology and Earth Systems Sciences, 9(6), 614-620,
2005.
62

Harding R.J., Hall R.L., Neal C., Roberts J.M., Rosier T.T.W. and Kinniburgh D.K.
Hydrological impact of broadleaf woodlands: Implications for water use and water quality.
Institute of Hydrology, 1992.
63

Reed D., Faulkner D., Robson A., Houghton-Carr H., Bayliss A., Flood estimation
handbook: Volume 1: Overview. Institute of Hydrology, 1999.
64

Nisbet T.R., Information Note: Water use by trees. Forest Research, 2005.

65

Nisbet T.R., Silgram M., Shah N., Morrow K., and Broadmeadows S., Woodland for Water:
Woodland measures for meeting Water Framework Directive objectives. Forest Research
Monograph, 4. Forest Research, Surrey. 2011.
66

Roberts, J. M., Forest transpiration: a conservative hydrological process? Journal of


Hydrology 66, 133141. 1983.
67

Calder I.R., Harrison J., Nisbet T.R. and Smithers R.J., Woodland actions for biodiversity
and their role in water management, Woodland Trust, 2009. (Peer reviewed).
68

Forestry Commission, A new focus for Englands woodlands strategic priorities and
progress. 1999.
69

Bird SB., Emmett BA., Sinclair PA., Reynolds B., Nicholason S., Jones T., Ponbren: The
effects of tree planting on agricultural soils and their functions. Centre for Ecology and
Hydrology Report, 2003.
70

Jackson B.M., Wheater H.S., Mcintyre N.R., Chell J., Francis O.J., Frogbrook Z., Marshall
M., Reynolds B., Solloway I., The impact of upland land management on flooding: insights
from a multiscale experimental and modelling programme, Journal of Flood Risk
Management, Volume 1 Issue 2, p71-80. 2008.
71

Alaoui A., Caduff U., Gerke H.H., Weingartner R., Preferential Flow Effects on Infiltration
and Runoff in Grassland and Forest Soils. Vadose Zone Journal, vol 10, p367 -377. 2011.
72

Nisbet T., Thomas H. and Shah N., Short Rotation Forestry and Water, In: Short Rotation
Forestry: Review of growth and environmental impacts. Ed McKay H., Forest Research,
Forestry Commission, 2011.
73

Calder I.R., Water use of eucalpt a review, In: Growth and water use of forest plantations,
Calder I.R., Hall R.L., Adlard P.G (eds). John Wiley and Sons, Chichester. 1992.
74

David, JS. And Ledger, DC. (1988) Runoff generation in a plough drained peat bog in
Scotland, Jnl of Hydrol. 99 187-199
75

Rutter A.J., Studies in the water relations of Pinus sylvestris in plantation conditions, I:
Measurements of rainfall and interception Journal of Ecology 51 191-203, 1963
76

UK forestry standard water guidelines forests and water, Forestry Commission 2011.

77

Robinson, M Changes in catchment runoff following drainage and afforestation


Journal of Hydrology 86 71-84, 1986
-149-

78

Robinson, M. Moore, RE. Nisbet, TR. Blackie, JR. Report no 133 From moorland
to forest: the Coalburn catchment experiment Institute of Hydrology. 1988
79

Archer, D. Newson, M. (2002) The use of indices of flow variability in assessing the
hydrological and instream habitat impacts of upland afforestation and drainage.
Journal of Hydrology. x 244-258
80

Calder, IR. Harrison, J. Nisbet, TR. Smithers, RJ. Woodland actions for biodiversity and
their role in water management Woodland Trust/Forest Research/Environment Agency. 2008.
81

Robinson M, Cognard-Plancq A-L, Cosandey C, David J, Durnard P, Fuhrer H-W, Hall R,


Henriques MO, Mark V, McCarthy R, McDonnell M, Martin C, Nisbet T, ODea P, Rodgers M,
Zollner A, Studies of the impacts of forests on peak flows and baseflow: a European
perspective. Forest Ecology and Management. 186: 85-97. 2003.
82

McCulloch J.S.G., & Robinson M., History of forest hydrology. Journal of Hydrology, 150:
189-216. 1993.
83

OConnell P.E., Beven K.J., Carney J.N., Clements R.O., Ewen J., Fowler H., Harris G.L.,
Hollis J., Morris J., ODonnell G.M., Packman J.C., Parkin A., Quinn P.F., Rose S.C.,
Shepherd M. and Tellier S., Review of impacts of rural land use and management on flood
generation: Impact study report. DEFRA R&D Technical Report FD2114/TR. 2004
84

Robinson, M. and Dupeyrat, A., Effects of commercial forest felling on streamflow regimes
at Plynlimon. Hydrological Processes 19:1213-1226. 2005.
85

Kirby C., Newson M.D. and Gilman K, Plynlimon research: The first tow decades, Institute
of Hydrology Report 109, Wallingford. 1991.
86

Neal C., Water quality of the Plynlimon catchments. Hydrology Earth Systems Science,
Special Issue, 1, 381-763. 1997.
87

CIFOR, Forests and floods: Drowning in fiction or thriving on facts? 2005

88

Robinson M, Cognard-Plancq A-L, Cosandey C, David J, Durnard P, Fuhrer H-W, Hall R,


Henriques MO, Mark V, McCarthy R, McDonnell M, Martin C, Nisbet T, ODea P, Rodgers M,
Zollner A,. Studies of the impacts of forests on peak flows and baseflow: a European
perspective. Forest Ecology and Management. 186: 85-97. 2003
89

Bates C.G. & Henry A.J., Forest and streamflow experiment at Wagon Wheel Gap,
Colorado, Mon. Weather Rev., Soppl., 30: 1-79. 1928.
90

Hornbeck J.W., Pierce R.S. & Federer C.A., Streamflow changes after forest clearing in
New England. Water Resources. Res., 6: pp1124-1132. 1970
91

Rothacher J. Increases in water yield following clear cut logging in the Pacific Northwest.
Water Resources. Res, 6: 653 658. 1970.
92

Swank W.T. and Crossley D.A. (Editors), Forest Hydrology and Ecology at Coweeta.
Ecological Studies No 66. Springer-Verlag, New York, 469 pp. 1988
93

Naden P., Crooks S. and Broadhurst P., Impact of climate change and land-use change on
the flood response of large catchments. 31st MAFF Conference of River and Coastal
Engineers, Keele University (3-5 July 1966), 2.1.1 2.1.16. 1996.
94

Calder I.R., Evaporation in the uplands. Wiley, Chichester. 148 pp. 1990

95

Hall R.L.,. Processes of evaporation from vegetation of the uplands of Scotland. Trans. R.
Soc. Edinburgh. Earth Sci., 78, pp 327-334. 1987
96

Hall R.L. and Harding R.J., The water use of the Balquhidder catchments: a process
approach. Journal of Hydrology, 145, pp 285-314. 1993.
97

Wright IR and Harding RJ, Evaporation from natural mountain grassland. Journal of
hydrology, 145, pp 267-283. 1993.

-150-

98

Haria A. and Price D.J., Evaporation from native pine woodlands in the Cairngorm
Mountains of Scotland. Hydrology and Earth Sciences, 4(3) pp 451-461. 1999.
99

Price D.J., Kaya Y. & Tingle S., Land-use change and flood attenuation the predicted
potential benefits of semi-natural woodland. 35th MAFF Conference of River & Coastal
Engineers, Keele, July 2000.
100

Francs F., Garca-Bartual R., Ortiz E., Salazar S., Miralles J., Blschl G., Komma J.,
Habereder C., Bronstert A., Blume T., Efficiency of non-structural flood mitigation measures:
"room for the river" and "retaining water in the landscape", CRUE Research Report No I-6
2008
101

Park J.S., Cluckie I. & King P., The Parrett Catchment Project (PCP): Technical Report on
the Whole Catchment Modelling Project. 2006
102

Marshall M.R., Frogbrook Z.L., Francis O.J., Reynolds B., McIntyre N.R. and Wheater
H.S., The impact of upland land management on flooding: preliminary results from a multiscale experimental programme. In: Proceedings of the British Hydrological Society 9th
National Hydrology Symposium, pp. 85-91. 2006.
103

Wheater H.S., McIntyre N., Jackson B.M., Marshall M.R., Ballard C., Bulygina N.S.,
Reynolds B. and Frogbrook Z.,. Flood Risk Science and Management Chapter 3 Multiscale
Impacts of Land Management on Flooding. Ed Pender G, Thorne C, Cluckie I and Faulkner
H. Blackwell Publishing Ltd. 2010
104

Wheater H., Reynolds B., McIntyre N., Marshall M., Jackson B., Frogbrook Z., Solloway I,
Francis O., Chell J., Flood Risk Management Research Consortium, Impacts of Upland Land
Management on Flood Risk: Multi-scale Modelling Methodology and Results from the
Pobntbren Experiment, December 2008, FRMRC Research Report UR16. 2008
105

Von Hoyningen-Huene J., Die interzeption des Niederschalages in landwirtschaftlichen


Pflanzenbestanden. Arbeitsbericht Deutscher verband fur Wasserwirtschaft and Kulturbau,
DVWK, Braunschwig, Germany, 1981.
106

Kozak, J. A., Ahuja, L. R., Green, T. R., and Ma, L., Modelling crop canopy and residue
rainfall interception effects on soil hydrological components for semi-arid agriculture,
Hydrology Process., 21, 229241, 2007.
107

Law, B.E., T. Arkebauer, J.L. Campbell, J. Chen, O. Sun, M. Schwartz, C. van Ingen, S.
Verma. Terrestrial Carbon Observations: Protocols for Vegetation Sampling and Data
Submission. Report 55, Global Terrestrial Observing System. FAO, Rome. 87 pp. 2008
108

Leaf Area Index FPAR MODIS (Shortname MOD15A2)


https://lpdaac.usgs.gov/products/modis_products_table (accessed 7 February 2012)
109

Price J.S., Heathwaite A.L. & Baird A.J. Hydrological processes in abandoned and
restored peatlands: an overview of management approaches. Wetlands ecology and
management 11, 65-83. 2003
110

Ramchunder S.J., Brown L.E. & Holden J., Environmental effects of drainage, drain
blocking and prescribed vegetation burning in UK upland peatlands. Progress in Physical
Geography, 33, 49-79. 200.
111

Labadz J, Allott T, Evans M, Butcher D, Billett M, Stainers S , Yallop A, Jones P,


Innerdales Mike, Harmon N, Maher K, Bradbury R, Mount D, OBrien H and Hart R, Peatland
Hydrology, Draft Scentific Review for IUCN UK Peatland Programmes Commission of
Inquiry on Peatlands, 2010
112

Soulsby C., Tetzlaff D., Rodger P., Dunn S. & Waldron S. Runoff processes, stream water
residence times and controlling landscape characteristics in a mesoscale catchment: An initial
evaluation. Journal of Hydrology, 325, 197-221. 2006
113

Rydin and Jeglum, The biology of peatland. Oxford University Press, Oxford. 2006.

-151-

114

Grayson R,, Holden J, & Rose R., Long-term change in storm hydrographs in response to
peatland vegetation change. Journal of Hydrology 389, 336-343. 2010.
115

Ballard C.E., McIntyre N., Wheater H.S., Effects of peatland drainage management on
peak flows. Hydrology Earth Systems Science, 8, 6533-6563, 2011.
116

Lewis W.K., Investigation of rainfall, runoff and yield of the Alwen and Brenig catchments.
Proc. Instn Civil Engrs 8: 17-51. 1957.
117

Robinson M, Farm drainage a cause for flooding? Agricultural and Advisory Service
Bulletin 159, MAFF, Oxford 10-12. 1985.
118

Robertson R.A., Nicholson I.A. and Hughes R., Runoff studies on a peat catchment. Proc.
2nd International Peat Congress, HMSO, London 161-166. 1968
119

Leeks G.J.L. and Roberts G., The effects of forestry on upland streams with special
reference to water quality and sediment transport. In: Good, JEG (ed), Environmental Aspects
of Plantation Forestry in Wales. Institute of Terrestrial Ecology Symposium 22: 9-24. 1987.
120

Acreman M., The effects of afforestation on the flood hydrology of the upper Ettrick valley.
Scottish Forestry, 39, 89-99. 1985
121

Holden J., Chapman P.J. and Labadz J.C., Artificial drainage of peatlands: hydrological
and hydrochemical process and wetland restoration. Progress in Physical Geography 28, 1,
95-123. 2004.
122

Holden J., Burt T.P., Evans M.G. and Horton M., Impact of land drainage on peatland
hydrology. Journal of Environmental Quality 35, 1764 1778. 2006.
123

Armstrong A., Worrall F. and Holden J., Monitoring grip-blocking techniques. (Whitendale
Trial). Draft Final Report for United Utilities. 2006.
124

Ballard C, McIntyre N and Wheater HS, Peatland drain blocking: Can it reduce peak flood
flows? BHS Third International Symposium, Managing Consequences of a Changing Global
Environment, Newcastle. 2010.
125

Geris J., Ewen J., ODonnell G. and OConnell P., Monitoring and modelling the pre-and
post-blocking hydrological response of moorland drains. BHS Third International Symposium,
Managing Consequences of a Changing Global Environment, Newcastle 2010.
126

JBA, Ripon land management project - Final Report. Department for Environment, Food
and Rural Affairs. 53pp. 2007.
127

Johnson R, Flood Planner: A manual for the natural management of river floods. Report for
WWF Scotland. 32 pp. 2007.
128

Scottish Government, Constructed Farm Wetlands, 2006


http://www.scotland.gov.uk/Topics/farmingrural/SRDP/RuralPriorities/Packages/SustainableFl
oodManagem/Constructionoffarmwetland (accessed 24/2/12)
129

Carty A.H., Sholz M., Heal K., Keohane J., Dunne E., Gouriveau F., Mustafa A.,
Constructed Farm Wetlands (CFW) Design Manual for Scotland and Northern Ireland,
Prepared for the Northern Ireland Environment Agency and the Scottish Environment
Protection Agency, 2008.
130

CIRIA, The SUDS Manual, C697, CIRIA, 2007

131

Wilkinson M.E., Quinn P.F. and Welton P., Runoff management during the September
2008 floods in the Belford catchment, Northumberland. Journal of Flood Risk Management 3,
p285-295, 2010.
132

Environment Agency, Belford Flood Alleviation Scheme, 2011. http://www.environmentagency.gov.uk/homeandleisure/floods/129472.aspx (accessed 22 February 2012)
133

Ackers J. and Bartlett J., Flood storage Chpt 10 In Fluvial Design Guide, Environment
Agency, 2010. http://evidence.environment-

-152-

agency.gov.uk/FCERM/en/FluvialDesignGuide/Chapter_10_Background.aspx (accessed 22
Feb 2012)
134

Hall M.J., Hockin D.L. and Ellis J.B. Design of flood storage reservoirs, B014. CIRIA and
Butterworth-Heinemann. 1993
135

Bullock A. and Acreman M., The role of wetlands in the hydrological society, Hydrology
and Earth Systems Sciences, 7(3), 358-389, 2003.
136

Natural Environment Research Council, Flood Studies Report, 1975.

137

Boorman D.B., Hollis J.M. and Lilly A., Hydrology of soil types: a hydrologically based
classification of the soils of the United Kingdom. Report 126, Institute of Hydrology. 1995.
138

J Poesen, J Nachtergaele, G Verstraeten, C Valentin, Gully erosion and environmental


change: importance and research needs, CATENA, Volume 50, Issues 24, 1, P 91-133,
2003
139

Lane S.N., Catchment-scale approaches to restoring rivers with coarse sediment


problems, Department of Geography, University of Durham. 2006.
(http://www.therrc.co.uk/pdf/References/Lane_2006.pdf accessed 16 February 2012).
140

J.A Martnez-Casasnovas, A spatial information technology approach for the mapping and
quantification of gully erosion, CATENA, Volume 50, Issues 24, 1, P293-308, 2003
141

Zhuo L., The effects of forest in controlling gully erosion. Erosion, Debris Flows and
Environmenta in Mountain Regions (Proceedings of Chegdu Symposium, IAHS Publication
209, 1992.
142

Raven E.K., Lane S.N., Bracken L.J., Understanding sediment transfer and morphological
change for managing upland gravel-bed rivers. Progress in Physical Geography, 34(1) 23-45.
2010.
143

Acreman M, Crook S, Glenny C, Hewitt N, Lucas F, Packman J, Pepper A, Perret L,


Scholey G, Tagg A, Tiller I and Woodhouse C, Hydrological Impact Assessment, Modelling
the impacts of floodplain restoration. EU Life, RSPB, 2002.
144

Blackwell M. and Maltby E., Ecoflood Guidelines, How to use floodplains for flood risk
reduction. EUR 22001, European Commission, Directorate- General Research, 2006.
145

Ahilan S, OSullivan J and Bruen M, Empirical Investigation of the Influence of Floodplain


Attenuation Effects on Flood Flow. Irish National Hydrology Conference, 2009.
146

Bullock A. and Acreman M., The role of wetlands in the hydrological cycle, Hydrological
and Earth Systems Sciences, 7(3), 358-389, 2003.
147

Potter K.M., Wheres the Space for Water? How floodplain restoration projects succeed.
Masters Dissertation for the Department of Civic Design, Liverpool University. 2006.
148

Environment Agency, River Habitat Quality: the physical character of rivers and streams in
the UK and Isle of Man. Environment Agency, Bristol. 1998.
149

Broadmeadow S. and Nisbet T.R., The effects of riparian forest management on the
freshwater environment: a literature review. Hydrology and Earth System Sciences, 8(3),
286-305, 2004.
150

Forest Research, The Robinwood Robinflood report: Evaluation of Large Woody Debris in
Watercourses, Forestry Commission (Wales), INTEREG IIIC, 2008.
151

Linstead C., And Gurnell A.M. Large woody debris in British headwater rivers: Physical
habitat and management guidelines. R&D Technical Report W185, School of Geography and
Environmental Sciences, University of Birmingham, 1999.
152

Gregory K.J., Gurnell A.M. and Hill C.T., The permanence of debris dams related to river
channel processes. Hydrological Sciences Journal des Sciences Hydrologiques, 20, 3, 9,
1985.

-153-

153

Thomas H. Modelling the Hydraulic Impact of Reintroducing Large Woody Debris into
Watercourses. The Journal of Flood Risk Management, DOI: 10.1111/j.1753318X.2010.01137.x, Accepted paper, 2012.
154

Thomas H. & Nisbet T. R., An assessment of the impact of floodplain woodland on flood
flows. Water and Environment Journal, 2006.
155

Nisbet T.R. and Thomas H., Restoring Floodplain Woodland for Flood Alleviation. Forest
Research DEFRA, 2008.
156

Jones M., Farming Floodplains for the future, Staffordshire Wildlife Trust, 2010.

157

Sartor J.F., Flood water retention by riverine and terrestrial forests. Trier University of
Applied Sciences, Schneidershof.
Aatz, M. and Musong, S. Two-dimensional dynamic calculation of natural water courses
(in German), Diploma thesis, Trier University of Applied Sciences (2005)
158

159

Chow V.T., Open-channel hydraulics. McGraw-Hill, 1959.

160

Barnes H.H., Characteristics of Natural Channels, US Geological Survey Water Supply


Paper 1849. USGS, United States Government Printng Office, Washington, 1967.
161

http://www.river-conveyance.net/ces/index.html (referenced 17 January 2011)

162

McGahey, C. & Samuels, P.G., Methodology for conveyance estimation in twostage straight, skewed and meandering channels, XXX IAHR Congress,
Thessaloniki, Volume C1 of Proceedings published by IAHR, 2003.
163

http://www.isisuser.com/ (referenced 17 January 2011)

164

Arcement G.J. and Schneider V.R., Guide for Selecting Manning's Roughness
Coefficients for Natural Channels and Flood Plains. United States Geological Survey
Water-supply Paper 2339. 1989
165

Lane S. N., Roughness time for a re-evaluation, Earth Surface Processes and
Landforms, 30, 251-2. 2005.
166

Whatmore S. and Langstrom C., Mannings n: putting roughness to work, in How well do
facts travel? Morgan M and Howlett P (eds.) Chapter 4, Cambridge University Press,
Cambridge. 2010.
167

Nisbet T.R. and Thomas H., Flooding and Flood Management Inquiry, Rural Affairs and
Environment Committee, Submission from Forestry Commission Scotland, The Role of
Woodland in Flood Control: A Landscape Perspective. Scottish Parliament Archives, 2007.
Accessible at:
http://archive.scottish.parliament.uk/s3/committees/rae/inquiries/flooding/ForestryCommission
Scotland.htm (accessed 16 January 2011)
168

Whatmore S. Understanding environmental knowledge controversies: the case of flood risk


management: Full Research Report, ESRC, RES-227-25-0018. Swindon. 2010.
169

Hardiman N., Cunningham R. & Johnstonova A., Technical review of NFRM techniques,
their effectiveness and wider benefits. RSPB, 2009.
170

Schultz S.D. and Leitch J.A., The feasibility of restoring previously drained wetlands to
reduce flood damage. Journal of Soil & Water Conservation, 58(1) 21-29, 2003.
171

Morris J., Hess T.M., Gowing D.J., Leeds-Harrison P.B., Bannister N., Wade M., Vivash
R.M., Integrated washland management for flood defence and biodiversity, English Nature,
2004.
172

Acreman M, Crook S, Glenny C, Hewitt N, Lucas F, Packman J, Pepper A, Perret L,


Scholey G, Tagg A, Tiller I and Woodhouse C, Hydrological Impact Assessment, Modelling
the impacts of floodplain restoration. EU Life, RSPB, 2002.

-154-

173

Riley A.L., Restoring Streams in Cities. A Guide for Planners, Policymakers and Citizens.
Island Press, 1998.
174

CIRIA, Beach recharge manual. C685, 2011.

175

University of Dundee & Aberdeen City Council, Aberdeen Beach Case Study, IMCORE,
2010
176

Pye K., Samantha S., Blott S., Sand dune processes and management for flood and
coastal defence, Part 2. Defra R&D Technical Report FD1392/TR, 2007.
177

Environment Agency & Royal Haskoning, Saltmarsh Management Manual, DEFRA, 2005.

178

Mller, I., Spencer, T., French, J.R., Leggett, D.J. & Dixon, M. Wave transformation over
salt marshes: a field and numerical modelling study from North Norfolk, England. Estuarine,
Coastal and Shelf Science 49: 411-426. 1999.
179

Mller, I. & Spencer, T. Wave dissipation over macro-tidal saltmarshes: effects of marsh
edge typology and vegetation change. Journal of Coastal Research Special Issue 36, 506521. 2002.
180

Cooper, N. Coastal Data Analysis: The Wash Study 2: Wave attenuation over inter-tidal
surfaces. Report to the Environment Agency. 2001.
181

Mller, I., Spencer, T., French, J.R., Leggett, D.J. and Dixon, M. The sea-defence value of
salt marshes: field evidence from north Norfolk. Journal of Chartered Institution of Water and
Environmental Management (CIWEM) 15: 109-116. 2001.
182

Reed D.J., Davidson-Arnott R. and Perillo G.M.E., Estuaries to coastal dunes: Chapter 5
Estuaries, coastal marshes, tidal flats and coastal dunes . 2009.
183

Rupp S. & Nicholls R.J., Managed Realignment of Coastal Flood Defences: A Comparison
between England and Germany. Flood Hazard Research Centre, 2002.
184

French J.R., Hydrodynamic Modelling of Estuarine Flood Defence Realignment as


Adaptive Management Response to Sea-Level Rise. Journal of Coastal Research, 2008.
185

OMReG database, http://www.abpmer.net/omreg/ (Accessed 20 February 2012)

186

Environment Agency, The Alkborough tidal defence scheme, Fact Sheet, 2006.

187

Leggett D.J., Cooper N., Harvey R., Coastal and estuarine managed realignment design
issues. C628 CIRIA, 2004.
188

SNH, A guide to managing coastal erosion in beach/dune systems.


http://www.snh.org.uk/publications/on-line/heritagemanagement/erosion/sitemap.shtml
(Accessed 24 February 2012)
189

Scyphers S.B., Powers S.P., Heck K.L., Byron D., Oyster Reefs as Natural Breakwaters
Mitigate Shoreline Loss and Facilitate Fisheries. PLoS ONE 6(8): e22396.
doi10.1371/journal.pone.0022396. 2011.
190

OConnell P.E., Beven K.J., Carney J.N., Clements R.O., Ewen J., Fowler H., Harris G.L.,
Hollis J., Morris J., ODonnell G.M., Packman J.C., Parkin A., Quinn P.F., Rose S.C.,
Shepherd M. and Tellier S., Review of impacts of rural land use and management on flood
generation: Impact study report. DEFRA R&D Technical Report FD2114/TR. 2004.
191

JBA Consulting, Delivery of Making Space for Water, HA6 & Ha7, Identification of
catchments sensitive to land use change. Environment Agency, 2008
192

JBA Consulting. Land Management CFMP Tool, Development of a software tool to


investigate the potential impact of changes in rural land use and land management on flood
generation. Environment Agency, 2008.
193

Broadmeadow & Nisbet T., Opportunity Mapping for Woodland to Reduce Flooding in
Yorkshire and the Humber Region. Forest Research, 2009.
194

Halcrow CRESS Allan Water Natural Flood Management Scoping Study. SEPA, 2011.

-155-

195

Pagella T. et al, POLYSCAPE: Multiple criteria GIS toolbox for negotiating landscape scale
ecosystem service provision. Presentation, 2009.
http://www.futureforest.eu/uploads/polyscape_future_forests_09.pdf
196

Watts K. & Eycott A. Biological and Environmental Evaluation Tools for Landscape
Ecology. Forest Research.
197

Hewett et al. Farm and Agriculture Risk Matrix (FARM) technical manual (Newcastle
University, 2006) http://www.ceg.ncl.ac.uk/theFARM/index.html
198

Odoni N. & Lane S., Assessment of the impact of upstream land management measures
on flood flows in Pickering Beck using OVERFLOW. Durham University, Forest Research
and Defra, 2010.
199

Lane, S.N., Reaney, S.M. and Heathwaite, A.L., 2009. Representation of landscape

hydrological connectivity using a topographically driven surface flow index. Water Resources
Research, 45, W08423, doi:10.1029/2008WR007336.
200

Porter R.A.I., An Evaluation of a Hydrological Model Used to Predict the Impact of Flow
Attenuation on Downstream Flood Flows. Masters thesis, Durham University. 2011 Available
at Durham E-Theses Online: http://etheses.dur.ac.uk/813/ (accessed 21 February 2012)
201

Leopold L.B. and Maddock T., The Hydraulic Geometry of Stream Channels and Some
Physiographic Implications. US Department of the Interior. 1953.
202

ADAS UK & Macaulay Land Use Research Institute, Provision of a Screening Tool to
Identify and Characterise Diffuse Pollution Pressures: Phase II. SNIFFER, 2006.
203

Beven K.J., Rainfall-Runoff Modelling: The Primer. Second Edition, Wiley-Black, 9780470714591, 2012.
204

Beven K.J., Changing ideas in hydrology the case of physically-based models. Journal of
Hydrology. 105: 157-172.
205

Bulygina N., McIntyre N., Wheater H., Conditioning rainfall-runoff model parameters for
ungauged catchments and land management impacts analysis. Hydrological Earth Systems
Science, 2009.
206

Moore, R.J.; Bell, V.A.; Cole, S.J.; Jones, D.A.. Rainfall-runoff and other modelling for
ungauged/low-benefit locations. Research Contractor: CEH Wallingford, Environment
Agency, Bristol, UK, 249pp. 2007
207

K) Louisianas 2012 Coastal Management Plan - Planning Tool


http://www.coastalmasterplan.louisiana.gov/leading-the-way/planning-tool/ (accessed 26
January 2012)
208

Kjeldsen T., Jones D., Bayliss A., Spencer P., Surendran S., Laeger S., Webster P.,
McDonald D., Improving the FEH Statistical Method. In: Flood & Coastal Management
Conference 2008. Environment Agency & Defra, 2008.
209

Kjeldsen T., Jones D., Bayliss A., Spencer P., Surendran S., Laeger S., Webster P.,
McDonald D., Improving the FEH Statistical Method. In: Flood & Coastal Management
Conference 2008. Environment Agency & Defra, 2008.
210

Parker, C., Thorne, C.R., and Clifford, N.J. Broad. A broad-scale assessment of sediment
dynamics in British rivers: Developing and assessing ST:REAM a reach based sediment
balance model. Earth Surface Processes and Landforms - In review.
211

ABPmer, The Met Office & Proudman Oceanographic Laboratory, Atlas of UK Marine
Renewable Energy Resources, Commissioned by DTI and produced by y (2008)
http://www.renewables-atlas.info
212

CIRIA, The Rock Manual, C683, CIRIA 2007.

-156-

213

Werritty A., Ball T., Spray C., Bonell M., Rouillard J., Archer N., Bowles C. and Moir H.,
Restoration Strategy: Eddleston Water Scoping Study. Dundee University and Cbec. 2010
214

Faulkner D. S., Rainfall Frequency Estimation. Flood Estimation Handbook Vol2.


Institute of Hydrology, 1999.
215

Institute of Hydrology, Design flood estimation in catchments subject to urbanisation.


Flood Studies Supplementary Report No. 5, Institute of Hydrology, Wallingford. 1979.
216

Hough M.N. and Jones R.J.A., The United Kingdom Meteorological Office rainfall and
evaporation calculation system: MORECS version 2.0 an overview. Hydrology and Earth
Systems Sciences, 1(2), 227-239, 1997.
217

Hough M.N. and Jones R.J.A., The United Kingdom Meteorological Office rainfall and
evaporation calculation system: MORECS version 2.0 an overview. Hydrology and Earth
Systems Sciences, 1(2), 227-239, 1997.
218

Nisbet T.R., and Thomas H., The role of woodland in flood control: a landscape
perspective Forest Research, 2008
219

Moore R.J., The probability-distributed principle and runoff production at point and basin
scales. Hydrological Sciences Journal, 30, 273 to 297. 1985.
220

Moore R.J., The PDM rainfall-runoff model. Hydrology and Earth System Sciences, 11(1),
483 to 499, 2007.
221

Bell V.A., and Moore R.J., A grid-based distributed flood forecasting model for use with
weather radar data: Part 2. Case studies Hydrology and Earth Systems Sciences, 2(3), 283298, 1998.
222

Bell V.A., Kay A.L., Jones R.G., Moore R.J., Development of a high resolution grid-based
river flow model for use with regional climate model output. Hydrology Earth Systems
Science, 11(1), 532-549, 2007.
223

Bayliss A.C., Catchment descriptors. Flood Estimation Handbook, vol 5. Institute of


Hydrology, Wallingford, UK. 1999.
224

Hough M., An historical comparison between the Met Office Surface Exchange SchemeProbability Distributed Model (MOSES-PDM) and the Met Office Rainfall and Evaporation
Calculation System (MORECS), Met Office 2003.
225

Robson A.J. and Reed D.W., Statistical procedures for flood frequency estimation. Flood
Estimation Handbook Vol 3., Institute of Hydrology, Wallingford, 1999.
226

USDA, National Engineering Handbook, Part 630 Hydrology. Chapter 15 Time of


concentration. United States Department of Agriculture, Natural Resources Conservations
Service, 2010.
227

Reed, Duncan, Faulkner, Duncan, Robson, Alice, Houghton-Carr, Helen, Bayliss, Adrian, ,
Flood estimation handbook: Vol. 4: Restatement and application of the flood studies report
rainfall-runoff method, Institute of Hydrology, 1999.
228

Beven K.J. and Binley A.M., The future of distribute models: Model calibration and
uncertainty prediction. Hydrological Processes, 6(3): 279-298, 1992.

-157-

Appendix A
Review of Jacobs Section 20 Review
(CREW, 2012)

Page | 0

Review of Jacobs Section 20 Review

Final Report
Project CD2011_07
20/01/2012

Page | 0

CREW 2012
The views expressed in this document are not necessarily those of CREW. Its members or agents accept
no liability for any loss or damage arising from the interpretation or use of the information, or reliance
upon views contained herein.
Whilst this document is considered to represent the best available scientific information and expert
opinion available at the stage of completion, it does not necessarily represent the final or policy
positions of the project funders.

Dissemination status
Unrestricted

Research contractors
This document was produced by:
1) Prof A Werritty and Dr T. Ball
School of the Environment, and
UNESCO Centre for Water Law, Policy and Science,
University of Dundee,
Dundee DD1 4HN,
Scotland, UK

2) Dr Scott Arthur

Institute for Infrastructure and Environment,


School of the Built Environment,
Heriot-Watt University,
Edinburgh EH14 4AS,
Scotland, UK

Centre of Expertise for Waters (CREW)


CREW is a hub which ensures that water research and expertise is available and accessible to the
Scottish Government and its agencies. This is designed to ensure that existing and new research and
expertise can feed into the development of water related policy in Scotland in a timely and effective
manner.

CREW Management
All queries related to this document should be directed to the CREW Facilitation Team

James Hutton Institute


Craigiebuckler
Aberdeen AB15 8QH
Scotland UK
Tel: +44 (0) 844 928 5428
Email: enquiries@crew.ac.uk
www.crew.ac.uk

Executive Summary
Project reference: Review of Jacobs Section 20 review and GIS tool (January, 2012)

Project contractors: University of Dundee and Heriot-Watt University

Background to research
Section 20 of the Flood Risk Management (Scotland) Act 2009 requires SEPA, by 22 December 2013, to
undertake national assessment of natural flood management (NFM), whether alteration (including
enhancement) or restoration of natural features and characteristics of any river basin or coastal area in
a flood risk management district could contribute to the management of flood risk for that district.

Objectives of research
To review the Jacobs section 20 review and GIS tool based on a set of pre defined questions.

Key findings and recommendations


NA

Key words
Flooding, Natural Flood Management, GIS, Land Use

1. INTRODUCTION
Guide questions to guide expert reviews of the Section 20 methodology were supplied by
SEPA.
1.1 Questions
1.1.1 Section 2 Literature review:
Q1) The literature review was limited to those studies that provided quantitative evidence of the
effects of NFM techniques on catchment scale flooding. In your view, is this the right approach
or have any important studies been missed?
Q2) In your view, are the studies identified in the literature review representative of Scottish
catchments?
Q3) Do you agree with conclusions and recommendations in sections 2.2.4, 2.3.4 and 2.4.4?
1.1.2 Section 3 GIS tool:
Q4) Do you agree with the outlined approach described in section 3.1?
Q5) Do you have any observations in relation to the testing of the model described in section
3.3?
Q6) In your opinion, is the preliminary rule set described in section 3.4 acceptable and are there
sufficient grounds for the rule set? Suggest, if possible, alternative ways to improve the rule
set.
Q7) In your opinion is the outline method for coastal floodplains correct?
Q8) Do you agree with the limitations specified in section 3.7 and recommendations in section
8.8?
Q9) Do you have any other observations from Section 3 that you would like to highlight?
Q10) Overall, would you recommend the use of the proposed GIS tool and the rule for use as
the screening tool to provide an indication of the effect of NFM techniques?
Q11) Do you have any recommendations for future work that would be beneficial to further
develop the rule set and the modelling tool?

2. REVIEW OF JACOBS SECTION 20 REVIEW AND GIS TOOL BY ALAN WERRITTY AND
TOM BALL, UNIVERSITY OF DUNDEE.
2.1 Overview
The task of providing a GIS tool to deliver the Section 20 requirements in the Flood Risk
Management (Scotland) Act 2009 was a very challenging one. As noted in Werrittys earlier
CREW report on the SAIFF Task and Finish group Land Management and Flooding paper, the
evidence base for effective and efficient NFM measures is still very fragile with many key issues
yet to be resolved. The most notable of these are:

Quantifying the effects of individual NFM measures in reducing flood risk across
catchments with diverse characteristics (effectiveness)
Determining how individual measures at a local scale contribute to flood risk reduction at
the catchment scale (scaling)
Reducing uncertainty by developing better physically-based rainfall-runoff models
(modelling)
Deploying measures within a catchment in an optimal way so that those installed
upstream do not adversely impact on communities downstream and vice-versa
(location)
Future-proofing the selection of measures to include potential impacts of climate
(changing risk)
Developing measures which, in terms of ecosystem services, provide other catchmentbased benefits in terms of enhanced water quality, protecting biodiversity and storing
carbon (multiple benefits).

The comments below should be read with these issues in mind by way of context.
2.2 Review:
Q1) The literature review was limited to those studies that provided quantitative evidence
of the effects of NFM techniques on catchment scale flooding. In your view, is this the
right approach or have any important studies been missed?
The literature review is exhaustive, authoritative and up to date. it is also extremely well
structured with a clearly specified context, identification of uncertainties (taken from
Environment Agency, 2007) and prefaces to each major group of measures (upland forest
cover, riparian woodland, floodplain woodland, benefits of upland drain blocking, benefits of
wetlands and floodplains) with an informed commentary. Each of the three major groups of
measures (upland forests and woodlands, upland drain blocking and wetlands and floodplains)
is reported on in a series of summaries. The quality of the reflective commentary in the
prefaces to each of the three major groups of measures is very impressive. The review is
written in very clear English with technical terms largely avoided or, where they are used,
suitably explained in footnotes.
Table 2.1 which itemises potential NFM measures and where they are likely to generate the
optimal benefit includes the following measures which are not addressed in the rest of the
literature review nor via the GIS tool:

Reducing the impact of sheep grazing


Blocking field drains (mentioned in 2.3.1.c but excluded from further consideration)
Land and soil management activities to delay surface water runoff
Restoring rivers and river channels
Re-meandering
Land-use change arable reversion

This is acknowledged in the statement (section 2.1.1) that other techniques will be considered
as the Section 20 methodology and GIS tool continues to be further enhanced. But for now the
Section 20 GIS tool has excluded any assessment of changes in agricultural practice most
notably reduced stocking densities and land-use change both of which could offer significant
NFM benefits. In future development of the Section 20 GIS tool, these omissions will need to be
addressed.
The rest of section 2 focuses on three groups of NFM measures relating to upland forests and
woodland, upland drain blocking and wetlands and floodplains. Each of these groups of
measures is commented upon in turn:

1.1 Upland forests and woodlands


This is a detailed and balanced account of a wide ranging literature. There are no major studies
with relevant quantitative findings that have been omitted. Section 2.2.1 which outlines key
issues that need to be noted in assessing the potential benefits of forest and woodland-based
NFM measures is especially useful with its emphasis on operational issues and local
environmental controls. My only minor criticism is that section 2.1.4 on the importance of
sediment fluxes in headwater reaches (reported in Raven et al., 2009 and Nisbet et al., 2011) is
not further developed in subsequent sections. As in subsequent sections, there is a useful
distinction between findings based on catchment studies and modelling.
The most robust quantitative findings relate to upland forests, but as the authors note (and
Table 2.5 records) there is a high level of uncertainty in terms of potential reduction in peak
flows. The authors explicitly state that there is no weighting of the findings reported in Table
2.5. In that three of the key studies (Poyo, Kamp and Iller Tables 2.4 and Table 2.5) are from
environmental settings far removed from those found in Scotland, this was unwise. Section
2.2.3.b on hydraulic roughness is very useful for both including tabulated values for Mannings n
(Table 2.7) plus photographs of roughness introduced by woody vegetation from Acrement and
Schneider (1989). But missing from the review is reference to Lanes (2005) commentary on
the use of Mannings n in modelling studies and Whatmore and Langstroms (2011) follow up
book chapter. Although both items focus on how to optimise the calculation of Mannings n in
open channels, they offer a valuable critique of the use of this parameter to estimate roughness
which is of wider interest. Table 2.5 usefully includes % forest cover.
The review of floodplain woodlands which includes a section on woody debris, (2.2.2.c often a
contentious NFM measure because of potentially blocking downstream bridges and culverts) is
well-judged. As the authors note, only the Laver, Pickering and Parrot studies provide usable
quantitative data on the potential impacts of floodplain woodlands.
1.2 Upland drain blocking
Again this is an exhaustive, balanced and well-crafted review, especially useful in firmly
rejecting the off-reported claim that upland peat bogs act as sponges with the potential to
significantly reduce downstream flood risk if appropriately managed. The preface to the
literature review (section 2.3.1) again provides a valuable reflective commentary by way of
context. Especially important in importing quantitative estimates of the impacts of drain blocking
is the discussion on former practice of ploughing of deep ditches prior to planting commercial
conifers. Although, following adherence to Forest and Water Guidelines (Forestry Commission
2003), this practice has since been abandoned, it would have unpinned the ground preparation
for planting at the Blacklaw Moss, Llanbrynmair and Coalburn catchment study sites all key
sites in providing quantitative estimates of impacts on runoff generation. In addition, as section
2.1.3.d makes clear, key information on the age, nature and spatial extent of upland drainage is
rarely available, further adding to the challenge of GIS modelling of upland drain blocking.
The key references for providing quantitative estimates of the impacts of drainage ditch blocking
have been identified. But the initial values against which reductions in peak flows and time to
peak are calibrated as ditches degrade and infill should be used with caution (especially the
Coalburn study). Subsequent refinements to land preparation practices prior to planting mean
that recent and present-day upland drainage will produce lower increases in peak flow and
modest changes in hydrograph shape. Accordingly, blocking recently excavated ditches may
have less of an impact than in the studies dating back to the 1970s and 80s.
As already noted, section 2.1.3.c explicitly excludes field drainage from the review. This
measure, designed to control the position of the water table, is found mainly in the lowlands on
arable or improved pasture where it is used to improve yields and the quality of grazing. By
controlling the volume and rate of runoff, field drains can potentially impact on downstream flood

risk. The precise nature of this impact is contested (Robinson, 1990), but it should be included
in any further developments of the Section 20 GIS tool.
1.3 Wetlands and floodplain
This is another thorough and detailed review with all the major sources incorporated. The
preface to the wetlands review (section 2.4.1) provides a valuable summary of the key findings
from the SNIFFER wetland typology. The associated Table 2.9 summarising the character of
each type of wetland (plus linked photographs) is useful, and needs to be cross-related to Table
3.1 which identifies potential NFM measures and associated GIS tool rule sets. It is good to
note the major reservations in terms of providing unqualified estimates of downstream benefits
of many of the measures reviewed in section 2.4.1. In particular, we endorse the statement that
there is less certainty about the potential benefits of different wetland types on flood risk
reduction and accept that this section provides a more speculative approach than other
sections.
The most important issue when using wetlands and floodplain to promote NFM is the
channel/floodplain interaction and the ratio of water stored (on the floodplain) and conveyed
down the channel. The cautionary advice (Section 2.4.3 c) that the Section 20 GIS tool should
only be used to provide a broad indication of possible benefits (and sometimes disbenefits) is
well judged as it the recommendation that should any measures be considered further they
must be more fully appraised by hydraulic modelling for the river and floodplain in question.
This points up the difficulty of devising a generic GIS tool to assess the NFM potential
floodplains with diverse properties in terms of areal extent, slope, connectivity with the channel,
and location and nature of any flood embankments. Especially important is the claim that
agricultural embankments can help reduce downstream flood risk by, once locally breached,
impounding significant volumes of water. Thus the often favoured measure of setting existing
flood embankments further back on the floodplain is not universally beneficial.
As a result of these qualifications, it is much more difficult to identify the quantifiable benefits of
wetlands and floodplains in terms of NFM. The particularity of each floodplain site and its local
connection with the channel can result in both enhanced and reduced downstream flood risk
depending on its areal extent, slope, and location and nature of existing flood embankments.
As the review makes clear, this is the most challenging group of NFM measures to incorporate
into a GIS-based assessment.
Q2) In your view, are the studies identified in the literature review representative of
Scottish catchments?
Each major group of measures is commented on separately:
1.4 Upland forests and woodlands
The section on upland forest reports many studies from around the world including classic
paired catchment projects in Colorado (Bates and Henry, 1928), New England (Hornbeck et al.,
1970), Oregon (Rochacher, 1970) and North Carolina (Swank and Crossley, 1988). Whilst
valuable in summarising the impacts of forests on the water cycle, these studies rarely provide
detailed information on flood runoff generation and are drawn from settings in which the tree
species, soils, topography and climate are markedly different from Scotland. As already noted,
the results from spatially-distributed process-based models under semi-arid, sub-alpine and
alpine climates in Poyo (Spain,) Kamp (Austria) and Germany (Iller) are used to compile key
findings in Table 2.2. The transferability of these findings (with no differential weightings) is
questionable. As was discovered by the Institute of Hydrology over 20 years ago, even the
water yields from coniferous forests and grasslands in Plynlimon (Wales) could not be replicated
in the Balquhidder (Scotland) due to the higher incidence of snow and contrasting plant and
scrub species in the latter (Hall, 1987).

1.5 Upland drain blocking


The key studies from which quantitative values in the rules sets are based are the Coalburn
(Lake District) and Blacklaw Moss studies (Renfrewshire). These are both likely to yield values
readily transferable to Scotland. But care should taken in selecting which values to transfer.
Findings specific to gripping (still practiced across the peatlands of Northern England to improve
habitat for grouse and sheep) are likely to be less relevant for Scotlands where much of the
focus on sporting estates is on deer management. In addition, since the Second World War, a
higher proportion of upland Scotland has been planted with conifers and broadleafs than in
England. With the passage of time, these forest drains will already be partially blocked and
infilled.
1.6 Wetlands and floodplain
As already noted, this literature is the most problematic in extracting key findings. Given the
prevalence of marshy grassland, montane grassland, wet heath and peat bog in Figure 2.7
(Relative abundance of wetland types in Scotland) it would be wise to privilege those studies
which explicitly focus on these wetland habitats. Paucity of studies on the impact of wetland
restoration places undue weight on the results from the Glendey wetland site reported by
Johnson and McQuat (2008). This is a very small demonstration site (0.0175 km 2) designed to
illustrate the combined impacts of drain blocking, restoration of a meandering channel and
planting of trees across the wetland. Given its function as a NFM demonstration site, it is not
possible to disentangle the impact of individual measures. Furthermore, the reliability of the
downstream runoff record has been questioned. Until these results have been subject to peer
review, they should be regarded as provisional.
Findings from other studies in the Broads, the Long Eau in Lincolnshire, the Quaggy River in
London, River Suir in SE Ireland, the River Cherwell in Oxfordshire and the Rivers Sow and
Penk in Staffordshire provide very site- specific changes in flood flows all of which help identify
potential benefits of wetland and floodplain-based measures. Given that many Scottish
floodplains are likely to replicate the conditions found at one or more of these sites, they
illustrate what may be achievable. But care should be taken not to over-interpret these findings.
The conclusion from the Farming Floodplains for the Future project (Jones, 2010) is worth
repeating:
the functionality of the main floodplains meant that beneficial results would only be achieved by
working at a whole catchment scale, focusing upstream in headwaters and on tributaries,
storing and attenuating flood flows much closer to their source (page 4).
Q3) Do you agree with conclusions and recommendations in sections 2.2.4, 2.3.4 and
2.4.4?
Each of these three summary sections seeks to identify the robust estimates of the impact of
specific NFM measures. Caught between the need, on the one hand to produce a single value,
or set of values for use in the GIS tool, whilst on the other hand reflecting current levels of
uncertainty, these recommended values are inevitably going to be questioned.
2.2.4 Forest Cover
The values reported are taken directly from Tables 2.4 and 2.5. We have already expressed
some reservations about lack of weighting and the transferability of some of these values and
the need to factor in the % of forest cover involved. With this is mind we would discount the
findings from the USA clearcut studies and the model-based results from the Poyo, Kamp and
Iller studies and note that major reductions in the 0.5% and 20% annual probability flows at
Pontbren are based on 100% forest cover. At Pontbren when the impacts of shelter belts only
are modelled, the flood risk benefits dramatically reduced. Accordingly, the 40% optimistic
reduction in 20% annual probability flows is thought to be too high and the most likely value of

25% should be reduced to 20%. The proposed 10% most likely attenuation for 0.5% annual
probability flows is credible.
2.3.4 Riparian woodland
We agree that the Laver and Pickering studies are the most robust in providing estimates for
flood risk reduction. But the 60 minute delay for upper and middle catchment locations strongly
reflects the local contexts. For example, the Pickering study involved both planting riparian
woodland and installing a large number of large woody debris dams. In modelling the impact of
these measures Odoni and Lane (2010) observe that:
the sizes of the peak flow and excess flood volume reductions are lower for
smaller events. This is an important property of the kinds of interventions being
discussed: as the size of the event increases, so the contribution that the
interventions make to reducing flood risk increases. It arises from the fact that
bigger events are more able to transfer water into the interventions .... and reflects
the general property of the Pickering Beck system viz. that the river is relatively
incised into floodplain material, reducing the natural attenuating effect of
floodplains on peak flows.
This illustrates the importance of local topography. They also add the caveat that location vis-vis the urban site to be protected is an important factor with interventions in the upper part of the
catchment generally preferred.
Nisbet et al. (2011) reporting on the Laver study also report that riparian woodland can be used
to desynchronise flood flows from specified tributaries (in their case the Laver and the Skell)
with reduction in downstream flood risk (in this case Ripon). This illustrates the need to take
into account downstream conveyance and contributions from other tributaries,
Given no other relevant studies available the proposed rule set is sound, but should be used
with due regard for location within the catchment.
2.4.4 Floodplain woodland
Again given the absence of other studies, the recommended rule set is endorsed, but again with
many caveats. As the Laver and Pickering studies demonstrate, local context can markedly
influence the impact of floodplain woodland and, as the photos in this section of the report
reveal, actual values for Mannings n are strongly influenced by the extent and nature of the
woody species present. The 300 mm increase in water level is thus likely to include much local
variation.
[Note: In seeking to evaluate the GIS tool, we sought to obtain access to the software
developed by Jacobs. Although the code was released by SEPA in late December, this was not
supported by access to the underlying data due to licensing issues. The answers to questions
4-11 are thus restricted to what can be gleaned from the report and are inevitably more partial
than we would have wished.]
Q4) Do you agree with the outlined approach described in section 3.1?
We commend the overall approach which builds on the findings reported in the literature review.
The limitations are clearly identified and the absence of measures involving channel
enhancement and floodplain reconnection noted. There is, however, one fundamental
weakness in the design of the tool the division of the catchment into equal sections based on
travel times down the main stem (upper, middle and lower) irrespective of where the
downstream outlet point Tp is located. If the tool is always used for a large catchment in which
there are clearly defined upland reaches (small tributaries, flowing through moorland and/or

forests into main stems with minimal floodplain development upper) followed downstream by
an emerging floodplain and larger tributaries (middle) which eventually gives rise to an
extensive floodplain in which the bulk of the flow is along the main stem (lower), then the tool is
appropriate. But if the downstream outlet point Tp is much further upstream (eg in which upland
catchment characteristics dominate) thus precluding this balanced mix of reaches, the results
will reflect the imposition of a tri-partite division which fails to reflect the actual characteristics of
the catchment. In addition, the arbitrary allocation of equal weighting (33% of average travel
time from source to outlet down the main stem) to each upper middle and lower reach will not
be appropriate for all catchments. The next iteration of the tool should allow users to define the
mix of upper, middle and lower segments of the catchment.

Q5) Do you have any observations in relation to the testing of the model described in
section 3.3?
Inevitably testing a model of this kind poses many challenges and given prior information on the
NFM opportunities afforded by the Enrick, this catchment has much to offer. But there are gaps
in the information provided which make it difficult to assess the robustness of the derived rule
sets in section 3.4. These are:

Only one flow gauge is identified and this is in the middle reaches so any flow
modelling results downstream cannot be compared with actual data.
The removal of the floodplain reach at Corrimony and Loch Mieklie (which is immediately
upstream of the only flow gauge) provides a valuable simplification of the catchment
characteristics but severely weakens the robustness of the results which are used to
derive the rule.
The quality of the DTM is not fully specified. How dense was the network of crosssections across the floodplain and channel? What was the extent of the LiDAR
coverage?

This points to a more fundamental problem the extent that the Enrick (or any single
catchment) can provide a robust calibration of the GIS tool set. The values reported for the
Enrick inevitably reflect a highly elongate catchment with a channel network with tributaries
steadily decreasing in size as they flow into the main stem in that sense it cannot be seen as
generic. Standard hydrologic theory tells us that the changing hydrograph along the main stem
will strongly reflect the shape of the catchment and drainage network. Were the catchment to
be roughly circular with most of the tributaries flowing into the main stem very close to the outlet,
the shape of the hydrograph along the main stem would be very different to that on the Enrick.
Similarly, if all the tributaries were to be of approximately the same size and equally spaced, the
main stem hydrograph would differ again. Clearly, the tool cannot be tested against all these
possibilities although we do recommend later that it be re-calibrated against the Eddleston
Water for which there is now a dense network of flow gauges and extensive 2D modelling of
floodplain inundation at the 0.05 and 10% flood probabilities But it would also be worth
undertaking further modelling with idealised channel networks (in quasi-circular and quasi-fishbone networks) to determine how far catchment shape and drainage network properties
condition the findings.
The development of rule sets based on key quantitative findings from the literature is clearly
developed and, with the reservations already noted on some of the values imported in terms of
forest cover see question 3 above generally works well. Clearly the weakest part of the
calibration of the GIS tool relates to the generic model of floodplain enhancement. Whilst the
methodology cannot be faulted, its heavy reliance on the Enrick site (with its lack of information
on the quality of the DTM, and the removal of part of the floodplain at Corrimony and Lock
Meiklie) raises serious questions as to the transferability of the findings. By the authors own
admission this is the least robust component of the GIS tool and should be subject to much
more scrutiny and testing before it is recommended for operational use.

Q6) In your opinion, is the preliminary rule set described in section 3.4 acceptable and
are there sufficient grounds for the rule set? Suggest, if possible, alternative ways to
improve the rule set.
Table 3.1 which cross references potential NFM measures (for which rule sets have been
developed) against a typology of wetlands provides an appropriate framework. But the forest
cover and riparian woodland rule sets are strongly conditioned by the arbitrary demarcation of
every catchment into 33% upper, 33% middle and 33% lower. As noted above (question 4)
this straight jacket needs to be loosened with consequential changes in how the rule sets are
used operationally. The current rule sets in Tables 3.2, 3.3 and 3.4 are contingent on the 33%
division of the catchment into three reaches.
As already noted (question 5) the derivation of the floodplain enhancement rule set strongly
reflects the topography and hydrology of the Enrick test catchment. Given the qualifications
already noted, top priority should be given to testing this rule further ideally against
catchments for which better flow data, and 2D modelling of flood inundation exist. Given the
limitations on the application of this particular component of the GIS tool noted by the authors, it
may not yet be possible to model this floodplain enhancement NFM measures without recourse
to site specific 1D/2D models. The use of an Efficiency Factor whilst conceptually valid, is
weakened by the lack of evidence and supporting case for the values reported in Table 3.8.
There is no reference to the literature review, nor any model results, to justify the values
proposed.
Q7) In your opinion is the outline method for coastal floodplains correct?
The outline method quantifies the potential for flood storage by an inundated area of coastline
subject to managed realignment. The difficulty in applying this approach to quantification of
flood risk reduction is the inherent nonlinearities that must be encompassed in an assessment
of flood peak reduction when applied to tidal surge conditions. This is especially true where the
realignment is provided by a breach in flood embankments rather than a wholesale removal,
usually the most cost-effective option. Thus, although the outline methods is not correct in a
scientific sense (to be so, it would need to incorporate surge characteristics that are highly
specific to the site in question, precluding generic assessment), it does represent an
assessment of theoretical maximum storage potential under optimal scenarios.
Q8) Do you agree with the limitations specified in section 3.7 and recommendations in
section 3.8?
The limitations identified in section 3.8 are generally well-specified. But the second limitation,
relating to the floodplain enhancement tool, fails to take into account undue reliance on results
specific to the topography and hydrology of the Enrick (addressed in question 6 above). We
strongly endorse the final limitation which suggests that lower reaches in large river systems
may require further consideration and some adjustment. As already noted, the rule sets for
lower reaches in Tables 3.2, 3.3 and 3.4 are contingent on the arbitrary demarcation of every
catchment into 33% upper, 33% middle and 33% lower reaches. Revisiting this demarcation
is likely to result in major adjustments to the values for lower reaches. In addition, issues
relating to the robustness of the rule set for floodplain enhancement will be most acute in the
lower reaches of large river systems. We also underline the statement that channel/floodplain
interactions and processes are invariably complex and outcomes are not always as expected.
Floodplain enhancement measures may well produce disbenefits as well as benefits, and these
will need to be thoroughly assessed, ideally via dynamic modelling.
Q9) Do you have any other observations from section 3 that you would like to highlight?

Our main concerns arising from Section 3 have already been itemised above. Here we flag up
the most significant ones:

undue reliance on calibration of rule sets from the Enrick catchment which inevitably is
not generic for other Scottish catchments
the straight jacket that all basins are assigned 33% proportions to upper, middle, and
lower reaches raises major issues in terms of the operational use of the GIS tool.

Q10) Overall, would you recommend the use of the proposed GIS tool and the rule for
use as the screening tool to provide an indication of the effect of NFM techniques?
As noted at the beginning of this commentary, the task to develop a GIS tool to deliver the
Section 20 requirements in the Flood Risk Management (Scotland) Act 2009 was a very
challenging one. Jacobs are commended for the thoroughness of their literature review and the
innovative way they have sought to translate a very partial evidence basis into an operational
GIS tool. They have skilfully circumvented the twin perils of over-interpreting a fragile evidence
base on the impacts of NFM measures, whilst at the same time developing a well-calibrated GIS
tool applicable to catchments the length and breadth of Scotland.
At this stage in the
development of flood risk science, there are few alternative approaches available. However, it
would be of great interest to compare the findings from the Jacobs GIS tool on the Allan Water
with those produced by the Centre for River EcoSystems Science/Halcrows Allan Water Natural
Flood Management Techniques and Scoping study report for SEPA (2011). This would begin to
address many of the issues raised above re undue reliance on the Enrick results. We much
regret that although the GIS tool code was released to us by SEPA, data licensing constraints
prevented us from running it on the Allan Water.

Q11) Do you have any recommendations for future work that would be beneficial to
further develop the rule set and the modelling tool?
There are many recommendations implicit in the earlier comments. Here we gather together
the key ones:

The forest cover and riparian woodland rule sets are strongly conditioned by the arbitrary
demarcation of every catchment into 33% upper, 33% middle and 33% lower. This
straight jacket needs to be loosened with consequential changes in how the rule sets are
used operationally. The next iteration of the tool should allow users to define the mix of
upper, middle and lower segments of the catchment.
The outcomes of the existing rule set for the GIS tool should be tested on the Eddleston
Water for which there is now a dense network of flow gauges and extensive 2D
modelling of floodplain inundation at the 0.05 and 10% flood probabilities.
Further modelling be undertaken with idealised drainage networks (in quasi-circular and
quasi-fish-bone networks) to determine how far catchment shape and drainage network
properties condition the outcomes of the existing rule set
The rule set relating to floodplain enhancement is very dependent on its calibration on
the Enrick Water. Top priority should be given to testing this rule further ideally against
catchments for which better flow data, and 2D modelling of flood inundation exist.
Initially this could be undertaken using the extensive database now available for the
Eddleston Water.
Given the limitations on the application of this particular component of the GIS tool, it
may not yet be possible to model floodplain enhancement NFM measures without
recourse to site-specific 1D/2D models.

The coastal inundation component of the tool be also subjected to site-specific testing in
relation to tidal dynamics and surges, looking particularly at the nature of the breach in
the site defences.
Table 2.1 itemises the full range of NFM measures. As explained in section 2.1.1 some
of these have yet to be included in the GIS tool. In due course the following measures
will need to be incorporated:
- Reducing the impact of sheep grazing
- Blocking field drains
- Land and soil management activities to delay surface water runoff
- Restoring rivers and river channels
- Re-meandering
- Land-use change arable reversion

References
Acrement G J and Schneider V R (1989) Guide for Selecting Mannings Roughness Coefficients
for Natural Channels and Floodplains. US Geological Survey Water Supply Paper 2339.
Bates C G and Henry A J (1928) Forest and streamflow experiment at Wagon Wheel Gap,
Colorado, Monthly Weather Review, Suppl. 30, 1-79.
Environment Agency (2007) Delivery of Making Space for Water: HA6 Catchment Scale LandUse Management: HA7 Land Management Practices, Final Report.
Forestry Commission (2003) Forest and Water Guidelines, Forestry Commission.
Hall R L (1987) Processes of evaporation from vegetation of the uplands of Scotland,
Transactions of the Royal Society of Edinburgh, 78, 327-334.
Hornbeck J W, Pierce R S and Federer C A (1970) Streamflow changes after forest clearing in
New England, Water Resources Research, 6, 1124-1132.
Johnson R, Watson M and McQuat E (2008) The Way Forward for Natural Flood
Management in Scotland, Report for Scottish Environment LINK
Jones M (2010) Farming floodplains for the future, Final Report, Staffordshire Wildlife Trust.
Lane S N (2005) Roughness time for a re-evaluation, Earth Surface Processes and
Landforms, 30, 251-2.
Nisbet N, Silgram M, Shah N, Morrow K and Boardmeadow S (2011) Woodland for Water:
Woodland Measures for meeting Water Framework Objectives, Environment Agency.
Odoni N A and Lane S N (2010) Assessment of the impact of upstream land management
measures on flood flows in Pickering Beck using Overflow. Appendix 12.2 to Final Report of
Slowing the flow at Pickering, Defra FCERM Multiple objective flood management
demonstration project RMP5455, Nisbet, T R, Marrington S, Thomas, H, Broadmeadow S and
Valantin G.
Raven E K, Lane S N, Ferguson R I and Bracken L J (2009) The spatial and temporal patterns
of aggradation in a temperate, upland, gravel-bed river, Earth Surface Processes and
Landforms, 34, 1181-1197

Rochacher J (1970) Increases in water yield following clear cut logging in the Pacific Northwest,
Water Resources Research, 6, 653-658.
Robinson M (1990) Impact of improved land drainage on river flows, Institute of Hydrology
Report 12, Institute of Hydrology.
Scottish Environment Protection Agency (2011) Allan Water Natural Flood Management
Techniques and Scoping study, report by Centre for River EcoSystems Science/Halcrows.
Swank W T and Crossley D A eds. (1988) Forest Hydrology and Ecology at Coweeta,
Ecological Studies No. 66, Springer- Verlag, New York.
Whatmore S and Langstrom C (2010) Mannings n: putting roughness to work, in How well do
facts travel? Morgan M and Howlett P (eds.) Chapter 4, Cambridge University Press,
Cambridge.

3) Review of Jacobs Section 20 review and GIS Tool by Scott Arthur & Nick Wallerstein,
Heriot-Watt University.

3.1) Review
Q1) The literature review was limited to those studies that provided quantitative evidence
of the effects of NFM techniques on catchment scale flooding. In your view, is this the
right approach or have any important studies been missed?
The approach undertaken is obviously driven by the questions that need to be addressed. In
this case it is assumed that the focus upon quantitative studies was initiated in order that a
practical tool (i.e. the Section 20 GIS Tool) could be developed for assessing the impact of NFM
scenarios upon flood attenuation through the catchment system. As such the approach adopted
was very effective.
With regard to the broader scope of this subject there are a significant number of other
management practices associated with NFM, which are not discussed in the literature review.
These are listed in the report however, (see Table 2.1) and comprise: blocking field drains;
reducing the impact of sheep grazing; restoring gully woodlands; land and soil management
activities, and; landuse change arable reversion. These are acknowledged in the report as
techniques that need to be addressed in future but which have not been considered as yet. This
might perhaps be due to the perceived lack of hard data associated with these measures.
However, several of these topic do have a significant body of literature associated with them,
especially relating to the impact of sheep grazing (reversion to arable land), and upland
agricultural landuse management. A selected list of relevant studies relating to these factors and
practices is supplied in a reference list at the end of this document.
Q2) In your view, are the studies identified in the literature review representative of
Scottish catchments?

The review is comprehensive with respect to the body of the literature cited. The main texts from
which quantitative data have been extracted are either directly associated with studies made in
Scottish catchments, or associated with other research conducted elsewhere in the UK in
catchments which have similar hydro-climatic conditions, landuse types and topography. The
review also discussed findings from a number of overseas studies, which were deemed to be
relevant to the UK.
A selection of other useful papers and reports, not cited in the literature review, which are
relevant to some of the NFM measure examined in this study are supplied in the list of
references at the end of this document. Three studies which are directly related to Scottish
catchments are highlighted in bold-italic in the list.
Q3) Do you agree with conclusions and recommendations in sections 2.2.4, 2.3.4 and
2.4.4?
Conclusions in section 2.2.4
The conclusions relating to Forest Woodland studies are consistent with the literature reviewed,
and the rule set values have been extracted from a reasonably broad range of literature which is
summarized in Table 2.5. Data obtained from a number of studies conducted overseas, which
were deemed relevant to the review, have been included in this Table but in reality there is
probably some disparity between those studies and the ones conducted in the UK (in terms of
hydro-climatic regime, etc.), which is likely to have introduce some uncertainty into the rule set.
The conclusions regarding Riparian Woodland are consistent with the findings obtained from the
two main sources of literature; the Pickering Beck and River Laver studies (refer to Table 2.6).
However, as the rule set values are only based upon the data extracted from these two studies,
the degree of uncertainty in the values is likely to be significant.
The conclusions relating to Floodplain Woodland are generally consisted with the findings
obtained from two simulation studies based upon the River Carey and River Laver (refer to
Table 2.8). It is not made clear in the summary that these studies were based upon model
simulations with extent and type of floodplain woodland simulated by manipulating mannings n
values rather than from field studies. Consequently, there is likely to be uncertainty associated
with transferring these outcomes to field conditions. From these studies a value of 300mm is
suggested for the increase in flood water level. However, actual values quoted in Table 2.8 are
50-270mm (River Cary study) and 610mm (River Laver Study).
Conclusions in section 2.3.4
The conclusions regarding Upland Drain Blocking are based upon data obtained from studies
associated with two catchments; the Coalburn, and Blacklaw Moss. In these two studies the
experimental procedure was to introduce ditching to observe the reduction in time-to-peak. The
assumption is made, in the absence of data relating to drain blocking, that the potential increase
in time-to-peak associated with blocking is likely to be of the same order as that caused by
ditching. It is through this rationale that the rule set values have been generated. It is considered
that these assumptions are very weak, introducing a serious degree of uncertainty. This issue,
along with the fact that there is a high level of uncertainty and inconclusive evidence inherent in
the available literature anyway, is made explicit in the text however.
Conclusions in section 2.4.4
These conclusions relate to wetlands and floodplains. The main conclusion is that, while there is
a general consensus that floodplain restoration and enhancement techniques have a positive

influence with respect to flood attenuation, it is very hard to draw general conclusions on the
magnitude of such effects. This conclusion is in agreement with the findings obtained from the
body of literature examined, (in which the results from more than ten studies are cited) because
the range of techniques and magnitude of physical adjustments made amongst these studies
are widely varied thus making it difficult to draw information from one study that is corroborated
by that of another.
Q4) Do you agree with the outlined approach described in Section 3.1?
The outlined approach, in general, appears to be clear in its objectives and the rationale for the
method adopted is sound. It acknowledges the difficulties which arise when devising a generic
NFM scenario-based Tool given the relative paucity of hard data which can be incorporated into
such a model.
There are two major components of the mechanics of the methodology which are of concern
however. The first relates to catchment areas upstream of 'Target Sites' (Tp points) along the
river network; the locations at which the user chooses to examine how upstream management
effects might impact the local hydrograph. The issue is associated with the fact that, given any
Tp, the upstream contributing catchment area is divided into thirds, representing Upper
(headwater), Middle and Lower catchment components, an operation which is undertaken to
enable the differences in influence had by some NFM measures between these three subcatchment area, to be accounted for in the model calculations. Incorporating this idea of subcatchment division into the Tool is sound in principal with respect to observations made in the
pertinent literature. However, division into thirds of, for example, the catchment area above a Tp
point which is located in the true headwater area of the complete catchment under study does
not make physical sense as all sub-catchments above that Tp point are, in effect, located in the
wider river headwaters. The second issue is that the effects had by each management option
selected upstream of a chosen Tp point are simply summed, a condition which is highly unlikely
in practice.
Q5) Do you have any observations in relation to the testing of the model described in
Section 3.3?
The approach adopted for testing the Tool, based upon the use of the River Enrick model, is
considered sound, as this catchment has a rich source of verified modelling data that was
directly applicable to this task. However, there is a question of transferability of the test
outcomes to other catchments/regions in Scotland (this is acknowledged in the text). The use of
the generic form of this model as the basis for generating the floodplain rule sets is reasonable,
given the lack of actual data, but it is still concerning that there are no ground-truth values
which can be applied. Given that this was a simplified model, it is of concern that so few
scenarios were tested to generate the rule-set (refer to Figure 3.5 in that text), there being only
Small (S) and Large (L) storage basin tests, and only one storage depth.
Q6) In your opinion, is the preliminary rule set described in Section 3.4 acceptable and
are there sufficient grounds for the rule sets? Suggest, if possible, alternative ways to
improve the rule set?
Generic conditions:
a) The fact that the Tool uses the SEPA 0.5% and 20% AEP national annual probability
mapping to generate design hydrographs is credible and demonstrates that consideration
has been given as to the ability for uptake of the Tool on a nationwide basis using the same,
currently available, baseline data.

b) The use of scenario 'banding' for hydrograph reduction (optimistic, most likely, and
pessimistic) associated with each NFM technique is a good idea, offering as it does
'sensitivity' bounds for the degree of attenuation associated with each measure. However,
given the limited numerical values available in the pertinent literature these bandings must
be considered with care, and must not be interpreted as a true measure of uncertainty
associated with the results.
c) There is a rule set, which applies to all NFM measures, whereby, if the area covered by a
particular measure is less than the total area available in any given catchment, the
hydrograph reduction potential for that measure is reduced via linear scaling between the
two areas. It is considered that this is a reasonable manipulation given that there appears to
be no data applicable to this issue in the pertinent literature.
Comments on specific rule-sets
Forest Cover
a) The rule set of 10% (most likely), 20% (optimistic) and 0% (pessimistic) attenuation values
for the 0.5% AEP, and 25% (most likely), 40% (optimistic) and 10% (pessimistic) values for
the 20% AEP (refer to Table 3.2 in report) appear reasonable given the data available in the
relevant literature (summarised in Section 2.4) and the model testing results.
b) The breakdown of hydrograph reduction by sub-catchment areas (i.e. upper, middle and
lower catchments), based upon the outputs from the generic model scenario testing (Section
3.3) is considered reasonable. However, this sub-catchment breakdown of hydrograph
reduction strictly only applies to the generic model so these values should be treated with
great care.
c) There is a rule within the model, which is considered reasonable, whereby, if the user
selects an area on the GIS map for forestation (via creation of a polygon) which intersects
with existing forest cover or indeed any other NFM cover (drain blocking etc.), the new area
identified has the existing cover removed before the rule set is applied.
Riparian Woodland
a) The rule set of 0.5% (most likely), 1% (optimistic) and 0% (pessimistic) hydrograph
attenuation values, which apply to both the 0.5% AEP and 20% AEP events (refer to Table
3.3) are in line with the pertinent literature associated with the Pickering and Laver studies
reviewed in Section 2.2.2.
b) The breakdown of hydrograph reduction values by sub-catchment, which have been based
upon the results from the generic Enrick model, are, as discussed in point b under Forest
Cover, susceptible to considerable uncertainty, and should be viewed with caution.
c) The same rule, defined in point c under Forest Cover, applies to Riparian Woodland, which
is again, considered to be a reasonable approach.
Upland Drain Blocking
The basis for the attenuation rule set associated with this management type (Figure 3.4)
appears to be weak, both in terms of the overall attenuation values and the reductions by
catchment sub-units. This issue is, to a degree, associated with conflicting findings in the
literature and also a considerably 'liberal' interpretation of the generic model results. The rule set

associated with Upland Drain Blocking should therefore be highlighted as a component of the
Tool, in its current form, which has significant uncertainty associated with it.
Floodplain Enhancement
For the range of techniques associated with this management option the Tool incorporates a
GIS layer that defines the floodplain extents, based upon the SEPA indicative mapping for the
estimated 0.5% AEP flood inundation extent, from which the user can select floodplain areas for
enhancement. There being no field data, a baseline rule set was derived by combining results
from a range of generic model runs. This baseline rule set is presented in Table 3.5, from which
the user selects, along with its sub-catchment location, a hydrograph reduction value based
upon the channel slope approximately within the range of that where the technique is to be
applied. This value is then adjusted according to the floodplain area selected, an 'enhancement
level' (the depth of floodplain inundation which is considered to bring about a significant
reduction in the design hydrograph), and an 'efficiency factor', which is designed to moderate
the output value according to its potential uncertainty. The variables required for input by the
user are therefore: the area of floodplain selected for application of the measure; the
enhancement level, and; a efficiency factor value. This methodology is utilised by all four
floodplain based management options these being: physically increasing floodplain storage;
increasing floodplain roughness using floodplain woodland; re-connecting the floodplain, and;
setting-back of embankments. All these techniques require the estimation of an enhancement
level. In the case of physically increasing floodplain storage capacity it is suggested that the
level of increase can be broadly estimated by dividing the increase in storage capacity created
by the relevant floodplain area, an assumption that seems reasonable. In the case of increasing
roughness using floodplain woodland it is acknowledged in the text that there is very limited
data, with only one value having been obtained from the literature; a potential increase in level
of 300mm. Table 3.6 presents rule set values of 100mm (pessimistic), 200mm (most likely) and
300mm (optimistic) for this technique, the only value having any physical basis being that of
300mm. These values are therefore considered to be highly unreliable. In the case of reconnecting the floodplain no data appears to have been available to make any judgment upon
enhancement levels such that recourse is taken to suggesting that 'equivalent' increases can be
applied based upon the (already weak, rule set values applied to the increasing roughness
option. As such therefore Table 3.7 presents values for the 20% AEP as per those in Table 3.6,
while values for the 0.5% AEP of -150mm (pessimistic), -50mm (most likely) and 100mm
(optimistic) appear to have no rational basis at all. The rule set values in Table 3.7 are also
applied to the setting-back of embankments option, again with no substantial justification.
Values computed for all of the floodplain enhancement options are then moderated by the
'efficiency factor', a percentage value that is designed as a means of modifying the hydrograph
reduction outputs according to their uncertainty. These values are presented in Table 3.8 where
pessimistic, most likely, and optimistic percentage rule set values are presented for the three
floodplain enhancement measures. This rule set appears to have no rational basis at all and
none is justified.
Consequently, it is considered that the rule set values offered for all floodplain techniques are
highly unreliable, there being three levels of degradation in the knowledge base. This is
because; first, the baseline rule set is derived from values obtained from a generic model;
second, the set of equivalent storage depth values, while having a reasonably sound basis for
the physically increasing floodplain storage capacity option, have a very limited basis in the
case of increasing floodplain roughness, and appear to have no physical basis at all in the case
of re-connecting the floodplain and setting-back of embankments, and; third, the efficiency
values appear to have been set with no justification other than simple guess work. Any results
obtained when considering these techniques in the Tool are therefore considered likely to be
extremely uncertain especially in the case of the 're-connecting' and 'setting-back' options and it

is quite possible that inclusion of these measures in any model run will greatly compromise the
overall predicted results.
It is strongly recommended that, while inclusion of this family of floodplain techniques is
important as management options, with the exception of the physically increasing the floodplain
storage capacity technique, they be excluded from the Tool until such time as any meaningful
data can be brought to bear on the calculations.
Wetlands
Multiple types of wetland (marshy, fen, etc.) are listed as potential independent candidates for
NFM options, and Table 2.10 in the literature survey indicates the NFM role that is likely to be
relevant for each. Within the Tool there is not a set of independent rules associated with each
type rather, suggestions are made as to which of the rule sets associated with the other options
(Forest Cover, etc.) might be used to simulate that type. This scheme is presented in Table 3.1.
This idea is sound in principal given the paucity of appropriate numerical data available to
characterise this parameter. However, given the range of NFM measures suggested for some of
the wetland types, the user is likely to be unsure as to which rule set is truly appropriate for the
catchment site under consideration. It is obviously desirable to make rule-sets as specific as
possible for different NFM measures and, although there may be many different wetland types
to consider, it might perhaps be better to collapse some of these into a smaller number of
broader categories, within which the range of sub-types might potentially benefit from the same
range of NFM options.
Q7) In your opinion is the outline method for coastal floodplains presented in Section 3.6
correct?
The approach outlined for assessing potential locations for managed realignment along the
open coast, if it has been interpreted correctly, is basic very basis. It is simple to implement,
needing just the SEPA national indicative flood mapping dataset loaded into say, ArcGIS, and
the Section 20 Tool does not necessarily have to be used at all. The approach does not actually
help the user to decide where defence breaching might be applicable, and is only correct in as
much as it would simply identify the landward area that would be inundated if breaching were to
be undertaken at a given location.
In simple terms this prospective component of the Tool does not offer any inbuilt, processbased, component that the user might use for testing breach location scenarios and as such it is
not in line with the Tools main functional utility.
The method outlined for estuarine situations has a more quantitative component, which offers a
way to make a broad-brush estimate of the potential change in water level that might occur due
to the introduction of user defined areas of managed realignment. The approach is considered
to be valid, although very crude and would need to be backed up by more intense modelling
studies. Once again, however, it does not necessarily have to be placed within the structure of
the Tool, and, does not have any inbuilt component that can be applied to help the user choose
as where suitable, locations of realignment might lie.
Q8) Do you agree with the limitations specified in Section 3.7 and are there any further
issues not identified?
The key limitations presented in Section 3.7 are all valid. However, a considerable number of
other limitations are not documented (although some are presented elsewhere in the report),
these being:

a) The effect had by multiple measures within a target catchment are purely additive in nature,
and this is not likely to be the case in practice as there will undoubtedly be interactions
between the various measures.
b) The issue associated with catchment partition upstream of any selected Tp point (refer to
Question 1).
c) The rule sets associated with the Floodplain restoration/enhancement/alteration techniques
are extremely tenuous and therefore, as tempting as it is to make the model as 'complete' as
possible by incorporating all possible factors, only the option associated with physically
increasing the floodplain storage capacity ought to be retain (given that it has the strongest
physical basis) and the others removed until field-based corroborative evidence becomes
available.
Q9) Do you agree with the recommendations made in Section 3.8 and do you have any
further recommendations, not identified in Section 3.8 that would be beneficial for
development of the model?
The recommendations are all valid with respect to improving the scientific validity and
functionality of the model.
However, a general observation that may be made is that a number of the recommendations
suggest the incorporation of new rule sets associated with other NFM techniques such as
channel restoration and enhancement. The idea of incorporating such factors is valid but it must
be stressed that they ought not be included until the uncertainty in rule set values associated
with other approaches, especially the Floodplain factors, is reduced significantly.
Referring to the concern raised under Question 1, regarding the validity of splitting the
catchment upstream of any target point into thirds, if this process has been interpreted correctly,
it is strongly urged that it be rectified in some form. A suggested approach would be to have the
GIS divide the entire study catchment into thirds (based upon travel time) and define the
hydrological reduction rule sets above any selected Tp point according to which overall
catchment sub-types are encompassed upstream; i.e. a Tp point located in the overall
catchment headwaters would take on the hydrological reduction conditions of only this sub-type,
while a Tp point located in the true middle reach would contain components of the true middle
and true upper catchment components.
Q10) Do you have any other observations from Section 3 that you would like to
highlight?
a) There is repeated mention of the use of practical experience in generating the rule sets, etc.
but no specifics are offered. If certain values/variables are currently present in the model
which are derived from such sources they should be made as explicit as possible, as
otherwise, there may be attributes which are effectively a 'black box' or which might be
simply 'fudge-factors'. It is quite acceptable to use practical/personal experience in the
absence of hard numbers, but knowledge of where, and how, such information is applied in
the Tool is crucial when it comes to others manipulating and improving on the model code.
b) It is stressed throughout the document that application of the Tool to other catchments is a
relatively simple task, a characteristic that is essential if it is to be applied to catchments
across Scotland. The manner in which this can be implemented is presented in Section 3.2
of the report). However, while this would appear a reasonably simple task, made all the
more compelling by the fact that a pilot catchment is already incorporated into the Tool for

the purposed of demonstration, its transferability is not entirely trivial. In practice, when
applying the Tool to any given target catchment, the necessary DEM data requires reconditioned in Arc, followed by manipulation using ArcHydro, along with updating parameter
values specific to that catchment within the Access database tables. In a practical sense
therefore the prospective user must have a skill-set which encompasses the use of these
various pieces of software.
c) The modified FEH approach utilised in the Tool should be fully documented in the report.
d) In the case of the example catchment currently available in the Tool, the DTM is based upon
NextMAP data but, of course, LiDAR might just as well been used if available. However, it is
strongly urged that a protocol be established as to the resolution of the topography data
which should be applied, if the approach were to be applied nationwide. This is because
ArcHydro is quite sensitive to terrain resolution when it defines the flow paths through a
DEM; information which will influences the calculation of slopes between Tp point and
consequently the apportioning of catchment areas (refer to 'Automation Scripts' in Section
3.2). Therefore, if LiDAR were to be utilised in one catchment and a different resolution of
data in another (or indeed two different resolutions of data in replications of the same
catchment) it is possible that inconsistencies might be introduced.
e) One point to considered, which is purely associated with the analysis presented in the
Section 20 report, is that the figures which display the results output from the River Enrick
test study are poor being both hard to read and interpret. The discussion of these graphs is
also very weak. This is not be a problem with the model per se but these outputs do
represents information that the user may well make reference to when interpreting their own
results.
Q11) Overall, would you recommend the use of the proposed GIS tool and the rule for
use as the screening tool to provide an indication of the effect of NFM techniques?
The approach is conceptually sound in general, and recognises the fact that only simple
processes can be simulated at present using, in general, rule sets based upon results obtained
from the pertinent literature. Also, the calculation processes are not top-heavy with respect to
any one individual sub-process. The use of an ArcGIS platform and associated Access
databases for data manipulation and calculation processes is considered to be appropriate as it
allows access to the Tools internal structure, enabling the user to update rule sets and internal
GIS layers relatively easily. The use of this platform is also appropriate as the software is likely
to be relatively familiar to in-house computer modelling staff.
It is therefore cautiously considered that the Tool could be applied as a broad-brush, high end
NFM scenario assessment tool, but the results must be interpreted intelligently, and be back up
by ground-truth data when targeting specific catchments. However, the Tool should not be
applied as of yet, until the three issues defined below are addressed.
The first, and most important is the concern over the sub-division of catchments above any
given Tp point into upper, middle and lower regardless of where the Tp point exists in the
wider catchment (refer to Question 1). As possible solution for dealing with this is issue offered
under Question 6.
Second, the rule sets associated with the 'increasing floodplain roughness by floodplain
woodland', 're-connecting the floodplain' and 'setting-back of embankments' Floodplain
enhancement techniques (Section 3.4) are considered to be too tenuous at present, and it is
strongly suggested that they be removed until such time as field based, or other data, becomes
available to substantiate these values.

Third, the fact that hydrograph reduction at a given Tp point is based upon the simple
summation of effects had by different measures in the upstream catchment is considered
extremely simplistic, and unlikely to be the case in reality. If any information might be gleaned
from the literature as to the impact had by combined measures this should be considered to
enable incorporation for potential adjustment factors. However, it is acknowledged that if only
tenuous information is available, the uncertainty associated with any adjustment factors might
outweigh that of the possible errors incurred by the simple addition process. Therefore, in the
light of this, it is considered that this issue should be flagged up as a serious limitation in the
current model structure, but does not render it completely unacceptable in its current form.
References relating to NFM techniques (by topic)
Upland Forest Cover
1. Calder, I.R. and Aylward, B. (2006) Forest and floods: Moving to an evidence-based
approach to watershed and integrated flood management. Water International, 31, 1-13.
2. Carroll, Z.L., Bird, S.B., Emmett, B.A., Reynolds, B. and Sinclair, F.L. (2004) Can tree
shelterbelts on agricultural land reduce flood risk? Soil Use and Management, 20, 357-359.
3. Johnson, R.C. (1983) Effects of forestry on suspended solids and bed load yields in the
Balquidder catchments. Journal of Hydrology, 145, 403-417.
4. Mount, N.J., Sambrook Smith, G.A. and Stott, T.A. (2004) An assessment of the impact of
upland afforestation on lowland river reaches: the Afon Trannon, mid-Wales.
Geomorphology, 64, 255-269.
5. Robinson, M. and Newson, M.D. (1986) Comparison of forest and moorland hydrology in an
upland area with peaty soils. International Peat Journal, 1, 49-68.
6. Stott, T.A., Ferguson, R.I., Johnson, R.C. and Newson, M.D. (1986) Sediment budgets in
forested and unforested basins in upland Scotland. In R.F. Hadley (ed.) Drainage Basin
Sediment Delivery
Upland Drain Blocking
1. Birnie, R.V. and Hulme, P.D. (1990) Overgrazing of peatland vegetation in Shetland.
Scottish Geographical Magazine, 106, 28-36.
2. Dunn, S.M. and Mackay, R. (1996) Modelling the hydrological impacts of open ditch
drainage. Journal of Hydrology, 179, 37-66.
3. Robinson M. (1980) The effect of pre-afforestation drainage on the streamflow and water
quality of a small upland catchment. Institute of Hydrology Report No. 73. Wallingford:
Institute of Hydrology.
Grazing (removal) and other upland landuse issues (not including gripping)
1. Bailey-Denton, J. (1861). On the discharge from underdrainage and its effects on the arterial
channels and outfalls of the country. Proceedings of the Institution of Civil Engineers, 21, 48
130.

2. Department for Environment, Food and Rural Affairs (2008) Catchment sensitive farming.
Available at: http://www.defra.gov.uk/farm/environment/water/csf/index.htm (Accessed 8
April 2008).
3. Gifford, G.F. and Hawkins, R.A. (1978) Hydrologic impact of grazing on infiltration: a critical
review. Water Resources Research, 14, 305-314.
4. Jackson, B.M., Wheater, H.S., Mcintyre, N.R., Chell, J., Francis, O.J., Frogbrook, Z.,
Marshall, M., Reynolds, B. and Solloway, I. (2009) The impact of upland land management
on flooding: insights from a multiscale experimental and modelling programme. Journal of
Flood Risk Management, 1, 71-80.
5. Jones, M. (1998) The impact of grazing and upland management on erosion and runoff.
R&D Technical Report P123. Swindon: Environment Agency.
6. Marshall, M.R., Francis, O.J., Frogbrook, Z.L., Jackson, B.M., McIntyre, N., Reynolds, B.,
Solloway, I., Wheater, H.S. and Chell, J. (2009) The impact of upland land management on
flooding: results from an improved pasture hillslope. Hydrological Processes, 23, 464-475.
7. Meyles, E.W., Williams, A.G., Ternan, J.L., Anderson, J.M. and Dowd, J.F. (2006) The
influence of grazing on vegetation, soil properties and stream discharge in a small Dartmoor
catchment, southwest England, UK. Earth Surface Processes and Landforms, 31, 622-631.
8. OConnell, P.E., Ewen, J., ODonnell, G.M. and Quinn, P.F. (2007) Is there a link between
agricultural land-use management and flooding? Hydrology and Earth System Sciences, 11,
96-107.

CREW Facilitation Team


James Hutton Institute
Craigiebuckler
Aberdeen AB15 8QH
Scotland UK
Tel: +44 (0) 844 928 5428
Email: enquiries@crew.ac.uk
www.crew.ac.uk

Appendix B
Summary Table of
Natural Flood Management
Measures

NFM
technique

Type

Conifer
woodland

Fluvial and
pluvial

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Upper and
middle
catchment

Canopy interception
reduces net rainfall and
evapotranspiration
reduces antecedent
conditions

Large body of literature, refer to Section


3.1.2.E. Literature focuses on outdated
practices and annual water yields.

Ability to modify runoff


generation in a
proportioned manner
with flow routing element
to estimate accumulated
effects

With moderate
uncertainty, complete
conifer forest cover
reduces runoff and may
offer 10-20% reductions in
flood peak when compared
to grazing land uses. The
pre-forest land use which
the forest is compared
against is significant.

Ability to modify runoff


generation in a
proportioned manner
with flow routing element
to estimate accumulated
effects

With high uncertainty


native deciduous trees
reduce runoff and may
reduce summer flood
flows by similar amounts to
conifers (10-20%) with
reduced effect in winter
months. The pre-forest
land use which the forest is
compared against is
significant.

Sub-type

Runoff
reduction

Assuming current best practice in place


(no drainage) there is agreement that
conifer forests delay and decrease runoff
but there is disagreement over the extent of
the effect (OConnell et al., 2004).
Consensus that the effects of forestation
diminish with both catchment size and
event magnitude. Location within catchment
is significant.

Research or data
need

Inclusion

Include within
screening for high
runoff areas and
priority for assessment
tool

There is no FEH/FSR catchment descriptor


for forest cover as no statistical significance
can be identified.
Broadleaf
woodland

Fluvial and
pluvial
Runoff
reduction

Upper and
middle
catchment

Canopy interception
reduces net rainfall and
evapotranspiration
reduces antecedent
conditions

Small body of literature, refer to Section


3.1.2.E. Literature focuses on conifer
plantations.
Compared to conifer forests there is more
uncertainty, partly due to the seasonal
fluctuations in leaf coverage but also due to
less research interest.
Consensus that the effects of forestation
diminish with both catchment size and
event magnitude. Location within catchment
is significant.
Jackson et al. (2008), whole catchment
forestation using native woodland over
pasture can reduce flood peaks by 10 to
54%. (This is one study and uses degraded
pasture for higher estimate).
OConnell et al. (2007) replacing conifers
with native woodland would increase
flood peaks by around 10%.
Robinson et al. (1986) replacing grassland
with native woodland would reduce flood
peaks by around 10%.
There is no FEH/FSR catchment descriptor
for forest cover as no statistical significance
can be identified.

Appendix B - 1

Could benefit from


additional studies to
support findings.
There would also be
benefit in more species
specific data.

Include within
screening for high
runoff areas and
priority for assessment
tool

NFM
technique

Type

Good
muirburn
practice
(compliance
with muirburn
code)

Fluvial & pluvial

Reducing
grazing
pressure on
pasture

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Upper
catchment

Improving ground
cover through the
provision of a mosaic
of heather cover. The
temporal and spatial
spread of burnt bare
ground ensures a
relatively consistent
ground cover. Good
ground cover increases
interception and slows
overland runoff rates.
May improve
antecedent moisture
condition.

Limited to annual interception rates (Nisbet,


2004) heather offers year round interception
of 16-19% of annual rainfall. No
information on interception capacity during
intense events.

Ability to modify runoff


generation in a
proportioned manner
with flow routing element
to estimate accumulated
effects. Literature gaps
would require reliance on
physical basis.

High uncertainty of
runoff reduction.
Absence of literature to
quantify effects, Dunn
(1986) reports a doubling
of flood peaks for whole
catchment burns.

National data of current


muirburn practice

Include within
screening for high
runoff areas and
priority for assessment
tool

Improving the
hydrological condition
of soils (increasing
infiltration)

Effects vary, some reports indicated up to


double runoff volumes at field scale for
degraded soils (Heathwaite et al., 1990)

Ability to modify runoff


generation in a
proportioned manner
with flow routing element
to estimate accumulated
effects

Moderate uncertainty of
runoff reduction,
magnitude quantifiable.

National data of soil


degradation due to
overgrazing

Include within
screening for high
runoff areas and
priority for assessment
tool

Would benefit from


additional studies to
support findings and to
provide guidance on
benefits for other soil
types.

Include within
screening for high
runoff areas and lower
priority for assessment
tool. In the absence of
inclusion within the
assessment tool
provide outline of
potential method to
facilitate local
assessment.

Sub-type

Runoff
reduction

Fluvial & pluvial


Runoff
reduction

Upper and midcatchment

Ramchunder et al (2009), relatively little is


known about the effects of muirburn. It is
speculated that catchment scale it does have
an effect.
Observations indicated that muirburn does
reduce infiltration, increase runoff and
extensive burns increase flood peaks
(Mallik et al, 1984; Conway & Millar, 1960;
Robinson, 1985; Dunn, 1986).

The more widely applicable method adopted


by the Environment Agency (Packman et al.,
2004) suggests a typical reduction of around
13% in percentage runoff for areas of
pasture, but in some areas up to 41%.
These figures are for English and Welsh
catchments.

Varies with local conditions


(soil, slope, land use
intensity etc)

Use of donor soil types to represent degraded


soils (Packman et al., 2004). No land
degradation data for Scotland.

Creation of
cross slope
tree shelter
belts

Fluvial & Pluvial


Runoff
reduction

Catchment wide
(Mainly sloping
intensive
upland areas)

In addition to benefits
of forest (increased
interception and
reduced antecedent
moisture) runoff from
upslope can be
infiltrated at the tree
shelter belt

Study of land use changes in the Pontbren


catchment indicate that infiltration at young
native woodland can be 60 times higher
than improved pasture, (Jackson et al,
2008). Using strategically paced cross slope
shelter belts could help reduce peak flows
in large catchments during large events
by 2 to 11%.

Appendix B - 2

Field scale effects of soil


degradation on runoff
generation range from 0 to
100%.

Work by Packman et al.


(2004) and the EAs CFMP
Tool offer an existing
means of quantifying
hydrological benefits.
A detailed distributed
hydrological model that
allows runoff generated
upslope to be infiltrated
on arrival at shelter belt.

Very high uncertainty, as


literature relating to cross
slope planting is only
available from one source.
May be very specific to
local conditions.
Effects appear to be
significant.

NFM
technique

Type

Creation /
restoration of
nonfloodplain
wetlands

Fluvial & Pluvial

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Principally
catchment
headwaters

It is commonly
misstated that wetlands
act as slow release
sponges. They
frequently act as
sources of flashy runoff
due to the high
antecedent moisture
condition but in some
cases can serve to
attenuate flows.

The literature contains conflicting


messages about the flood reducing effect of
wetlands, some report increases other report
decreases, (Bullock and Acreman, 2003)

Detailed continuous
simulation hydrological
model which includes an
accurate groundwater
modelling component.
Would require a large
amount of data and
technical input to
accurately represent a
single wetland.

Very high degree of


uncertainty. Effects can be
positive and negative.

Requires detailed site


specific data of each
wetland to facilitate
accurate assessment

Include within
screening of high runoff
areas and lower priority
for assessment tool. In
the absence of
inclusion within the
assessment tool
provide outline of
potential method to
facilitate local
assessment.

Ability to modify runoff


generation in a
proportioned manner
with flow routing element
to estimate accumulated
effects

Moderate uncertainty of
runoff reduction,
magnitude quantifiable.

Requires data of
existing land
management practices

Include within
screening of high runoff
areas and priority for
assessment tool

Requires detailed
mapping of existing
drains

Include within
screening of high runoff
areas and lower priority
for assessment tool

Sub-type

Runoff
reduction*/delay
, attenuate

Most wetlands were identified as WRAP


class 5 within the FSR, this is the class with
the highest percentage runoff, (Boorman et
al., 1995).
Many bogs do not act to delay flow into
streams and the sponge analogy is
incorrect (Rydin and Jeglum, 2006)
Each wetland is unique.

Reducing soil
compaction
in arable
areas,
improving
soil texture
and reducing
bare earth in
wetter
seasons

Fluvial & pluvial

Mid-catchment

Runoff
reduction

Improving soils and


ground cover
(increasing
interception, infiltration
and overland flow
rates)

Infiltration rates can be eight times less in


ploughed ground to mature pasture (Holtan
and Kirkpatrick, 1950)
The EAs CFMP Land Management Tool
suggests a reduction in flood flows of up
to 7% when best practice in place. The flow
reductions decrease with event magnitude.
The effects are highly dependent on soil
types, soil condition and land use.

Work by Packman et al.


(2004) and the EAs CFMP
Tool offer an existing
means of quantifying
hydrological benefits.

Use of donor soil types to represent degraded


soils (Packman et al., 2004)
Changing
agricultural
field drainage

Fluvial (&
pluvial)
Runoff
reduction
(Rapid shallow
groundwater)

Upper and midcatchment

Field drains provide an


improved route for
water to reach
watercourses (but can
conversely lower
antecedent moisture
levels)

OConnell et al. (2007) states that field


drainage can speed or slow time to peaks by
2-3 times in both directions.
IH113 reported that drainage does not alter
the hydrograph volume. At a field scale the
peak rate was reduced in clay soils and
increased in permeable soils. At a
catchment scale arterial channels
accelerate runoff. Baseflow is generally
increased with drainage, but effect on peak
flows is site specific.

Appendix B - 3

Varies with local conditions


(soil, slope, crop cover
intensity etc)

Complex distributed
hydrological model with
shallow groundwater
component allowing the
representation of
modified antecedent
conditions and the
accelerated shallow
water flow where pipes
are present.
Arterial ditches would
require a flexible flow
routing model that allows
the presence of known
arterial ditches to be
represented.

Very high uncertainty of


effect which can be
positive and negative.
Arterial channels have an
increasing effect with event
magnitude.
Requires knowledge of the
drainage network.

NFM
technique

Type

Upland drain
blocking

Fluvial

Floodplain
reconnection

Fluvial

Location

Principal action

Literature

Upper
catchment

Drains provide an
improved route for
water to reach
watercourses (but can
conversely lower
antecedent moisture
levels). Upland drain
blocking is a complex
issue which requires
consideration of soil
type, topography and
proximity to
watercourse.

In floodplains,
mostly lower
and middle
parts of
catchments
where
floodplains are
sufficiently wide
to merit flood
protection
embankments

Extends catchment
time to peak by slowing
travel time through the
catchment. Can allow
the desynchronisation
of sub-catchment
peaks. Often
associated with river
restoration which can
also slow down the
flow of water.

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Price et al (2003) efficiency depends on


Complex distributed
depth, spacing and hydraulic conductivity. hydrological model with
shallow groundwater
Ramchunder et al (2009) noted drain
component allowing the
direction (cross or downslope) is important.
representation of
Downslope drains have a larger effect
modified antecedent
than cross slope.
conditions and the
Robinson et al. (1988) suggests a 15 to 20% accelerated shallow
shortening (~1hr) of the time to peak for
water flow where pipes
the Blacklaw catchment following introduction are present.
of drains.
Arterial ditches would
General agreement that upland drains can
require a flexible flow
serve to lower antecedent moisture
routing model that allows
conditions but conversely serve to
the presence of known
improve connectivity to watercourses.
arterial ditches to be
Ballard et al (2010) suggests that the
represented.
quantification of effects is at the limit or
beyond current capabilities.

A very high degree of


uncertainty, can reduce or
increase flood peaks
depending on local
circumstances.

Requires detailed
mapping of existing
upland drains

Include within
screening of high runoff
areas and lower priority
for assessment tool

General consensus that floodplain


reconnection can reduce flood flows,
Bullock and Acreman (2003).

A very high degree of


uncertainty, normally
reduces downstream flood
peaks depending on local
circumstances but can
cause increases.

Requires detailed
topographic data to
enable modelling.

Include within
screening of
channel/floodplain
areas and lower priority
for assessment tool. In
the absence of
inclusion within the
assessment tool
provide outline of
potential method to
facilitate local
assessment.

Sub-type

Runoff
reduction,
delay/desync

Attenuation ,
delay/desync

Disagreement on magnitude of effects but


this is probably due to local variation,
Blackwell et al. (2006).
Acreman et al. (2002) reports a 50 to 150%
increase in flood flows for watercourses
which are entrenched and completely
disconnected from the floodplain. A smaller
10 to 15% reduction for entrenched
watercourses.
Ratio of hydrograph volume to floodplain
volume is significant, Ahilan (2009) and
Potter (2006).
Bullock and Acreman (2003) identified that in
some circumstances increasing
conveyance by restoring floodplains can
increase downstream flood risk.
Rapid estimation of desynchronisation
benefits is not feasible, Odoni and Lane
(2010). The case for desynchronisation
reductions is speculative and unproven.

Appendix B - 4

Preferable to use a
distributed or semidistributed hydrological
model which
incorporates a detailed
hydraulically based flow
routing engine

Each case is unique and


must be modelled.
Realistically achievable
upper limits appear to be
10-20% reduction in peaks

Requires development
of an appropriate
hydrological model
which includes a
detailed hydraulic flow
routing engine (1D-2D,
diffuse wave)

NFM
technique

Type

Creation of
washlands

Fluvial

Creation of
constructed
farm
wetlands or
ponds

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Mostly in lower
and middle
parts of
catchments
where
floodplains are
wide

Increases the
availability of storage of
flood water within the
catchment, thus
attenuating or delaying
the flood peak.
Engineer arrangements
can serve to reserve
volume for flood peak.

Morris et al (2004) reported that washlands


can have a significant role to play in flood risk
management. For ecological benefits it is
necessary for the washland to be kept wet
thus taking up some washland storage.

Preferable to use a
distributed or semidistributed hydrological
model which
incorporates a detailed
hydraulically based flow
routing engine

Uncertainty varies with the


level of engineering used
to control the inlet and
outlet to the washland.
Effects are similar to
floodplain reconnection but
marginally higher benefits
can potentially be accrued
via more efficient use of
available storage.

Requires detailed
topographic data to
enable modelling.

Include within
screening of
channel/floodplain
areas and lower priority
for assessment tool. In
the absence of
inclusion within the
assessment tool
provide outline of
potential method to
facilitate local
assessment.

On minor
watercourses
and ephemeral
flow routes

Increases the available


storage within the
catchment helping to
attenuate flows.

Wilkinson et al (2010) present that an array of


flood storage areas can be used to attenuate
flows within a catchment.

Preferable to use a
distributed or semidistributed hydrological
model which can allow
the hydraulic
representation of the flow
controls.

Uncertainty varies with the


level of engineering used
to control the inlet and
outlet to the storage area.
Additional uncertainty
introduced through design,
large number of
interventions and need for
maintenance. Effects are
similar to traditional large
flood storage areas but
marginally lower benefits c
accrued due to less
efficient use of storage.

Requires detailed
topographic data

Sub-type

Attenuation,
delay/desync

Fluvial (and
pluvial)
Attenuation,
delay/desync

Literature will be similar to floodplain


reconnection with greater effects for some
engineered washland configurations.

The implementation of flood storage requires


the careful design and consideration of
operation issues (Ackers & Bartlett, 2010;
Hall et al., 1993).
SAIFF (2011) states that constructed flood
storage (particularly online) should not be
considered as natural flood management.

Appendix B - 5

Can be represented
using existing models

Requires development
of an appropriate
hydrological model
which includes a
detailed hydraulic flow
routing engine (1D-2D,
diffuse wave)

Could benefit for


additional studies to
support findings

Dont include within


screening and not a
priority for assessment
tool. In the absence of
inclusion within the
assessment tool
provide outline of
potential method to
facilitate local
assessment.

NFM
technique

Type

Afforestation
of floodplains

Fluvial

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

In floodplains,
mostly lower
and middle
parts of
catchments with
existing
floodplains

Extends catchment
time to peak by slowing
travel time through the
catchment. Can allow
the desynchronisation
of sub-catchment
peaks. Often
associated with river
restoration which can
also slow down the
flow of water.

All literature relates to modelling studies.

Preferable to use a
distributed or semidistributed hydrological
model which
incorporates a detailed
hydraulically based flow
routing engine

A very high degree of


uncertainty, normally
reduces downstream flood
peaks depending on local
circumstances but can
cause increases.

Requires detailed
topographic data to
enable modelling.

Include within
screening of
channel/floodplain
areas and lower priority
for assessment tool. In
the absence of
inclusion within the
assessment tool
provide outline of
potential assessment
method to facilitate
local assessment.

Sub-type

Attenuation,
delay/desync

Not suitable for


downstream of
flood receptors

Thomas and Nisbet (2006) reported that


forestation of a 2.2km reach of floodplain
on the River Parret reduced the average
flow velocity of large floods by 78%.
Nisbet and Thomas (2008) report a 0.3%
reduction in 1 in 100yr flow on the River
Laver using 40Ha of floodplain forest. The
delay was identified as 1-3.5min/Ha, with
the greatest delay for areas with the highest
flow velocity.

Each case is unique and


must be modelled.
Effects should increase
with magnitude.
Realistically achievable
upper limits appear to be
10-20%

All studies use increased Mannings values


within a hydraulic model to represent
floodplain roughness.

Requires development
of an appropriate
hydrological model
which includes a
detailed hydraulic flow
routing engine (1D-2D,
diffuse wave)

Ratio of hydrograph volume to floodplain


volume is significant, Ahilan (2009) and
Potter (2006).
Backwater effects can result in an increase
in upstream flood risk, Nisbet and Thomas
(2008) and Thomas and Nisbet (2006).
Rapid estimation of desynchronisation
benefits is not feasible, Odoni and Lane
(2010). The case for desynchronisation
reductions is speculative and unproven.
Reach
restoration

Fluvial
Attenuation or
delay/desync

Catchment
wide, typically
lower reaches
where naturally
sinuous rivers
have been
straightened

Can slow flows and


increase flow path
length thus extending
time to peak. Often
increases connectivity
with floodplain.

Acreman et al. (2002) reports a 10 to 15%


reduction for entrenched watercourses
where floodplain connection can be
improved. Totally disconnected floodplains
can increase flood flows by 150%.

Appendix B - 6

Preferable to use a
distributed or semidistributed hydrological
model which
incorporates a detailed
hydraulically based flow
routing engine

For channels which are


highly entrenched benefits
can be very significant.
Effects smaller for less
entrenched channels.
Local effects are very
significant.

Include within
screening of
channel/floodplain
areas and low priority
for assessment tool. In
the absence of
inclusion within the
assessment tool
provide outline of
potential assessment
method to facilitate
local assessment.

NFM
technique

Type

Creation of
riparian
woodland

Fluvial

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Catchment wide

Slows in channel flows


by roughening channel
banks and providing a
supply of large woody
debris for formation of
debris dams. Slowed
flow encourages
floodplain linkage

Greatest benefit in narrow watercourses


found in upper and mid catchment (Odoni
and Lane, 2010).

Assessment requires a
distributed hydrological
model that has an
integrated detailed
hydraulic flow routing
engine that is capable of
representing the
increased channel
roughness and should
allow for channelfloodplain flow linkage

A high degree of
uncertainty particularly for
cases relying on improving
floodplain connection.
Effects are marginal and
there are indications that
incorrect woodland
placement can increase
risk.

Guidance on the
natural spacing of large
woody dams and how
the Mannings
roughness values vary
with slope, channel
size and vegetation
type.

Include within
screening and low
priority for assessment
tool. In the absence of
inclusion within the
assessment tool
provide outline of
potential assessment
method to facilitate
local assessment.

Catchment data would


bring efficiencies i.e.
streampower maps and
morphological pressures
database

Geomorphological aspects
are very uncertain
depending on local
circumstances.

Sub-type

Delay/desync or
attenuation

The Robinwood study (2008), Gregory et al


(1985) reported that the effect would
decrease with event size, Odoni and Lane
(2008) report the opposite. The decrease in
effect is taken to be more plausible due to
observed data.
There is general agreement within the
literature that riparian woodland increases
the supply of large woody debris within the
channel. The effect of riparian woodland is
modelled by Robinwood (2008), Odoni and
Lane (2010) and others by increasing
Mannings roughness of banks and channel
within flow routing model. Gregory et al
(1985) reported that it was necessary to
reduce roughness with flow to match
observations.

Closure to the issue of


where delay induced
by large woody debris
increases or decreases
with flow.

Rapid estimation of desynchronisation


benefits is not feasible (Odoni and Lane,
2010). The case for desynchronisation
reductions are speculative and unproven.
Gully
woodland
planting

Fluvial
Delay/desync or
attenuation

Upper
catchment
gullies or high
energy
watercourses
(potential
gullies)

Slows in channel flows


by roughening channel
banks and providing a
supply of large woody
debris for formation of
debris dams. Helps
stabilise sources of
sediment that can lead
to deposition and
problems with channel
capacity downstream.

Very little literature specific to hydrological


effect of foresting gullies, refer to riparian
woodland literature.
Gullies can be a major supply of sediment
within a catchment (Martinez-Casasnovas,
2003).
Forestation of gullies can reduce
sediment yield by 92.5% (Zhuo, 1992).
A study of the Upper Wharfedale reported
that the reforestation of steep gullies within
the catchment would reduce the coarse
sediment supply by 85% and offer a
substantial reduction in flood risk (Lane,
2006)

Assessment requires a
distributed hydrological
model that has an
integrated detailed
hydraulic flow routing
engine that is capable of
representing the
increased channel
roughness and should
allow for channelfloodplain flow linkage
Ability to modify runoff
generation in a
proportion manner with
flow routing element to
estimate accumulated
effects

Appendix B - 7

Refer to riparian woodland


for hydrological benefits.

Include within
assessment via
approaches for riparian
woodland and the
management of
instabilities.

NFM
technique

Type

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Placed large
woody debris
and boulders

Fluvial

Mid catchment

Attenuation or
delay/desync

In channel,
upstream of
flood receptors

Slowing in channel
flows and also
increasing floodplain
connectivity

Modelling work undertaken by Odoni and


Lane (2010) in Pickering Beck identified a
potential 7.5% reduction in flood flows for
an observed flood event of unknown return
period using 100 strategically placed large
woody debris dams.

Assessment requires a
distributed hydrological
model that has an
integrated detailed
hydraulic flow routing
engine that is capable of
representing the
increased channel
roughness and should
allow for channelfloodplain flow linkage

Very high uncertainty.


Placed boulders and
woody debris can be
washed out or buried by
sediments during flood
events.

Requires guidance on
the selection of
Mannings roughness or
other modelling
technique

Include within
screening and low
priority for assessment
tool. In the absence of
inclusion within the
assessment tool
provide outline of
potential assessment
method to facilitate
local assessment.

Effects can be accurately


modelled using existing
hydraulic models

Relatively low degree of


uncertainty of effect
(assuming a stable
geomorphological system)

Improve quality of
artificial structures
database

Include within
screening of
channel/floodplain
areas and not included
within assessment tool.
Provide outline of
potential assessment
method to facilitate
local assessment.

Sub-type

Gregory et al (1985) reported that 36% of


natural debris dams were destroyed or
relocated within a 12 month period. Natural
debris dams were estimated to slow the flood
wave by 10 minutes over an observed 4km
reach.
Reach
restoration

Fluvial
Convey/protect

At or
immediately
downstream of
receptors where
the channel is
artificially
constricted

Can assist in
conveying water past
receptors

Increasing conveyance of artificially


constricted channels is a common flood risk
management activity (Riley, 1998).
Local effects are very significant.
Can increase downstream flood risk.

Appendix B - 8

Large number of
interventions required to
generate significant
impact.

Local effects are very


significant.

NFM
technique

Type

Managing
channel
instabilities

Fluvial

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Catchment wide
within the river
channel and at
flood receptors

Reduces erosion and


stabilises mobile
sediments where
measures are
implemented and
reduces consequent
blockage of
downstream
watercourse near flood
receptors

A study of the Upper Wharfedale reported


that the reforestation of steep gullies within
the catchment would reduce the coarse
sediment supply by 85% and offer a
substantial reduction in flood risk (Lane,
2006)

Current best approach is


a (judgement based)
study by a professional
with geomorphological
expertise.

Site specific. Reduction in


flood risk can be
significant. Uncertainty is
variable depending on
cause of instability and the
available predictive tools
(e.g., success less certain
if giving a river room to
move; more certainty of
success if stabilising a
head cut that is known to
provide a major source of
sediment)

Completion of the
sediment budget
modelling

Include within
screening via sediment
budget modelling being
developed by the
hydromorphology
team. Do not include
within assessment tool
but provide outline of a
potential methodology
for local assessment.

Sub-type

Protection

(When flooding
can be
attributed to
geomorphologic
al activity such
as channel
aggradation)
Instabilities can
be natural or
caused by
humans.

May involve sediment


being stored within the
channel or on the
floodplain or being
moved to downstream
reaches

Selkirk Flood Protection Scheme (Scottish


Borders Council, 2010) reported that it was
necessary to address the cause of
deposition where it was occurring within a
modified reach on the Long Philip Burn in
addition to reducing the supply of sediments

Can reduce risk of


channel aggradation (a
process which can
trigger avulsion into
property) or channel
entrenchment (a
process which can
lead to additional
sediment input and
floodplain
disconnection)

Use of SUDS

Urban drainage
Reduce, delay
and attenuate

Entire urban
drainage
system

By increasing
attenuation, infiltration
and mimicking natural
processes the impact
of urban development
on the water
environment can be
reduced. Reduction of
discharges to
greenfield rates

Catchment data would


help identify causes of
instability principally
sediment budget data
should show where there
are zones of aggradation
and supply. Other
datasets are also likely to
be of use e.g.,
streampower and
morphological pressures
databases.

Development of new
models, or validation of
existing dynamic
models (eg CAESAR)
that are able to predict
rates of aggradation,
rates of channel
movement etc.

Hydraulic impact of
estimated or observed
aggradation can be
assessed using existing
models
Predicting rate of
aggradation or incision,
or locations of avulsion,
requires complex and
little-tested 2D dynamic
modelling and carries
high uncertainty.

The use of SUDS is well document and


understood (CIRIA, 2007) and SUDS are now
standard practice in all developments across
Scotland.
The retrofitting of SUDS to existing systems
can be expensive and technically challenging
(SNIFFER, 2006)

Appendix B - 9

Existing assessment
techniques are
appropriate

Well designed and


maintained systems are
known to be effective with
a low degree of uncertainty

Dont include within


screening. (Local
assessment better
placed to identify
opportunities.

NFM
technique

Type

Using high
water
demand
vegetation to
reduce
groundwater
levels

Groundwater

Beach
management
(beach
recharge
schemes and
shingle
management
)

Coastal

Artificial/
biogenic
reefs and
detached
breakwaters

Coastal

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Entire aquifer
(benefits for
locations where
groundwater is
close to the
surface)

High water demand


vegetation will reduce
the recharge of
groundwater. Some
species may draw on
water directly from the
groundwater system.

There is consensus within the literature that


high water demand can reduce the recharge
to aquifers due to increased
evapotranspiration.

Existing groundwater
modelling techniques are
appropriate for making
assessments

The impact of land use


change on the seasonal
recharge to aquifers can
be made with a moderate
level of confidence. The
confidence reduces with
aquifer complexity and for
flood mechanisms driven
by shorter storm durations.

National mapping of
groundwater flood risk
would be beneficial

Dont include within


screening (local
assessment better
placed to identify
opportunities). Provide
outline of potential
assessment method to
facilitate local
assessment.

Typically
exposed
coastal areas
where beaches
are eroding. If
the source of
material is
related by
coastal
processes to
the eroding
area then this
approach is
termed
recycling.

Artificial addition of
sand or gravel material
to maintain beach
levels. Increased width
and reduced slope of
beach dissipates wave
energy. Generally not
possible in many areas
as depends on the
availability of large
quantities of suitable
material from local
sources e.g. port and
harbour dredging.
Often needs hard
engineering structures
to keep material in
place and often needs
to be replenished as
material erodes away
over time.

CIRIA (2011) reports that beach recharge


projects are frequently undertaken to
supplement existing beaches to manage
coastal erosion and flooding.

Existing techniques are


appropriate.

Provided appropriate study


is undertaken the effects
can be reasonably well
predicted.

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

Exposed
coastal areas

Establishment of
biogenic reefs e.g.
oyster and mussel
beds to dissipate wave
energy.

The use of artificial reefs and near shore


breakwaters is well documented, such
structures are frequently used to reduce
coastal erosion.

Effects can be assessed


using existing techniques

Provided appropriate study


is undertaken the effects
can be reasonably well
predicted.

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

Sub-type

Reduce
recharge

Protect

Protect

Detached breakwaters
reduce wave energy
reaching the shoreline
but do not completely
isolate natural beach
processes.

Calder (1992) presented a case where the


evapotranspiration rate for a high water
demand plantation exceeded the available
rainfall by more than 50%. It was concluded
that the plantation was importing water from
other areas by drawing down the aquifer.

Beach recharge has been used at Aberdeen


and St Andrews West Strand.

Biogenic oyster reefs can be used to protect


eroding shorelines in lower energy systems
(Cyphers et al., 2011).

Appendix B - 10

NFM
technique

Type

Sand dune
restoration
(e.g. dune
fencing and
thatching,
marram
grass
planting).

Coastal

Standline
and beach
management
techniques

Coastal

Creation/
restoration of
intertidal
area
including
mudflats and
saltmarsh

Coastal

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Exposed
coastal areas
behind sandy
beaches were
sand dunes
would naturally
form

Creates a defence
between the beach and
land to dissipate
largest waves and a
direct defence to high
water levels

Defra FD1302 (2007) provides an overview


on the use of sand dunes to manage coastal
flood risk. It identifies that maintaining sand
dunes in good condition is an effective means
of reducing coastal erosion and flood risk.

Existing assessment
techniques are
appropriate

Provided appropriate study


is undertaken the effects
can be predicted with
moderate uncertainty

High amenity
beaches.

Includes hand picking


of litter rather then
mechanical removal of
litter leaving strandline
in situ to reduce
erosion and improve
opportunity for
vegetation cover to
form.

It is understood that this approach is being


trialled by West Sands Partnership, Fife and
that University of St Andrews and Abertay
University are undertaking research in this
area. No literature has been identified
however (pers comms) it is suggested that
the early results are very promising in that
retaining strandlines do increase the stability
of the shoreward edge of dune systems.

No known assessment
techniques. The validity
of the approach would
need to be assessed
using the judgement of
an experienced
geomorphologist

Due to absence of
literature there is a high
degree of uncertainty
about the magnitude of
effects.

Areas where
flats would
naturally form,
typically
estuarine

Creates an improved
buffer to dissipate
wave energy.

Mller et al. (1999) reports that saltmarshes


are more effective at dissipating wave
energy than bare sand or mud flats,
dissipating up to twice the energy.

Site specific wave run-up


modelling using currently
available software

Relatively low degree of


uncertainty (ignoring
geomorphology issues)

Sub-type

Protect

Attenuate

Protect

Research or data
need

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

SNH (2000) provides guidance on a range of


techniques which can be used to restore
dune systems. 6000cum of recycled sand
was used to restore degraded dunes at St.
Andrews West Sands in 2010.

Areas of
reclaimed land
and
uneconomic
flood defences.
Areas of
degraded
mudflats and
saltmarsh

Appendix B - 11

Up-to-doubling of wave
energy dissipation.

Inclusion

There is an absence of
literature detailing the
effectiveness of the
altered maintenance
approach

Local assessment best


placed to undertake
screening for
opportunities and more
detailed assessments.

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

NFM
technique

Type

Managed
realignement
, creation/
restoration of
intertidal
area
including
mudflats and
saltmarsh

Coastal

Location

Principal action

Literature

Model requirements

Effect and uncertainty

Research or data
need

Inclusion

Typically mid
and upper
estuaries.

Provides increased
volume for the storage
of estuarine surges.
The increased flooded
area also increases the
resistance to flow
further helping to slow
propagation up
estuary.

FHRC (2002) there is consensus between


experts is that the method is effective at
reducing estuarine surge.

Effects can be modelled


using existing
hydrodynamic models

Moderate uncertainty

Baseline hydraulic
models of all estuaries
linked to potentially
vulnerable zones would
be beneficial.

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

Effects can be modelled


using existing
hydrodynamic models

Moderate uncertainty

Existing coastal surge


analysis software is
thought to be capable of
representing the effect

High degree of uncertainty


due to absence of
literature.

Sub-type

Delay/attenuate

Areas of
reclaimed land
and
uneconomic
flood defences.

Allows natural
processes to continue
with no active
intervention. Ensures
natural morphological
processes are
sustained.

Regulated
tidal
exchange

Coastal
Attenuate

Typically
estuarine
Areas of land
claim and
uneconomic
flood defences.

Provides increased
volume for the storage
of estuarine surges.
Can form buffer to
dissipate wave energy.

Scott et al (2011) identifies 51 managed


breach, regulated exchanges and defence
removal projects in the UK between 1990 and
2011 which create in total 1300Ha of coastal
retreat.

A reduction of 150mm was


reported in the Hull
Estuary. Large areas of
retreat will be most
effective although location
and connectively will be
important factors too.

Coastal retreat at Alkborough (EA, 2008) is


estimated to reduce the 1 in 200yr water
level in the surrounding Humber Estuary by
approximately 150mm.
The approach of no active intervention has
been adopted as components of shoreline
management plans, for example the Fife
Shoreline Management Plan (Fife Council,
2011). By allowing erosion to occur this
provides a supply of sediment to other
sections of the coastline.
FHRC (2002) there is consensus between
experts is that the method is effective at
reducing estuarine surge.
Scott et al (2011) identifies 51 managed
breach, regulated exchanges and defence
removal projects in the UK between 1990 and
2011 which create in total 1300Ha of coastal
retreat.

Include within
screening. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

Large areas of retreat will


be most effective although
location and connectively
will be important factors
too.

How the regulated exchange is operated can


be significant. Extent, level relative to mean
sea level and location within estuary will be
important
Removal of
artificial
channels

Coastal
Delay/attenuate

Typically
estuarine
Areas where
artificial
channels have
been
created/maintai
ned for shipping
or drainage
purposes.

These channels may


provide conduits for
coastal surges to
rapidly propagate
inland/upstream/up
estuary

No previous studies or projects identified


within the literature.

Dredging of channels
can be detrimental to
surrounding flats

Appendix B - 12

Effects unknown.

A dataset of artificially
sustained channels
would be beneficial

Not included within


screening, local
assessment best
placed to identify
opportunities. Dont
include within
assessment tool.
Provide outline of
methodology which can
be used for local
assessment.

Appendix C
Other considerations

Draft

Neil Nutt

Blockage risk
It was identified that culverts and bridges with openings greater than approximately 6m are less
prone to blockage during extreme flood events (Rigby et al., 2002)1. The study was based on
four forested catchments within Australia following a large flood event. Despite the blockage
condition of 92 culverts being observed no correlations between any of the investigated factors
could be identified.
A more recent study of blockage risk at 140 screened culverts in Belfast (Wallerstein et al.,
2010)2 found that organic material such as leaves, twigs and branches form the bulk of material
found to be captured at culvert screens. The study identified that there was a strong seasonal
trend for blockage caused by small organic material such as leaves, with the highest risk being
in November and correlating with the autumn leaf drop from deciduous trees.
It is proposed in Johnson (2007)Error! Bookmark not defined. that the use of floodplain
woodland can serve to trap floating debris helping to reduce the risk of blockage at downstream
culverts. Nisbet et al (2011)Error! Bookmark not defined. that short rotation forestry, and in
particular short rotation coppice, may be particularly effective at entrapping floating debris
before it forms blockages at culverts.
In summary, while there is little scientific information on the subject, it would seem prudent that
due consideration is given to the impact of increasing the supply of organic material might have
to the blockage of constricted downstream structures or those which are fitted with a screen.

Reservoir and loch management


It is accepted that modifying the operational strategy of existing reservoirs and lochs is at the
periphery of what might be considered natural flood management, however such practices could
be considered aligned with the Acts definition which includes the enhancement of an existing
function for the purpose of flood risk management. In the right circumstances changes to
reservoir and loch management may be considered a sustainable approach.
The Water of Leith Flood Prevention Scheme is reported to benefit from the changed
management philosophy of three existing reservoirs located within the catchment (Arup, 2003)3.
The Victorian reservoirs, which were constructed for industrial water supply, had notches cut

Rigby E.H., Boyd M.J., Roso S., Silveri P., Davis A., Causes and effects of culvert blockage
during large storms, Forbes Rigby Consultants & Wollongong City Council, 2002.
2

Wallerstein N., Arthur S., Sisinnggih D, Towards predicting flood risk associated with debris at
structures, (unpublished), 2010.
3

ARUP Water, Water of Leith Flood Prevention Scheme, Hydrological and hydraulic modelling
Design Report, Volume 1. City of Edinburgh Council, 2003.

Appendix C - 1

within their spillways so that reservoir storage could be reserved for the flood peak. The use of
the three reservoirs is understood to reduce the 1 in 200yr plus climate change flow from
approximately 108cumecs to approximately 76cumecs significantly reducing the extent and high
of traditional direct defences downstream.
It is understood that Scottish Borders Council proposes to modify the existing outlet control
structure at St Marys Loch to both reduce rural flooding along the Yarrow Water and to provide
compensatory storage for the areas of active floodplain that will be lost through the construction
of direct defences as part of the Selkirk Flood Protection Scheme (Halcrow, 2011)4. The
reduction of the lochs normal operating level by approximately 250mm and the raising of the
existing flood spillway by 250mm will significantly reduce rural flood damages along the Yarrow
Water and will more than compensate for the loss in floodplain volume as a consequence of
direct defences within the town of Selkirk. The estimated reduction in flood flows in the Yarrow
Valley range between 10-30% with the effect reducing by event magnitude. The reduction in
water storage available for release as freshets or compensation flow is overcome via the
replacement of a manually controlled release with an automated system.

Backwater effects
A number of natural flood management measures have been highlighted as having a potential
impact on upstream flood risk through the creation of backwater effects. A Forest Research
paper on restoring floodplains for flood alleviationError! Bookmark not defined. found for one
model based case study the backwater could extent upstream between 120-330m upstream of
introduced forest floodplain but that the effect would be primarily dependent on channel gradient
in addition to the increase in water depth. The potential to increase water levels upstream
highlights the need to give due consideration to the placement of in flow obstructions (such as
floodplain woodland, in channel interventions or riparian vegetation). Unfortunately there is an
absence of appropriate literature discussing what the key variables controlling the extents are.
Therefore a look at the issue from first principles is presented below to provide an initial
overview.
The rate of change of flow depth within a gradually varied flow is defined by Equation 1. Where
the Froude Number ( ) is defined in Equation 2 and the hydraulic gradient (
approximated using Mannings Equation (Equation 3).

) can be

Equation 1: The general differential equation of the water surface profile for gradually varied flow

Halcrow Group Limited, Selkirk Flood Protection Scheme; St Marys Loch Flood Storage
Option; Feasibility Study. Scottish Borders Council, 2011.

Appendix C - 2

The Froude Number (


detailed in Equation 2.

) is the local flow velocity ( ) divided by the wave celerity ( ) as

Equation 2: Froude Number for an open channel

The Mannings equation can be used to express the hydraulic gradient (


Mannings roughness(

), flow (

), wetted perimeter (

) and flow area (

) in terms of
).

Equation 3: Mannings equation

By rearrangement and integration of Equation 1 the change in distance between two depths can
be expressed as shown in Equation 4.

Equation 4: The distance between two depths in a backwater profile

By considering the channel geometry, channel slope and flow to be constant in space and time
it can be concluded that the backwater effect diminishes asymptotically with the extent being
dependent on channel slope, roughness, flow and deviation above the normal depth as follows:

The backwater extent is inversely related to channel slope, that is shallower slopes will
generate longer back water extents

The backwater extent increases with flow and roughness

Further discussion and analysis are presented in the following draft paper.

Appendix C - 3

Blanger first correctly described the


backwater effect within an open channel
where the flow is gradually varied in 1828
(Chanson, 2009). The M1 backwater profile
is presented in Figure 1.

The backwater effect generated


by floodplain roughening:
Working from first principles
Neil Nutt CEng MICE, Halcrow Group
Limited, Edinburgh

The rate of change of flow depth within a


gradually varied flow is defined by Equation
1. Where the Froude Number ( ) is
defined in Equation 2 and the hydraulic

Abstract
There is currently great interest in the use of
natural processes to manage flood risk.
One such process is the use of floodplain
woodland to increase floodplain roughness
and thus help attenuate floods. Concern
has been raised that the elevated water
levels (backwater) generated by floodplain
roughening could propagate a significant
distance upstream thus increasing upstream
flood risk. In an absence of published
material on the extent of backwaters this
paper will investigate the nature of these
backwater profiles via first principles.

gradient ( ) is defined by Mannings


Equation (Equation 3).

Introduction
There is increasing interest in the use of
natural processes to manage flood risk
(Scottish Government, 2009; Pitt, 2008).
Woodland for Water (Nisbet et al, 2011)
discusses the practice of using vegetation
roughen floodplains thus serving to slow the
flow of flood water and increasing the depth
of flood water that can be held on the
floodplain. The elevated water profile
(backwater) backwater created by this
mechanism has been highlighted as having
the potential to extend a significant distance
upstream leading to an increase in
upstream flood risk (Thomas & Nisbet,
2006; Nisbet & Thomas, 2008). However,
there is an absence of literature on the
subject on the factors which influence the
magnitude and extent of these backwaters.
It is proposed to investigate the factors
which influence the extent and magnitude of
backwater effects from first principles.

Draft

Figure 1: The M1 backwater profile

Equation 5: The general differential equation


of the water surface profile for gradually
varied flow

The Froude Number (

) is the local flow

velocity ( ) divided by the wave celerity ( )


as detailed in

Equation 6: Froude Number for an open


channel

The Mannings equation can be used to


express the hydraulic gradient (
of Mannings roughness(
wetted perimeter (

) in terms

), flow (

),

) and flow area (

).

Neil Nutt

(Figure 3) and flow (Figure 4) on the


backwater length can be investigated.
Equation 7: Mannings equation

By rearrangement and integration of


Equation 1 the change in distance between
two depths in a backwater profile can be
expressed as shown in Equation 4.

Equation 9: Numerical integration using


Simpsons rule

The test range for variables was set based


on typical values often found within British
catchments as detailed in Table 1.
Variable range
0.0002<S<0.01

Equation 8: The distance between two


depths can be expressed as follows

Within natural river systems there is the


potential for the channel geometry, slope,
flow and roughness to vary both spatially
and temporally. For the purpose of this
study these variables will be fixed as
constants. It will also be assumed that:
1. The gradient of the channel is mild,
that is the slope is such that the
normal depth is greater than the
critical depth.
2. The channel is a wide uniform
rectangular channel such that the
wetted perimeter reduces to the
channel width
3. The flow depth is increased to a
value greater than the normal depth
(
)
4. Since deviations from the normal
depth will return to the normal depth
asymptotically it will be assumed
that the backwater effect will
terminate when the flow depth
returns to 1mm above the normal
depth (
)

Source
(Leopold and Maddock,
1953)
0.015<n<0.2
Conveyance
Estimation System
(Samuels et al., 2002)
0.01<Q<5cumecs Broad range in
absence of guidance
from literature
0.05<dy<0.5m
(Thomas & Nisbet,
2006; Nisbet &
Thomas, 2008)
Table 1: Variable ranges and sources used
in the analysis

Figure 2: Graph showing the effect of


varying Mannings Roughness on backwater
length for a wide uniform rectangular
channel with Q=1cumec/m and S=0.002m/m.

Using Simpsons rule (Equation 9) to


numerically integrate Equation 4 the effect of
Mannings roughness (Figure 2), slope

Appendix C - 2

To aid initial screening of floodplain


roughening opportunities it would be
reasonable to consider a simplified system
based on valley slope (assuming n<0.05,
Q<1cumec/m & dy<0.5m). Figure 5 shows
that a simplified system based on only slope
can provide a rapid means of estimating
backwater lengths for screening purposes.

Figure 3: Graph showing the effect of varying


slope on backwater length for a wide uniform
rectangular channel with Q= 1cumec/m and
n=0.03.

Figure 5: Graph showing the effect of


varying valley slope on backwater length for
a wide uniform rectangular channel with
Q=1cumec/m, n=0.05 and dy=0.5m.

Figure 4: Graph showing the effect of


varying flow on backwater length for a wide
uniform rectangular channel with
QS=0.002m/m and n=0.03.

Discussion
The analysis shows that floodplain
roughening on shallow gradient rivers has
the greatest potential to generate an
extensive backwater. Increasing roughness
and flow also serve to increase backwater
length but their effects are less marked.
The magnitude by which the water depth is
increased over the normal depth does serve
to increase the backwater length, however
the extent is largely insensitive to small
changes.

It is accepted that this paper presents a


much-simplified overview of the backwater
effect. However, it should be noted that
presenting a more detailed formulation,
such as using a more complex channel
section or by allowing localised variation in
slope, roughness and flows to assess the
backwater effect would simply allow an
understanding of the represented system
rather than provide an insight into the
general underlying principles. By
understanding the underlying relationships it
is proposed that it will be feasible to
translate the findings to more complex
systems.

Conclusion
Through the consideration of first principles,
this study has found that the backwater

Appendix C - 3

effect diminishes asymptotically with the


extent being dependent on slope,
roughness, deviation above the normal
depth and flow. The extent of a backwater
can be estimated (with a reduced accuracy)
when screening for floodplain roughening
opportunities by using a simplified
relationship based on valley slope only.
While this study does serve to provide basic
guidance it cannot, and should not, replace
detailed modelling of the actual system.

Pitt M., (2008) The Pitt Review: Learning


lessons of the 2007 floods. The Cabinet
Office.
Samuels P., Bramley M.E. and Evans E.P.,
(2002) A New Conveyance Estimation
System, DEFRA.
Thomas, H. and Nisbet, T. R. (2007), An
assessment of the impact of floodplain
woodland on flood flows. Water and
Environment Journal, 21: 114126.
doi: 10.1111/j.1747-6593.2006.00056.x

References
Blanger J.B., (1828) "Essai sur la Solution
Numrique de quelques Problmes Relatifs
au Mouvement Permanent des Eaux
Courantes." ('Essay on the Numerical
Solution of Some Problems relative to
Steady Flow of Water.') Carilian-Goeury,
Paris, France (in French).
Chanson, H., (2008) "Jean-Baptiste Charles
Joseph Blanger (1790-1874), the
Backwater Equation and the Blanger
Equation." Hydraulic Model Report No.
CH69/08, Div. of Civil Engineering, The
University of Queensland, Brisbane,
Australia, 40 pages
Leopold L.B. and Maddock T., (1953)The
Hydraulic Geometery of Stream Channels
and Some Physiographic Implications.
United States Department of the Interior.
Nisbet T.R., Silgram M., Shah N., Morrow
K., Broadmeadow S., (2011) Woodland for
water: Woodland measures for meeting
Water Framework Directive objectives.
Forest Research.
Nisbet T.R., Thomas H., Restoring
Floodplain Woodland for Flood Alleviation,
DEFRA, 2008.

Appendix C - 4

Appendix D
Requirements on SEPA in relation to Natural Flood
Management under the FRM Act

Annex 1: FRM Act Section 20 requirements


Assessment of possible contribution of alteration etc. of natural features and characteristics
(1) SEPA must, by 22nd December 2013 or such other date as the Scottish Ministers may
direct, assess whether alteration (including enhancement) or restoration of natural features and
characteristics of any river basin or coastal area in a flood risk management district could
contribute to the management of flood risk for the district.
(2) For the purposes of this Act, natural features and characteristics include such features and
characteristics which can assist in the retention of flood water, whether on a permanent or
temporary basis, (such as flood plains, woodlands and wetlands) or in slowing the flow of such
water (such as woodlands and other vegetation), those which contribute to the transporting and
depositing of sediment, and the shape of rivers and coastal areas.
(3) SEPA must
(a) by such date as the Scottish Ministers may direct, review and where appropriate update its
assessment under subsection (1), and
(b) by the end of the period of 6 years beginning with that date, and of each subsequent period
of 6 years (or, in each case, such lesser period as the Scottish Ministers may direct) review and
where appropriate update the latest assessment reviewed under this subsection.
(4) Each assessment under subsection (1), and each assessment updated after review under
subsection (3) must
(a) take into account the flood risk assessment, any flood hazard maps and flood risk
maps and any flood risk management plan for the time being applicable to the flood risk
management district, and
(b) refer to a map showing where alteration (including enhancement) or restoration of natural
features and characteristics of any river basin or coastal area in the district could contribute to
management of flood risk in the district.
(5) The map referred to in subsection (4)(b) must be prepared at a scale which SEPA considers
will assist in
(a) the identification of measures under section 27(4)(b), and
(b) the inclusion of information in local flood risk management plans under section 34(3)(b)(ii)
and (4)(c)(i).
(6) SEPA must make available for public inspection copies of
(a) any assessment under this section for the time being applicable to each flood risk
management district, and
(b) the map to which the assessment refers.

Annex 2: Associated provisions under the FRM Act


Section 28 (1): In setting objectives and identifying measures under section 27(4) SEPA must take account of, so far as relevant(iii) any assessment done under section 20 in relation to the district
must consider, so far as is appropriate, both structural and non-structural measures as means
of achieving objectives
Section 28(2): for the purpose of 1(b), a measure is structural if it involves flood protection
works
Section 28 (3): In considering structural measures under subsection (1)(b), SEPA must
consider measures that seek to reduce, slow, or otherwise manage flood water by altering
(including enhancing) or restoring natural features and characteristics.
Section 28 (4): The measures considered in pursuance of subsection (3) must include
measures that consist of carrying out any alteration or restoration of natural features and
characteristics identified as being capable of contributing to the management of flood risk in an
assessment done under section 20 in relation to the district
Section 44 (3)(b)(ii): the supplementary part of a local flood risk management plan must include
information about how implementing the measures identified in the relevant part of the district
plan will restore, alter or enhance natural features and characteristics
Section 49 (2)(a)(iii) specifies one of the roles of district advisory group to advice SEPA on 20
assessment
Section 50 (3)(a)(iii) specifies one of the roles of sub-district advisory groups to advice SEPA
and local authorities on section 20 assessment.
Flood protection works - do whole list means any operation on land for the purpose of protecting
any land from flooding including any works that involve - the alteration (including enhancement)
or restoration of natural features and characteristics of any river basin or coastal areas.
Schedule 1, Part 1: Matters to be included in every FRM plan
Paragraph 5: A description of:
- (a): in relation to each measure included in the plan under paragraph 1 (b), the
reasons for identifying the measure, and
- (b) in relation to any alteration (including enhancement) or restoration of a
natural feature or characteristic in the flood risk management district which (i) is identified in an assessment under section 20,
(ii) could contribute to the management of flood risk, and
(iii) is not to be carried out by a measure included in the plan under paragraph
1(b),
the reasons why no such measure has been identified.

Annex 3: Policy intention for Section 20


Source: http://www.scottish.parliament.uk/parliamentarybusiness/Bills/16275.aspx
The Convener: Group

10 is on natural features and natural characteristics. Amendment 9, in the


name of the minister, is grouped with amendments 10, 11, 17, 67, 18 to 20, 68, 69, 21 to 24, 85,
35, 37, 38, 40 to 42, 53, 57, 58, 88, 61, 63 and 64.
This extensive group of amendments reflects the long and detailed
conversations that have taken place on natural flood management, which has been a key topic of
discussion for the committee. There are a lot of points for me to address, so I will have to spend
a little time on the group. I ask members to bear with me.
Roseanna Cunningham:

Amendments 9 to 11, 17 to 24, 35, 37, 38, 40 to 42, 53, 57, 58, 61, 63 and 64 relate directly to
the assessments of natural features that will be made under section 16. I will lodge separate
amendments to later sections that deal with wider issues concerning the role of natural flood risk
management and its relationship to sustainable flood risk management. I will refer to those
amendments when I address the amendments that committee members have lodged.
We all appreciate that the evidence base supporting the use of more natural approaches to
managing flooding is still limited and evolving. However, available information suggests that a
number of techniques show significant promise. Furthermore, we should aim to take advantage
of the added benefits that can be gained from adopting more natural approaches to managing
flooding, which include environmental and social benefits.
The Government is in the process of developing a long-term research strategy to co-ordinate
investment in improving our understanding of natural flood management techniques. We have
also set up a stakeholder group to advise the Government on natural flood management. Those
important steps will contribute to the cultural shift that, as the committee has rightly highlighted,
must occur.
The Scottish Government has worked closely with stakeholders on considering the role of
natural flood management and the committee's recommendations on section 16. The
amendments to which I will speak have been drafted in close consultation with stakeholders,
including SEPA, Scottish Natural Heritage and Scottish Environment LINK.
Amendments 18, 23 and 24 mean that the assessments that will be prepared under section 16 will
consider not only natural features but how alterations or restoration of natural characteristics of
river basins and coastal areas could contribute to managing flood risk. The reference to river
basins indicates that the emphasis is clearly on taking a catchment-based approach.
Amendments 19 and 20 replace the examples of natural features that are set out in the bill with a
new set of examples that covers natural features and characteristics of river basins and coastal
areas. In setting out those examples, we have given particular consideration to ensuring that they
express some of the key concepts and aims of natural flood management, including using natural
features and characteristics to assist in the retention or slowing of flood water. I stress that the
examples that are set out in the amendments are not intended to form an exhaustive list. In
compiling them, the Government liaised closely with key stakeholders, including Scottish
Environment LINK and SEPA.

Amendment 17 sets a dateDecember 2013by which assessments must be prepared. That


date has been chosen to ensure that assessments can be co-ordinated with other important steps
in the flood management process and with work that is being carried out under the Water
Environment and Water Services (Scotland) Act 2003.
Amendments 21 and 22 are consequential on the introduction of a date and ensure that, when
preparing an assessment, SEPA will give consideration to those maps and plans that are
available when the assessment is made.
Amendments 9, 10, 11, 37, 38 and 53 are consequential amendments that ensure that the new
terminology that is introduced in section 16 is applied elsewhere in the bill, including in section
9, which deals with flood risk assessments.
Amendment 35 gives SEPA the power to obtain information, documents and assistance under
section 37 when it is preparing an assessment.
Amendments 57 and 58 make consequential changes to section 79, which as introduced would
give SEPA a more limited power to obtain information when performing its functions under
section 16. The new wider-ranging power will ensure that SEPA can obtain not only information
about land but any other information or assistance to support the preparation of assessments,
which could include seeking assistance from local authorities.
Amendment 40 applies to the bill the same definitions of a body of surface water, a loch and a
watercourse as are used in the Water Environment and Water Services (Scotland) Act 2003. That
is part of the integration that I mentioned.
Amendment 41 amends the definition of a sub-basin.
Amendment 42 changes the definition of a wetland to include wetlands that do not depend
directly on bodies of surface water or groundwater for their water needs. That will ensure that all
wetland features, including rainwater-fed wetlands that might help to slow run-off, are captured
by the definition of wetland in the bill. We are trying to include as much as possible.
Amendments 61, 63 and 64 make consequential changes to the index of defined terms in
schedule 4.
Amendment 88, which Rhoda Grant lodged, would amend the definition of flood protection
work for the whole bill by replacing the reference to
"the sowing or planting of vegetation or forestry"
with a reference to
"the alteration ... or restoration of natural features".
Amendment 88 would make the definition slightly narrower than that in the bill.
The amendments that I have lodged to section 16 will ensure that
"the alteration ... or restoration of natural ... characteristics of any river basin or coastal area"
5

are recognised as flood protection measures when they contribute to managing flood risk. I am
confident that those amendments, combined with the definition of flood protection work, which
covers
"any work of construction, alteration, improvement, repair, maintenance, demolition or
removal",
will be wide enough to encompass
"the alteration ... or restoration of natural features and characteristics of any river basin or coastal
area".
For those reasons, I do not support Rhoda Grant's amendment 88 and I ask her not to move it.
Amendment 69 was lodged by Elaine Murray. I agree fully that the assessments that are prepared
under section 16 should be subject to consultation with local authorities, other responsible
authorities and other key organisations, including SNH. That is why the bill requires SEPA to
have regard to the advice of advisory groups when preparing assessments. I assure the member
that the bodies that her amendment names will need to be involved in the preparation of
assessments and that SEPA will be required to have regard to their advice. To initiate a separate
consultation exercise would risk undermining the important role that is envisaged for advisory
groups. For those reasons, I do not accept amendment 69. The bill already makes adequate
provision in that respect, therefore I ask the member not to move the amendment.
Amendment 85, which was lodged by Peter Peacock, does not relate to the content of
assessments that are prepared under section 16. Instead, it relates in part to how the information
that is garnered from assessments is used when SEPA sets objectives and measures to manage
flood risk for inclusion in flood risk management plans. I would first like to clarify that the bill
already requires SEPA to have regard to section 16, along with various other important factors,
when it identifies objectives and measures.
The other aspects of Peter Peacock's amendment would require SEPA to select the most
sustainable measures and to give reasons for selecting measures that are inconsistent with things
that are identified through assessments that are made under section 16. On the first point, I will
later discuss a proposed Government amendment that will require SEPA to select measures that
will achieve objectives in the most sustainable way. On the second point, I fully support the
principle that SEPA should be required to set out reasons for selecting particular measures and I
believe that that principle should apply to all measures, not only to those that relate to section 16.
I ask Peter Peacock not to move amendment 85, on the understanding that I will bring forward
an amendment at stage 3 to require SEPA to set out the reasons behind the selection of measures.
I move amendment 9.
Peter Peacock: I welcome all

the minister's amendments in the group and my colleagues'


comments on them. I will focus on amendment 85. As the minister rightly acknowledged, natural
flood management was one of the issues that got the most attention during the committee's
flooding inquiry and in its stage 1 report. There is no doubt that the committee is attracted to
using more natural approaches to flood risk management. Those approaches might be more
sustainable than hard engineering solutions, with the additional benefit that they can create or
restore habitat, which relates to other aspects of Government policy. The committee argued that,
6

where natural approaches can be used to remove the peaks of floods and reduce the need for hard
engineering solutions downstream, they should be consideredI stress the word "considered".
The stage 1 report is clear that natural flood management techniques will not work in every
circumstance and are therefore not a panacea. The committee was equally clear that, as others
have mentioned, a change of culture is required in the consideration of future flood management
approaches. If we have a system for developing flood risk management plans that is dominated
by engineersestimable people though they are, and I include John Scott in thatwe should not
be surprised if the subsequent debate is dominated by engineering solutions or arrangements for
flood management.
Recommendation 15 in the committee's stage 1 report calls for measures that are likely to cause
the shift in culture that the committee feels to be necessary, but which are not prescriptive about
the use of natural flood management techniques. The committee wants to
"require responsible authorities to consider what contribution natural flood management
approaches could make",
and recommends that
"Such an amendment should stipulate that, where natural flood management approaches are
assessed as being able to make such a contribution but are not proceeded with, authorities must
set out the reasons for that decision."
Amendment 27, which the minister mentioned and which we will debate later, is the
Government's way of addressing the issue. However, it does not go far enough and it is not
explicit enough. Amendment 85 would deliver the committee's unanimous recommendation, and
I hope that it embraces the purpose of the Government's amendment 27.
Amendment 85 is both reasoned and reasonable and I urge the committee to support it. It seeks
to ensure that natural flood management techniques are considered, but it would not require
action using those techniques to be taken above any other action. Under amendment 85, when
the contribution that natural techniques could make was considered and it was decided not to use
them, the reasons for the decision would have to be set out. The amendment would not require
natural flood management techniques always to prevail, which is an understandable concern of
the Government. It would provide for the necessary considerations to bring about the change in
culture to which the committee has referred.
I noted the minister's comments about the intention to return at stage 3 with an amendment to
take care of the reporting aspects. I welcome that, but it does not deal with the central point of
amendment 85of which reporting is a partwhich is the requirement to consider natural
approaches, although not necessarily to implement them. I stress that point, which would give
discretion to SEPA and local bodies.
Liam McArthur: The

committee deliberated the presumption in favour of natural processes and


features for quite a while, and the evidence that we took indicated that there was confusion about
the implications of that presumption. Peter Peacock's amendment 85 would get round some of
those difficulties. I acknowledge the intention of amendment 27 and what the minister said about
amendments that she may lodge at stage 3, but amendment 85 would implement a central
recommendation from the committee's stage 1 report. It would facilitate the cultural shift for
which we called; it would not prescribe or proscribe particular approaches but would ensure that
7

hard engineering is not the approach with which local authorities and SEPA naturally kick off
the debate about flood risk management.
I am inclined to support amendment 85. Natural flood management was central to
the committee's report and, if amendment 85 would help to deliver that, it is important.
Notwithstanding the minister's intention to introduce amendments at stage 3, amendment 85 is
worthy of support.
John Scott:

Roseanna Cunningham: Committee members

have raised a number of points and there have been

some interesting conversations.


On amendment 85, I have already indicated that I would be prepared to lodge a stage 3
amendment to require SEPA to set out the reasons behind its selection of measures. We will
come to other amendments that deal with other aspects of amendment 85, but I point out that
section 24 already requires SEPA to consider section 16 in setting objectives and measures.
Many provisions in the bill interrelate with one another, and amendments to one section are not
necessarily the only amendments that relate to a particular issue. I said right at the start that there
are other amendments that relate to natural flood processes and I have said clearly that the
Government has been guided all the way through the process by our consultation with a wide
variety of stakeholders. Therefore, the amendments that we have lodged are not capricious in any
way.

Appendix E
Examples Effect of land use on runoff generation
Initial estimation of routing model coefficients

1.

Effect of land use on runoff generation - example

The proposed runoff generation model has been tested for a range of land use types
over a range of rainfall depths. The testing has been undertaken using HOST class
19 as this is a relatively common soil class that is susceptible to degradation by land
use. It should be noted that this analysis represents the effect of land use on the
runoff generated by a single cell within the proposed hydrological model and does
not provide a complete statement on the flood flow generated by a catchment.
The variation in initial soil moisture with land use has been represented via the
modification of PROPWET using the following equation:

PROPWET 0.5365Log10 SAAR PE LAI 4.70 r 0.79 2.003E 5 0.998


In cases with high water use the PROPWET has been constrained to a minimum
value of 0.3. This decision is based on the absence of high leaf area index and high
potential evapotranspiration from the development of the above equation.
The variation in soil moisture capacity has been varied with soil degradation using
the method of analogue degraded HOST classes presented by Packman et al.
(2004). In the case of the HOST class 19, this degrades to HOST class 22.
The following land use scenarios have been investigated:
Scenario

LAI for
PROPWET

LAI for
r
interception

BFI
(HOST)

PE

SAAR

Baseline

PROPWET
from donor
catchment
6
6

0.47 (19)

6
6

70
80

0.47 (19)
0.47 (19)

500mm/yr
500mm/yr

1200mm/yr
1200mm/yr

80

0.47 (19)

500mm/yr

1200mm/yr

60

0.47 (19)

500mm/yr

1200mm/yr

60

0.47 (19)

500mm/yr

1200mm/yr

40

0.32 (22)

500mm/yr

1200mm/yr

40

0.32 (22)

500mm/yr

1200mm/yr

70

0.47 (19)

500mm/yr

1200mm/yr

Conifer forest
Broad leaf forest
(summer)
Broad leaf forest
(winter)
Unimproved
pasture
Semi improved
pasture
Improved
pasture
Potatoes (winter
bare earth)
Conifer forest
(with
HoyningenHunene canopy)

The testing is presented for a range of rainfall depths (15mm to 150mm) and it is
intentional that no guidance has been provided on what event rarity these depths
represent. As discussed in the FEH vol 2 the rarity of a given rainfall depth is related
to the locality, storm area and storm duration. The range of rainfall depths has been
selected as the lower end is broadly representative of a frequent flood event in a
small catchment through to a relatively rare flood event in a large catchment at the
upper end.
The results of the testing are presented in Figures 1 to 6 for the selected range of
rainfall depths. These results are then summarized in Figure 7 and 8 to provide an
indication of the impact land use has on runoff generation. In summary the analysis
indicates that the impact of soil degradation increases with event magnitude, while
the impact of forestation decreases. The analysis introduces an interesting question
of that to compare the results to. While a baseline scenario has been presented
using PROPWET from a selected donor catchment it is cannot be stated that this
donor PROPWET is representative without further work.
It should be stressed that these results are provided as simply a guide to the general
trends which the proposed method might produce. There is still a large requirement
for calibration and verification (and in the case of PROPWET a method selected)
therefore the magnitude of effects reported by the final tool could be significantly
different.

Runoff generation for a range


of land cover conditions
HOST class 19 - 15mm rainfall depth
Conifer forest

1.8
Broad leaf forest summer
Runoff depth per time interval (mm)

1.6
Broad leaf forest winter

1.4
Semi improved pasture

1.2
Improved pasture
(degraded soil)

1.0

Unimproved pasture

0.8
Potatoes (winter bare
earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

0.6
0.4
0.2
0.0
0

10

15

Time interval (unitless)

Figure 1: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 15mm rain storm

Runoff generation for a range


of land cover conditions
HOST class 19 - 30mm rainfall depth
Conifer forest

4.0
Runoff depth per time interval (mm)

Broad leaf forest summer

3.5
Broad leaf forest winter

3.0
Semi improved pasture

2.5

Improved pasture
(degraded soil)
Unimproved pasture

2.0
1.5

Potatoes (winter bare


earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

1.0
0.5
0.0
0

10

15

Time interval (unitless)

Figure 2: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 30mm rain storm

Runoff generation for a range


of land cover conditions
HOST class 19 - 60mm rainfall depth
Conifer forest

9.0

Broad leaf forest summer

Runoff depth per time interval (mm)

8.0

Broad leaf forest winter

7.0

Semi improved pasture

6.0
Improved pasture
(degraded soil)
Unimproved pasture

5.0
4.0

Potatoes (winter bare


earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

3.0
2.0
1.0
0.0
0

10

15

Time interval (unitless)

Figure 3: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 60mm rain storm

Runoff generation for a range


of land cover conditions
HOST class 19 - 90mm rainfall depth
Conifer forest

16.0
Runoff depth per time interval (mm)

Broad leaf forest summer

14.0
Broad leaf forest winter

12.0

Semi improved pasture

10.0

Improved pasture
(degraded soil)
Unimproved pasture

8.0
6.0

Potatoes (winter bare


earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

4.0
2.0
0.0
0

10

15

Time interval (unitless)

Figure 4: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 90mm rain storm

Runoff generation for a range


of land cover conditions
HOST class 19 - 120mm rainfall depth Conifer forest
Broad leaf forest summer

25
Runoff depth per time interval (mm)

Broad leaf forest winter

20

Semi improved pasture


Improved pasture
(degraded soil)
Unimproved pasture

15

Potatoes (winter bare


earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

10

0
0

10

15

Time interval (unitless)

Figure 5: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 120mm rain storm

Runoff generation for a range


of land cover conditions
HOST class 19 - 150mm rainfall depth Conifer forest
Broad leaf forest summer

30
Runoff depth per time interval (mm)

Broad leaf forest winter

25

Semi improved pasture


Improved pasture
(degraded soil)
Unimproved pasture

20
15

Potatoes (winter bare


earth) (degraded soil)
Forest (Hoyningen-Huene
interception)
Baseline

10
5
0
0

10

15

Time interval (unitless)

Figure 6: Runoff depth generated for each time interval (13 interval storm) for a range
of land use types for a 150mm rain storm

Peak runoff depth per time interval (mm)

Runoff generation for a range of land cover conditions HOST class 19


30mm

Baseline
Conifer forest

25mm

Broad leaf forest summer

20mm

Broad leaf forest winter


Unimproved pasture

15mm
Semi improved pasture

10mm

Improved pasture (degraded


soil)
Potatoes (winter bare earth)
(degraded soil)
Forest (Hoyningen-Huene
interception)

5mm

mm
mm

20mm

40mm

60mm

80mm

100mm 120mm 140mm 160mm

Storm rainfall depth (mm)

Figure 7: Peak runoff depth for a range of land use types over a range of rain storm
depths

Baseline

130%

Conifer forest

120%

Broad leaf forest summer

110%

Broad leaf forest winter

100%

Unimproved pasture

baseline

Peak runoff depth per time interval percent of

Runoff generation for a range of land cover conditions HOST class 19


140%

Semi improved pasture

90%

Improved pasture (degraded


soil)
Potatoes (winter bare earth)
(degraded soil)
Forest (Hoyningen-Huene
interception)

80%
70%
60%
mm

20mm

40mm

60mm

80mm

100mm 120mm 140mm 160mm

Storm rainfall depth (mm)

Figure 8: Percentage change in the peak runoff depth for a range of land use types
over a range of rain storm depths

2.

Effect of land use on runoff generation - example

Initial testing of the flow routing model has been undertaken for a collection of
Scottish gauged catchment with reported observed catchment lag times in the Flood
Estimation Handbook vol 4 (1999)i as presented in Table 1. The aim of this initial
testing phase was to demonstrate that the flow routing model can demonstrate
reasonable potential in matching the observed data.
Catchment ID

Name

Observed lag
(hrs)

Number of
observed events

19001
Almond
8.2
5
20001
Tyne
10.8
10
21030
Megget Water
6.9
4
77002
Esk
7.3
46
79006
Nith
7.7
25
84002
Calder
3
4
84008
Rotten Calder Wtr
5.1
7
84012
White Cart Water
7.6
6
85002
Endrick Water
5.7
4
Table 1: Observed catchment lag times published in the FEH vol 4 for a selection of
Scottish catchments

The first step was to identify an appropriate initial estimate of the sinuosity factor,
which would be taken to be a universal constant for both the channel and land flows.
The initial estimate was made by investigating the impact of DTM grid resolution on
the drainage path length. In the absence of whole catchment coverage by LiDAR it
was necessary to resort to the 5m NextMap DTM. The selected catchment area was
the River Teviot at Hawick being a moderate sized catchment with no open
waterbodies. The effects of changing grid resolution on the maximum drainage path
length (LDP) and the average (DPLBAR) are presented in Table 2. The analysis
assumes that the extrapolated drainage path length using a 1m grid resolution is
representative of the actual drainage path length. It is acknowledged that the
analysis is sensitive to both the function used for the extrapolation and the selection
of 1m being representative of the true answer, which could be argued to be only
reached with a resolution of 0m however this could indicate an unpractical answer of
infinity.

Grid
resolution
(m)

LDP
DPLBAR
Fitted
LDP
Calculated
Fitted
DPLBAR
Calculated function sinuosity DPLBAR
function sinuosity
LDP (m)
(m)
(unitless)
(m)
(m)
(unitless)

(below
(below
1
resolution)
36979
1
resolution)
16456.49
5
34539.5
34727.4
1.064837
16363.2
16211.12
10
33872.4
33757.68
1.095425
16110.9
16027.26
25
32900.9
32475.79
1.138664
15823.1
15662.45
50
31755.5
31506.08
1.17371
15371.3
15251.32
100
30112
30536.37
1.210982
14532.6
14669.9
150
29643.3
29969.12
1.233903
14019.3
14223.76
200
29350.4
29566.65
1.2507
13585.2
13847.64
250
29281.2
29254.48
1.264046
13214.8
13516.28
350
28814.1
28783.75
1.284718
12789.2
12941.22
400
27770.6
28596.94
1.29311
12310.9
12684.8
500
29420.3
28284.76
1.307382
13127
12216.18
Table 2: The effect of DTM resolution on the estimated drainage path lengths

1
1.015136
1.026781
1.050697
1.079021
1.121786
1.156972
1.188396
1.217531
1.271634
1.297339
1.347106

The effect of grid resolution on sinuosity is presented graphically in Figure 9. There


is notable variation between longest drainage path (LDP) and average drainage path
(DPLBAR). It is hypothesised that the LDP is more representative of the channel
sinuosity since the longest drainage path will be predominated by in channel
distance. The DPLBAR indicates a lower sinuosity factor which would be more
representative of a combined land and channel flow, however it is suspected that the
smoothing of the NextMap results in this factor being understated. It is therefore
recommended to adopt the LDP sinuosity factor at this time (this was not realised
initially and a mean of the two values was first adopted hence the selection of 1.1 for
a resolution of 25m).

Initial overview of the effect of analysis grid size on sinuosity


1.6

Indicative sinuosity (unitless)

1.4
1.2
1
LDP

0.8

DPLBAR

0.6
0.4
0.2
0
1

10

100

1000

Analysis grid resolution

Figure 9: The effect of grid resolution on the sinuosity factor

The catchments presented in Table 1 were tested within the time to outlet model on a
25m grid derived from NextMap using a representative 6hr RMed event which was
locally weighted by SAAR. A broad range of permutations of universal land and
channel roughnesses were run through the model with the objective of identifying a
combination of universal roughness values which maximised the Rsq value while
maintaining a modelled lag time divided by observed lag of unity. All model runs were
run with a universal sinuosity factor of 1.1. It should be noted that the mean of the
time to outlet rather than median was used to estimate the lag time for each
catchment, this was on the basis of computational efficiency within ArcGIS and it is
recognised that future analysis should be based on the median to discount outliers. A
100% runoff was assumed universally, this has since been superseded by a
recommendation of basing runoff on HOST. The threshold of channel formation was
set at 0.1cumecs based on a visual comparison of the RMed flows and the OS 25K
river network, this assumption will need to be revisited if the percentage runoff is
changed.
The outcome of this stage of the flow routing demonstration process was that for a
25m grid using a sinuosity of 1.1 the optimum universal channel and land
roughnesses are 0.045 and 0.1. The average channel roughness (0.045) correlates
well with the channel roughnesses contained within the literature and the land
roughness (0.1) is comparable to the shallow surface flow roughness values
presented by USDA. The results of this analysis are presented in Figure 10 and
Figure 11.

Average modelled lag divided by observed for a range of land and channel
roughnesses using a 25m grid and a sinuosity factor of 1.1
0.2

1.4-1.5

0.1

Land roughness (n)

1.3-1.4
1.2-1.3
1.1-1.2
1-1.1
0.9-1
0.8-0.9
0.7-0.8
0.6-0.7
0.5-0.6

0.02

0.03

0.05
0.05

0.04

Channel roughness (n)

Figure 10: Contour plot of the mean modelled catchment lag divided by observed for
the nine test catchments for a range of channel and land roughnesses

Best fit for universal land and channel roughness


(25m grid, n,land = 0.1, n,channel=0.045, sinuosity = 1.1)
12

Modelled lag (hrs)

10

8
y = 0.9787x
R2 = 0.7093
6

0
0

10

12

Observed lad (hrs)

Figure 11: Plot comparing the modelled and observed catchment lag with the optimal
universal land and channel roughnesses using a 25m grid and a sinuosity of 1.1

The next stage of the analysis was to test the impact of modifying individual land
classes on the models ability to predict lag times. In this analysis the channel

roughness and sinuosity were kept constant along with the land roughness for all
land use classes except the class which was being investigated in that run. After
investigating the effect of changing the roughness of a land class it was returned to
the default value (0.1) prior to testing the influence of another class. The results of
the analysis are presented in Figure 12 and Figure 13.
Methodical analysis of the effect of changing the roughness for
each LCM2000 class(n,channel = 0.045, S=1.1, n,land = 0.1 with
only 1 class being varied)
11
0.78

Rsq (modelled v observed)

0.77
0.76
0.75
0.74
0.73
0.72
0.71
0.7
0.69
0.68
0

0.05

0.1

0.15

0.2

Land roughness (n)

0.25

21
41
42
43
51
52
61
71
81
101
102
121
131
161
171 & 172
201
211

Figure 12: Plot showing the effect of varying individual land classes on the models
ability to predict the lag for the selected nine test catchments
Best fit catchment lag
12

Modelled lag time (hrs) __

10

y = 1.0309x
2

R = 0.884
8
S1.1,nL0.1,nC0.045
S1.1,nC0.045,Lv1

S1.15,nC0.045,Lv1
Linear (S1.15,nC0.045,Lv1)

0
0

10

12

Observed lag time (hrs)

Figure 13: Plot showing the modelled against observed catchment lag for the nine test
catchments using the best performing calibration coefficients

The provision roughness values for each land class are presented in Table 3.
LCM class

Location of
maxima

Adopted land
roughness
(provisional)

11
-0.627
0.1
21
0.323
0.3
41
0.147
0.15
42
0.157
0.15
43
0.256
0.1
51
-0.427
0.05
61
0.218
0.15
71
(minima, diverging)
0.05
81
0.221
0.15
101
(minima, diverging)
0.1
102
(minima, diverging)
0.05
121
(minima, diverging)
0.05
131
(minima, diverging)
0.05
161
(minima, diverging)
0.05
171 & 172
(minima, diverging)
0.025
201
(minima, diverging)
0.1
211
(minima, diverging)
0.1
Table 3: The provisional land roughness for each land class based on the analysis of
nine Scottish catchments

It should be noted that not all LCM classes were available within the nine test
catchments. In some cases some LCM classes only appear in one test catchment,
hence the optimisation process is biased. Indeed the small number of catchments
compared to the large number of variables does suggest that there is considerable
risk that the calibration is just bending the model to match the tested cases and that
the selected coefficients would perform poorly for other catchments.
Recommendations:

Additional work on sinuosity should be undertaken, it may be the case


that there is existing literature on the subject. Even if existing literature is
in place it is recommended that additional study is undertaken using
additional catchments and an improved DTM (LiDAR)

The analysis should be repeated using locally derived RMed depths for
storm durations which are representative of the catchment. At the same
time the percentage runoff should be based on HOST data not a 100%
runoff.

The modelled average catchment lag should be based on median rather


than mean due to the influence of outliers.

The impact of varying land roughness with HOST class should also be
investigated to determine whether HOST is a better predictor of land
roughness. It may be the case that both HOST and LCM class should be
used to define land roughness

Reed, Duncan, Faulkner, Duncan, Robson, Alice, Houghton-Carr, Helen, Bayliss, Adrian,
Flood estimation handbook: Vol. 4: Restatement and application of the flood studies report
rainfall-runoff method, Institute of Hydrology, 1999.

You might also like