ArticlePDF Available

An integrated 1D–2D hydraulic modelling approach to assess the sensitivity of a coastal region to compound flooding hazard under climate change

Authors:

Abstract and Figures

Coastal regions are dynamic areas that often lie at the junction of different natural hazards. Extreme events such as storm surges and high precipitation are significant sources of concern for flood management. As climatic changes and sea-level rise put further pressure on these vulnerable systems, there is a need for a better understanding of the implications of compounding hazards. Recent computational advances in hydraulic modelling offer new opportunities to support decision-making and adaptation. Our research makes use of recently released features in the HEC-RAS version 5.0 software to develop an integrated 1D–2D hydrodynamic model. Using extreme value analysis with the Peaks-Over-Threshold method to define extreme scenarios, the model was applied to the eastern coast of the UK. The sensitivity of the protected wetland known as the Broads to a combination of fluvial, tidal and coastal sources of flooding was assessed, accounting for different rates of twenty-first century sea-level rise up to the year 2100. The 1D–2D approach led to a more detailed representation of inundation in coastal urban areas, while allowing for interactions with more fluvially dominated inland areas to be captured. While flooding was primarily driven by increased sea levels, combined events exacerbated flooded area by 5–40% and average depth by 10–32%, affecting different locations depending on the scenario. The results emphasise the importance of catchment-scale strategies that account for potentially interacting sources of flooding.
This content is subject to copyright. Terms and conditions apply.
ORIGINAL PAPER
An integrated 1D–2D hydraulic modelling approach to assess
the sensitivity of a coastal region to compound flooding
hazard under climate change
Ulysse Pasquier
1
Yi He
1
Simon Hooton
2
Marisa Goulden
3
Kevin M. Hiscock
4
Received: 26 December 2017 / Accepted: 21 August 2018 / Published online: 30 August 2018
ÓThe Author(s) 2018
Abstract
Coastal regions are dynamic areas that often lie at the junction of different natural hazards.
Extreme events such as storm surges and high precipitation are significant sources of
concern for flood management. As climatic changes and sea-level rise put further pressure
on these vulnerable systems, there is a need for a better understanding of the implications
of compounding hazards. Recent computational advances in hydraulic modelling offer new
opportunities to support decision-making and adaptation. Our research makes use of
recently released features in the HEC-RAS version 5.0 software to develop an integrated
1D–2D hydrodynamic model. Using extreme value analysis with the Peaks-Over-
Threshold method to define extreme scenarios, the model was applied to the eastern coast
of the UK. The sensitivity of the protected wetland known as the Broads to a combination
of fluvial, tidal and coastal sources of flooding was assessed, accounting for different rates
of twenty-first century sea-level rise up to the year 2100. The 1D–2D approach led to a
more detailed representation of inundation in coastal urban areas, while allowing for
interactions with more fluvially dominated inland areas to be captured. While flooding was
primarily driven by increased sea levels, combined events exacerbated flooded area by
5–40% and average depth by 10–32%, affecting different locations depending on the
scenario. The results emphasise the importance of catchment-scale strategies that account
for potentially interacting sources of flooding.
Keywords Flooding Hydraulic modelling Storm surge Sea-level rise Compound
hazard Extreme value analysis
1 Introduction
1.1 Flooding hazard in a changing climate
Floods are significant and regular threats to a great number of people worldwide. In
Europe, flooding represents the most costly natural hazard (Whitfield 2012) with damages
on the rise as population grows in flood-prone areas (Barredo 2009) and human activities
Extended author information available on the last page of the article
123
Natural Hazards (2019) 98:915–937
https://doi.org/10.1007/s11069-018-3462-1(0123456789().,-volV)(0123456789().,-volV)
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
lead to land-cover changes (He et al. 2013). Recent severe disruptions in the UK during the
2013/2014 and 2015/2016 winters were reminders of the devastating potential of such
extreme floods. While there is still much uncertainty in attributing a climate signal to a
possible trend in extreme events (Wilby et al. 2008), climate models suggest that climate
change could lead to more frequent and intense precipitation in certain regions (Wang et al.
2017), thereby increasing flood hazard. On the other hand—as the Intergovernmental Panel
on Climate Change (IPCC) reported (Church et al. 2013)—there is a high level of confi-
dence that sea levels will continue to rise throughout and beyond the next century.
Moreover, changes in mean sea level (MSL) are fundamental drivers for extreme sea levels
(Mene
´ndez and Woodworth 2010), thereby putting further pressure on coastal regions.
While the development of flood defences and forecasting has prevented a significant
increase in coastal flooding (Stevens et al. 2016), these trends highlight the need for better
preparedness and an improved understanding of future hazards.
Coastal environments are vulnerable systems that can act as the interface for different
hazards. Groundwater, pluvial (surface water), fluvial (river), tidal and coastal sources of
flooding can all exist in areas near the sea, which also often host dense population centres.
As presented by Wong et al. (2014) there is ample research on the risks coastal regions face
and therefore the importance of adaptive measures. More recently, increasing attention has
been dedicated to compounding extreme events (e.g. Kew et al. 2013; van den Hurk et al.
2015). Coinciding hazards, such as storm surges and precipitation, can lead to impacts that
would otherwise not have been observed had they occurred separately and can therefore
have significant implications for flooding risk. A number of studies have looked to
determine the dependence between these hydrological extremes (e.g. Zheng et al. 2014),
including in the UK (Svensson and Jones 2002). While a significant dependence is not
always found (Klerk et al. 2015), it remains highly uncertain how the climate will influence
this relation in the future. Wahl et al. (2015) for example, observed in the USA a change
towards storms surges that also promote high rainfall. The threat of combined events from
different origins underlines the importance of adopting a holistic stance in assessing flood
hazard.
1.2 Integrated flood modelling
There has been in recent decades a paradigm shift towards a broader catchment-scale
approach for flood risk management in Europe, as demonstrated by the European Union’s
Water Framework Directive (2000) and Floods Directive (2007). Integrated strategies that
identify synergies at the river basin level, notably between rural and urban areas, have
gained increasing support (Rouillard et al. 2015). Isolated actions to mitigate flooding run
the risk of leading to unwanted outcomes. For example, a flood alleviation measure taken
at a location in a catchment can have downstream impacts that should be taken into
account. An integrated approach is moreover justified when sources of flooding are varied,
originate from different hydrological processes and interact with each other. The lack of
adequate information on these interactions remains an important hurdle for decision-
making.
There is a need for modelling methods to follow the above trends to be able to provide
information required for planning. Hydrodynamic models solve equations of fluid motion
to replicate the movement of water and are widely used to assess flooding risk. The
simplest and most common practice is to use one-dimensional (1D) models that treat flow
one-dimensionally along the river channel. This assumption is appropriate in many
123
916 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
situations but may not be suitable for flood mapping in areas where flow is expected to
spread, such as in wide floodplains (Ne
´elz and Pender 2009). Alternatively, while two-
dimensional (2D) models can provide more detailed results and have gained in popularity,
they remain computationally and data intensive and therefore difficult to apply to large
areas. Recent advances and software developments offer new opportunities to help meet
the goals of integrated approaches by allowing for linkages between 1D and 2D models
(Teng et al. 2017). Coupled 1D–2D models can dynamically represent coastal, urban, river
and floodplains interactions and are therefore well suited to assess the impact of flooding
from different sources. While—as was shown in the previous section—there has been an
increasing number of studies looking at the impact of combined events on flooding, 1D–2D
hydraulic models remain relatively new tools in this field that are subject to more inves-
tigation (Webster et al. 2014).
This paper aims to present a modelling methodology to assess the sensitivity of a coastal
area to the combination of fluvial, tidal and coastal sources of flooding. The fitness for use
of an integrated 1D–2D hydraulic modelling approach is to be evaluated in the context of
the Broads National Park in the UK. The aim of this study is to provide a modelling
framework for simulating compound modelling scenarios. In this study, we are not pro-
viding a comprehensive probabilistic flood risk assessment framework. Finally, an aim of
the modelling design is to understand the implications of portraying interacting sources of
flooding from opposite ends of a river sub-catchment.
2 Study area: the Broads, UK
Located on the eastern coast of England, the Norfolk and Suffolk Broads is Britain’s
largest designated wetland. The network of rivers and shallow lakes—or ‘‘broads’’—covers
a total area of 303 km
2
at the downstream end of the 3200 km
2
Broadland Rivers
Catchment (Fig. 1). The low-lying national park holds importance for natural conservation,
navigation, recreation and tourism, as well as for its cultural features. Land use is mostly
shared between coastal and floodplain grazing marshes, fens and arable land. The Broads
are bounded by several urban centres, namely, Norwich, Lowestoft and Great Yarmouth,
where the River Yare flows into the North Sea.
The Broads Authority was established in 1988 to coordinate the management of land
and water in the Broads because of its special landscape. While offering many economic
and environmental opportunities, water also presents considerable risks. The Broads have a
long history of flooding driven by its low elevation and proximity to the sea. The 1953
storm had severe impacts in East Anglia, as it did throughout much of the North Sea coasts.
The event led to significant investments in flood protection and forecasting. Most recently,
the Broads Flood Alleviation Project has been responsible for the improvement and
maintenance of the 240 km of flood defences that exist in the Broads. The scheme has been
successful in limiting inundation, and defences coped well during the largest storm surge
since 1953 in December 2013. As climatic conditions change and sea level rises, the
Broads are however anticipated to face further pressures and there remains uncertainty
over the best strategic line to follow to manage flood risk.
Flood management in the Broads is a challenging task due to the area’s complex
hydrology and range of potential flooding sources. In the context of the Broads, coastal
flooding—or the ingress of water inland directly from the sea—is differentiated from tidal
flooding, caused by the propagation of the tidal wave upriver. Although coastal flooding
can have devastating consequences (Wu et al. 2015), tidal flooding is still the main concern
123
Natural Hazards (2019) 98:915–937 917
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
in many parts of the Broads as low gradients along the key rivers allow the tidal influence
to travel throughout much of the area. Major floods have also occurred due to heavy
rainfall, for example in 1959 and 1968. Past studies in the catchment have found that
fluvial floods and surge events occurred independently (Mantz and Wakeling 1979). There
remains however a risk of combined river and tidal flooding in the Broads. Extreme sea
levels can indeed coincide with high river flows or prevent proper drainage to cause
flooding, for example on the River Bure (Environment Agency 2009). While they can
exacerbate the impact of inundation, little research has focused on combined events and
how they could affect the Broads in the future with projections of climate change and sea-
level rise (SLR).
3 Data and methods
3.1 Environmental conditions
3.1.1 Sea level
Tide gauge data of sea level between 1964 and 2015 were obtained from the British
Oceanographic Data Centre. The observations were made in Lowestoft (52°28023.055600N,
1°4500.8100E), approximately 10 km south of Great Yarmouth. The east coast of England
experiences a semidiurnal tidal regime. Chart datum at Lowestoft is located 1.50 m below
ordnance datum (OD, at Newlyn). Sea level was recorded every 60 min prior to 1992 and
every 15 min after 1992, with fewer than 3% missing data in the whole dataset.
A critical driver for flood hazard in coastal areas is peak sea level during extreme events
that may occur, for instance, when a large storm surge coincides with high spring tide. The
Fig. 1 The Broads National Park is part of the Broadland River Catchment in eastern England. The majority
of the area within the Broads’ administrative boundaries lies below sea level
123
918 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
historical sea level data at Lowestoft were analysed with extreme value statistics to
determine the probability of occurrence of extreme sea levels. Block maxima and Peaks
Over Threshold (POT) are the primary approaches for extreme value analysis (EVA), and
both have been used in the past to analyse sea levels (Webster et al. 2014; Haigh et al.
2016). POT however allows for more control over which events are included in the
extreme value distribution and has been found to perform better than the more traditional
Block Maxima method in previous flood frequency studies (Arns et al. 2013, Bezak et al.
2014). An average of 1.92 extreme values per year were thereby extracted that exceeded a
level of 1.90 m above ordnance datum (maOD), corresponding to the 99.7th percentile of
high tide peak sea levels (Fig. 2).
Due to the thermal expansion of water, melting glaciers and vertical land movement,
relative sea level has been rising at Lowestoft at a rate of 2.70 ±0.40 mm a
-1
in the
second half of the twentieth century (Wahl et al. 2013). A simple additive method was used
to detrend the data and remove yearly changes in MSL with 2015 serving as the reference
year. Moreover, the chosen peaks were declustered using a 48-h window to ensure only
independent events were retained. A Generalised Pareto (GP) distribution was fitted to the
remaining sea levels to determine return periods relative to the year 2015. The GP dis-
tribution has the distribution function
FxðÞ¼11kx
a

1=k
ð1Þ
where the distribution’s parameters a, the scale parameter, and k, the shape parameter, are
determined with the maximum likelihood estimation method. The fit of the distribution
was evaluated with plotting positions using the Gringorten formula, which is widely
recognised for GP distributions (Chen and Sign 2017).
3.1.2 River discharge
Daily mean river flow data at Horstead Mill (52°43025.867200N, 1°21014.874500E) on the
River Bure between 1974 and 2015 were obtained from the National River Flow Archive.
In the same way that sea levels were analysed, the POT method was used to determine the
probability of extreme discharge. The GP distribution provided a better fit than a gener-
alised extreme value distribution, which was tested using annual maxima of river flow. The
mean residual life plot, an exploratory technique described by Saeed Far and Abd. Wahab
(2016), here helped identify an appropriate threshold. An average of 2.20 extreme values
−2
0
2
1970 1980 1990 2000 2010
Year
Sea level (maOD)
Fig. 2 Sea level relative to ordnance datum at Lowestoft, UK between 1964 and 2015. Red points represent
sea level peaks above a defined threshold (blue, dashed horizontal line) chosen to fit a Generalised Pareto
distribution and derive extreme return levels
123
Natural Hazards (2019) 98:915–937 919
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
per year were extracted that exceeded a level of 6.83 m
3
s
-1
, corresponding to the 99th
percentile of river discharge levels (Fig. 3). An extreme value of 30.80 m
3
s
-1
in 1981
particularly stood out from other peaks corresponding to an event that saw approximately
70 mm of rainfall in Norfolk between 25 April 1981 and 27 April 1981.
3.2 Hydrodynamic model: HEC-RAS
3.2.1 Model structure and domain
A 1D–2D hydraulic model was developed with the HEC-RAS software to map flooding
extent and depth under different extreme scenarios. HEC-RAS is a free modelling tool
developed by the United States Army Corps of Engineers (USACE). Among its many
applications, the software is well tested for flood mapping in both coastal (e.g. Ray et al.
2011) and fluvial (e.g. Javaheri and Babbar-Sebens 2014) environments as well as to assess
the impacts of climate change (e.g. Shrestha and Lohpainsankrit 2016). Previously limited
to 1D models, a new version of HEC-RAS (version 5.0) was released in 2016 allowing for
full 2D modelling and linkages between 1D and 2D features. While other tools such as
Flood Modeller, developed by CH2 M, or MIKE FLOOD, developed by the Danish
Hydraulic Institute (DHI), also offer the possibility to combine 1D and 2D models, HEC-
RAS is the non-commercial software that has not previously been applied to the Broads.
Moreover, although the new 2D capabilities of HEC-RAS offer opportunities for flood
mapping, the model still requires testing for different applications (Vozinaki et al. 2017).
The new HEC-RAS version was used, for example, by Quiroga et al. (2016) and Patel et al.
(2017) to simulate past fluvial floods. Due to its recent release however, few studies are yet
to apply HEC-RAS version 5 in coastal regions.
The Broads is a hydrologically complex and highly engineered area. The main rivers
that make up the wetland—namely, the River Bure, River Yare and River Waveney—are
narrow and constrained by high levees. These defences protect over 21,000 ha in the
Broads and over 1700 properties. In many parts of the Broads, the flood banks are sig-
nificantly higher than the wide floodplains they protect. Much of the Broads floodplain has
a low elevation gradient and lies below sea level. A failure in the defences can therefore
lead to widespread flooding. An accurate representation of the study area’s elevation is a
fundamental requirement in hydraulic modelling. A composited digital terrain model
(DTM) derived from light detection and ranging (LIDAR) data was obtained from the
Environment Agency. The DTM had a resolution of 2 m by 2 m with a vertical accuracy
0
10
20
30
1980 1990 2000 2010
Year
Discharge ( m3s1)
Fig. 3 River discharge at Horstead Mill between 1974 and 2015. The points represent discharge peaks
above a defined threshold (blue, dashed horizontal line) chosen to fit a Generalised Pareto distribution and
derive extreme return levels
123
920 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
of ±5 cm and provided a good coverage of the study area. River bathymetry is also an
important input to the hydraulic model. As LIDAR data are poor at representing under-
water elevations, river surveys from the Broads Authority conducted between 2011 and
2015 were used to correct the DTM within river channels. Moreover, information from the
Environment Agency on flood defences in the area ensured that the latest levee heights
were included in the DTM.
The 1D–2D hydraulic model shown in Fig. 4was built in HEC-GeoRAS, the ArcGIS
extension for HEC-RAS. Cross sections of the river channels were drawn approximately
every 30–50 m from one river bank to the other, forming the model’s main 1D feature. A
common method for out-of-bank flood modelling and mapping is to extend the model’s
cross sections into the floodplain. This technique is however not suitable for flood mapping
in wide floodplains, which are common throughout the Broads. Instead, the floodplain is
represented as a series of flood cells, called storage areas in HEC-RAS, where water can
spill into from the rivers. The storage areas are separated by high ground and connected to
the river cross sections in the HEC-RAS model with lateral structures, in this case, the
Fig. 4 HEC-RAS model domain. Storage areas and 2D areas are used to represent overbank flow in
upstream and downstream portions of the model domain, respectively. Observations of river levels and
discharge are available at different gauges: F1 (Horstead Mill), T1 (Great Yarmouth), T2 (Burgh Castle), T3
(Haven Bridge), T4 (Three Mile House), T5 (Acle Bridge) and T6 (Hoveton Broad)
123
Natural Hazards (2019) 98:915–937 921
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
flood defences on both sides of the rivers. Water will flow into the storage areas if the river
level surpasses the corresponding height of the flood defence. Storage areas are 1D features
represented using a volume-elevation table calculated with the DTM data and can provide
satisfactory accounts of floodplain flow with little computational demands. More detail is
however required in urban areas and where flow is likely to spread significantly as is the
case at the downstream end of the study area. 2D flexible meshes were therefore set up and
dynamically linked to the river cross sections in Great Yarmouth and the large low-lying
area called the Halvergate Marshes. The mesh size varied between 10 m and 50 m and
aligned to capture high ground features such as flood defences, roads, and railway tracks. A
2D domain is appropriate at the coast as it has the added benefit of being capable of
portraying flooding occurring directly from the sea—in case of the overtopping of defences
(coastal flooding)—and how it may interact with other sources of flooding.
The hydraulic model covers a 260 km area from the mouth of the River Yare in Great
Yarmouth to Horstead Mill, approximately 40 km upstream on the River Bure. Portions of
the River Bure’s tributaries—namely the River Ant and the River Thurne—are also
included. The location of a flow gauge at Horstead Mill was chosen for the upstream
boundary of the model. As a predominantly tidally influenced area, gauges in the Broads
primarily measure river levels, and their locations are presented in Fig. 4. Land-cover data
were obtained from the EDINA Environment Digimap Service as supplied by the Centre
for Ecology and Hydrology (CEH) for the year 2015 (Fig. 5). The original classification
was simplified to represent the main land uses across the HEC-RAS 2D areas. The large
floodplains of the Broads consist first and foremost of grassland and grazing marshes. Land
used for arable crops and horticulture tends to be located on the higher ground and make up
most of the rest of the area. The most significant urban area is Great Yarmouth on both
sides of the River Yare.
Fig. 5 Land-cover map of the downstream end of the Broads near Great Yarmouth in 2015 (Data obtained
from EDINA Environment Digimap Services)
123
922 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
3.2.2 Unsteady flow analysis
Flood events were simulated in HEC-RAS under unsteady flow conditions. The HEC-RAS
model solves the full Saint-Venant equations for the conservation of mass and momentum:
of
otþou
oxþov
oy¼0ð2Þ
ou
otþo
ox
u2
h

þo
oy
uv
h

¼n2ug ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u2þv2
p
h2gh of
oxþuf þo
qoxhsxx
ðÞþ
o
qoyhsxy

ð3Þ
ov
otþo
ox
v2
h

þo
oy
uv
h

¼n2vg ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u2þv2
p
h2gh of
oyþvf þo
qoyhsyy

þo
qoxhsxy

ð4Þ
where his the water depth (m), uand vare the specific flow in the xand ydirections
(m
2
s
-1
), fis the surface elevation (m), gis the gravitational acceleration (m s
-2
), nis the
Manning’s resistance, qis the water density (kg m
-3
), fis the Coriolis parameter and sxx,
sxy and syy are the components of the effective shear stress (Quiroga et al. 2016). While
HEC-RAS offers the option of solving the diffusion-wave approximation of the equations
in two dimensions, this method cannot be used for the propagation of waves in tidally
influenced conditions. The full momentum equations were therefore chosen. A computa-
tional time step of 10 s was selected based on the guidelines proposed by the Courant–
Friedrichs–Lewy condition:
C¼VDT
Dx1OrDTDx
Vwith C¼1:0ðÞ ð5Þ
where Cis the Courant Number, Vis the flood wave velocity (m s
-1
), DTis the com-
putational time step (s) and Dxis the average cell size (m). The performance of the model
was tested with the Nash–Sutcliffe Efficiency (NSE) coefficient defined as:
1Pn
t¼1Qt
mQt
o

2
Pn
t¼1Qt
oQo

2ð6Þ
where Qt
oare observations at time tand Qt
mare modelled values.
The HEC-RAS model boundary conditions consisted of a stage hydrograph downstream
and a flow hydrograph upstream. The observed sea level can be considered as the sum of
MSL, an astronomical tide component and a non-tidal residual (Pugh 1996). The tidal
component is the response of sea level to astronomical forces such as the relative position
of the moon and the sun, and can be isolated with a harmonic analysis of sea levels. What
remains when the MSL is also removed is termed the non-tidal residual and primarily
represents the meteorological impact on sea level from a surge.
An average storm surge shape was determined by identifying the 20 highest storm
surges since 1964 at Lowestoft (Fig. 6a). Ideally, local storm surge models can be used to
reconstruct more physically realistic conditions in the definition of synthetic events (e.g.
Villatoro et al. 2014). The chosen method of generalisation was however described by the
Environment Agency (McMillan et al. 2011) as providing a reasonable means to derive a
design surge profile. Although the averaging leads to a smoothed profile, the resulting
123
Natural Hazards (2019) 98:915–937 923
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
storm surge shape is similar to the rest of the sample (Fig. 6a) and can be considered
representative of historical events. Moreover, by choosing the non-tidal residuals and not
total sea level peaks to determine an average storm surge shape, large storm surges that
may have occurred during low tide are also taken into account. An extreme sea level event
stage hydrograph for a target maximum level can thereby be recreated using this average
surge shape, a base tidal prediction and MSL (Fig. 6b).
The skew surge is the difference between the predicted astronomical high tide and the
nearest experienced high water. Since meteorological processes are independent of tidal
forces, a surge can occur at any stage of the tide. Other studies have performed a joint
probability analysis to form a probability distribution of total sea levels from the distri-
bution of skew surges and peak tide levels (McMillan et al. 2011). The assumption was
made here that the storm surge peak coincided with the mean high predicted tide. This
method, also used by Webster et al. (2014), was justified by analysing past extreme storm
surge events that led to flooding concerns in the study area, which tended to occur at or
near high tide.
An analogous method was applied to create synthetic flow hydrographs. The hydro-
graph shape of the last 20 most important storms in terms of flow at Horstead Mill on the
River Bure was analysed to produce an average event shape. Due to limited data avail-
ability, upstream boundaries at the River Yare and internal boundaries at the tributaries of
the River Bure were assumed to be proportional to the discharge rate at Horstead Mill
based on their relative drainage areas. This is a common method used for ungauged
catchments (Webster et al. 2014) that assumes similar hydrogeological characteristics.
Drainage areas were determined in ArcGIS using 30 m by 30 m resolution Shuttle Radar
Topography Mission (SRTM) data (Table 1). Initial conditions for both stage and dis-
charge are taken directly from the boundary data.
4 Results and discussion
4.1 EVA and scenario definition
Exploratory semi-structured interviews were conducted with a set of 11 stakeholders to
identify priorities, interests and to help base the definition of scenarios on local knowledge.
Stakeholders were chosen from professionals with extended knowledge of the Broads, and
a
−0.5
0.0
0.5
1.0
−20 −10
Time from storm surge peak (h)
Normalised non−tidal residual
b
−1
0
1
2
3
−20 −10
01020 01020
Time from storm surge peak (h)
Sea level (maOD)
Fig. 6 aAverage surge shape (red, dotted) estimated from the 20 largest surges at Lowestoft between 1964
and 2015. bSynthetic total sea level (black) derived from the surge residual (red, dotted) and the
combination of a base astronomical tide) and the 2015 mean sea level (blue, dashed)
123
924 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
active residents with a long-lasting interest in the area’s overall management. Specific
experience in flood management varied greatly as participants covered a wide range of
sectors such as farming, angling, environmental protection, engineering and coastal
management. The interviews confirmed the importance of tidal and coastal sources of
flooding in the Broads and highlighted vulnerable locations such as—but not limited to—
Great Yarmouth or several protected areas. One of the main recurring statements
emphasised in the interviews was a concern for the risk of combined events. More
specifically, the occurrence of a storm surge during high river discharge was identified as a
worry for different stakeholders. Although the small sample of participants does not allow
for statistically significant conclusions, this information was used to guide modelling
choices and define future scenarios.
A comparison of the available data on past peak sea levels, non-tidal residuals and
discharge shows that these events do not tend to occur simultaneously (Fig. 7). However,
Fig. 7also shows that it is physically possible for the peak of the storm surge to occur
during a high discharge event and therefore near peak flow.
The EVA served to find return levels of both extreme sea level and extreme discharge to
define representative downstream and upstream boundary conditions, respectively. The
purpose of the EVA was not to provide a robust probabilistic assessment of flooding risk
from different or combining sources. Without an analysis of the probability of joint
occurrence of high tide and extreme storm surge, it was not possible to assign return levels
to entire extreme sea level events. The EVA performed on total sea levels however did
provide return levels for the peak of recreated extreme events.
The GP distribution performed relatively well to describe both extreme sea level
(Fig. 8a) and extreme discharge (Fig. 8b). It should be noted that the most extreme values
were found above the fitted distribution curves. These events corresponded to the
December 2013 storm surge and a peak river flow in April 1981. Both occurrences were
0
10
20
30
1980 1990 2000 2010
Year
Discharge ( m3s1)
Fig. 7 The timing of the 40 highest non-tidal residuals (red points) decomposed from sea level data at
Lowestoft, UK compared to river discharge at Horstead Mill between 1974 and 2015
Table 1 Drainage area of
upstream and internal boundaries
for the HEC-RAS model used to
estimate flow hydrographs rela-
tive to the River Bure
River Drainage area at model boundary (km
2
)
Bure 336.54
Ant 145.24
Thurne 119.35
Spix 59.94
Yare 1392.57
Waveney 891.43
123
Natural Hazards (2019) 98:915–937 925
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
verified using data from other nearby gauges, and it was therefore decided not to discard
them as recording errors. These points were by far the most extreme observations and did
not provide strong evidence against the choice of the GP distribution function compared to
other tested distribution functions. The lack of data is a common issue in EVA. More
investigation using other sources of data (such as news reports if they exist) that extend
past the recorded data period would allow for more confidence in this estimation.
Evidence suggests that changes in MSL are the primary factor leading to an increase in
extremes sea levels (Mene
´ndez and Woodworth 2010). Relative MSL (RMSL) is not only
rising, but has also been found to accelerate at various rates around the world, with a trend
of 4.4 ±1.1 mm a
-1
estimated at Lowestoft from 1993 to 2011 by Wahl et al. (2013). It
indeed remains highly uncertain how climate change will impact local storm surge pat-
terns. A linear increase in RMSL was assumed to determine future conditions and return
levels up to the year 2100. Uncertainty moreover resides in current projections of the rate
of SLR in the twenty-first century. Pfeffer et al. (2008) found that accelerated sea-level rise
between 0.8 m and 2 m up to 2100 was physically plausible depending on glaciological
conditions. To account for such possibilities, extreme scenarios of 1 m and 2 m MSL rise
by 2100 were also considered.
While seasonal precipitation changes are expected in the UK, notably with an increased
proportion of heavy precipitation events occurring during winter months, current projec-
tions do not show significant changes in annual precipitation in East Anglia (Jenkins 2009).
Moreover, little is known on the intensity of extreme precipitation events in coming
decades and therefore which trajectory river discharge will also follow. Patterns of extreme
river discharge were therefore assumed to the same up to 2100 as in 2015 in the presented
scenarios. This assumption is moreover warranted by the much greater influence of tidal
processes in the Broads.
The chosen scenarios are presented in Table 2. They included three scenarios of
100-year return peak sea levels under different MSL rise pathways. As explained in
Sect. 3.2.2, only the peak sea level is assigned a 100-year return period as opposed to the
entire event. Each storm surge event was then also combined with a simultaneous 100-year
return river discharge to test the sensitivity of the study area to coinciding extreme events.
The timing of events can have significant impacts on flooding occurrence and extent. It is
therefore important to note that previous studies have found it most likely for these types of
events to not coincide with up to several days separating the different extremes (Klerk et al.
2015). With these caveats taken into account, the proposed scenarios provide a basis to
assess the sensitivity of the Broads to compound flooding.
a
2.0
2.5
3.0
3.5
4.0
4.5
10 20 50 100 200 500
Return Period (Years)
Sea Level (maOD)
b
0
25
50
75
25 25
10 20 50 100 200 500
Return Period (Years)
River Discharge ( m3s1)
Fig. 8 Return levels at the reference year 2015 for asea level at Lowestoft, UK expressed in relation to
ordnance datum and briver discharge at Horstead Mill. The dashed lines represent the 95% confidence
intervals
123
926 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
4.2 Calibration and validation
The HEC-RAS model was calibrated and validated with storm surge events from October
2014 and December 2013, respectively. The calibration parameter used was the Manning’s
nroughness coefficient. Data on past flooding inundation extent in the Broads are lacking
in both availability and accuracy. While there have not been major flooding events since
1953, localised defence failures have been observed during extreme storm surge events.
Spencer et al. (2015) provided an account of the impact of the December 2013 storm surge
along the Norfolk coast. Tidal flooding was however also observed further inland due to
overtopping and reported in parts of the Broads (Broads Authority, 2014). As there is no
record of the spatial footprint of this inundation, the validation process was carried out
using river levels at different stations on the Bure and the Yare (Fig. 6), as well as reports
from the Broads Authority, news articles, dated photos, and local accounts of flooding.
Descriptions of the local environments and recommended ranges obtained from Chow
(1959) served to make initial benchmarks for Manning’s nvalues. The model’s calibration
was performed on the Manning’s nwithin river channels to reach final values as shown in
Table 3. A roughness coefficient was also applied to land classes out of the river banks in
the 2D modelling domain. These values were not used during the model’s calibration as
Table 3 Manning’s nin river
channels after calibration Land cover Manning’s nroughness coefficient
River Bure 0.045
River Ant 0.045
River Thurne 0.045
River Yare—Great Yarmouth 0.04
River Yare—Breydon Water 0.025
River Yare—Upper 0.03
River Waveney 0.04
Table 4 Manning’s nfor differ-
ent land classes Land cover Manning’s nroughness coefficient
Arable and horticulture 0.05
Broadleaf woodland 0.15
Fen, marsh and swamp 0.07
Improved grassland 0.035
Urban areas 0.2
Table 2 Scenario names
Upstream boundary—
river flow
Downstream boundary—sea level
2100–4 mm a
-1
MSL rise
1:100 peak sea level event
1 m MSL rise
1:100 peak sea
level event
2 m MSL rise
1:100 peak sea
level event
Base 2100Q0 1mQ0 2mQ0
1:100 event 2100Q100 1mQ100 2mQ100
123
Natural Hazards (2019) 98:915–937 927
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
flood extent data were not available (Table 4). In tidally influenced rivers, the inertial terms
in the momentum equation are important and rivers levels are not highly sensitive to
adjustments in the roughness coefficient (USACE 2016). Theta is a weighting factor that
ranges between 0.6 (more accurate) and 1.0 (more computationally stable) applied to the
finite difference approximations when solving the unsteady flow equations. A Theta value
of 0.6 was used to improve the accuracy in the representation of the propagating tidal
wave, which did not decrease the model’s stability.
As expected, the model performed well at recreating river levels near the model’s
downstream boundary condition in Great Yarmouth at Haven Bridge (Fig. 9a) with an
NSE of 0.92. The model also performed well upstream on both the River Bure and the
River Yare, at the Three Mile House (Fig. 9b) and Burgh Castle (Fig. 9c) gauges,
respectively. It should be noted that the instrument at Three Mile House was unable to
measure the river level during the peak of the tide on 06/12/2013. The NSE remained
relatively high at 0.84. The gauge at Burgh Castle is a flood warning monitoring station
only and due to the position of its pressure sensor instrument, it therefore does not measure
any levels below 0 maOD. Still, the model produced a good fit to both the level of the
peaks and their timing at Burgh Castle. The model’s performance decreased upstream of
the River Bure. At Acle, once the tidal wave had propagated, the NSE dropped to 0.67 and
there was a slight shift in the timing of the tide (Fig. 9d). The modelled peak river level
remained within 0.03 m of the observed value. Nearly 40 km from the sea, the error
increased further upstream towards Hoveton Broad, where the model overestimated the
a
−1.00
0.00
1.00
2.00
River level (maOD)
d
0.25
0.50
0.75
1.00
River level (maOD)
b
0.00
0.50
1.00
1.50
2.00
River level (maOD)
e
0.20
0.40
0.60
0.80
03−12−2013 05−12−2013 07−12−2013
Date
River level (maOD)
c
−1.00
0.00
1.00
2.00
03−12−2013 05−12−2013 07−12−2013
Date
River level (maOD)
Fig. 9 Observed (black) and modelled (red dashed) river levels during the December 2013 storm surge at
aHaven Bridge, bThree Mile House, cBurgh Castle, dAcle, and eHoveton Broad
123
928 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
river level by a maximum of 0.1 m. While river levels were high during this event, the
defences were largely successful in holding back the water from the floodplains. This was
also the case in the model’s recreation of the event, where only localised flooding was
visible at moorings located near Berney Arms, which allowed water to flow into Halver-
gate Marshes.
4.3 Hydrodynamic simulations
Model results derived from simulations in HEC-RAS were exported to ArcGIS and R for
analysis. The maximum flooding depth from each simulation run can be found in Fig. 10.
The inundation extent shown in these profiles represents an aggregation of the overall runs
rather than a specific simulation time. The profiles should therefore be differentiated with
the extents occurring during maximum sea level, since flooding is dynamic and its timing
Fig. 10 Maximum flooding depth in the Broads between Great Yarmouth and Horstead Mill on the River
Bure under different extreme scenarios (simulation names from Table 3). a2100Q0, b2100Q100, c1mQ0,
d1mQ100, e2mQ0, f2mQ100
123
Natural Hazards (2019) 98:915–937 929
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
varies across various locations. Extreme sea levels cause flooding both downstream and
upstream in the Broads when assuming a linear mean SLR up to 2100 (Fig. 10a). The
largest affected area is Halvergate Marshes, where water is able to flow throughout the
large floodplain located north of Breydon Water. Elevated roads and railway tracks are
well captured by the model’s 2D mesh and slow the propagation of the flood wave.
Flooding is minimal in the more densely populated Great Yarmouth as there is almost no
overtopping of high defences. With the exception of Halvergate Marshes, flood walls and
levees are successful in preventing extensive flooding. Upstream of Ranworth Broads, the
floodplains are unprotected and consist mostly of marshes that are well connected to the
river. While buildings near the riverbanks in the towns of Horning and Hoveton are
affected, the flood depth remains relatively low. As Fig. 10b shows, combining this event
with a 1:100 return river discharge has significant consequences on flooding on the
upstream boundary of the tidal Bure. Impacts downstream remain limited. As SLR has
been observed to accelerate in the last decades, a linear increase in RMSL over the next
century is a conservative assumption. Scenarios representing an accelerated rise leading up
to 1 m and 2 m increase in MSL are shown in Fig. 10c–f.
The topology of the rivers and floodplains in the Broads causes flooding to occur rapidly
and spread significantly when a defence is overtopped. Figure 10 shows that certain areas
are susceptible to lower thresholds of embankment failure, thereby flooding first and
highlighting potential vulnerabilities. A notable observation from the scenarios with a 1 m
and 2 m RMSL rise is the increased impact on Great Yarmouth. Not only are more tidal
defences overtopped, but coastal waters are also able to flow into the town directly from
the sea and cause more flooding at some simulation time steps. These interacting sources of
flooding lead to an important increase in impacted buildings (Table 5). While a 2 m
increase in MSL by 2100 is still considered unlikely and would require a drastic accel-
eration of SLR, this scenario is useful to highlight the area’s sensitivity. For example, the
model showed flooding outside of some of the left banks of the Bure only during scenarios
2mQ0 and 2mQ100. The main urban zone in the study area is Great Yarmouth, located
near the coast. Sea level is therefore the main driver for the number of flooded buildings.
Other towns located farther upstream in the Broads are also affected. Centres of activity for
tourism and sailing in Horning and Hoveton lie in close proximity to the River Bure, and
several buildings in both towns are susceptible to flooding in all scenarios.
While flooding occurs in all the presented scenarios, both extent and depth vary greatly
between the different simulations. Depth is important to consider for risk management as it
is used in determining flood damage. Figure 11 shows the density of flooded 2-m cells by
depth in all six scenarios. Although the flooding extent was already high in scenario
2100Q0, most of the flooding occurred at low depths between 0 m and 0.5 m, meaning
actual damages would be limited or easier to cope with (Fig. 11a). The maximum density
Table 5 Number of buildings affected by flooding under different extreme scenarios in the model study area
Scenario Number of buildings flooded Proportion of buildings flooded (%)
2100Q0 702 16.78
2100Q100 892 21.32
1mQ0 1285 30.72
1mQ100 1389 33.21
2mQ0 1635 39.09
2mQ100 1797 42.96
123
930 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
shifts towards 0.5 m and 1 m for scenario 1mQ0 (Fig. 11b) and increases considerably to
over 2 m for scenario 2mQ0 (Fig. 11c).
Both Table 5and Fig. 11 emphasise that increasing RMSL has a significant impact on
inundation extent and depth in the Broads. While sea level is indeed the main driver for
flooding in the study area, the results also show that coinciding high river flows can
exacerbate these impacts. The average depth of cells below 5 m in depth increased from
0.82 m to 1.08 m (Fig. 11a), from 0.92 m to 1.16 m (Fig. 11b) and from 1.9 m to 2.09 m
(Fig. 11c) for the three scenario pairs, respectively. A similar pattern can be observed for
the total area of the flooding in each scenario. For both average depth and inundation area
however, the influence of high discharge decreases as the maximum sea level increases.
Average flood depth increases by 40% from scenarios 2100Q0 to 2100Q100, while it
increases by 5% from scenarios 2mQ0 to 2mQ100. Similarly total inundated area increases
0.0
0.5
1.0
1.5
Depth (m)
Density
0.00
0.25
0.50
0.75
1.00
1.25
Depth (m)
Density
0.0
0.2
0.4
0.6
0.8
543210
543210
543210 Depth (m)
Density
a
b
c
Fig. 11 Kernel density plots of flooded cells by depth for scenarios a2100Q0 (blue), 2100Q100 (red, dotted
line), b1mQ0 (blue), 1mQ100 (red, dotted line) and c2mQ0 (blue), 2mQ100 (red, dotted line)
123
Natural Hazards (2019) 98:915–937 931
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
by 32% from scenarios 2100Q0 to 2100Q100 compared to a 10% rise from scenarios
2mQ0 to 2mQ100.
The simulated compound events did not have significant added consequences in Great
Yarmouth on either flooding extent or depth, compared to unique events of extreme sea
level. The longitudinal profile of the modelled rivers indeed shows that the influence of the
combined extreme discharge decreases going downstream (Fig. 12). Near the mouth of the
River Yare, the extreme discharge has almost no impact on the water level in all three
envisaged cases. Figure 12 also shows that the difference in water level between Q0 and
Q100 events is greater for a lower MSL. Upstream areas are much more affected. The
flooded area of broadleaf woodland, which occurs mostly upstream of Ranworth Broads
along the River Bure, is highly influenced by the occurrence of a combined event (Fig. 12,
Table 6). The Bure Broads and Marshes are well connected to the river, and the
encroachment of water is therefore not a direct concern or a rare occurrence.
The deeper upstream flooding observed in Fig. 10b, c and d remains significant as it can
lead to longer residence times of saline waters. Large areas of improved grassland, notably
used for grazing, are predisposed to flooding under each scenario, with arable and horti-
culture land classes also highly impacted (Table 6). There are moreover several protected
areas, such as sites of specific interests (SSSI), located in the Broads. A topic for future
research would be the impact of extreme events on salinity in the Broads. Salinity can
cause damage to agricultural land and therefore lead to significant economic losses as well
as representing a threat to sensitive species. Studying the impact of combined events may
lead to counter-intuitive results as several processes affect salinity. Indeed, high river flows
Table 6 Area flooded by land-cover class (km
2
)
Scenario Broadleaf
woodland
Arable and
horticulture
Improved
grassland
Fen, marsh and
swamp
Urban Sub-
urban
2100Q0 6.31 1.01 23.27 6.92 1.45 0.16
2100Q100 8.91 2.16 33.14 8.65 1.60 0.31
1mQ0 7.93 3.34 35.86 8.08 4.59 0.56
1mQ100 10.41 6.73 47.89 9.14 4.69 0.74
2mQ0 12.83 14.22 61.52 10.02 6.73 1.58
2mQ100 14.22 15.92 63.26 10.09 6.77 1.86
a
Bure Yare
0
1
2
3
4
0 20000 40000
Distance from upstream boundary (m)
Water level (m)
b
Bure Yare
0
1
2
3
4
020000 40000
Distance from upstream boundary (m)
Water level (m)
Fig. 12 Longitudinal profile view of maximum water levels along the River Bure and River Yare from the
model’s upstream boundary to its downstream boundary, the North Sea. aMaximum water levels for
scenarios 2100Q0 (red) and 2100Q100 (blue, dashed). bMaximum water levels for scenarios 1mQ0 (red)
and 1mQ100 (blue, dashed)
123
932 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
add freshwater to the system, while surges push saline water upstream into the Broads.
River salinity and conductivity can be simulated in HEC-RAS’s water quality module.
A significant benefit of the described 1D–2D approach in portraying overtopping is the
use of specific lateral structures for flood defences to guarantee that maximum crest heights
were accounted for, regardless of the chosen mesh resolution. It is a fundamental
requirement for 2D cells in HEC-RAS to be set up such that cell edges (or ‘‘faces’’) align
with high ground or structures impeding the movement of water. This task can be difficult
for narrow flood defences, even with a relatively fine resolution of 2 m. Cells that are too
large or that are not adequately oriented can cause issues with the model’s calculations,
leading water to incorrectly ‘‘leak’’ through natural or man-made barriers. The results in
such cases are fragmented and therefore produce unrealistic outputs of flooding extents.
The Broads is a highly engineered area with many embankments protecting large expanses
of land from rivers. It was therefore essential to use lateral structures between 1D and 2D
domains that capture the height of defences for their entire lengths. Until computational
capabilities increase to allow for extremely fine mesh resolutions, this study finds that a
1D–2D method remains the most feasible approach for the geographical location in
question.
The HEC-RAS 1D–2D model was able to highlight vulnerabilities and weak points
within the study area as well as account for complex interactions between different sources
of flooding. The model structure could still be improved by including building footprints in
the 2D mesh to better represent the flow of water in urban areas. Such levels of accuracy
were however not necessary to assess the overall sensitivity of the case study area and the
fitness for use of the HEC-RAS model version 5.0. Further developments for the model
could moreover be to include other parts of the Broads that currently lie outside the
modelling domain. Areas in the River Yare, Waveney, Thurne and Ant basins, as well as in
Lowestoft have experienced flooding in the past.
Several important considerations should be made when interpreting the results derived
from the presented hydraulic model. The first is that while flood defence infrastructure can
fail in a number of ways, only the overtopping of defences was considered here. The
erosion and breaching of dunes, embankments and walls are a common concern in coastal
regions (Hall et al. 2015). Although these processes can be simulated in HEC-RAS and can
be useful to represent catastrophic or ‘‘what if’’ scenarios, their impacts fell outside of the
scope of this study.
A more comprehensive study of flooding risk would moreover need to incorporate
processes of wind and waves, which were omitted in this simplified hydraulic modelling
framework. Wind is a key parameter that plays a role in the dynamics of both waves and
surges and can therefore have important consequences on coastal flooding. With the
necessary data, the EVA and the scenarios used for simulations could therefore be refined
by setting up local wave and storm surge models (e.g. Villatoro et al. 2014). Similarly, the
lack of available discharge data was also a limitation for this work. A hydrological model
could be used in future research to determine more accurate upstream boundaries for the
HEC-RAS hydraulic model. A hydrological model would moreover make it possible to
account for projected changes in temperature and precipitation in the Broadland catchment
to better understand the impact of these climatic changes on flooding hazard.
This study highlighted the potential for multiple extreme events occurring simultane-
ously to exacerbate flooding risk in the Broads. Validating the proposed modelling
framework to assess the sensitivity of the Broads, the aim of this research was however not
to understand the probabilities of co-occurrence of these events. The assumption was made
that peak river discharge and peak sea level occurred simultaneously in scenarios where
123
Natural Hazards (2019) 98:915–937 933
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
both events occurred. While it helped in interpreting the created scenarios, this assumption
may not be representative of likely events in the Broads. Past studies in other regions, such
as the Netherlands, have, for example, shown a dependency between discharge peaks and
water levels, but with a lag time of several days (Klerk et al. 2015). More analysis should
be performed to determine the dependency between discharge peaks and sea levels in the
East coast of England. Moreover, understanding the types of weather patterns associated
with different events could provide some useful insights into flooding hazard in the region.
As the timing of events can have significant consequences not only of flooding extent but
also on the usefulness of flood mitigation strategies, joint probabilities should be carefully
considered to make robust planning recommendations on flood risk management.
5 Conclusions
This study has looked to evaluate the sensitivity of a complex coastal environment to
different sources of flooding, using the new tools made available in HEC-RAS version 5.0.
A 1D–2D approach was found to be appropriate for flood mapping in this context, accu-
rately reproducing the flow of water in both large floodplains and urban areas while
reducing computational requirements. Lower simulation run times moreover made it
possible to cover a larger area from the coast and to 40 km inland where tidal and fluvial
processes interact. The proposed approach is particularly relevant to low-lying and low-
gradient regions like the Broads, which are prone to tidal flooding and where the tidal
boundary extends far upstream. There will continue to be more opportunities for 2D
modelling in the UK as the coverage of fine-resolution LIDAR data grows.
Hydraulic models are not only sensitive to topographical data but also to the choice and
fundamental design of boundary conditions. With extremes being the primary cause of
flooding in the Broads and in many regions around the world, it is important to capture the
hydrological conditions occurring during these events. The GPD function was used to
determine return levels of sea level and river discharge to create synthetic extreme events
under future conditions of SLR. Important assumptions were made to create simplified
synthetic events as the interest of this work was to assess the sensitivity of the Broads to
extreme flooding and the potential for the modelling framework to map out maximum
flooding extents. Peak river discharge and sea level were thereby designed to occur at the
same time. Similarly, the storm surge peak coincided with the highest point in the tide
cycle. For a more comprehensive assessment of flood risk, further research should look into
the significance of the timing of these events as well as the joint probability of their
occurrence. The proposed model however helps to understand the Broads’ sensitivity to
different sources of flooding. Storm surges are, and are likely to continue to be, the main
drivers for flooding in the Broads as RMSL rises over the next century. While there is still
uncertainty in the pattern of future precipitation with climate change, this study has shown
that high discharge could exacerbate the flooding caused by storm surges.
While the described hydraulic model can be expanded to cover a larger portion of the
Broads, this case study highlights the potential for 1D–2D modelling in assisting decision-
making. This methodology indeed allows for the consideration of urban coastal areas,
requiring a high amount of detail, as well as vast inland rural zones. It is moreover suited to
dynamically represent interacting sources of flooding and potential combined extreme
events. The presented approach is therefore a step towards helping meet the requirements
of integrated catchment management as well as flood alleviation and adaptation.
123
934 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Acknowledgements The first author is supported by the Natural Environment Research Council as part of a
CASE partnership with the Broads Authority. The authors would like to thank the editors of the Special
Issue in ‘‘Advances in extreme value analysis and Application to Natural Hazards’’, Dr Ivan Haigh and Dr
Thomas Wahl, for their valuable comments and for the organisation of the 2017 EVAN Conference.
Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 Interna-
tional License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution,
and reproduction in any medium, provided you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license, and indicate if changes were made.
Funding The Corresponding Author (Ulysse Pasquier) received a PhD studentship from the Natural
Environment Research Council as part of a CASE partnership with the Broads Authority. Award Number:
NE/L002582/1
References
Arns A, Wahl T, Haigh ID, Jensen J, Pattiaratchi C (2013) Estimating extreme water level probabilities: a
comparison of the direct methods and recommendations for best practise. Coast Eng 81:51–66
Barredo JI (2009) Normalised flood losses in Europe: 1970–2006. Nat Hazard Earth Syst Sci 9(1):97–104
Bezak N, Brilly M, S
ˇraj M (2014) Comparison between the peaks-over-threshold method and the annual
maximum method for flood frequency analysis. Hydrol Sci J 59(5):959–977. https://doi.org/10.1080/
02626667.2013.831174
Broads Authority (2014) Reflections on the December Tidal Surge and how this relates to adapting to
environmental changes in the Broads. http://www.broads-authority.gov.uk/__data/assets/pdf_file/0006/
426597/Reflections-on-the-December-Tidal-Surge-and-How-This-Relates-to-Adaption-to-
Environmental-Change-in-the-Broads.pdf. Accessed 28 Sept 2017
Chen L, Sign VP (2017) Generalized beta distribution of the second kind for flood frequency analysis.
Entropy 19(6):254. https://doi.org/10.3390/e19060254
Chow VT (1959) Open-channel hydraulics. McGraw-Hill, New York
Church JA, Clarks PU, Cazenave A, Greogry JM, Jevrejeva S, Levermann A, Merrifield MA, Milne GA,
Nerem RS, Nunn PD, Payne AJ, Pfeffer WT, Stammer D, Unnikrishnan AS (2013) Sea level change.
In: Stocker TF, Qin D, Plattner GK, Tignor M, Allen SK, Boschung J, Nauels A, Xia Y, Bex V,
Midgley PM (eds) Climate change 2013: the physical science basis. contribution of working group I to
the fifth assessment report of the intergovernmental panel on climate change. Cambridge University
Press, Cambridge and New York, pp 1137–1216
Environment Agency (2009) Broadland rivers catchment flood management plan: summary report
December 2009. Environment Agency. http://www.broads-authority.gov.uk/news-and-publications/
publications-and-reports/conservation-publications-and-reports/water-conservation-reports/36.-Broadland-
Flood-Management-Plan-2009.pdf. Accessed 28 Sept 2017
Haigh ID, Wadey MP, Wahl T, Ozsoy O, Nicholls RJ, Brown JM, Horsburgh K, Gouldby B (2016) Spatial
and temporal analysis of extreme sea level and storm surge events around the coastline of the UK. Sci
Data 3:160107
Hall JW, Dawson RJ, Wu XZ (2015) Analysing flood and erosion risks and coastal management strategies
on the Norfolk coast. In: Nicholls RJ, Dawson RJ, Day SA (eds) Broad scale coastal simulation: new
techniques to understand and manage shorelines in the third millennium, 1st edn. Springer, Dordrecht,
pp 233–254
He Y, Pappenberger F, Manful D, Cloke HL, Bates P, Wetterhall F, Parkes B (2013) Flood inundation
dynamics and socioeconomic vulnerability under environmental change. In: Hossain F (ed) Vulnera-
bility of water resources to climate. Climate vulnerability, vol 5. Elsevier Sciences, pp 241–255.
https://doi.org/10.1016/B978-0-12-384703-4.00508-6
Javaheri A, Babbar-Sebens M (2014) On comparison of peak flow reductions, flood inundation maps, and
velocity maps in evaluating effects of restored wetlands on channel flooding. Ecol Eng 75:132–145
Jenkins G (2009) Exeter: met office hadley centre. UKCP09 Briefing report, UK Climate projections. http://
ukclimateprojections.metoffice.gov.uk/media.jsp?mediaid=87868&filetype=pdf. Accessed 28 Sept
2017
123
Natural Hazards (2019) 98:915–937 935
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Kew SF, Selten FM, Lenderink G, Hazelger W (2013) The simultaneous occurrence of surge and discharge
extremes for the Rhine delta. Nat Hazards Earth Syst Sci 13:2017–2029. https://doi.org/10.5194/nhess-
13-2017-2013
Klerk WJ, Winsemius HC, van Verseveld WJ, Bakker AMR, Diermanse FLM (2015) The co-incidence of
storm surges and extreme discharges within the Rhine–Meuse Delta. Environ Res Lett 10:035005.
https://doi.org/10.1088/1748-9326/10/3/035005
Mantz PA, Wakeling HL (1979) Forecasting flood levels for joint events of rainfall and tidal surge flooding
using extreme value statistics. Proc Inst Civ Eng 67:31–50
McMillan A, Batstone C, Worth D, Tawn J, Horsburgh K, Lawless M (2011) Coastal flood boundary
conditions for UK mainland and islands. Environmental Agency, Bristol
Mene
´ndez M, Woodworth PL (2010) Changes in extreme high water levels based on a quasi-global tide-
gauge data set. J Geophys Res Oceans 115:C10011. https://doi.org/10.1029/2009JC005997
Ne
´elz S, Pender G (2009) Desktop review of 2D hydraulic modelling packages. DEFRA/Environment
Agency. http://evidence.environment-agency.gov.uk/FCERM/Libraries/FCERM_Project_Documents/
SC080035_Desktop_review_of_2D_hydraulic_packages_Phase_1_Report.sflb.ashx. Accessed 28 Sept
2017
Patel DP, Ramirez JA, Srivastava PK, Bray M, Dawai H (2017) Assessment of flood inundation mapping of
Surat city by coupled 1D/2D hydrodynamic modeling: a case application of the new HEC-RAS 5. Nat
Hazards 89(1):93–130
Pfeffer WT, Harper JT, O’Neel S (2008) Kinematic constraints on glacier contributions to 21st-century sea-
level rise. Science 321(5894):1340–1343
Pugh DJ (1996) Tides, surges and mean sea level. A handbook for engineers and scientists. Wiley,
Chichester
Quiroga VM, Kure S, Udo K, Mano A (2016) Application of 2D numerical simulation for the analysis of the
February 2014 Bolivian Amazonia flood: application of the new HEC-RAS version 5. RIBAGUA-
Revista Iberoamericana del Agua 3:25–33
Ray T, Stepinski E, Sebastian A, Bedient PB (2011) Dynamic modeling of storm surge and inland flooding
in a Texas coastal floodplain. J Hydraul Eng 137(10):1103–1111
Rouillard JJ, Ball T, Heal KV, Reeves AD (2015) Policy implementation of catchment-scale flood risk
management: learning from Scotland and England. Environ Sci Policy 50:155–165
Saeed Far SS, Abd. Wahab AK (2016) Evaluation of peaks-over-threshold method. Ocean Sci Discuss.
https://doi.org/10.5194/os-2016-47
Shrestha S, Lohpaisankrit W (2016) Flood hazard assessment under climate change scenarios in the Yang
River Basin, Thailand. Int J Sustain Built Environ. https://doi.org/10.1016/j.ijsbe.2016.09.006
Spencer T, Brooks SM, Evans BR, Tempest JA, Moller I (2015) Southern North Sea storm surge event of 5
December 2013: water levels, waves and coastal impacts. Earth Sci Rev 145:120–145. https://doi.org/
10.1016/j.earscirev.2015.04.002
Stevens AJ, Clarke D, Nicholls RJ (2016) Trends in reported flooding in the UK: 1884–2013. Hydrol Sci
61:50–63. https://doi.org/10.1080/02626667.2014.950581
Svensson C, Jones DA (2002) Dependence between extreme sea surge, river flow and precipitation in
eastern Britain. Int J Climatol 22:1149–1168
Teng J, Jakeman AJ, Vaze J, Croke BFW, Dutta D, Kim S (2017) Flood inundation modelling: a review of
methods, recent advances and uncertainty analysis. Environ Model Softw 90:201–216
USACE (2016) HEC-RAS river analysis system user’s manual version 5.US Army Corps of Engineers.
http://www.hec.usace.army.mil/software/hec-ras/documentation/HEC-RAS%205.0%20Users%
20Manual.pdf Accessed 28 Sept 2017
van den Hurk B, van Meijgaard E, de Valk P, van Heeringen K-J, Gooijer J (2015) Analysis of a com-
pounding surge and precipitation event in the Netherlands. Environ Res Lett 10:035001. https://doi.
org/10.1088/1748-9326/10/3/035001
Villatoro M, Silva R, Mendez FJ, Zanuttigh B, Pan S, Trifonova E, Losada IJ, Izaguirre C, Simmonds D,
Reeve DE, Mendoze E, Martinelli L, Formentin SM, Galiatsatou P, Eftimova P (2014) An approach to
assess flooding and erosion risk for open beaches in a changing climate. Coast Eng 87:50–76. https://
doi.org/10.1016/j.coastaleng.2013.11.009
Vozinaki A-EK, Morianou GG, Alexakis DD, Tsanis IK (2017) Comparing 1D and combined 1D/2D
hydraulic simulations using high-resolution topographic data: a case study of the Koilaris basin,
Greece. Hydrol Sci J 62(4):642–656. https://doi.org/10.1080/02626667.2016.1255746
Wahl T, Haigh ID, Woodworth PL, Albrecht F, Dillingh D, Jensen J, Nicholls RJ, Weisse R, Wo
¨ppelmann
G (2013) Observed mean sea level changes around the North Sea coastline from 1800 to present. Earth
Sci Rev 124:51–67. https://doi.org/10.1016/j.earscirev.2013.05.003
123
936 Natural Hazards (2019) 98:915–937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
Wahl T, Jain S, Bender J, Meyers SD, Luther ME (2015) Increasing risk of compound flooding from storm
surge and rainfall for major US cities. Nat Clim Change 5:1093–1097. https://doi.org/10.1038/
nclimate2736
Wang G, Wang D, Trenberth KE, Erfanian A, Yu M, Bosilovich MG, Parr DT (2017) The peak structure
and future changes of the relationships between extreme precipitation and temperature. Nat Clim
Change 7:268–274. https://doi.org/10.1038/nclimate3239
Webster T, McGuigan K, Collins K, MacDonald C (2014) Integrated river and coastal hydrodynamic flood
risk mapping of LaHavre River Estuary and Town of Bridgewater, Nova Scotia, Canada. Water
6:517–546. https://doi.org/10.3390/w6030517
Whitfield PH (2012) Floods in future climates: a review. J Flood Risk Manag 5:336–365. https://doi.org/10.
1111/j.1753-318X.2012.01150.x
Wilby RL, Beven KJ, Reynard NS (2008) Climate change and fluvial flood risk in the UK: more of the
same? Hydrol Process 22:2511–2523. https://doi.org/10.1002/hyp.6847
Wong PP, Losada IJ, Gattuso JP, Hinkel J, Khattabi A, McInnes KL, Saito Y, Sallenger A (2014) Coastal
systems and low-lying areas. In: Field CB, Barros VR, Dokken DJ, Mach KJ, Mastrandrea MD, Billir
TE, Chatterjee M, Ebi KL, Estrada YO, Genova RC, Girma B, Kissel ES, Levy AN, MacCracken S,
Mastrandrea PR, White LL (eds) Climate change 2014: impacts, adaptation, and vulnerability. Part A:
global and sectoral aspects. contribution of working group II to the fifth assessment report of the
intergovernmental panel on climate change. Cambridge University Press, Cambridge and New York,
pp 361–409
Wu XZ, Hall JW, Liang Q, Dawson RJ (2015) Broadscale Coastal Inundation Modelling. In: Nicholls RJ,
Dawson RJ, Day SA (eds) Broad scale coastal simulation: new techniques to understand and manage
shorelines in the third millennium, 1st edn. Springer, Dordrecht, pp 213–232
Zheng F, Westra S, Loenard M, Sisson SA (2014) Modeling dependence between extreme rainfall and storm
surge to estimate coastal flooding risk. Water Resour Res 50:2050–2071. https://doi.org/10.1002/
2013WR014616
Affiliations
Ulysse Pasquier
1
Yi He
1
Simon Hooton
2
Marisa Goulden
3
Kevin M. Hiscock
4
&Ulysse Pasquier
u.pasquier@uea.ac.uk
1
Tyndall Centre for Climate Change Research, School of Environmental Sciences, University of
East Anglia, Norwich Research Park, Norwich NR4 7TJ, UK
2
Broads Authority, Yare House, 62-64 Thorpe Road, Norwich NR1 1RY, UK
3
School of International Development, University of East Anglia, Norwich Research Park,
Norwich NR4 7TJ, UK
4
School of Environmental Sciences, University of East Anglia, Norwich Research Park,
Norwich NR4 7TJ, UK
123
Natural Hazards (2019) 98:915–937 937
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center
GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers
and authorised users (“Users”), for small-scale personal, non-commercial use provided that all
copyright, trade and service marks and other proprietary notices are maintained. By accessing,
sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of
use (“Terms”). For these purposes, Springer Nature considers academic use (by researchers and
students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and
conditions, a relevant site licence or a personal subscription. These Terms will prevail over any
conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription (to
the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of
the Creative Commons license used will apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may
also use these personal data internally within ResearchGate and Springer Nature and as agreed share
it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not otherwise
disclose your personal data outside the ResearchGate or the Springer Nature group of companies
unless we have your permission as detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial
use, it is important to note that Users may not:
use such content for the purpose of providing other users with access on a regular or large scale
basis or as a means to circumvent access control;
use such content where to do so would be considered a criminal or statutory offence in any
jurisdiction, or gives rise to civil liability, or is otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association
unless explicitly agreed to by Springer Nature in writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a
systematic database of Springer Nature journal content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a
product or service that creates revenue, royalties, rent or income from our content or its inclusion as
part of a paid for service or for other commercial gain. Springer Nature journal content cannot be
used for inter-library loans and librarians may not upload Springer Nature journal content on a large
scale into their, or any other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not
obligated to publish any information or content on this website and may remove it or features or
functionality at our sole discretion, at any time with or without notice. Springer Nature may revoke
this licence to you at any time and remove access to any copies of the Springer Nature journal content
which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or
guarantees to Users, either express or implied with respect to the Springer nature journal content and
all parties disclaim and waive any implied warranties or warranties imposed by law, including
merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published
by Springer Nature that may be licensed from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a
regular basis or in any other manner not expressly permitted by these Terms, please contact Springer
Nature at
onlineservice@springernature.com
... The adjusted catchments' boundar- Fig. 1 Methodological framework utilised within this study. After Pasquier (2020) ies ( Fig. 2) were delineated using the 30 × 30 m resolution Shuttle Radar Topography Mission digital elevation model (Pasquier et al. 2018). ...
... Flooding within the Broadland catchment has been priorly simulated in the contexts of extreme river discharge and tidal surges, alongside associated climatic changes. One such study, Pasquier et al. (2018) explored inundation within the Broads region from Great Yarmouth northwards. Pasquier et al. (2018) scenario 2mQ100 (a 2-metre mean SLR 100-year storm surge event in parallel with a 100-year river discharge event) showed greater flooding through the Filby Broad than the 2070 compound scenario (Fig. 5c), although, 2mQ100 scenario is based in 2100 with 2 m sea level rise, while the sea level rise exhibited to the 2070 in this study only documents an elevation of 0.55 m from 1990 (Table 2). ...
... One such study, Pasquier et al. (2018) explored inundation within the Broads region from Great Yarmouth northwards. Pasquier et al. (2018) scenario 2mQ100 (a 2-metre mean SLR 100-year storm surge event in parallel with a 100-year river discharge event) showed greater flooding through the Filby Broad than the 2070 compound scenario (Fig. 5c), although, 2mQ100 scenario is based in 2100 with 2 m sea level rise, while the sea level rise exhibited to the 2070 in this study only documents an elevation of 0.55 m from 1990 (Table 2). Nevertheless, this elevation produces significant propagation and expansion of inundation along the Waveney limb of the Broads in each surge-related scenario, as well as its bordering tributaries, due to its shallow elevation profile and proximity to the surge-dominant core Broadland (Figs. 2 and 5b). ...
Article
Full-text available
Flooding represents the greatest natural threat to the UK, presenting severe risk to populations along coastlines and floodplains through extreme tidal surge and hydrometeorological events. Climate change is projected to significantly elevate flood risk through increased severity and frequency of occurrences, which will be exacerbated by external drivers of risk such as property development and population growth throughout floodplains. This investigation explores the entire flood hazard modelling chain, utilising the nonparametric bias correction of UKCP18 regional climate projections, the distributed HBV-TYN hydrological model and HEC-RAS hydraulic model to assess future manifestation of flood hazard within the Broadland Catchment, UK. When assessing the independent impact of extreme river discharge and storm surge events as well as the impact of a compound event of the two along a high emission scenario, exponential increases in hazard extent over time were observed. The flood extent increases from 197 km² in 1990 to 200 km² in 2030, and 208 km² in 2070. In parallel, exponential population exposure increases were found from 13,917 (1990) to 14,088 (2030) to 18,785 (2070). This methodology could see integration into policy-based flood risk management by use of the developed hazard modelling tool for future planning and suitability of existing infrastructure at a catchment scale.
... Finally, a hazard assessment based on flood modelling leading to a probabilistic hazard for specific areas at a municipal scale is advisable. The modelling should be carried out using a 1D, or 2D hydrodynamic model [40]. The advantage of these models is the ability to predict water depth, flow velocity, and flood extent over the whole area under study. ...
Article
Full-text available
Mining contributes to an increased tax revenues, export earnings, employment opportunities, infrastructure development, and transfer of technology to the host countries. However, there is a high risk of socio-economic enclaves or environmental damages caused by mining operations which include the release of heavy metals into water and soil resources, air pollution, land degradation, and depletion of vegetation. Mineral exploration is often preoccupied with integrating relevant datasets and utilising various techniques to discover subtle surface indications of mineralisation and usually results in huge disturbances of large surface area and underlying strata including aquifers. These issues with mineral exploration are pronounced in developing countries where the primary goal of many investors is to look for the occurrence of mineral of interest and its economic viability with little or no attention to the associated hydrogeological risk. Hydrogeological Risk Assessment, even though crucial to the development of sound environmental impact assessment for a proposed mineral exploration project, is often not included in mineral exploration activities. This could potentially result in a significant health risk and environmental impact which could negatively impact various communities within the mining areas. In this research, an overview of the various stages involved in evaluating and assessing the hydrogeological risk from the initial exploration stage has been provided. A framework that will enable a short-term to long-term integrated groundwater resources management system into mineral exploration projects have also been developed. This framework would provide valuable information to stakeholders to verify whether a proposed project is likely to be debilitating to hydrogeology in a locality.
... However, the available river discharge data for the current case are limited. Then, by following the approach of Pasquier et al. (2019), a comparison of the available data on past storm surges and discharge has been carried out. ...
Article
Full-text available
This paper aims to emphasize the importance of simulating river hydrodynamics for all the stages of river mouth deviations, since their potential influence on the flood hazard assessment. The case study of the Pescara River mouth deviation in the Abruzzo Region, Italy, has been considered. Only the realized steps of the project have been analyzed, for which the sea level at the mouth has turned out to not remarkably change for the considered layouts. Consequently, flooded areas and flow velocity do not significantly vary between the analyzed configurations. The adopted modeling chain has proved to be a suitable tool to support engineers, contractors, and controlling authorities in the different project phases.
... Furthermore, it is crucial to assess the downstream impacts of a dam collapse to identify high-risk areas and formulate emergency evacuation strategies for the residents. In this regard, hydraulic numerical models of one-dimensional (e.g., Pilotti et al. 2011;Guido et al. 2023), two-dimensional (e.g., Tsakiris and Bellos 2014;Yilmaz et al. 2023), and integrated 1D/2D models (e.g., Pasquier et al. 2019;Jibhakate et al. 2023) are commonly employed in the literature. Although 1D models are faster compared to 2D models, they cannot adequately represent flood wave diffusion on floodplains having complex topographic conditions and might be disadvantageous for the cases which water flow path is ambiguous (Ahmad and Simonovic 1999). ...
Article
Full-text available
One of the most perilous natural hazards is flooding resulting from dam failure, which can devastate downstream infrastructure and lead to significant human casualties. In recent years, the frequency of flash floods in the northern part of Nicosia, Cyprus, has increased. This area faces increased risk as it lies downstream of the Kanlikoy Dam, an aging earth-fill dam constructed over 70 years ago. In this study, we aim to assess potential flood hazards stemming from three distinct failure scenarios: piping, 100-year rainfall, and probable maximum precipitation (PMP). To achieve this, HEC-HMS hydrologic model findings were integrated into 2D HEC-RAS hydraulic models to simulate flood hydrographs and generate flood inundation and hazard maps. For each scenario, Monte Carlo simulations using McBreach software produced four hydrographs corresponding to exceedance probabilities of 90%, 50%, 10%, and 1%. The results indicate that all dam breach scenarios pose a significant threat to agricultural and residential areas, leading to the destruction of numerous buildings, roads, and infrastructures. Particularly, Scenario 3, which includes PMP, was identified as the most destructive, resulting in prevailing flood hazard levels of H5 and H6 in the inundated areas. The proportion of inundated areas in these high hazard levels varied between 52.8% and 57.4%, with the number of vulnerable structures increasing from 248 to 321 for exceedance probabilities of 90% and 1%, respectively. Additionally, the number of flooded buildings ranged from 842 to 935, and 26 to 34 km of roads were found to be inundated in this scenario. These findings revealed the need for authorities to develop comprehensive evacuation plans and establish an efficient warning system to mitigate the flood risks associated with dam failure.
... The proposed embankments here necessitate thorough evaluation to ascertain their effectiveness and potential areas for improvement. While factors like tidal activities, waves, and precipitation are accounted for in the planning, emerging challenges such as sea-level rise (SLR) and land subsidence (LS) due to climate change and urbanization, respectively, pose additional risks (Pasquier et al., 2019;Prasetyo et al., 2019). These evolving factors could diminish the efficacy of structural solutions like embankments, underscoring the need for a comprehensive approach to coastal flood management incorporating SLR and LS considerations. ...
Article
Full-text available
Jakarta's coastal region, characterized by dense settlements and industrial zones, faces significant flood risks due to land subsidence and rising sea levels, exacerbated by human activities and climate change. This study evaluates the effectiveness of the National Capital Integrated Coastal Development (NCICD) project, which proposes embankments along Jakarta's coast for enhanced flood resilience. Focusing on the Cengkareng Drain estuary, two-dimensional HEC-RAS simulations were employed. This study incorporates various factors such as precipitation, tides, waves, rising sea levels, and land subsidence to model extreme flood scenarios. The analysis reveals key strategies to improve flood management. Firstly, introducing new pumping stations and augmenting existing ones can reduce inundation by 43.16% under present conditions and significantly up to 53.16% by 2050. Secondly, regular dredging to sustain channel capacity is crucial. Additionally, mitigating local land subsidence through stringent law enforcement and heightened public awareness is vital for long-term flood control. These findings offer actionable insights into advancing Jakarta's flood management strategies in the face of evolving environmental challenges.
... Several studies on compound floods have been published (e.g., Ikeuchi et al. 2017;Kumbier et al. 2018;Pasquier et al. 2019;Hendry et al. 2019;Yin et al. 2021;Toyoda et al. 2021). For example, Ikeuchi et al. (2017) conducted a flood simulation to investigate the simultaneous occurrence of storm surges and river flows in Bangladesh during Cyclone Sidr. ...
Article
Full-text available
This study evaluated the compound flood risk of 11 different-sized rivers in the estuaries of the Ise and Mikawa Bays, Japan using an integrated framework of atmosphere–ocean–river developed in this study (one-way coupling). First, the framework was developed by incorporating the river channel into a coupled model of surge-wave-tide to include the interaction of the storm surge runup and river flow. In addition, the framework was validated by the Typhoon Trami (2018)-induced meteorological field, discharge, and storm surge with high accuracy. Second, the time difference between the storm surge and discharge at the estuary (ΔT) was investigated, assuming six typhoons with different tracks and similar distributions of intensity and precipitation using Typhoon Hagibis (2019) as a case study. The ΔT was highly positively correlated with the length of the river channel (correlation coefficient: 0.90). Moreover, the smaller rivers were more prone than large rivers to simultaneous storm surges and high river flow. The average ΔT for the smaller rivers was 180.4 min (normalized S.D. = 0.31) with a minimum of 15 min in the most severe case, while the average ΔT for the large-scale rivers was 614.1 min (normalized S.D. = 0.39). We clarified that the storm surge and high river flow occurred simultaneously (within 15 min) in the most severe river case (Yagyu River). These results infer that small rivers have a more significant impact on the co-occurrence of storm surge and high-river flow than large-scale rivers.
Chapter
This paper presents a comprehensive general overview of the extensive literature available in the field of application of artificial intelligence (AI) in flood prediction. The initial approach involves examining the sources responsible for flood occurrence and the various components associated with them. Subsequently, it explores the potential of different remote sensing platforms and sensors in quantifying the diverse flood variables originating from these sources and components. These quantified variables are then utilized as inputs for training, testing, predicting, and validating AI models. A concise explanation of the AI concept, along with its subfields such as machine learning (ML) and deep learning (DL) algorithms has also been provided. Furthermore, a brief overview of prospective AI algorithms is provided. Moving forward, we summarize various types of AI methodologies and provide relevant references in the context of natural hazards, with a specific focus on floods. Finally, we discuss advancements in AI and highlight how it enhances flood modeling. Additionally, it presents insights into potential future applications in this field, emphasizing the continuous progress and the immense potential of AI in flood prediction. The findings suggest that the field of the chapter’s theme is highly relevant but there is significant lack of work in this area of research, and it should be one of the most important thrust area in the field of flood disaster prediction.
Preprint
Coastal zones are particularly vulnerable to flooding. Several climatic and state variables may drive the occurrence of such events, e.g., storm surges, sea level rise, heavy rainfall, and high river and groundwater levels. The co-occurrence of such events, i.e. compound or cascading effects, has been shown to escalate flooding impacts and extent, but the contribution of groundwater is routinely overlooked. Here, we apply an integrated hydrological/hydrodynamic/groundwater model to investigate underlying causes and compound effects in a Danish Wadden sea catchment. Two models were developed: a long-term model and an overbank-spilling model. The long-term model was calibrated and used to simulate 30-year periods. Extreme value analyses were carried out for sea levels, precipitation, simulated river water stages, and groundwater levels. The co-occurrence of extremes was used to identify compound effects on high river-stage incidents (as a flood proxy). The overbank-spilling model was then used for simulating flooding for a subset of the largest river stage events identified from the long-term model. The analysis showed that the river-stage events were closely correlated to the sea level extremes, but that the largest river-stage events were almost exclusively compounded by precipitation or groundwater, or both. High groundwater tables seem to correlate to the flooding events with the largest spatial extent, as well as prolonged extreme events where either precipitation or sea level were elevated during long periods. Thus, this study shows that there is a general need to acknowledge the potential effect of groundwater levels on the resulting flooding on terrain in coastal zones.
Article
Floods have been causing the world’s costliest weather-related catastrophes and their magnitude and frequency are projected to increase even further due to climate change. Current flood risk quantification procedures include the use of complex and highly uncertain hydrologic-hydraulic models for hazard mapping and computationally-tedious manipulations for vulnerability evaluation—hindering urban centers climate resilience planning. Adopting a novel approach that bypasses such time-consuming procedures, this study presents a deep learning-based rapid and accurate flood risk prediction tool, RAPFLO, to directly relate flood risk characteristics (level, extent, and likelihood) to their main drivers (e.g., climate, topography, and land cover). The approach employed to develop RAPFLO is generic in nature and the associated methodology is not site-dependent. To demonstrate its utility, RAPFLO is deployed on the City of Calgary, Canada, and is used to reproduce the fluvial flood risk across the city between the years 2010 and 2020. RAPFLO efficiently replicated the risk level with an overall accuracy of 80 % and the risk likelihood with a coefficient of determination of 0.96. Subsequently, RAPFLO was employed for predicting future fluvial flood risk from the year 2025 to 2100 under the RCP 8.5 climate scenario. RAPFLO presents a valuable computationally efficient, accurate, and rapid decision support system that empowers city managers and infrastructure operators to devise effective climate resilience strategies considering different climate projections and future what-if scenarios.
Article
Full-text available
Estimation of flood magnitude for a given recurrence interval T (T-year flood) at a specific location is needed for design of hydraulic and civil infrastructure facilities. A key step in the estimation or flood frequency analysis (FFA) is the selection of a suitable distribution. More than one distribution is often found to be adequate for FFA on a given watershed and choosing the best one is often less than objective. In this study, the generalized beta distribution of the second kind (GB2) was introduced for FFA. The principle of maximum entropy (POME) method was proposed to estimate the GB2 parameters. The performance of GB2 distribution was evaluated using flood data from gauging stations on the Colorado River, USA. Frequency estimates from the GB2 distribution were also compared with those of commonly used distributions. Also, the evolution of frequency distribution along the stream from upstream to downstream was investigated. It concludes that the GB2 is appealing for FFA, since it has four parameters and includes some well-known distributions. Results of case study demonstrate that the parameters estimated by POME method are found reasonable. According to the RMSD and AIC values, the performance of the GB2 distribution is better than that of the widely used distributions in hydrology. When using different distributions for FFA, significant different design flood values are obtained. For a given return period, the design flood value of the downstream gauging stations is larger than that of the upstream gauging station. In addition, there is an evolution of distribution. Along the Yampa River, the distribution for FFA changes from the four-parameter GB2 distribution to the three-parameter Burr XII distribution.
Article
Full-text available
Surat city of India, situated 100 km downstream of Ukai dam and 19.4 km upstream from the mouth of river Tapi has experienced the largest flood in 2006. The peak discharge of about 25,770m3/s released from the Ukai dam was responsible for a disaster. To assess the flood and find inundation in low lying areas, simulation work is carried out under the 1D/2D couple hydrodynamic modeling. 299 cross sections, 2 hydraulic structures and 5 major bridges across the river are considered for 1D modeling, whereas a topographic map at 0.5 m contour interval was used to produce a 5 m grid and SRTM (30 & 90 m) grid has been considered for Surat and the Lower Tapi Basin. The tidal level at the river mouth and the release from the Ukai dam during 2006 flood is considered as the downstream and upstream boundaries respectively. The model is simulated under the unsteady flow condition and validated for the year 2006.The simulated result shows that 9th August was the worst day in terms of flooding for Surat city and a maximum 75-77 % area was under inundation. Out of seven zones, the West zone had the deepest flood and inundated under 4-5m. Furthermore, inundation is simulated under the bank protection work (i.e. levees, Retaining Wall) constructed after the 2006 flood. The simulated results show that the major zones are safe against the inundation under 14,430 m3/s water releases from Ukai dam except for the West zone. The study shows the 2D capability of new HEC-RAS 5 for flood inundation mapping and management studies.
Article
Full-text available
In this paper we analyse the spatial footprint and temporal clustering of extreme sea level and skew surge events around the UK coast over the last 100 years (1915–2014). The vast majority of the extreme sea level events are generated by moderate, rather than extreme skew surges, combined with spring astronomical high tides. We distinguish four broad categories of spatial footprints of events and the distinct storm tracks that generated them. There have been rare events when extreme levels have occurred along two unconnected coastal regions during the same storm. The events that occur in closest succession (<4 days) typically impact different stretches of coastline. The spring/neap tidal cycle prevents successive extreme sea level events from happening within 4–8 days. Finally, the 2013/14 season was highly unusual in the context of the last 100 years from an extreme sea level perspective.
Article
Full-text available
Climate change is expected to increase both the magnitude and frequency of extreme precipitation events, which may lead to more intense and frequent river flooding. This study aims to assess the flood hazard potential under climate change scenarios in Yang River Basin of Thailand. A physically-based distributed hydrological model, Block-wise use of TOPMODEL using Muskingum-Cunge flow routing (BTOPMC) and hydraulic model, HEC-RAS was used to simulate the floods under future climate scenarios. Future climate scenarios were constructed from the bias corrected outputs of three General Circulation Models (GCMs) for 2020s, 2050s and 2080s. Results show that basin will get warmer and wetter in future. Both the minimum and maximum temperature of the basin is projected to increase in future. Similarly average annual rainfall is also projected to increase in future, higher in near future and lower in far future. The extreme runoff pattern and synthetic inflow hydrographs for 25, 50 and 100 year return flood were derived from an extreme flood of 2007 which were then fed into HEC-RAS model to generate the flood inundation maps in the basin. The intensity of annual floods is expected increase for both RCP 4.5 and 8.5 scenarios. Compared to the baseline period, an additional 60 km² area of basin is projected to be flooded with the return period of 100 years. The results of this study will be helpful to formulate adaptation strategies to offset the negative impacts of flooding on different land use activities in Yang River Basin.
Article
Full-text available
Two extreme wave analysis models, namely Peaks-Over-Threshold (POT) and Generalized Pareto Distribution (GPD), were developed in order to improve the POT model and highlight merits and limitations of the two models. Studies have shown that the POT model was not equipped with a suitable approach to determine a true threshold value. This paper proposed an approach to specify the most suitable threshold value for the POT model, which is called Hybrid method. In addition, until now the MIR (minimum ratio of residual correlation coefficient) criterion has been used as a goodness-of-fit method in the POT model. However, the examinations on the method represented that MIR is not always a stable approach in determining true distribution function. This paper proposed an alternative approach instead of the MIR criterion method, it is called Norm of Residuals, and its credibility was examined by the Chi-Square test. The results drawn from this study also demonstrated that the Hybrid method completely matched with the POT model, and the threshold obtained by this method is credible, moreover, the Norm of Residuals method is completely stable in determining the best fitting distribution for the POT model.
Article
Full-text available
Llanos de Moxos are vast plains in the Bolivian Amazonia that are continually flooded by the Mamore river. The flood lasts for several days affecting important cities like Trinidad, drowning people, drowning cattle and swamping arable land. Because of the cloudy skies, remote sensing observations are limited to some areas and few days. Thus, there is huge uncertainty about characteristics of flood events and possible consequences. Two-dimensional (2D) numerical simulation proved to be an important tool for understanding flood events. The HEC-RAS model is one of the most popular hydraulic models. In 2014 a new version of HEC-RAS (HEC-RAS-v5) was released including 2D capabilities. The present study applied the new HEC-RAS-v5 to simulate the February 2014 flood event in the Bolivian Amazonia. The flood simulated shows good performance when compared with satellite image of the flood event. In addition, the simulation provides information like water depth, flow velocity and a temporal variation of the flood. Specific locations where water begins to overflow were identified. Over most of the flooded area the water velocity is lower than 0.25ms−1. During first ten days of the flood the flood extent increases rapidly. The flood depth allows identifying areas exposed to different hazard levels. The west plain of the Mamore river is the most exposed to the flood; it shows bigger flood extent, longer flood duration and deeper water depth. The flood that threatens the city of Trinidad originates in two locations; one located 32km at the north and other located 10km at the south west. The flood from the north gets close to Trinidad twelve days after it begins to overflow, while the flood from the south gets close to Trinidad seven days after it begins to overflow. Although the flood from the north is deeper than the flood from the south, the flood from the south begins flooded before the north. Thus, water borne and vector borne diseases may originate at the south earlier than the north. The city of San Javier gets covered by flood five days after the water begins to overflow. The study shows the applicability and the value of the 2D capabilities of the new HEC-RAS for flood studies.
Article
This paper reviews state-of-the-art empirical, hydrodynamic and simple conceptual models for determining flood inundation. It explores their advantages and limitations, highlights the most recent advances and discusses future directions. It addresses how uncertainty is analysed in this field with the various approaches and identifies opportunities for handling it better. The aim is to inform scientists new to the field, and help emergency response agencies, water resources managers, insurance companies and other decision makers keep up-to-date with the latest developments. Guidance is provided for selecting the most suitable method/model for solving practical flood related problems, taking into account the specific outputs required for the modelling purpose, the data available and computational demands. Multi-model, multi-discipline approaches are recommended in order to further advance this research field.
Article
Theoretical models predict that, in the absence of moisture limitation, extreme precipitation intensity could exponentially increase with temperatures at a rate determined by the Clausius-Clapeyron (C-C) relationship. Climate models project a continuous increase of precipitation extremes for the twenty-first century over most of the globe. However, some station observations suggest a negative scaling of extreme precipitation with very high temperatures, raising doubts about future increase of precipitation extremes. Here we show for the present-day climate over most of the globe, the curve relating daily precipitation extremes with local temperatures has a peak structure, increasing as expected at the low-medium range of temperature variations but decreasing at high temperatures. However, this peak-shaped relationship does not imply a potential upper limit for future precipitation extremes. Climate models project both the peak of extreme precipitation and the temperature at which it peaks (T peak) will increase with warming; the two increases generally conform to the C-C scaling rate in mid- and high-latitudes, and to a super C-C scaling in most of the tropics. Because projected increases of local mean temperature (T mean) far exceed projected increases of T peak over land, the conventional approach of relating extreme precipitation to T mean produces a misleading sub-C-C scaling rate.
Article
The development of high spatial resolution Digital Elevation Models takes place via the use of GeoEye-1 stereo-pair imagery, providing high accurate geometry representation of a complex riverine system. The combination of geographic information systems with hydraulic models facilitates the exploitation of the satellite extracted topographic information throughout the cross-section extraction process. 1D HEC-RAS and combined 1D/2D HEC-RAS model are adjusted making use of the produced high-resolution input. Several hydraulic simulations are effectuated in order to test how significantly DEM resolution affects hydraulic modelling results, regarding also the model dimensionality. The ability of the combined 1D/2D model, based mainly on the high accuracy input data, provides the accurate estimation of the flood hazard area. Flood prone areas could take advantage of high accurate results and facilitate the effective management of extreme events and sufficient decision making.