You are on page 1of 19

Journal of Hydrology 314 (2005) 139157

www.elsevier.com/locate/jhydrol

Identifiability of distributed floodplain roughness


values in flood extent estimation
M.G.F. Wernera,b,*, N.M. Hunterc, P.D. Batesc
a
Delft University of Technology, P.O. Box 5048, 2600 GA Delft, The Netherlands
b
Delft Hydraulics, P.O. Box 177, 2600 MH Delft, The Netherlands
c
School of Geographical Sciences, University of Bristol, University Road, Bristol BS8 1SS, UK
Received 18 February 2004; revised 28 February 2005; accepted 18 March 2005

Abstract
In this paper we explore the potential for identifying roughness values for distributed land use types using a comprehensive
calibration data set of the 1995 flood event in the River Meuse, including gauged levels, flood extent maps and distributed flood
plain level observations. The reach studied is modelled using an integrated 1D2D hydrodynamic model, with floodplain flow
modelled in the 2D domain. Detailed information on floodplain land use types is aggregated to form one, two or five classes of
floodplain roughness. Sensitivity analysis of model performance against the calibration data shows that as the number of
floodplain classes increases, sensitivity to these roughness values decreases, given allocation of prior roughness values on the
basis of constituent land use types and associated roughness values found from literature. Evaluating the identifiability of the
roughness in these classes using the Generalised Likelihood Uncertainty Estimation (GLUE) method confirms this insensitivity.
As a consequence, application of complex formulae to establish roughness values for changed floodplain land use would seem
inappropriate, and evaluation of such changes within a probabilistic framework is suggested.
q 2005 Elsevier B.V. All rights reserved.

Keywords: Floodplain inundation; Calibration; Parameter uncertainties; Distributed roughness values

1. Introduction regulation, river normalisation, building of groynes,


and dredging (Hesselink, 2002). Although some of
Anthropogenic influences on lowland rivers with these changes have improved discharge capacity, the
extensive floodplains such as the river Rhine, have in constriction of floodplains has, in particular, resulted
the last centuries resulted in major changes in the in higher flood stages. Sedimentation on the con-
riverine landscape, including constriction of flood- stricted floodplain leads to still higher flood stages
plains by construction of embankments, discharge (Asselman and van Wijngaarden, 2002) and projected
changes in river regime as a consequence of enhanced
climate change are foreseen to raise levels even
* Corresponding author. Address: Delft Hydraulics, P.O. Box
177, 2600 MH, Delft, The Netherlands. Tel.: C31 15 285 8807;
further (Middelkoop et al., 2001). Traditionally, the
fax: C31 15 285 8582. engineering response to these higher flood stages has
E-mail address: micha.werner@wldelft.nl (M.G.F. Werner). been to strengthen and raise embankments, but more
0022-1694/$ - see front matter q 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhydrol.2005.03.012
140 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

recently public acceptance of these engineering data of higher accuracy will result in more accurate
measures has diminished (Wijbenga et al., 1994), roughness estimation. Recent work has shown the
and the focus is shifting to flood control as an use of airborne scanning laser altimetry for establish-
integrated part of river restoration (Pedroli and ing reach scale distributed roughness coefficients
Dijkman, 1998). In this context, a much broader (Asselman et al., 2002; Mason et al., 2003; Cobby
range of measures are considered, attempting to et al., 2003), but more traditional surveys of floodplain
combine sustainable measures such as nature devel- land use may be equally suitable. Estimating the
opment and floodplain lowering (Duel et al., 2002). change in roughness due to change in floodplain
Given the importance of safety, proposed measures vegetation then follows the same principle as
need to be evaluated on the impact on flood risk, and estimating the roughness for the unchanged situation.
various decision support systems have been devel- Applying this procedure in evaluating the impact
oped to support this type of evaluation (Silva and van on flood levels due to land use change, however, relies
de Langemheen, 1997; Counsell et al., 2000). In these, on the validity of a reductionist approach in
impacts may be assessed through some form of hydrodynamic modelling as it assumes that the bulk
hydrodynamic modelling, requiring that measures be roughness of the reach is the sum of roughness due to
expressed in terms of changes in model structure and distributed vegetation, and this assumption is at best
parameterisation. While translating measures affect- doubtful (Beven, 2000). Even for finite-element 2D
ing floodplain topography is relatively straightfor- hydrodynamic models, considered as the most com-
ward, translating measures that affect floodplain plex model practicably applicable to dynamic reach
vegetation and therefore roughness is much less scale flood stage prediction, comparison to laboratory
apparent. experiments shows these have shortcomings (Wilson
Much empirical work has been undertaken in et al., 2002), and as a result some calibration and
researching roughness characteristics of floodplain validation of models will always be required before
land use types. Research work has included laboratory these can be reliably used in prediction (Sellin et al.,
experiments using stiff rods to represent vegetation 2003). The result of the calibration is, however, that
(Sellin et al., 1993; Nepf, 1999) or aluminium strips the roughness factors no longer represent roughness
(Sellin et al., 2003), as well as actual vegetation such due to vegetation only, but are bulk representations of
as submerged grass (Wilson and Horritt, 2002), momentum losses not explicitly modelled. This holds
various types of aquatic vegetation (Stephan and particularly for 1D modelling where turbulent
Gutknecht, 2002), and even coniferous trees (Kouwen momentum interchange is not explicitly accounted
and Fathi-Moghadam, 2000). Based on such exper- for (Ackers, 1992), but applies equally to 2D
imental work, analytical and numerical relationships modelling where turbulent closure models are used
have been derived to estimate roughness values for (Wilson et al., 2002).
application in modelling based on the vegetation The inverse problem tackled in calibration shows
present on the floodplain (Kutija and Hong, 1996; that floodplain roughness parameters are difficult to
Klopstra et al., 1997). Simpler methods for evaluating identify given the data typically available at the reach
floodplain roughness values are tables of roughness scale (Khatibi et al., 1997), with roughness values
coefficients for different roughness types (Chow, across a wide range giving equivalent, or near
1959) or through matching physical characteristics equivalent calibration results (Aronica et al., 1998;
to photographs to establish a compound roughness Horritt and Bates, 2001; Aronica et al., 2002). This
coefficient (Acrement and Schneider, 1989). For a paper addresses the problem of identifiability of
complete overview of methods in estimating flood- distributed roughness values. Given a highly detailed
plain roughness, see Sellin et al. (2003). data set available from the 1995 flood event for a
Establishing an effective roughness coefficient at a 30 km of the Meuse river in the Netherlands, as well
reach scale using all these methods requires some as a detailed map of floodplain land use, the utility of
knowledge of the vegetation present on the floodplain, the calibration data in constraining uncertain rough-
and therefore depends on the availability of spatial ness is investigated. Twenty types of land use are
data with sufficient detail, and potential vegetation distinguished in the original land use data, and these
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 141

are aggregated to three levels: (a) a single floodplain between the gauging stations of BorgharenDorp and
land use class, (b) two floodplain land use classes, and Grevenbricht, with the gauging station at Elsloo
(c) five floodplain land use classes. For each of these internal to the modelling domain (see Fig. 1a). This
aggregated classes, the sensitivity of model cali- reach consists of a meandering channel with an
bration to heterogeneity in roughness specification is approximate width of 100 m, and a floodplain of up to
investigated. Finally the utility of the different types 3 km in width, constrained by the high embankments
of calibration data in constraining prior roughness of the Juliana canal to the east and the main river dyke
distributions is explored. running along the BelgianDutch border to the west.
Land use in the reach is primarily agricultural with
some 65% of the area being fields and pastures. An
2. Study site and available data overview of land use in the reach in Table 1 (only land
use types U1% of area are given) shows that the
The reach of the River Meuse considered in this floodplain also contains a number of urban settle-
study is just downstream (North) of the town of ments. The range of roughness values are given as
Maastricht in the Netherlands, covering about 30 km estimated from the tables provided in Chow (1959)

Fig. 1. Meuse river between BorgharenDorp and Grevenbricht.


142 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

Table 1
Land use, roughness values and aggregated classes

Land use type Area (%) Mannings-n (m1/3 sK1) Single class Two classes Five classes
(ids) (ids) (ids)
Min. Max.
Isolated sand/gravel pit 4.69 mc mc mc
Main channel 6.27 0.025 0.06 mc mc mc
Urban area 8.52 fp fp2 fp4
Weeds without structure 1.75 0.035 0.08 fp fp1 fp3
Weeds with structure 4.01 0.035 0.08 fp fp1 fp3
Cultivated fields 34.63 0.025 0.05 fp fp1 fp3
Pasture 24.53 0.025 0.05 fp fp1 fp1
Marginal grassland 5.10 0.025 0.05 fp fp1 fp1
Bare soil 1.58 0.02 0.04 fp fp1 fp1
Hardwood woodland 2.45 0.08 0.2 fp fp2 fp5
Cultivated woodland 3.97 0.08 0.2 fp fp2 fp5

and French (1986). Despite the dominance of pasture, was processed using the snake algorithm described in
the significant areas of urban area, woodland, and Horritt (1999), with a spatial accuracy estimated at
weeds would be expected to add to roughness 1 pixel in the image (w12:5 m). However, due to
characteristics. back scattering of the radar energy from the wind
During the January 1995 flood event, the floodplain roughened water surface as well as from particular
in the reach suffered extensive inundation. An land use types, some areas are misclassified in the
exceptionally complete dataset was collected by the SAR image, and the air photo is considered as more
Dutch Water Authorities during this event, making accurate (Hunter et al., 2003). Both images show
four types of calibration data available, as summarised inundation beyond the main river embankment to the
in Table 2. Fig. 2a shows the levels recorded at the west of Elsloo. This is believed to result from
three gauges, with the peak at BorgharenDorp being saturated areas due to excess rainfall and not from
recorded at 45:71 m on the morning of the 31st of overtopping of the embankment and was not used in
January, where the peak flow was rated as 2746 m3/s, model evaluation.
estimated as having a return period of 63 years. The The final set of data available was from a detailed
event showed the same triple peak seen in many ground survey of maximum flood levels distributed in
adjacent basins and was quite prolonged when the (mainly eastern) floodplain. These point data were
compared to other events in the Meuse such as the
Table 2
slightly larger December 1993 flood event (Wind Summary of observed data sources
et al., 1999). Although the two surveys of flood extent,
one from a low altitude airborne survey on 27 January, Data type Time Description
and one from the ERS-1 satellite overpass on 30 Gauges Hourly Gauged levels at BorgharenDorp,
January at 10:33, were some 70 h apart, due to the Esloo and Grevenbricht
Hourly Discharge at BorgharenDorp (derived
elongated event the hydraulic conditions in the through stage-discharge transformation
different samples are likely to be very similar, with Vector 27-01-1995 Air photo survey. Image mosaic
the discharge at BorgharenDorp estimated at 2645 converted to shoreline by
and 2631 m3/s, respectively. Fig. 1b, however, shows Dutch Water Authorities
30-01-1995 ERS-1 SAR image converted into a
the shorelines indicating flood extent from these two
(10:33) shoreline using statistical active
sources, and considerable classification differences contour model (Horritt, 1999)
can be seen. The estimated spatial accuracy of the Point Maximum Ground survey of maximum water
digitised air photo was of the order of 25 m (Hunter level levels on the floodplain. Thirty-four
points within the model extent
et al., 2003). The flood extent shoreline from the
were used
satellite borne Synthetic Aperture Radar (SAR) image
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 143

(a) (b)
6 3000
Qpeak=2746 m3s 1
Level above local reference (m)

5 2500

Discharge (m3s1)
4 2000

3 1500

ERS1 overpass
2 BorgharenDorp (40m) 1000

Air survey
Elsloo (35m)
Grevenbricht (28m)
1 500

0 0
0 48 96 144 192 240 288 336 384 432 480 0 48 96 144 192 240 288 336 384 432 480
Time (hours) Time (hours)

Fig. 2. (a) Gauged levels at BorgharenDorp, Elsloo and Grevenbricht. Levels are given to local reference to increase readability. (b) Discharges
derived for the BorgharenDorp gauge. The time of overpass for the air survey and ERS-1 satellite are indicated.

considered to be of great value in giving a detailed channel computational points through mutual
view on hydraulic variability in the floodplain. Thirty- volumes of the connected cell and the adjoining
four points were available in the model extent from channel section(s) as shows in Eq. (1)
BorgharenDorp and Grevenbricht, and the surveyed
dDxDyhi;j
levels at these points are considered to be of good C Dyuhi;j K uhiK1;j C Dxvhi;j
accuracy. dt
Cross-section data for the reach were available at K vhi;jK1 C Qk K QkK1 Z 0 1
approximately 500 m intervals, as well as a detailed where (uh)iK1,j is the x unit discharge at the interface
digital elevation model of the floodplain for the reach. between cells iK1 and i and QkK1 is the discharge
The digital data included surveyed heights of major from main channel reach segments kK1 to k. The
embankments such as dykes and roads in the area. level h is the depth in cell i, j.
Fig. 3 gives an impression of the connection
between the 1D and the 2D domain. The division
3. Approach between the two domains follows the principle of the
horizontal divided channel method except that here
3.1. Model description the division between main channel and floodplain is
not a zero shear interface. This accounts for turbulent
A hybrid 1D2D hydrodynamic code was applied momentum interaction between main channel and
to the reach between BorgharenDorp and Greven- overbank flow, and the roughness value at this non-
bricht. The model resolves overbank flow in a 2D zero shear interface is defined here as that applied as
raster domain that allows easy integration with main channel bed roughness.
available GIS formats common for digital elevation Using this hybrid code, a model for the reach
models. Flow is resolved through solving the full between BorgharenDorp and Grevenbricht was
shallow water equations in two dimensions in the 2D developed, using discharges observed at Borgharen
domain and one dimension in the 1D domain. Dorp as the upstream boundary, and recorded water
Numerically, a staggered grid is used, with fluxes levels at Grevenbricht as the downstream boundary.
defined across grid cell interfaces and water levels at Cross-sections for application in the 1D main channel
each grid centre point. 1D main channel flow is were derived as level-width tables giving an average
equally resolved using a staggered grid, and the two geometry for river sections some 550 m in length (see
domains are coupled at grid cells overlying main Werner et al., 2000 for details). Digital terrain models
144 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

Fig. 3. Impression of connection between 1D and 2D domains (after Frank et al., 2001). Water level in grid cells i, j are indicated by z. u and v
are fluxes across cell boundaries. The main channel is shown as a grey line with black circles indicating 1D calculation points. These share the
water level z of the cells they fall in.

with a grid resolutions of 25, 50, and 100 m were individually. The last three columns in Table 1
aggregated from high-resolution laser altimetry data. indicate how the land use types were aggregated.
It was found that performance of models with the Prior ranges of Mannings-n roughness were selected
100 m grid deteriorated significantly when compared for each aggregate class based on the constituent land
to the 25 m grid, while this was much less so for the use types as given in Chow (1959) and Table 3, with
50 m grid. The same result was found for this reach by some relaxation on the limits of the main channel
Bates and de Roo (2000), and the 50 m resolution was roughness as good model performance was found at
used throughout. In aggregating the data, special values tending towards the suggested minimum.
attention was given to the various line elements in the
floodplain, particularly the high embankments to the
3.2. Evaluating model performance
east and west. Point level data at regular intervals
from cadastral surveys were available for the tops of
The four datasets available for evaluating model
these embankments (roads and dykes), and as these
performance are of two distinct types. The flood
were considered to be of a higher accuracy than the
extent map data from the air survey and SAR image
laser altimetry data, the levels along the embankment
are essentially spatially distributed binary classifi-
lines were enforced in the grid.
cation data (Aronica et al., 2002). Comparison of
Roughness in the 2D floodplain domain can be
modelled flood extent data against the observed flood
applied to the same resolution as the digital terrain
extent is achieved by comparing the area correctly
model itself, thus allowing easy change between
classified as wet or dry against the area incorrectly
different levels of floodplain roughness heterogeneity.
classified. Here the Fh2i statistic previously used by
Three levels of disaggregation were considered in
Aronica et al. (2002) is used
using the distributed land use data (see Fig. 1). In the
first level, a single, uniform floodplain roughness was Aobs h Amod
considered. For the second level, the floodplain F h2i Z (2)
Aobs g Amod
roughness was considered in two main groups: (i)
high roughness stiff land use types (urban and where Aobs is the observed area of inundation and
woodlands) and (ii) lower roughness flexible land Amod the modelled area of inundation. In effect, the
use types (cultivated fields, pastures, grasslands, and statistic determines the ratio between the number of
weeds). In the final level of disaggregation, the main grid cells classified correctly as being either dry or wet
land use types in the reach were considered and the number of cells classified incorrectly. Eq. (2)
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 145

Table 3
Range of Mannings-n roughness values (in m1/3 sK1) for each aggregate land use class

Class Single class Two classes Five classes


Main channel 0.0160.040 (0.022) 12% 0.0160.040 (0.022) 12% 0.0160.040 (0.022) 12%
Floodplain 1 0.0200.200 (0.040) 88% 0.0200.080 (0.029) 72% 0.0250.050 (0.030) 32%
Floodplain 2 0.0500.200 (0.140) 16% 0.0200.050 (0.030) 34%
Floodplain 3 0.0350.080 (0.040) 6%
Floodplain 4 0.0500.200 (0.100) 9%
Floodplain 5 0.0500.200 (0.100) 7%
Number runs 150 250 500

Bracketed values indicate optimal parameter values. Percent area of each class is indicated in italics.

can then be reformulated as Likelihood Estimator (HMLE) measure (Sorooshian


Pn D1M1 and Dracup, 1980; Gupta et al., 1998) was used in
iZ1 Pi
F h2i Z Pn P n P comparing observed and calculated levels at gauges
D1M1 C
iZ1 Pi iZ1 Pi
D1M0
C niZ1 CPD0M1
i
Pn
(3) 1 ^
tZ1 wt ht q; Y K ht
2
HMLE Z n  Qn 1=n (5)
where PD1M1
i obtains a value of 1 for cells classified as tZ1 wt
inundated both in observed and modelled data, PD0M1 i
obtains a value of 1 for cells observed as dry but where ht is the observed level at time t and h^t q; Y the
predicted to be wet, and PD1M0 obtains a value of calculated level for given parameters q and data Y.
i
1 cells observed as wet but classified as dry. The weights wt assigned to time t are computed as
Following Hunter et al. (2003), the measure is wt Z ht2lK1 , where l is an unknown transformation
amended to additionally penalise over-prediction of factor to stabilise the variance. l is estimated
the flood extent: iteratively for each observed and calculated series
Pn P compared (see Sorooshian et al., 1993 for a complete
h2i
D1M1
iZ1 Pi K niZ1 PD0M1
i description of the procedure used here). The HMLE
F Z Pn D1M1 C
Pn D1M0
P n D0M1
(4)
iZ1 Pi iZ1 Pi iZ1 Pi measure was derived primarily for comparing com-
puted against observed discharge series, and the more
Comparison of flood extent with the air photo
or less constant bias of the level series with reference
was restricted to the reach for which it is available
to the absolute datum was found to make the
(BorgharenDorp to Elsloo), while for the SAR
estimation procedure of l instable. To increase
image comparison was limited to upstream of the
stability, both observed and calculated levels were
narrow floodplain section directly between Greven-
first transformed to a local datum at each gauge. This
bricht and Elsloo to minimise the effects of the
transformation was found to have negligible effect on
imposed down stream boundary levels on the
the resulting HMLE values.
results.
The distributed floodplain observation points are
Attenuation in the reach during the 1995 event is
compared using the Mean Absolute Error measure
limited due to the elongated nature of the hydrograph,
and differences in calculated levels at these gauges Pm
jh^ q; Y K hk j
between model runs are primarily in the form of a MAE Z kZ1 k (6)
n
negative or positive bias. Errors in the time series are
therefore highly correlated with time, making some of where h^k q; Y is the calculated level at point k, hk is
the more widely used objective functions such as the the observed level and m is the number of points used.
NashSutcliffe model efficiency unsuitable as these Similar to model comparison using the SAR image,
assume errors are normally distributed with zero mean points directly upstream of Grevenbricht were not
and constant variance (Beven and Freer, 2001). To included in the comparison as these were again
account for autocorrelation of errors and possible considered too close to the downstream boundary,
changing variance the Heteroscedastic Maximum resulting in a total of 31 points being used in
146 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

the comparison. For lower roughness parameter sets, parameters. It also provides useful information on the
model results did not predict flooding at all of these value of the different types of calibration data in
points. Where this occurred the level of the DEM cell constraining these uncertainties (Freer et al., 1996). In
in which the point falls is used in the comparison. this Monte-Carlo based technique, an ensemble of
Although this imposes an upper bound on the MAE parameter sets is sampled from prior distributions of
measure, it implicitly allows for discrepancies the constituent parameters and the model is run with
between the DEM and the actual terrain (Hunter each set. The acceptability of each run is assessed by
et al., 2003). comparing the predicted with the observed values
using a chosen likelihood measure (Aronica et al.,
3.3. Sensitivity analysis 2002), where the likelihood measure is obtained here
using one or more of the selected objective functions.
Sensitivity analysis is a widely used tool in Likelihood measures derived using different objective
modelling to determine how the variation in model functions can be easily combined analogous to
performance can be apportioned to different par- combining likelihood values from different data sets
ameters (Crosetto et al., 2000). Here a univariate through application of Bayesian updating (Lamb
sensitivity analysis was applied, where response of et al., 1998; Franks et al., 1998)
model performance due to variation in parameter
L Yjqi L0 qi
values was determined for parameters individually, Lqi jY Z PN 1 (7)
and interaction between parameters was ignored. jZ1 L1 Yjqj L0 qj
For each of the three levels of floodplain roughness where L0(qi) is the prior likelihood measure for
aggregation, an optimal set of parameters was parameter set qi, L1(Yjqi) is the likelihood measure
selected such that good performance in terms of the derived by comparing model output using parameter
selected objective functions was found. Where set q to observations Y using a selected objective
different objective functions gave good results in function, L1(qijY) is the posterior likelihood of
different parts of the parameter space, the parameter parameter set qi given observations Y, and N is the
set was selected with a preference of good perform- number of parameter sets. Optionally, a behavioural
ance in comparing levels predicted at the gauges, as threshold can be applied to divide runs with
the gauged levels are considered of higher reliability acceptable performance from those with unacceptable
than the two flood extent observations. The values of performance (Beven and Binley, 1992). Parameter
these optimal parameter sets are given in brackets in sets that perform below the threshold in the selected
Table 3. In the subsequent sensitivity analysis, one objective function are then given a likelihood of 0 and
parameter was varied across its range in 1015 likelihood values for the remaining sets are rescaled.
increments, while the remaining parameters were
kept at the optimal values. For each variation, the
model was run and resulting performance against the
four types of calibration data evaluated. 4. Results

3.4. Uncertainty analysis 4.1. Sensitivity analysis

To elaborate on the sensitivity analysis where the Fig. 4 shows the values of the HMLE, MAE, and
influence of individual parameters on model perform- Fh2i measures for sensitivity tests for the main channel
ance was explored, an uncertainty analysis was and a single floodplain roughness class (see Table 3
undertaken. The uncertainty analysis is set within for parameter ranges and default parameters). Com-
the context of the generalised likelihood uncertainty parison of observed to calculated levels at the gauges
estimation method (GLUE, see Beven and Binley, of BorgharenDorp and Elsloo using the HMLE
1992). This helps to establish the effect uncertainty in measure shows clear sensitivity to both main channel
individual parameters has on uncertainty of model and floodplain roughness. Values for the gauge at
predictions, while considering interaction between Elsloo are more sensitive to main channel roughness
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 147

(a) (b)
1.0 1.0

0.9 0.0 0.9 0.0

HMLE () / MAE (m)


HMLE() / MAE (m)
0.8 0.2 0.8 0.2

()
()

0.7 0.4 0.7 0.4

<2>
<2>

0.6 0.6 0.6 0.6

F
F

0.5 0.8 0.5 0.8

0.4 1.0 0.4 1.0

0.3 1.2 0.3 1.2


0.016 0.020 0.024 0.028 0.032 0.036 0.040 0.02 0.05 0.08 0.11 0.14 0.17 0.20
nmc nfp

Fig. 4. Sensitivity analysis for single floodplain roughness class. (a) Main channel roughness. (b) Floodplain roughness. C and ! are the HMLE
values for BorgharenDorp and Elsloo, respectively, C are the MAE results for the distributed floodplain points, B and $ are the Fh2i results
for comparison with the SAR image and the Air survey, respectively.

than BorgharenDorp, while for floodplain roughness somewhat erratic for the SAR image because of the
this is reversed. This different response is due to the patchy nature of the SAR shoreline towards the
location of each of the gauges. Elsloo is upstream of a downstream end of the reach (Fig. 1) As dyke
large bend in the river with the floodplain narrowing, thresholds overtop, some of these patches are
making main channel flow more dominant, while predicted as being inundated, therefore increasing
BorgharenDorp is upstream of a relatively wide performance. A further increase leads again to over-
floodplain, with floodplain flow gaining in import- prediction and a drop in performance.
ance. The sensitivity of the distributed floodplain Increasing the disaggregation of roughness values
observations (MAE measure) again shows a very to two classes gave a very similar response for
similar pattern. The best simulations from all three sensitivity to main channel roughness and is not
gauge level comparisons are in the same region of the shown here. Sensitivity to the first floodplain rough-
roughness ranges. ness class (dominant pastures and fields) is lower
A similar pattern is shown in the response of the when compared to the single floodplain class, but
binary flood map comparison expressed by the Fh2i zooming-in in Fig. 4b using the same floodplain
measure. Even at low roughness values, the floodplain roughness range as in Fig. 5a reveals the response to
is filled completely, with flood extent bounded by the be very similar. Sensitivity to the second roughness
steep walls of the flood defence dykes to the west and class (Fig. 5b) is absent, with performance for all
the canal to the east. This is particularly so for the measures being the same for all roughness values. A
reach upstream of Elsloo, and as the air photo covers further disaggregation to five floodplain roughness
only this reach, the Fh2i measure remains fairly classes results in negligible sensitivity in any of the
constant with increasing roughness. For comparison roughness classes within the ranges specified for each
with the SAR image performance improves with class (Fig. 6).
increasing roughness to show an optimum in the same
region as the comparisons for the level data. Both 4.2. GLUE analysis
measures start to decline, however, with a further
increase in roughness. This decline is easily explained In the GLUE analysis, parameter sets were
when the individual flood maps are studied. These sampled from the prior distributions. The sample
show that as the levels increase, dykes in some sizes were set to 100 runs for models using a single
sections start to overtop, and as the Fh2i measure floodplain roughness class, 250 runs for models using
introduced in Eq. (4) penalises for over-prediction, two floodplain roughness classes, and 500 runs for
this causes the measure to drop. This decline is models with five floodplain roughness classes.
148 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

(a) (b)
1.0 1.0
0.9 0.0 0.9 0.0

HMLE () / MAE (m)

HMLE () / MAE (m)


0.8 0.2 0.8 0.2
F<2> ()

F<2> ()
0.7 0.4 0.7 0.4
0.6 0.6 0.6 0.6
0.5 0.8 0.5 0.8
0.4 1.0 0.4 1.0
0.3 1.2 0.3 1.2
0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.05 0.08 0.11 0.14 0.17 0.20
nfp Class 1 nfp Class 2

Fig. 5. Sensitivity analysis for two floodplain roughness classes. See Fig. 4 for legend.

For each of the parameter sets sampled, likelihood dotty plots in Figs. 79. These plots show the
values were established using all four of the available likelihood value of each individual parameter set in
calibration data types. This was through successive the (multi-dimensional) parameter response surfaces
application of Eq. (6), with the sequence being (i) as projected onto a single parameter axes, and
comparison of levels at gauges, (ii) comparison of although these reveal little explicit detail of parameter
flood extent to airphoto, (iii) comparison of flood interaction (Blazkova et al., 2002), they do show the
extent to SAR image, and finally (iv) comparison of spread of fit and resulting likelihood value across the
levels at distributed floodplain points. Resulting parameter range. The vertical spread of likelihood
likelihood values for each parameter set are given in values can also be used to infer parameter interaction,

(a) (b)
HMLE/MAE

HMLE/MAE
0.9 0.0 0.9 0.0
F<2> ()

F<2> ()

0.8 0.1 0.8 0.1


0.7 0.2 0.7 0.2
0.6 0.3 0.6 0.3
0.5 0.4 0.5 0.4
0.02 0.03 0.04 0.05 0.02 0.03 0.04 0.05
nfp Class 1 nfp Class 2

(c) (d)
HMLE/MAE

0.9 0.0 0.9 0.0 HMLE/MAE


F<2> ()

F<2> ()

0.8 0.1 0.8 0.1


0.7 0.2 0.7 0.2
0.6 0.3 0.6 0.3
0.5 0.4 0.5 0.4
0.035 0.050 0.065 0.080 0.050 0.080 0.110 0.140 0.170 0.200
nfp Class 3 nfp Class 4

(e)
F<2> Air survey
HMLE/MAE

0.9 0.0
F<2> SAR image
F<2> ()

0.8 0.1
0.7 0.2 MAE floodplain points
0.6 0.3 HMLE Borgharen Dorpa
0.5 0.4 HMLE Elsloo
0.050 0.080 0.110 0.140 0.170 0.200
nfp Class 5

Fig. 6. Sensitivity analysis for five floodplain roughness classes.


M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 149

(a) (b)
2.5 2.5

102 ()

102 ()
2.0 2.0

1.5 1.5
likelihood

likelihood
1.0 1.0

0.5 0.5

0.0 0.0
0.016 0.024 0.032 0.040 0.02 0.08 0.14 0.20
nmc (m1/3s1) nfp (m1/3s1 )

Fig. 7. Dotty plots for single floodplain roughness class. (a) Main channel, (b) floodplain.

as is clearly seen in Figs. 7 and 8. The main channel sensitivity of model performance to these parameters.
roughness shows high likelihood values in the same Figs. 1012 show that for a single floodplain rough-
region as would be expected from the sensitivity ness class, the marginal posterior distribution is well
analysis. However, some runs with the same main constrained when compared to the uniform prior
channel roughness show low likelihoods indicating (horizontal line) for both main channel and floodplain
significant interaction with the floodplain roughness. roughness. This holds equally for the main channel
For very low main channel roughnesses this inter- roughness with two floodplain roughness classes, but
action is smaller, which is intuitively correct as flow the posterior distribution of the first roughness class is
would be more in-bank for these low roughness constrained only slightly, while for the second class
values. no improvement on the prior is made. With five
As the number of floodplain roughness classes floodplain roughness classes, some improvement on
increases, the interaction between main channel the prior range is shown only for the first floodplain
roughness and floodplain roughness seems to be less roughness class, confirming the results of the
(Figs. 8 and 9). With the increase in floodplain sensitivity analysis.
roughness classes, good model fits can be seen across To explore the consequence of introducing a
the prior parameter ranges, reflecting the decrease in behavioural threshold on posterior roughness

(a) (b) (c)


1.5 1.5 1.5
102 ()

102 ()

102 ()

1.0 1.0 1.0


likelihood

likelihood

likelihood

0.5 0.5 0.5

0.0 0.0 0.0


0.016 0.024 0.032 0.040 0.02 0.04 0.06 0.08 0.05 0.10 0.15 0.20
nmc (m1/3s 1) nfp class 1 (m1/3s 1) nfp class 2 (m1/3s 1)

Fig. 8. Dotty plots for two floodplain roughness classes. (a) Main channel, (b and c) floodplain.
150 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

(a) (b) (c)


0.8 0.8 0.8
()

102 ()

()
0.6 0.6 0.6
2

2
10

10
0.4 0.4 0.4
likelihood

likelihood

likelihood
0.2 0.2 0.2

0.0 0.0 0.0


0.016 0.024 0.032 0.040 0.02 0.03 0.04 0.05 0.02 0.03 0.04 0.05
1/3 1
nmc (m s ) nfp class 1 (m1/3s1) n class 2 (m1/3s1)
fp

(d) (e) (f)


0.8 0.8 0.8
102 ()

()

()
0.6 0.6 0.6
2

2
10

10
0.4 0.4 0.4
likelihood

likelihood

likelihood
0.2 0.2 0.2

0.0 0.0 0.0


0.035 0.050 0.065 0.080 0.05 0.10 0.15 0.20 0.05 0.10 0.15 0.20
n class 3 (m1/3s1)
1/3 1 1/3 1
n class 4 (m s ) n class 4 (m s )
fp fp fp

Fig. 9. Dotty plots for five floodplain roughness classes. (a) Main channel, (bf) floodplain.

distributions (Beven and Binley, 1992), the top would more appropriately be defined on the basis of
performing 20% runs for each level of roughness explicit criteria in view of the models application, the
aggregation are selected and the remaining 80% selection serves for the purpose of illustration. Fig. 13
discarded. Although this threshold is arbitrary, and shows the full range of values for each parameter in

(a) (b)
0.25 0.25

0.20 0.20
likelihood ()

likelihood ()

0.15 0.15

0.10 0.10

0.05 0.05

0.00 0.00
0.016 0.024 0.032 0.040 0.02 0.08 0.14 0.20
nmc (m1/3s1) nfp (m1/3s1)

Fig. 10. Posterior parameter distributions for single floodplain roughness class. (a) Main channel, (b) floodplain.
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 151

(a) (b) (c)


0.25 0.25 0.25

0.20 0.20 0.20


likelihood ()

likelihood ()

likelihood ()
0.15 0.15 0.15

0.10 0.10 0.10

0.05 0.05 0.05

0.00 0.00 0.00


0.016 0.024 0.032 0.040 0.02 0.04 0.06 0.08 0.05 0.10 0.15 0.20
nmc (m1/3s1) nfp class 1(m1/3s1) nfp class 2 (m1/3s1)

Fig. 11. Posterior parameter distributions for two floodplain roughness classes. (a) Main channel, (b and c) floodplain.

the preserved parameter sets, when compared to the for the first roughness with two floodplain roughness
prior range. This again shows that the ranges are classes, while for the remaining floodplain roughness
constrained considerably for the main channel and classes, parameter sets with values almost across the
perhaps the first floodplain roughness class. This holds full prior range are retained.
152 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

(a) (b) (c)


100 100 100 0.080 0.200 0.200
prior parameter range (%)

prior parameter range (%)

prior parameter range (%)


0.192 0.048

80 80 80 0.044

60 60 60
0.028 0.051
40 40 40
0.082 0.024
0.023
20 20 0.019 20 0.020
0.018 0.063
0.022 0.020 0.021 0.020 0.036 0.052 0.052
0 0 0
mc fp mc fp1 fp2 mc fp1 fp2 fp3 fp4 fp5

Fig. 13. Constrained parameter ranges after applying threshold. (a) Single, (b) two, and (c) three floodplain roughness class(es).

5. Discussion the resulting values of H, and although the absolute


changes are modest, relative changes show the value
5.1. Impact of floodplain roughness aggregation of the gauge data and the distributed floodplain points
on model performance to be greatest. The relative reduction of entropy using
all the data is 90.7, 92.2 and 92.5% for a single, two
For the four different types of observations used in and five floodplain roughness classes, respectively,
constraining roughness distributions in both main but the differences are too small to suggest an increase
channel and floodplain, comparison of the best of total overall uncertainty with increasingly complex
performing parameter set of those sampled shows parameterisation.
that for none of these, a clear improvement in Reduction of the Shannons entropy index for
performance is achieved with increasing floodplain comparison of the flood maps is found to be very
roughness disaggregation. This would suggest for the sensitive to how the binary comparison is evaluated.
model used to simulate the given event in this reach The values in Table 4 have been derived using the
even the comprehensive data set available here still formulation as in Eq. (4), where over-prediction of
does not provide sufficient detail on the distributed flood extent is penalised (Hunter et al., 2003). If the
nature of floodplain flow, and can only be used as a original formulation of Aronica et al. (2002) is used
descriptor of bulk flow at the reach scale. (Eq. (3)) then discrimination between runs is much
To evaluate the value of updating likelihood less pronounced and as a consequence reduction in
weights using each of the available data sets, the Shannons entropy is minimal. Amending the Fh2i
Shannon entropy measure H (Beven and Binley, measure to additionally penalise under-prediction by
1992; Franks et al., 1998) was calculated, where a adding PD1M0 to the numerator of Eq. (4) results in
i
lower entropy value indicates more structure and
runs where the area behind the dykes is inundated
therefore less uncertainty
being assigned higher likelihood, while for low
X
H ZK Lqi log2 Lqi (8) Table 4
i Shannon entropy measure
where i is the model realisation index. The value of H Likelihoods after H
is the result a discrete summation, and the number of updating using data Single class Two classes Five classes
realisations must be taken into account when type
comparing entropy values between simulations, mak- Uniform prior 7.229 7.966 8.966
ing such comparison difficult. The entropy value is Gauges 6.936 7.722 8.709
again used to compare the value of calibration data SAR image 6.860 7.646 8.633
Air survey 6.787 7.574 8.570
sets in lowering entropy values for simulations with
Floodplain points 6.556 7.346 8.294
the same number of model realisations. Table 4 shows
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 153

roughness parameter sets where sections of the For the code used here, sensitivity to floodplain
floodplain between the dykes remain dry obtain low roughness for the single floodplain class is very
likelihood values. As a consequence, good performing apparent, but this sensitivity is dominated by the
parameter sets for this measure shift to the rougher selection of the prior roughness range. This has been
part of the friction parameter space, resulting in selected to be quite wide to reflect the full range of the
Shannons entropy value increasing rather than constituent land use types. However, the posterior
decreasing on updating likelihoods using the flood- range of roughness values showing good results
maps after the gauge data. This confirms that optima complies more or less with the prior that would be
for different objective functions may result in different selected had only dominant land use type being
parameter optima (Gupta et al., 1998), and illustrates considered. Establishing a more precise roughness
the impact of selecting even small variations in value even for these dominant land use types than the
objective function on calibration results. suggested range given in tables from literature is,
however, shown to be difficult. The correlation
5.2. Identifiability of distributed floodplain roughness between roughness values of each roughness class
and resulting parameter set likelihood values is given
Both the sensitivity analysis and the results of the in Table 5. As would be expected, the main channel
GLUE method show the identifiability of distributed roughness parameter shows good correlation with the
floodplain roughness to be relatively low. In a number likelihood values for all cases. Although low, some
of other studies, where a single floodplain roughness correlation is found for the dominant land use types at
is used, low sensitivity to floodplain roughness values higher levels of heterogeneity, but what is surprising
has been shown (Romanowicz et al., 1996; Horritt and is the low influence of the roughness class in which
Bates, 2001), but this is not a conclusion universally the urban areas are incorporated (last class for both
applicable. More significant sensitivity to floodplain two and five floodplain roughness classes). Some 9%
roughness for a similar reach has been shown (Horritt, of the reach is urban, and although these are slightly
2000), as well as in more two-dimensional polder type elevated when compared to the surrounding flood-
inundation events (Aronica et al., 2002; Hesselink plain, the position with respect to the flow direction
et al., 2003). Floodplain friction sensitivity would be (see Fig. 1) would be expected to influence results
expected to be linearly proportional to the area more, but the results imply the influence to be too little
inundated and non-linearly proportional to the to have distinctive influence on the bulk roughness
velocity (cDxv2), therefore having a strong relation- characteristics of the whole reach.
ship with how dominant floodplain flow is for the The dotty plots in Figs. 79 show that the effective
reach. The complex topography of natural floodplains main channel roughness becomes increasingly well
will, therefore, have a dominant effect on the defined as the number of floodplain roughness classes
contribution of the floodplain to conveyance of flow increases. This is in contrast to the apparent
as opposed to a storage element with impact only on identifiability of the floodplain roughness values
flood wave attenuation. This topography also strongly with the increase in heterogeneity, which at five
influences sensitivity to different likelihood measures, floodplain roughness classes show almost similar
as seen in the case here where the embankments performance across the full range of prior roughness
strongly influence flood extent, with resulting sensi-
tivity to comparison of flood extent maps being low. Table 5
Besides the characteristics of the reach, however, the Correlation between parameter values and likelihoods
model code used also has an influence on the Parameter Single class Two classes Five classes
sensitivity. Cell-inundation type codes show lower nmc 0.660 0.720 0.724
sensitivity than finite-element codes (Horritt and nfp class 1 0.537 0.297 0.140
Bates, 2001). Indeed a model of the same reach of nfp class 2 0.058 0.109
the Meuse river with a cell-inundation code shows nfp class 3 0.077
nfp class 4 0.032
very low sensitivity to floodplain roughness (Hunter
nfp class 5 0.011
et al., 2003).
154 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

values. Although this effect is counter-intuitive, it 2D depth averaged code with explicit accounting of
may be explained by looking closely at the ranges of turbulence. A comparison of depth averaged vel-
the roughness classes in relation to how dominant the ocities across the floodplain in such a code to
particular class is in determining bulk roughness laboratory observations by Wilson et al. (2002),
values at the reach scale. In the case of only one however, shows this may be poor even in such
roughness value in the floodplain, the prior range is complex codes, and calibrated roughness factors
chosen to be relatively wide. The constrained remain bulk representations of momentum
distribution shows a tendency to the lower roughness dissipation.
values being in the range that would have been
selected for the dominant land use types. The spread
in the main channel roughness can be explained by 6. Conclusions
effective main channel roughness values compensat-
ing for badly selected floodplain roughness values, This paper shows that despite a very comprehen-
and still giving reasonable performance. In the case of sive calibration dataset being available, the utility of
five floodplain roughness classes, the ranges of this data in constraining distributed floodplain rough-
roughness values for the dominant land use classes ness parameters is relatively low. Although the
are much smaller. Within these ranges, there is too sensitivity to floodplain roughness is shown to be
little information content in the calibration data to reasonable when only a single floodplain roughness
constrain the values further than the prior range, but class is used, the constrained range of values giving
there is little room left for compensating ill chosen good model performance is found to be in the order of
roughness values on the floodplain with channel the range of roughness values for the dominant land
parameters, therefore leaving the effective main use type as given in tables from literature (cf. Chow,
channel roughness values better defined. 1959). Applying the values from literature to
The cell-inundation code used by Hunter et al. distributed land use types neither increases model
(2003) for the same reach also shows posterior main performance, nor does the calibration data constrain
channel roughness values to be higher than those the suggested ranges. A positive view of this result
found here, while resulting model performance is would be that to calibrate models for floodplain
comparable to that found here. In that code, however, inundation such as used, here attention need to be
no account is taken of turbulent momentum exchange given only to the main channel roughness values,
between main channel and floodplain. The code used while those on the floodplain can be set somewhere in
here also does not take this explicitly into account, but the suggested range. Evaluating the effect of flood-
the non-zero shear interface between main channel plain land use change could then follow the same
and floodplain acts as a proxy turbulence model. The procedure, while keeping the main channel roughness
comparable performance confirms that calibrated to the value calibrated. The interaction between main
roughness factors describe not only momentum loss channel and floodplain as seen in the reach studied
due to resistance of flow to vegetation, but also here, however, suggests that independent treatment is
incorporate all other turbulent momentum losses not flawed.
explicitly described. As such, roughness factors The problem of this limited ability to identify more
cannot be seen as being independent of the layout precise distributed roughness values stems from these
and sinuosity of floodplain and main channel, the roughness values being more than simply the effect of
position of land use types with respect to floodplain floodplain vegetation on flow, but the bulk effect of
flow, and the resolution with which the complex turbulent momentum interaction between main chan-
topography is represented. Although the model used nel and floodplain, discrepancies in elevation data,
here solves the full shallow water equations, assump- crossing flows, and all other momentum losses not
tions made in modelling the turbulent interaction explicitly accounted for in the model code. The
between floodplain and main channel impact the relative contribution of these and vegetation rough-
relative roughness sensitivities, and the results found ness to the bulk roughness values is neither clear nor is
would need to confirmed through application of a full the combination linear. To compound the problem,
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 155

effective roughness values for the individual flood- and Sustainability, European Union Joint Research
plain land use type may be influenced by backwater Centre (Ispra, Italy) is thanked for giving kind
effects from adjacent, rougher land use types. The permission to use the SAR Image.
calibrated values of main channel roughness are,
moreover, not independent of the roughness values in
the floodplain. The balance of these contributions may References
also shift if one contributor is changed, and as such
evaluating the effect of land use change by simply Ackers, P., 1992. Hydraulic design of two-stage channels.
amending the roughness due to vegetation is not as Proceedings of the Institution of Civil Engineers: Water,
straightforward as it would seem. In this light the Maritime and Energy 96, 247257.
Acrement, Jr., G., Schneider, V., 1989. Guide for selecting
utility of complex formulations of roughness with mannings roughness coefficients for natural channel and
parameterised vegetation characteristics (cf. Klopstra floodplains. Technical Report WSP2339, USGS.
et al., 1997) would seem limited, and estimations Aronica, G., Hankin, B., Beven, K., 1998. Uncertainty and
using the well established tables would suffice. equifinality in calibrating distributed roughness coefficients in
The low sensitivity to roughness values for land a flood propagation model with limited data. Advances in Water
Resources 22 (4), 349365.
use types other than the dominant land use would Aronica, G., Bates, P., Horritt, M., 2002. Assessing the uncertainty
suggest that changing these will have little impact on in distributed model predictions using observed binary pattern
the bulk roughness characteristics of the reach, given information within GLUE. Hydrological Processes 16,
the uncertainty in the value of the roughness for the 20012016.
dominant land use type. Changing the dominant land Asselman, N., van Wijngaarden, M., 2002. Development and
application of a 1D floodplain sedimentation model for
use type would on the other hand lead to a change in the River Rhine in the Netherlands. Journal of Hydrology 268,
bulk roughness characteristics, but this change may 127142.
not be due to change in roughness of the individual Asselman, N., Middelkoop, H., Straatsma, M.R.M., Jesse, E.V.P.,
land use type alone. It would rather reflect a change 2002. Assessment of the hydraulic roughness of river flood
in the balance of roughness values across different plains using laser altimetry. In: Dyer, F., Thoms, M., Olley, J.
(Eds.), The Structure, Function and Management Implications
land use types, resulting in a different pattern of of Fluvial Sedimentary Systems IAHS Publication, vol. 276.
floodplain storage, conveyance and backwater International Commission On Continental Erosion and Unesco,
effects. Alice Springs, pp. 381388.
To substantiate the results found for the reach Bates, P., de Roo, A., 2000. A simple raster-based model for flood
studied here with the selected model code, these inundation simulation. Journal of Hydrology 236, 5477.
Beven, K., 2000. Uniqueness of place and process representations in
conclusions will need to be assessed with a full 2D hydrological modelling. Hydrology and Earth System Sciences
model code including explicit turbulence modelling, 4 (2), 203213.
despite the added computational burden this will Beven, K., Binley, A., 1992. The future of distributed models:
bring. The results do suggest that evaluation in a Model calibration and uncertainty prediction. Hydrological
probabilistic setting would help, but the problem of Processes 6, 279298.
Beven, K., Freer, J., 2001. Equifinality, data assimilation, and
independence of roughness values for distributed land uncertainty estimation in mechanistic modelling of complex
use types need to be tackled before these can be environmental systems using the glue methodology. Journal of
independently changed. Hydrology 249, 1129.
Blazkova, S., Beven, K., Kulasova, A., 2002. On constraining
TOPMODEL hydrograph simulations using partial saturated
area information. Hydrological Processes 16, 441458.
Acknowledgements Chow, V., 1959. Open Channel Hydraulics. McGraw-Hill, New
York.
The authors wish to thank the Dutch Ministry of Cobby, D., Mason, D., Horritt, M., Bates, P., 2003. Two-
Public Works and Water Resources; Directie Limburg dimensional hydraulic flood modelling using a finite-element
and the Institute for Inland Water Management and mesh decomposed according to vegetation and topographic
features derived from airborne scanning laser altimetry.
Waste Water Treatment (RIZA), for providing both Hydrological Processes 17, 19792000.
the hydrological and topographic data, as well as the Counsell, C., Gammal, E.E., Samuels, P., 2000. Demonstration of a
air photo. Ad de Roo of the Institute of Environment decision support system for assessing the impacts of engineering
156 M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157

works and environmental change on flood riskthe Thames Klopstra, D., Barneveld, H., J.M. van Noortwijk, J., van Velzen, E.,
catchment study. In: Bronstert, A., Bismuth, C., Menzel, L. 1997. Analytical model for hydraulic roughness of submerged
(Eds.), European Conference on Advances in Flood Research. vegetation. In: Proceedings of the 27th IAHR Congress. IAHR,
Potsdam, Germany, pp. 8798. San Francisco, USA, pp. 775780.
Crosetto, M., Tarantola, S., Saltelli, A., 2000. Sensitivity and Kouwen, N., Fathi-Moghadam, M., 2000. Friction factors for
uncertainty analysis in spatial modelling based on gis. coniferous trees along rivers. Journal of Hydraulic Engineering
Agriculture, Ecosystems and Environment 81, 7179. 126 (10), 732740.
Duel, H., Baptist, M., Geerling, G., Smits, A., van Alphen, J., Kutija, V., Hong, H., 1996. A numerical model for assessing the
2002. Cyclic floodplain rejuvenation as a strategy for both additional resistance to flow induced by flexible vegetation.
flood protection and enhancement of the biodiversity of the Journal of Hydraulics Research 34 (1), 99114.
River Rhine. In: King et al. (Ed.), Proceedings of the Lamb, R., Beven, K.S.M., Myrab, S., 1998. Use of spatially
Conference on Environmental Flows and Ecohydraulics. distributed water table observations to constrain uncertainty in a
IAHR, Cape Town. rainfall-runoff model. Advances in Water Resources 22 (4),
Frank, E., Ostan, A., Coccato, M., Stelling, G., 2001. Use of an 305317.
integrated one dimensional two dimensional hydraulic model- Mason, D., Cobby, D.M., Horritt, M., Bates, P., 2003. Floodplain
ling approach for flood hazard and risk mapping. In: friction parameterization in two-dimensional river flood models
Falconer, R., Blain, W. (Eds.), Proceedings of the First using vegetation heights from airborne scanning laser altimetry.
Conference on River Basin Management. WIT Press, South- Hydrological Processes 17, 17111732.
ampton, UK, pp. 99108. Middelkoop, H., Daamen, K., Gellens, D., Grabs, W., Kwadijk, J.,
Franks, S., Gineste, P., Beven, K., Merot, P., 1998. On Lang, H., Parmet, B., Schadler, B., Schulla, J., Wilke, K., 2001.
constraining the predictions of a distributed model: the Impact of climatic change on hydrological regimes and water
incorporation of fuzzy estimates of saturated areas into resource management in the rhine basin. Climatic Change 49,
the calibration process. Water Resources Research 34 (4), 105128.
787797. Nepf, H., 1999. Drag, turbulence, and diffusion in flow through
Freer, J., Beven, K., Ambroise, B., 1996. Bayesian estimation of emergent vegetation. Water Resources Research 35 (2),
uncertainty in runoff prediction and the value of data: an 479489.
application of the glue approach. Water Resources Research 32 Pedroli, B., Dijkman, J., 1998. River restoration in European
(7), 21612173. lowland river systems. In: Loucks, D. (Ed.), Restoration of
French, R., 1986. Open Channel Hydraulics. McGraw-Hill, New Degraded Rivers: Challenges, Issues and Experiences. Kluwer
York. Academic Publishers, New York, pp. 221227.
Gupta, H., Sorooshian, S., Yapo, P., 1998. Toward improved Romanowicz, R., Beven, K., Tawn, J., 1996. Bayesian calibration of
calibration of hydrological models: multiple and noncommen- flood inundation models, In: Floodplain Processes. Wiley, New
surable measures of information. Water Resources Research 34 York pp. 493533 (Chapter 15).
(4), 751763. Sellin, R., Irvine, D., Willets, B., 1993. Behaviour of meandering
Hesselink, A., 2002. History makes a river: morphological changes two-stage channels. Proceedings of the Institution of Civil
and human interference in the River Rhine. PhD Thesis. Faculty Engineers: Water, Maritime and Energy 101, 99111.
of Geographic Sciences, Utrecht University. Sellin, R., Bryant, T., Loveless, J., 2003. An improved method for
Hesselink, A., Stelling, G., Kwadijk, J., Middelkoop, H., 2003. roughening floodplains on physical river models. Journal of
Inundation of a dutch river polder, sensitivity analysis of a Hydraulics Research 41 (1), 314.
physically based inundation model using historic data. Water Silva, W., van de Langemheen, W., 1997. Landscape planning of
Resources Research 39 (9), 1234. the river rhine in the Netherlandsintegrated flood protection.
Horritt, M., 1999. A statistical active contour model for SAR In: RIBAMOD: River Basin Modelling, Management and Flood
image segmentation. Image and Vision Computing 17, Mitigation, Proceedings of the First Workshop. European
213224. Community, EAR18019EN, pp. 355357.
Horritt, M., 2000. Calibration of a two-dimensional finite element Sorooshian, S., Dracup, J., 1980. Stochastic parameter estimation
flood flow model using satellite imagery. Water Resources procedures for hydrological rainfall-runoff models: correlated and
Research 36 (11), 32793291. heteroscedatic errors. Water Resources Research 16 (2), 430442.
Horritt, M., Bates, P., 2001. Predicting floodplain inundation: raster- Sorooshian, S., Duan, Q., Gupta, V., 1993. Calibration of rainfall-
based modelling versus the finite-element approach. Hydro- runoff models: application of global optimization to the
logcial Processes 15, 825842. sacramento soil moisture accounting model. Water Resources
Hunter, N.M., Bates, P., Horritt, M., de Roo, A., Werner, M., Research 29 (4), 11851194.
(2003). Utility of different data types for calibrating flood Stephan, U., Gutknecht, D., 2002. Hydraulic resistance of submerged
inundation models within a GLUE framework. Hydrology and flexible vegetation. Journal of Hydrology 269, 2743.
Earth System Sciences Submitted. Werner, M., Ververs, M., van Haselen, C., Daamen, K.,
Khatibi, R., Williams, J., Wormleaton, P., 1997. Identification Muerlebach, M., 2000. A comparison of methods for generating
problems of open-channel friction parameters. Journal of cross-sections for flood modelling using detailed floodplain
Hydraulic Engineering 123 (12), 10781088. elevation models. In: Bronstert, A., Bismuth, C., Menzel, L.
M.G.F. Werner et al. / Journal of Hydrology 314 (2005) 139157 157

(Eds.), European Conference on Advances in Flood Research, Wilson, C., Bates, P., Hervouet, J.-M., 2002. Comparison of
Potsdam, Germany, pp. 7386. turbulence models for stage-discharge rating curve prediction
Wijbenga, J., Lambeek, J., Mosselman, E., Nieuwkamer, R., in reach scale compound channel flows using two-dimen-
Passchier, R., 1994. River flood protection in the Netherlands. sional finite element methods. Journal of Hydrology 257,
In: White, W., Watts, J. (Eds.), River Flood Hydraulics. Wiley, 4258.
New York. Wind, H., Nierop, T., de Blois, C., de Kok, J., 1999. Analysis of
Wilson, C., Horritt, M., 2002. Measuring the flow resistance of flood damages from the 1993 and 1995 meuse floods. Water
submerged grass. Hydrological Processes 16, 25892598. Resources Research 35, 34593465.

You might also like