Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Nat Hazards (2012) 63:417–447

DOI 10.1007/s11069-012-0155-z

ORIGINAL PAPER

Numerical modelling of wave overtopping-induced


erosion of grassed inner sea-dike slopes

Tuan Thieu Quang • Hocine Oumeraci

Received: 13 March 2011 / Accepted: 19 March 2012 / Published online: 31 March 2012
Ó Springer Science+Business Media B.V. 2012

Abstract Recent field wave overtopping experiments on grassed sea-dikes indicate that a
moderate-quality grass cover can surprisingly resist severe overtopping rates. This high
erosion-resistant capacity against wave overtopping of grass covers is owing to the root
reinforcement of grass in the subsoil. Inspired by this finding, the present study makes first
attempt to develop a numerical model of wave overtopping-induced erosion of the inner
slope of grassed sea-dikes, considering the effect of the root reinforcement. The critical
velocity for grass erosion is described as a function of the root cohesion and thus decreases
with the erosion depth. This depth-dependent strength of grass allows for a better
description of the nature of grass erosion, especially eligible for modelling erosion initiated
from a naturally or artificially weak spot in a grass slope. Wave overtopping on grass
slopes exhibits high turbulence with entrained air bubbles, for which the bed shear stress
determined according to the conventional approach for ordinary open-channel flows is
underestimated. To resolve this, it is assumed that the structure of wave overtopping
resembles that in a bubbly turbulent wall jet, so that the bed shear stress can be determined
in connection with the degree of flow turbulence. Model validation against the data from
field overtopping experiments conducted in the Netherlands shows that main features of
grass erosion on landward dike slopes are successfully simulated. Numerical experiments
on various aspects of grass erosion are carried out which give some new insights into the
nature of wave overtopping-induced grass erosion. Also, if the process of breach initiation
of a grassed dike can be divided into subprocesses, erosion of the grass turf and erosion of
the bare clay cover, an attempt is successfully made to roughly anticipate the breach
initiation time of grassed sea-dikes composed of a sand core and a grassed clay cover.

T. Thieu Quang (&)


Faculty of Marine and Coastal Engineering, Water Resources University (WRU), 175 Tay Son,
Dong Da District, Hanoi, Vietnam
e-mail: Tuan.T.Q@wru.edu.vn

H. Oumeraci
Department of Hydromechanics and Coastal Engineering, Leichtweiss Institute for Hydraulic
Engineering and Water Resources (LWI), Technical University of Braunschweig, Beethovenstrasse
51a, 38106 Braunschweig, Germany
e-mail: H.Oumeraci@tu-bs.de

123
418 Nat Hazards (2012) 63:417–447

Keywords Wave overtopping  Grass erosion  Breach initiation  Root reinforcement 


Grassed sea-dikes  Critical velocity for grass erosion

1 Introduction

Breach initiation of grassed sea-dikes due to wave overtopping starts as grass starts to
erode until the cover layer has failed to protect the sand core from the attack of the flow.
Severe wave overtopping during storm can induce breach initiation and subsequently dike
failure due to breach development at the inner slope. In the context of a changing global
climate, it is desirable to build sea-dikes that are more overtopping resistant and sustain-
able. Sea-dikes with inner slopes covered by grass are technically and economically
suitable for this purpose and thus have attracted lots of research attention recently.
Recent large-scale and extensive field wave overtopping experiments on grass dike
slopes carried out under the frame work of European projects such as Floodsite, ComCoast,
and Eurograss (see e.g. Akkerman et al. 2007; Van der Meer 2008; Geisenhainer and
Oumeraci 2008) have revealed many fundamental problems related to breach initiation and
breaching processes of grassed sea-dikes. One of the most noticeable findings is that a
grassed dike slope, owing to the root reinforcement, can surprisingly withstand severe
wave overtopping, which is in contrast with our present understanding of the resistive
capacity of grass slopes against wave overtopping.
Breach initiation or failure of grass and clay cover can be described through a combination
of several causal modes. Superficial erosion, infiltration-induced turf set-off (shallow grass
turf sliding), and clay cover (planar) sliding were attributed to be the main causes of sea-dike
initial failures in the past in the Netherlands and Germany (see e.g. Schüttrumpf and
Oumeraci 2004). Due to the effect of grass root reinforcement, the last two appear to be very
unlikely (at least in the case of mild inner slopes as indicated by Young 2005 and Akkerman
et al. 2007). In terms of grass erodibility, grass erosion as the result of (continuous and
discrete) sediment transport is of importance. Other failure mechanisms such as seepage-
induced slip circle and heat-cut erosion as described in Schüttrumpf and Oumeraci (2004),
rather observed during the process of breach development, fall beyond this research scope.
The present article concerns the development of a numerical model of breach initiation
of grassed sea-dikes due to wave overtopping with a special focus on the process of grass
erosion. The process addressed is limited to the cause by superficial erosion only. In
essence, an attempt has been made to include the effect of the bubbly flow structure of
wave overtopping and the depth-dependent resistance of grass root-permeated soil.
The article is organized as follows. Section 2 covers the basic model formulations for
modelling grass erosion, including the elaboration of the depth-dependent critical velocity
for grass erosion and the local bed shear stress exerted by turbulent wave overtopping. The
model validation against field experiment data of grass erosion is discussed in Sect. 3. A
series of numerical experiments to investigate various aspects of grass erosion due to wave
overtopping are carried out in Sect. 4. Finally, conclusions are given in Sect. 5.

2 Modelling of wave overtopping-induced erosion of grassed sea-dike slopes

Grass has long been used as a protective cover for dike slopes against erosion induced by
flow. However, present knowledge of grass erodibility against flow is still very limited and

123
Nat Hazards (2012) 63:417–447 419

largely empirical. Processes related to this phenomenon are not yet well understood.
Complex botanic aspects of grass together with uncertainties associated with physico-
chemical properties of cohesive subsoil make the erodibility of grass root-permeated soil
very erratic. It is even more challenging when addressing the overtopping resistance of a
dike as whole. This is due to, for examples, inhomogeneity of grass and soil, interaction
between soil and water leading to complex failure mechanisms, and existence of weak
spots vulnerable for failure initiation (Schüttrumpf and Oumeraci 2004).
In practice, probably because of these uncertainties, most guidelines such as CEM
(2002) or EurOtop (2007) recommend a so-called mean allowable overtopping discharge in
the range of 1–10 l/s per unit length for grass-slope dikes. However, recent field test studies
(see e.g. Akkerman et al. 2007; Van der Meer 2008) have demonstrated that grass slopes
can resist a surprisingly high overtopping rate, that is approximately about 75 l/s/m or
more for a moderate grass cover. This suggests that the present wave overtopping criteria
for grass seem rather conservative, regardless of the grass cover quality and wave over-
topping conditions.
The erosion resistance discussed in this study represents a homogeneous segment of
sea-dike and is restricted to erodibility aspects of grass root-permeated soil. The capacity
of a grass cover to resist against wave overtopping is ensured by the root reinforcement in
the top soil layer of around 20 cm from the surface. It appears from field experiments that
grass sods contribute mostly to the erosion resistance, whilst it is less significant from the
grass sward (leafy part).
Van den Bos (2006) developed a conceptual model for the erosion depth induced by
wave overtopping on grass slopes. The model, based on the exponentially time-resolving
scour concept, describes the maximum scour depth of an assumed vulnerable (bare clay)
spot as a function of overtopping parameters such as the maximum flow velocity and the
overtopping time, a flow turbulence parameter, and a soil parameter dependent on the
quality of both grass and soil. The overtopping parameters for a given storm are estimated
following the statistical properties and existing empirical formulations of wave overtop-
ping. However, the model uncertainty largely lies in the determination of the grassed soil
parameter.
The present criterion based on acceptable mean overtopping rates alone is shown
insufficient to present the erodibility of grass slopes. In a recent attempt, Dean et al. (2010)
proposed the concept of erosional equivalences of grass slopes subject to wave overtop-
ping, taking into account various parameters of wave overtopping including the overtop-
ping duration. The proposed erosional indices are as follows: (1) excess velocity, (2) excess
shear stress, and (3) excess work, all based on the wave overtopping velocity and the
critical velocity for grass erosion. The work index is found to fit the best the stability curve
of Hewlett et al. (1987) developed under steady overflow conditions. Similarly, Van der
Meer et al. (2010) used a parameter called the cumulative hydraulic load, like the shear
stress index proposed by Dean et al. (2010) but without the overtopping duration, to
evaluate the various damage states of an eroding grass slope. However, in these studies, the
determination of the critical velocity for grass erosion is rather erratic and not physically
related to the grass characteristics. Moreover, the velocity of the wave overtopping flow
substantially varies along a dike slope, whilst the critical velocity for grass erosion is not
the same everywhere. Therefore, the proposed indices are not quantitative, and the use of
either of these to represent a damage state of a grass slope as a whole is not yet suitable.
Following the same concept of excess shear stress, this section addresses an approach
for numerical modelling of the time-dependent dike breach initiation by surface erosion of
grass at the inner slope. In essence, as shown in the following subsections, it is the

123
420 Nat Hazards (2012) 63:417–447

modelling of the critical bed shear stress for grass erosion and the local bed shear stress
exerted by turbulent wave overtopping.

2.1 Grass and clay erodibility in terms of excess bed shear stress

Generally speaking, the overall erosion of grassed clay occurs as a combination of two
modes: surface (continuous) erosion and mass (discrete) erosion. At present, the latter
process is poorly understood as it is highly stochastic (Winterwerp and Van Kesteren
2004). If the considered time scale is sufficiently large (viz. final situation rather than
momentary state is of interest), it is reasonable to address these two as a whole through a
modelling approach for surface erosion in combination with (non-stochastic) triggering
mechanisms for mass erosion to occur. Herein, it is assumed that mass erosion partly
occurs through bed sliding whenever surface erosion makes the bed slope steeper than the
angle of internal friction of the bed material.
The rate of soil erosion can most appropriately be quantified in terms of the excess bed
shear stress as follows (see Partheniades 1965 and also Winterwerp and Van Kesteren 2004
for a review):
dzb Mgs
Egs ¼ qs ¼ ðsb  sgs;c Þ sb [ sgs;c
dt sgs;c ð1Þ
Egs ¼ 0 sb  sgs;c

in which Egs (kg/m /s) is the erosion rate per unit bed area, Mgs (kg/m2/s) is the erosion
2

coefficient of grass root-permeated soil (calibrated on the basis of overall erosion), qs is the
bed mass density, zb is the bed level, sb is the instantaneous turbulent bed shear stress, and
sgs,c is the mean critical (threshold) bed shear stress for erosion of a root-permeated soil.
For ordinary open-channel flows, the bed shear stress sb can generally be related to the
instantaneous depth-averaged overtopping velocity u according to:
 u 2 1
sb ¼ qg ¼ qfw u2 ð2Þ
C 2
where q is water density, g is gravitational gravity, u is depth-averaged velocity, C is
Chezy coefficient, and fw is bed friction coefficient.
The critical bed shear stress can also be related to the critical velocity ugs,c, following a
similar manner:
1
sgs;c ¼ qfw u2gs;c ð3Þ
2
where ugs,c is the critical velocity of root-permeated soil.
Observations from recent field and large-scale wave overtopping experiments on grass-
slope dikes indicate that the flow on the inner dike slope is highly turbulent with a
considerable amount of entrained air bubbles (see e.g. Fig. 6a). The flow structure here
deviates substantially from that in ordinary open-channel flows, and thus, Eq. (2) is likely
to underestimate the bed shear stress. To determine the bed shear stress in this extraor-
dinary situation, the present study assumes that the characteristics of wave overtopping on
the inner dike slope resemble those in an air-entrained turbulent wall jet. It is delineated in
Sect. 2.3 that in a turbulent wall jet or turbulent hydraulic jump, the maximum horizontal
velocity is found near the bottom and the bed shear stress should instead be related to this
near-bed velocity.

123
Nat Hazards (2012) 63:417–447 421

To quantify the rate of erosion according to Eq. (1), Mgs and sgs,c must also be known.
Existing laboratory experiments on erosion of cohesive (mostly mud) beds suggest Mgs (kg/
m2/s) to vary in the range between 0.01.10-3 and 0.5.10-3 (Winterwerp and Van Kesteren
2004). It should be noted that this value range of Mgs corresponds to soils of low critical
shear stress (say up to several m/s2/m2 only). For root-permeated and compact soils, the
critical shear stress is about two orders of magnitude higher (up to hundreds of kg m/s2/m2,
see Verheij 1995). A much larger value of Mgs would therefore be expected for this study.
Several approaches exist to estimate the critical velocity ugs,c and the critical bed shear
stress sgs of grass slopes such as by Verheij (1995) and Hoffmans and Verheij (1997).
However, most of these assume a constant critical velocity for erosion for a given grass cover
quality, which seems inappropriate. In fact, preceding studies on the reinforcement of root-
permeated soils (e.g. Wu et al. 1979) indicate that the soil shear strength is significantly
enhanced by the presence of roots in the soil and that the reinforcement can be related to the
root density or more specifically to the root area ratio (RAR). The root density decays rapidly
over the depth from the surface, and the resistance of root-permeated soil must therefore
follow the same trend. As a result, the soil resistance is found very large near the surface and
quickly reduces to that of bare soil at a depth level of negligibly small root density (several
decimetres below the surface for dike grasses). This depth-dependent behaviour of the grass
erosion resistance was reported in preceding laboratory studies on erosion of grass slopes.
Smith (1993) and Kruse (1994a, b) (see also Sprangers 1999) reported from their large-scale
experiments that the erosion rate at the top 0–5-cm surface layer (containing 60 % of the total
grass sods) is about one order of magnitude lower than that at a deeper layer at depth of
5–10 cm (containing only 20 % of the total grass sods), under the same hydraulic conditions.
The strong erosion resistance of the surface grass layer was also observed in recent field
overtopping experiments (see e.g. Akkerman et al. 2007), viz. grass erosion could only occur
when a surface layer of 5 cm thick (initial damage) was removed from the tested slope.
The above argument lends strong support to the assumption that the critical velocity for
erosion of a grass cover is a function of (soil) depth. Besides modelling grass erosion more
properly, the depth-dependent strength of grass allows us to model localized erosion
resulting from a natural weak spot (see Sects. 2.4, 3). The concept of (soil) depth-
dependent critical velocity of root-permeated soil is introduced in the present model, for
which the formulations are elaborated in Sect. 2.2.

2.2 Depth-dependent critical velocity for erosion of root-permeated clay

In general, the critical velocity for erosion of cohesive sediment is poorly determined as it
is governed by a large number of uncertain and mutually influenced processes. Most
available formulations are therefore rather sediment specific and not yet applicable for
generic use. For the present purpose, it is probably most suitable to adopt the generic
expression of Mirtskhoulava (1991) as influences from soil properties and flow parameters
are well parameterized. This formulation was simplified by Hoffmans and Verheij (1997)
for cohesive sediments and further modified by Van der Meer et al. (2007a) to include the
effect of grass root reinforcement:
 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
8:8h 1
ugs;c ¼ 0:64 log Dgda þ ð0:6Cf þ Cr Þ ð4Þ
da q

where ugs,c is the critical velocity for erosion of grass-permeated soils, h is the flow depth,
q is the water density, D ¼ ðqs  qÞ=q is the relative density of clay, qs is the saturated

123
422 Nat Hazards (2012) 63:417–447

clay density, da (= 0.003–0.005 m) is the size of clay detaching aggregates, Cf (=0.035c) is


the fatigue rupture strength of clay, and c is the clay cohesion. Here, the effect of grass root
reinforcement is included through the root cohesion Cr (Van der Meer et al. 2007a).
Though Eq. (4) has not been verified against experimental data of grass erosion, Van der
Meer et al. (2007a) claimed that this gives reasonable estimates for moderate-quality grass
covers with Cr = 5 kN/m2. In the present research, we advocate this straightforward
inclusion of the root cohesion because of its realistic behaviours. On the other hand, unlike
Van der Meer et al. (2007a), we consider Cr herein as a depth-varying parameter.
It is important to realize that Cr in general is a function of the root density that decreases
rapidly over the depth from the soil surface. Therefore, Eq. (4) is indeed the desired
expression for the depth-dependent critical velocity of root-permeated soils if the variation
in depth of the root cohesion Cr can properly be determined. The determination of the root
cohesion as a function of soil depth is discussed in the following.
Studies on root reinforcement of grassland plants have been pursued at length in the past
decades because of its importance to the bank slope stability. The basic principle of root
reinforcement is according to the well-known (simplified) formulation by Wu et al. (1979):
Ar
CrW ¼ 1:2 tr or CrW ¼ 1:2:RAR:tr ð5Þ
A
where CW 2
r is the root cohesion in kN/m estimated according to Wu et al. (1979), tr is the
2
root tensile strength (kN/m ), and Ar/A or RAR is the root area ratio defined as the total area
of roots Ar per unit area A of the crossing shear plane.
To provide an approximation of the magnitude of the root cohesion estimated by
Eq. (5), we assume that c = 11.0 kN/m2, tr = 20.106 N/m2 (average tensile strength of
typical Dutch dike grasses), and RAR = 0.066 % and 0.047 % at depths of 2.5 and 5 cm,
respectively (for an average-quality grass slope according to VTV (2007)). From Eqs. (5)
and (4), the critical velocities for grass ugs,c under the average overtopping thickness
h = 0.2 m are 6.7 and 5.7 m/s for depths of 2.5 and 5 cm, respectively. The critical
velocity us,c for the bare clay is 0.86 m/s (with Cr = 0), which is significantly smaller than
those for the grassed clay.
In Eq. (4), the root cohesion Cr can generally be one to two orders of magnitude larger
than 0.6Cf. This hints at that the erosion resistance (the critical velocity) for grass slopes is
in fact largely determined by the root reinforcement but not by the soil strength. This
character complies well with observations by Sprangers (1999) and from recent field
overtopping experiments (see e.g. Van der Meer 2008). Therefore, for the application of
Eq. (4) to the present model, a proper determination of the root cohesion as a function of
the soil depth is very essential.
On the whole, two factors that mostly affect the determination of the depth-dependent
root cohesion are the variation with depth of the root area ratio RAR and the basic
assumption of the Wu et al. (1979) model. We, respectively, address these problems in
Sects. 2.2.1 and 2.2.2.

2.2.1 Depth-dependent root area ratio RAR

In general, many factors such as botanic composition (species and age), climate, nutrient,
and soil conditions all have complex effects on the root density and the decay of the root
density over depth. Therefore, the root area ratio curve for a grass slope is rather site
specific, representing the local conditions.

123
Nat Hazards (2012) 63:417–447 423

In a study on sea-dike grassland, Sprangers (1999) found that the root density decreases
exponentially with depth. Young (2005) re-analysed the data of Sprangers (1999) and
proposed a regression curve for the root area ratio RAR as read below:
RAR ¼ a þ ðb  aÞRðddoÞ ð6Þ
where RAR (in %) is the root area ratio at a considered depth d (in cm), d0 (=1.5 cm) is a
reference depth of the surface layer, a (=0) and b (=2,67) are the root parameter values at
depths of 50 cm and d0, respectively, and R (=0,8) is a factor representing the degree of decay.
Similarly, Stanczak (2008) claimed the same trend for the root area ratio of grass
samples taken from a sea-dike in Leybucht (Germany) with a = 0 (at depth of 17 cm),
b = 1.58, R = 0.75, and d0 = 2 cm.
In the present study, we use an alternative expression for the decay over depth of the
root area ratio:
RAR ¼ RAR0 ebðdd0 Þ ð7Þ
in which RAR0 is the averaged root area ratio over the reference depth layer d0 = 2.5 cm
from the surface (hereafter referred to as the reference root area ratio for brevity), and b is a
root density decay (shape) parameter.
The exponential functions in Eqs. (6) and (7) are in fact identical with b = -lnR. How-
ever, the latter has a fixed base (exponent varies) that enables more accurate tuning.
Figure 1 plots the data of Sprangers (1999) and Stanczak (2008) together with the
regression curve based on the function in Eq. (7) with RAR0 = 1.87 % and b = 0.30.
The Dutch guidelines VTV (2007) classifies grass covers into 4 different categories of
‘‘very poor’’, ‘‘poor’’, ‘‘moderate’’, and ‘‘good’’ by the number of roots found at several
depth levels. Given the mean root diameter of 0.13 mm for Dutch dike grassland (Sprangers
1999), the categories of grass cover quality now can be expressed in terms of the root area
ratio as shown in Fig. 2. The mean curve of the root area ratio for each category can also be
approximated according to the proposed expression in Eq. (7). The two curves shown in
Fig. 2 for poor and moderate grass covers are plotted with RAR0 = 0.04 % (±0.01) and
0.065 % (±0.01), respectively, and a common value of b = 0.12 (±0.02). In the present
mode, these curves are used by default in the absence of field data.
The root area ratio shown between Figs. 1 and 2 differs by one order of magnitude. The
main reason for this marked difference is probably due to the mean root diameter used in
the latter case (the root area ratio is proportional to the root diameter by a power of 2). It is
noted that a larger root area ratio due to larger roots does not necessarily means a larger
root cohesion because the root tensile strength tends to be inversely proportional to the root
size (see Sect. 2.2.2). The mean root tensile strength Tr = 20.5 MPa associated with the
mean root diameter of 0.13 mm is commonly assumed for Dutch dike grassland (see
Hoffmans et al. 2008), whilst Stanczak (2008) reported a much smaller mean strength
Tr = 5 MPa measured from his samples. It appears that the use of a single mean root size
and a single mean root strength in the determination of the root cohesion may be inap-
propriate. The effect of different root sizes on the root cohesion is treated in Sect. 2.2.2.

2.2.2 Root mobilized strength

The second factor, which considerably influences the determination of the root cohesion,
lies in the fundamental assumptions of Wu et al. (1979) model itself.

123
424 Nat Hazards (2012) 63:417–447

Fig. 1 Decay over depth of the root area ratio RAR

Fig. 2 Categories of grass cover quality in terms of root area ratio derived from VTV (2007) with mean
grass root diameter of 0.13 mm

123
Nat Hazards (2012) 63:417–447 425

In general, the Wu et al. (1979) model, hereafter referred to as Wu’s model for brevity,
considerably overestimates the actual root cohesion. This is because the model assumes
that all roots break simultaneously. In other words, the tensile strength of all roots is fully
and simultaneously mobilized at the moment of shear failure. In reality, roots are not the
same in terms of tensile strength and thus break progressively. Also, permeated roots in the
soil do not have the same diameter. Large roots (say larger than 3–4 mm) do not form tight
binding network in the soil like small roots and thus they tend to be pulled out rather than
broken during shear failure (see e.g. Tosi 2007). The reinforcement part contributed by
large roots is mainly through root-soil bond strength (surface friction), which is small
compared to the mobilized tensile strength of thin roots. The latter aspect seems insig-
nificant for dike grassland being considered which typically has tiny roots in the range
between 0.1 and 1.5 mm. Therefore, we can disregard the effect of unbroken large roots
and consider the Wu’s model overestimation in the light of the former aspect, that is, roots
break progressively during failure.
Generally speaking, the cohesion estimated by Wu’s model should be regarded as the
maximum or potential cohesion when the root tensile strength of all roots is fully and
simultaneously mobilized. In the present study, we introduce a mobilized strength coef-
ficient lr, as defined in Eq. (8), to account for the partial strength mobilization of grass
roots in the determination of the root cohesion.
Cr
lr ¼ ð8Þ
CrW
where lr (\1) is the root mobilized strength coefficient.
Substitution of Eq. (8) into Eq. (4) yields the following expression for the determination
of the critical velocity of grass slopes:
 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
8:8h 1
ugs;c ¼ 0:64 log Dgda þ ð0:6Cf þ lr Crw Þ ð9Þ
da q

The main purpose of introducing coefficient lr in Eq. (9), together with the simplicity of
Wu’s model, is to ease the difficulty in determining the actual root cohesion Cr, provided
that lr can be reasonably estimated. Here, it is assumed that lr does not change over depth.
The determination of lr is addressed in the following.
At first, the magnitude of lr can be estimated from preceding studies. In a study on
stability of stream bank reinforced by riparian vegetation, Pollen and Simon (2005) found
that Wu’s model can overestimate the actual root reinforcement between 20 and 50 % in
case the driving force is in excess of the root strength. The overestimation increases (up to
one order of magnitude) with the increase of the deficit in the driving force compared to the
root strength. Operstein and Frydman (2000) reported an average overestimation of Wu’s
model for the root cohesion of herbal trees of about 80 % compared to the data from their
measurements in situ. These suggest that lr may vary in the range between 0.20 and 0.80,
depending on soil conditions and botanic properties of vegetation. For grasses, as roots are
more uniform in terms of tensile strength, a narrower range of lr is therefore expected.
Several approaches exist which can be used for determining the root cohesion and thus
the root mobilized strength coefficient lr according to Eq. (8). Discussion amongst these
approaches falls beyond this research scope. We adopt herein the approach of Pollen and
Simon (2005), which is straightforward to implement into the present model. More
complicated approach is not desirable since this does not help improve the reliability of
Eq. (9).

123
426 Nat Hazards (2012) 63:417–447

Pollen and Simon (2005) applied the approach of fibre bundle models (FBM), com-
monly used in the science of composite materials to describe the progressive breaking of
grass roots during failure. This simple algorithm assumes that during failure, intact roots
have to take additional burden redistributed from already broken roots, which in the end
leads to failure of the whole root system at a much smaller load.
For modelling superficial grass erosion, we assume erosion occurs only when the
driving stress (bed shear stress) exceeds the grass strength (see Eq. 1). Therefore, the
mobilized strength coefficient lr in this case is determined based on the assumption that all
roots break eventually.
Let us consider a unit soil bed area crossed through by a grass bundle consisting of N
classes of fibres of different diameters and tensile strengths. By definition, from Eq. (8) and
the root cohesion according to the FBM approach, the mobilized strength coefficient lr can
be determined as follows:
CrFBM Pmax
l¼ ¼ ð10Þ
CrW P
N pd 2
1:2 nr;i 4r;i tr;i
i¼1

in which CFBM
r and Pmax (N) are the root cohesion and the maximum force withstood by the
grass bundle determined according to the FBM approach of Pollen and Simon (2005),
respectively; nr,i is the number of root in i-th root class which has a class diameter dr,i and
class tensile strength tr,i (N/m2).
Alternatively, lr can be defined if CW r is calculated through the strength of the mean-
diameter root instead of a sum of strength of individual roots:
CrFBM Pmax Pmax
l¼ ¼ N  ¼ ð11Þ
CrW P pd 2 1:2Ar :tr;m
1:2 nr;i 4r;i :tr;m
i¼1

where tr,m is the tensile strength corresponding to the mean root diameter, and Ar is the
total root area in a unit bed soil area considered.
The resulting value of lr can be different between Eqs. (10) and (11). However, the use
of Eq. (11) is more practical since in most cases, only the RAR and the strength associated
with the mean-diameter root are given (see e.g. Hoffmans et al. 2008).
The determination of Pmax and thus lr according to Eq. (10) or Eq. (11) requires two
input curves: the root diameter distribution curve (field data) for a considered vegetation
species and the root size, and tensile strength curve. The latter is available for grass from
studies on bank stability because this relationship can reasonably be established, although
it can markedly differ from species to species. It is more erratic, unfortunately, for the
former case, and thus, field data are needed for most situations. There has been very little
information on the root size distribution for dike grassland species so far.
The above facts indicate that at present, the determination of the mobilized strength
coefficient for dike grassland is only exact to a certain extent. Together with the unknown
reliability of Eq. (9) it is therefore better to judge the final value of lr through model
validation against data of grass erosion (see Sect. 3). Here, as shown in the following, we
can at least determine the possible range of lr for dike grassland using Eq. (10) through a
sensitivity analysis by varying the root characteristics.
First, we address the root size distribution. It is assumed that the variation of the grass
root size follows a Weibull distribution, in which the frequency of occurrence of roots
smaller than diameter dr reads:

123
Nat Hazards (2012) 63:417–447 427

Fig. 3 Examples of root size distribution of Bermuda grass fitted with Weibull distribution

k
Fð\dr Þ ¼ 1  eðdr =drm Þ ð12Þ
where dr is a root diameter and F(\dr) is the frequency of occurrence of roots smaller than
dr, drm is the mean root diameter of the considered grass, and k is a shape factor repre-
senting the degree of root size spreading around the mean value (the two curve parameters
k and dm can be determined through a fitting procedure).
It should be noted that the Weibull distribution is widely used for grain or thread size
distribution in the science of materials because of its flexibility. Larger k (k [ 1.0) means
the root size is more narrowly distributed around the mean value and vice versa (k \ 1.0);
k = 2 when the root size is Rayleigh distributed.
To illustrate the application of the proposed Weibull distribution for this purpose, the
root size data from three samples of a Bermuda grass collected at a dike testing site in
Vietnam are used (see WRU-CE 2009). The fittings of these data with the Weibull curves
and the corresponding pair of parameters (k and dm) are shown in Fig. 3, generally yielding
a good agreement. It follows that in this case, the root size of dike grassland does not
deviate much from the mean size, so k [ 1 (k = 1,30 * 1,40 as shown in Fig. 3) can be
selected.
Second, a certain relationship is required for determining the root tensile strength tr,i
from the root diameter dr,i for each root size class i in Eq. (10). It has been reported in
numerous studies (see e.g. Operstein and Frydman 2000; Tosi 2007; De Baets et al. 2008)
that the root tensile strength decreases exponentially with the increase in the root diameter
according to:
tr ¼ adrb ð13Þ
where a and b (b \ 0) are species-dependent root parameters; a represents a reference
strength whilst b represents the degree of variation of strength with diameter.
The diameter-strength relation for a number of grass species is readily available or can
be derived from reported data in the literature. Figure 4 shows this relation for 8 grass

123
428 Nat Hazards (2012) 63:417–447

Fig. 4 Tensile strength versus root diameter derived from data of Cheng et al. (2003)

species derived from the data of Cheng et al. (2003). The relations for 5 other grass species
were reported in De Baets et al. (2008) (see also Table 1 for a summary).
Little information is known about the root characteristics of Dutch dike grassland except
the mean diameter of 0.13 mm evaluated according to the data of Sprangers (1999) (see
Young 2005). However, most quantitative data of grass erosion to be used for model
validation are available for Dutch dike grassland only; hence, assumptions about the root
diameter-strength curve have to be made. It is very likely from Fig. 4 (see also Table 1)
that common grassland species that form dense turfs on the ground such as Bermuda,
Manila, and Bahia do have a smaller value of b compared to clump grasses (e.g. Vetiver
and White Clover). It is assumed that, for Dutch dike grassland, the tensile strength
associated with the mean diameter tr,m = 20.5 MPa (Van der Meer et al. 2007a; Hoffmans
et al. 2008) and b = -0.57 as is common for turf-forming grassland species, we yield from
Eq. (13) a = 6.40. For species-rich and low-strength (tr \ 50 MPa) grassland, the average
values a = 14.5 and b = -0.41 can be guessed from Fig. 4.
Given the root size distribution and the diameter-strength relationship, the procedure for
determining the mobilized coefficient lr for each grass is as follows. It is assumed that the
total number of roots in the considered bundle is 1,000 roots of various diameters ranging
between 0.05 mm and 2.5 mm (this number of roots is found to be large enough not to
affect the results). The diameter range is divided into N = 50 root diameter classes. The
number of roots in each class nr,i and the class root strength tr,i associated with the mean
class diameter dr,i are determined from Eqs. (12) and (13), respectively. The maximum
force Pmax and subsequently lr in Eqs. (10) and (11) can now be determined using the
aforementioned FBM algorithm of Pollen and Simon (2005). The resulting mobilized
coefficient for various types of grass is shown in Table 1. Figure 5 plots the variation of lr
against b for various k values representing different degrees of root size deviation around
the mean diameter.
It can be seen from Fig. 5 that there exist strong correlations between lr and the two
parameters representing the root characteristics b and k, viz. lr is linearly proportional to

123
Table 1 Species-dependent root mobilized strength coefficient lr determined using FBM
Grass name Mean root diameter a b lr by Eq. (10) with k = lr by Eq. (11) with k =
dr (mm)
0.50 0.75 1.00 2.00 3.00 0.50 0.75 1.00 2.00 3.00
Nat Hazards (2012) 63:417–447

Vetiver 0.66 ± 0.32 48.0 -1.00 0.42 0.40 0.37 0.45 0.52 0.22 0.22 0.22 0.39 0.51
Bermuda 0.99 ± 0.17 14.5 -0.57 0.60 0.60 0.59 0.57 0.62 0.49 0.49 0.49 0.52 0.61
Manila 0.77 ± 0.67 14.5 -0.57 0.60 0.58 0.56 0.57 0.62 0.43 0.43 0.43 0.52 0.61
White Clover 0.91 ± 0.11 22.4 -1.64 0.21 0.22 0.23 0.31 0.4 0.16 0.16 0.17 0.29 0.42
Bahia 0.73 ± 0.07 14.5 -0.57 0.59 0.58 0.56 0.57 0.62 0.42 0.42 0.42 0.52 0.61
Late Juncellus 0.38 ± 0.43 14.5 -0.57 0.57 0.52 0.48 0.57 0.63 0.29 0.29 0.31 0.52 0.62
Dallis 0.92 ± 0.28 17.8 -1.05 0.41 0.41 0.40 0.43 0.50 0.29 0.29 0.29 0.38 0.50
Common Cetipede 0.66 ± 0.05 17.8 -1.05 0.40 0.38 0.35 0.43 0.51 0.21 0.21 0.21 0.38 0.50
Stipa tenacissima 0.88 ± 0.45 24.34 -0.61 0.58 0.57 0.56 0.55 0.61 0.44 0.44 0.44 0.50 0.60
Brachypodium retusum 0.77 ± 0.67 45.05 -0.61 0.59 0.57 0.55 0.55 0.61 0.41 0.41 0.41 0.50 0.60
Helictotrichon filifolium 0.78 ± 0.44 14.51 -1.08 0.39 0.38 0.37 0.43 0.5 0.24 0.24 0.24 0.38 0.50
Piptatherum miliaceum 0.37 ± 0.27 11.49 -1.77 0.15 0.17 0.2 0.31 0.4 0.03 0.05 0.09 0.29 0.42
Avenula bromoides 0.24 ± 0.09 4.77 -1.52 0.18 0.2 0.25 0.37 0.45 0.03 0.05 0.10 0.33 0.46
Dutch dike grassland 0.13 6.40 -0.57 0.50 0.43 0.48 0.60 0.66 0.19 0.24 0.36 0.64 0.77
429

123
430 Nat Hazards (2012) 63:417–447

Fig. 5 Variations of mobilized coefficient lr with root parameters b and k computed according to (a) Eq.
(10) and (b) Eq. (11)

the decrease in variation of the root strength (smaller absolute values of b), and lr also
increases as the deviation from the mean diameter of the root size decreases (larger
k values). However, the correlation between lr and b is worse for small k (\1) values.
Also, the strength parameter a is found to have negligible effect on the mobilized coef-
ficient lr. These tendencies agree well with the general observation that the variation in the
root size and root strength is mainly responsible for the overestimation of the root cohesion
according to Wu’s model.
Between the results of lr computed from Eqs. (10) and (11), for small k values (k \ 1),
the latter (Fig. 5b) shows a larger scattering compared to the former (Fig. 5a). Hence, the
use of Eq. (11), although more practical, is only appropriate when the root size is narrowly
distributed around the mean diameter (k [ 1).
In general, Fig. 5 or Table 1 can be used to estimate the value range of lr for a
considered grass species, provided that the possible values of b and k are known. The tested
root samples with various diameters reported in Cheng et al. (2003) and WRU-CE (2009)
suggest that k roughly varies between 1 and 2 for some common turf-forming grassland
species (Bermuda, Manila, and Bahia). It should be noted that this is only a very rough
estimate since the tested samples from Cheng et al. (2003) and WRU-CE (2009) are neither
representative nor sufficiently large (these were indeed not aimed for this purpose). For
Vetiver, it is stated in Cheng et al. (2003) that 50–66 % of roots have a diameter smaller
than 1 mm, which hints at k = 0.75 * 1.0 with the mean root size of 0.66 mm.
With guessed values k = 1–2 and b = -0.57, the mobilized coefficient lr for Dutch
dike grassland ranges between 0.36 and 0.64 (see Table 1). The final k value for this case is
judged in Sect. 3 according to model validation against data of grass erosion from field
experiments.

123
Nat Hazards (2012) 63:417–447 431

2.3 Bed shear stress in a turbulent wall jet

Though no measurement has been made so far, it can be concluded through visual
observations that the turbulence in the overtopping flow on the inner grass dike slopes is
much higher than that in an ordinary open-channel flow (see also Hoffmans et al. 2008).
This also means that Eq. (2) underestimates the bed shear stress and consequently leads to
an under-prediction of grass erosion by Eq. (1). Therefore, a more appropriate approach for
modelling the bed shear stress is needed. To this end, the actual turbulent flow structure of
wave overtopping on the inner dike slope must be known in detail, which is unfortunately
not available yet. To resolve this, it is assumed in this study that the flow structure in wave
overtopping on the inner dike slope resembles those in a bubbly turbulent wall jet whereby
the bed shear stress can more properly be determined.
Once a turbulent jet exists, the current profile is substantially modified, that is, the
maximum velocity is found very close to the bed (see e.g. Hager 1992). This is in sharp
contrast with the logarithmic profile of ordinary open-channel flows without jet, where the
maximum velocity is at the surface. Therefore, in a turbulent jet, the shear stress exerted on
the bed is much larger and thus more sediment is brought into transport. As a consequence,
a scour hole, which is a noticeable morphologic feature in dike breaching, develops around
the jet (see e.g. Tuan 2007). Also, a turbulent jet is the major driving mechanism for head-
cut erosion in cohesive beds (see Robinson 1996).
During a sufficiently large overtopping event, the presence of a turbulent jet in the form
of a hydraulic jump can visually be observed at transitions between the sloping part with
the horizontal such as at the inner berm, the foot of the inner slope, or breach plunging
pool. At other locations on the slope (rather than at transitions), the overtopping flow
structure is assumed to resemble that in a turbulent wall jet. There are several reasons that
support the assumption about this similarity. First, as previously mentioned and as shown
in Fig. 6, wave overtopping on the dike slope is highly turbulent with a significant amount
of entrained air bubbles as the result of its high shear characteristics (high Froude number).
Second, intermittent wave overtopping is similar to that generated behind the slot of a
sluice gate that opens and shuts periodically. The flow exits behind a sluice gate is a type of
turbulent wall jet commonly encountered in hydraulic engineering practice.
The effect of turbulent jet on the erosion of grass slopes has been reported from recent
field overtopping experiments on grass-slope dikes (e.g. Akkerman et al. 2007; Van der
Meer 2008; WRU-CE 2009), in which severe scour holes at the end of the inner slope and
head-cut erosion in bare clay beds were clearly observed (see e.g. Fig. 7).
It is noted that the actual effect of air bubbles on the bed shear stress and thus on grass
erosion is not understood yet. However, this effect is partly considered through the
interactions between the air bubble diffusion/advection and momentum transfer, which
result in the modification of turbulence scale and flow structure of a bubbly turbulent wall
jet (see e.g. Chanson 2006).
The above observations confirm the necessity of incorporating the influence of the
turbulent jet in the modelling of breach initiation. The ultimate goal is to quantify the bed
shear stress exerted by the jet, for which the characteristics of the near wall (or bottom
boundary) layer are needed. In the present 1-D flow model, the current profile and thus the
near wall layer could not be computed in detail, and hence, a parametric approach is
applied for the refinement of the flow parameters in the jet region. To this end, the
parametric approach of Ohtsu et al. (1990) as reported in Chanson and Brattberg (2000) is
adopted, whereby the current profile is determined as a function of the depth-averaged
velocity and the hydraulic properties of the turbulent jet. At a slope transition, the

123
432 Nat Hazards (2012) 63:417–447

Fig. 6 Turbulent wall jet: (a) wave overtopping flow on dike slope by wave simulator (picture from Van
der Meer et al. 2007b) and (b) behind the slot of a sluice gate (principle sketch)

hydraulic jump properties (if jump exists) are automatically captured as part of the flow
conditions according to an approach addressed in Tuan (2007).
Although the flow structure in a hydraulic jump and that in a turbulent wall jet is similar
(see Chanson and Brattberg 2000), in the present model, the determination of the bed shear
stress is still subdivided in two cases: one for the slope, where the flow structure is assumed
to resemble that in a turbulent wall jet, and one for transitions, where hydraulic jumps may
take place. However, the former is just a special case of the latter.
As is common for shear flows, the velocity profile in a hydraulic jump is specified in
two layers: one near wall or boundary layer and one shear or mixing layer as shown in
Fig. 6b. The re-circulating (roller) region, if exists, is treated as a dead layer. Considering
mass conservation, the maximum velocity um at the top of the boundary layer (see Fig. 6b)
can be related to the depth-averaged velocity u as follows (see Tuan 2007 for more details):

123
Nat Hazards (2012) 63:417–447 433

Fig. 7 Severe scour hole observed at the toe of the landward slope showing the strong effect of turbulent
wall jets: (a) test conducted in the Netherlands (Van der Meer 2008), (b) test conducted in Vietnam (WRU-
CE 2009)

u
um ¼ pffiffi  k  1:60 ð14Þ
p 1
rd 2g þ 1þa
u
um ¼ pffiffi  k [ 1:60 ð15Þ
p 1 1
rd 2g erfðgðrd  1ÞÞ þ 1þa

where rd is the dimensionless thickness of the bottom boundary layer, a (=1/7) and g
(=0.44) are constants, u is the depth-averaged flow velocity of wave overtopping, k is a
dimensionless distance (k B 1.60 means locations within jump and vice versa), and erf()
is the error function.
It is noted that all the depth-averaged flow parameters (depth and velocity) used for the
refinement of the flow field in Eq. (14) are computed from the module of wave overtopping
hydrodynamics (see Tuan and Oumeraci 2010).
It is assumed that the flow is hydraulically rough, the bed shear stress can now be
determined at the top of the bottom boundary layer according to the law of the wall as
follows (see also Ribberink 1998):
1
sb ¼ qfc u2m ð16Þ
2
where fc is the friction coefficient estimated using the rough-wall friction formulation.
 2
j
fc ¼ 2 ð17Þ
ln db =z0
where j (=0.4) is von Karman coefficient, z0 = ks/30 is the zero velocity level, and ks is an
apparent roughness (ks = da in Eq. (4) for clay transport or a user-defined parameter).

123
434 Nat Hazards (2012) 63:417–447

Equations (14) to (17) can now be used to determine the bed shear stress under various
degrees of flow turbulence. A smaller boundary layer thickness means a stronger turbulent
flow and vice versa. When this layer approaches the surface, the bed shear stress reduces to
that of ordinary open-channel flows (i.e. rd approaches 1.0 in Eq. (15)).
It can be seen from a comparison between Eqs. (16) and (2) that, because um  u and
fc  fw, the bed shear stress within the jet region is substantially larger than that outside
the jet. As a result, more erosion is expected and in the case of hydraulic jump, a scour hole
will develop around the jump.
In applying the above formulations of a turbulent hydraulic jump to the case of wave
overtopping at an arbitrary location on the slope (not at a transition), we realize that there is
no transition to a common channel flow downstream, to which the bottom boundary layer
has to grow to adapt as delineated in Fig. 6b. This hints at wave overtopping on the slope is
like a turbulent wall jet with a small and weakly growing near wall layer. Moreover, the re-
circulating region in wave overtopping does not exist. Therefore, the determination of the
bed shear stress for locations on the slope slightly differs from the above procedure for
hydraulic jump in that only Eq. (15) for the maximum near-bed velocity um (without re-
circulating region), and a fixed value of the dimensionless boundary thickness can be used.
As a first guess, the range rd = 0.25–0.35 can be derived from the boundary layer
thickness at the toe of a turbulent jet (at depth d1, see Fig. 6b) observed from various
laboratory air-water hydraulic jump experiments on a horizontal bed (see Chanson and
Brattberg 2000 for a review). If we assume that this is applicable to the case of a sloping
bed, a smaller value of rd (rd B 0.25) might be expected.

2.4 Bed changes and schematization of weak spots

For modelling changes of root-permeated clay or compacted bare clay beds due to
superficial erosion, it is assumed that eroded bed materials are transported by the flow like
loose particles. As the actual transport rate of bed materials is negligibly small compared to
the flow transport capacity for loose particles, all eroded materials are carried away without
sedimentating. This means that only bed erosion takes place and thus the bed level change
at a computational node is mainly related to the sediment transport rate at that node.
From Eq. (1), discretized bed changes for this particular situation read:
1 1 
zb ðxi ; tj þ DtÞ ¼ zb ðxi ; tj Þ  Egs ðxi ; tj ÞDt  w zb ðxiþ1 ; tj Þ  2zb ðxi ; tj Þ þ zb ðxi1 ; tj Þ
qs 2
ð18Þ
where Egs(xi, tj) is the erosion rate according to Eq. (1), and w is a (calibrated) bed
smoothing factor to avoid spurious bed changes in case of rapid scour development around
a wall jet.
The grassed dike slope considered herein is modelled as three distinct material layers in
successive order from the surface: grass root-permeated clay, bare clay, and sand. The
layer of grass root-permeated clay (or grass layer for brevity) extends from the slope
surface down to the level of the so-called grass working depth or grass base (see details in
Sect. 3). The remaining of the clay cover underneath the grass base and above the sand core
is the bare clay layer. The boundary lines of these three layers must be specified by users as
input for model computation.
Because of the multi-layer bed, a layer detecting algorithm is required, so as bed
changes of various materials can simultaneously be simulated. At a computational time

123
Nat Hazards (2012) 63:417–447 435

step, a layer thickness function is activated to detect the type of material being exposed to
the flow for each computational cell and thus allow appropriate sediment transport func-
tions to be used. In modelling erosion of the grass layer, this procedure is also required for
determining the eroded depth and the corresponding critical velocity for grass erosion at
this bed level.
Weak spots in grass and clay cover are unavoidable under field conditions, such as
caused by natural irregularity of grass growth, animal burrowing holes, cracks, and
damages. Mechanisms for failure initiation from these types of weak spots are different.
Here, only a localized weak (natural or damaged) spot in the grass slope is considered and
defined in the model as a segment where the critical velocity for grass erosion (determined
at the surface) within the segment is smaller than that of the surroundings. This also means
that, as the critical velocity for grass erosion decreases with the erosion depth, the actual
grass thickness (from the surface down to the grass base) of the weak spot is smaller than
that of the surroundings. In this way, a damaged spot (having a small top layer of grass
removed) can be schematized as a depressed surface profile portion and a natural weak spot
(weakly growth grass and often invisible at the surface) can be described as a portion with
a shallower grass base whilst the surface profile is unchanged (see Fig. 10 for illustration of
these two types of weak spot).

3 Model validation—grass erosion ensued from a weak spot

As aforementioned, a number of field overtopping experiments have recently been con-


ducted at sea-dikes in Groningen, the Netherlands (Akkerman et al. 2007). Herein, the test
case of grass erosion initiated from an artificial weak spot is used for model validation. For
the sake of comparison with various other cases throughout this study, this is addressed as
the reference case hereafter.
The tested dike section has an outer slope of 1/4 and an inner slope of 1/3. The length of
the inner slope was about 15 m covered by moderate grass according to the quality cat-
egories by VTV (2007). The tested wave conditions at the dike toe used for generating
overtopping events by the wave simulator were Hs = 2.0 m and Tp = 5.7 s. It was shown
that no erosion could be observed on the slope even at the average overtopping rate of 50 l/
s/m. Erosion only initiated when an initial damage of 1 m long and 5 cm deep on the slope
was introduced. The erosion depth after 6 h was reported about 10 cm with deep gully
formation behind the initial damage.
Wave overtopping events in the experiment were generated on the dike crest, whilst
these are simulated in the present model using the wave time-series at the inflow boundary
(possibly closest to the dike toe). Therefore, to avoid validation for the wave overtopping
hydrodynamics, we make use of the existing GWK (large-scale) experiments (see Geise-
nhainer and Oumeraci 2008), which fortunately bears a reasonable resemblance of the
experimental conditions to the field experiments (although the wave parameters at the dike
toe are somewhat different). The present model has already been validated against the
wave overtopping hydrodynamics from these large-scale experiments (see Tuan and
Oumeraci 2010 for details). A summary of the tested parameters, which were also used as
input for model computation, is given in Table 2. Only one necessary change was to raise
the upstream water level in order to meet the average discharge of 50 l/s/m like in the field
experiment. The measured 2-h wave time-series (TMA-spectra) at the nearest wave gauge
(about 40 m from the dike toe) were used as inflow wave conditions and re-circulated to
cover the entire 6-h storm duration.

123
436 Nat Hazards (2012) 63:417–447

Table 2 Reference case:


Parameter Unit Value
parameters for computation of
grass erosion
Hydraulics
Wave height Hm0 at toe m 0.90
Peak period Tp s 5.0
Surge level m 5.15
Storm duration h 6.0
Chezy coefficient: grass slope (foreshore) m0.5/s 30 (60)
Inland water level m Dry bed
Computed average discharge q l/s/m 50.0
Computed number of overtopping events Event 2,390
Computed largest volume per wave Vmax m3/m 2.24
Computed maximum depth-average m/s 6.2
velocity on slope umax
Dike geometry
Crest level m 5.80
Dike slopes: inner (outer) – 1/3 (1/4)
Standard grass base (working depth) m 0.30
Clay cover thickness m 0.80
Grass and clay properties
Clay cohesion c kN/m2 24
Mean root tensile strength tr,m 103 kN/m2 20.5
Reference RAR0 (at depth 2.5 cm) % 0.065
Root decay parameter b – 0.12
Mobilized strength coefficient lr – 0.60
Grass erosion coefficient Mgs kg/m2/s 5.0E-03

To simulate grass erosion for this case, Dutch dike grassland parameters for average
grass as addressed in Sects. 2.2.1 and 2.2.2 were applied (see also Table 2). The clay
parameters were from soil inspection data given in SBW report (Van der Meer et al. 2007a).
Figure 8 illustrates the depth-dependent critical velocity for erosion of this grass dike
slope, determined according to Eq. (9) for the average overtopping flow depth h = 0.10 m,
as function of the mobilized strength coefficient lr (=0.4, 0.5, and 0.6 within the possible
range, see Sects. 2.2.2 and Table 1) and the root decay parameter b (=0.12 and 0.30). It
appears that the critical velocity of grass with lr = 0.4–0.6 at the reference depth of
2.5 cm is approximately in the same range as that determined by the method of Verheij
(1995) (ugs,c is around 4.2 m/s for moderate-quality grass, see also Hoffmans et al. 2008).
At the surface, this value is expected to be higher, that is up to 6.8 m/s with lr = 0.6 (not
shown in Fig. 8), which is only exceeded by very few waves at the rate of 50 l/s/m during a
6-h storm. However, at 5 cm deep below the surface, the critical velocity for grass erosion
is markedly reduced. This explains why, in the experiment, grass erosion could not occur
until a weak spot was created by removing the top 5-cm grass layer from the slope.
Also, it is noted that the root decay parameter b has tremendous effects on the grass
strength through a gradient over depth of the root density. To describe the grass layer, the
so-called grass working depth or grass base is necessary and defined as a depth boundary at
which the soil strength or the critical velocity for erosion has reduced to that governed by
soil only (negligible effect from root reinforcement). The grass working depth is strongly
determined by b, for example, it is 40 cm for the present considered grass with b = 0.12

123
Nat Hazards (2012) 63:417–447 437

Fig. 8 Depth-dependent critical velocity of grass root-permeated clay

(fitted with the average grass according to VTV 2007), whilst only 20 cm (see also Fig. 8)
for grass with b = 0.30 (fitted with the data of Sprangers 1999 and Stanczak 2008).
The effect of b on grass erosion is further discussed in Sect. 4.4.
The model results of wave overtopping parameters such as the number of overtopping
events, largest overtopping discharge per wave Vmax, and maximum depth-averaged flow
velocity on slope umax are shown in Table 2. It is learnt that time sequencing and inter-
action between events complicate the hydrodynamics of wave overtopping on the slope. A
large overtopping volume wave does not necessarily induce a large overtopping flow
velocity on the slope. Also, the maximum velocity can occur at an arbitrary position on the
slope (i.e. not necessarily at the end of the inner slope as theoretically suggested),
depending on the initial momentum of events, event overlap, and the slope roughness.
For modelling grass erosion, a trapezoidal damaged spot in the grass slope is sche-
matized instead of a rectangular one as was artificially initiated in the field experiment
(1 m long by 5 cm deep). The model results of the time development of grass erosion
compared with the final situation (after 6 h) measured from the experiment are plotted in
Fig. 9. The root mobilized strength lr = 0.60 and the erosion coefficient for grass
Mgs = 5.10-3 (kg/m2/s) were found to give the best results. It follows that the model
predicted correctly the erosion depth (about 10 cm) but not the formation of the erosion
gully downstream. However, it was reported in the experiment that the gully was formed
due to the flow concentration behind the damaged spot, which cannot be described by the
present one dimensional model. An upstream migration of the scour hole (like head-cut
erosion) was instead predicted in this case. This is because the flow tends to accelerate at
the (following) upstream side of the erosion spot. The present model, which rests on the
principle of surface erosion, systematically predicts an upward progression of the erosion.
It can be noticed from Fig. 9 that the speed of erosion increases rapidly at the end,
which might be explained by the depth-dependent behaviour of the grass strength, that is,
the strength decreases with the erosion depth. Overall, the erosion of grass turf tends to
grow in surface area (length of erosion) much faster than in depth.

123
438 Nat Hazards (2012) 63:417–447

Fig. 9 Reference case: measured and computed grass erosion initiated from an artificial weak spot

The above artificially initiated weak spot of 5 cm deep can alternatively be schematized
as a hidden weak grass spot (viz. invisible at the surface) through reducing the grass
strength within this area to that at 5 cm below the surface (in the model this corresponds to
a spot with a 5-cm thinner grass layer). This approach can be used to schematize an
arbitrary natural weak spot in grass slopes. The final situation of grass erosion computed
with this alternative schematized spot, in comparison with that from the former case, is
shown in Fig. 10. It follows that the results produced from these two schematized spots are
comparable. Of course, the erosion is somewhat larger for the case of dug spot since the
overtopping flow was locally accelerated by the bed of the spot itself.

4 Numerical experiments

In this section, the characteristics of grass erosion induced by wave overtopping are
investigated through a series of numerical experiments based on several test scenarios. The
model parameters validated against the data from the field experiments as well as the
material (grass and clay) properties as described in Sect. 3 are used for all simulations.

4.1 Effect of incident wave period and seaward slope steepness

Despite the same given average overtopping rate, different conditions of wave overtopping
at the upper end of the inner slope may result in a different extent of grass erosion. In this
study, we consider modifications of the wave overtopping conditions at the upper end of
the inner slope, compared to the reference test case in Sect. 3 (see Table 2), due to a more

123
Nat Hazards (2012) 63:417–447 439

Fig. 10 Comparison of erosion between schematized weak spots: (a) trapezoidal dug spot, (b) hidden weak
grass spot

Table 3 Wave overtopping characteristics (storm surge duration of 6 h)


Parameter Scenario

Reference case Outer slope 1/6 Tp = 7.0 s

Dike slope: inner (outer) 1/3 (1/4) 1/3 (1/6) 1/3 (1/4)
Wave height Hs (m) 0.90 0.90 0.90
Wave period Tp (s) 5.0 5.0 7.0
Average overtopping discharge q (l/s/m) 50 50 50
Storm surge level (m) 5.15 5.39 4.85
Number of overtopping events (event) 2390 2070 1910
Relative total overtopping time Fcd (-) 0.46 0.57 0.38
Largest volume per wave Vmax (m3/m) 2.24 2.50 3.28
Maximum velocity at crest umax, crest (m/s) 4.30 4.00 4.93
Maximum erosion depth after 6 h (m) 0.13 0.18 0.10

gentle seaward slope (1/6 vs. 1/4) and a longer wave (Tp = 7.0 s vs. 5.0 s). It is noted that
these modifications are ensued by upstream conditions only (i.e. either by incident wave or
upstream geometric conditions), the inner slope remains unchanged. This means the
investigation focuses on the effect of the initial momentum of wave overtopping at the start
of the inner slope. Since the average discharge q = 50 l/s/m was kept the same, the storm
surge level in these cases had to be adjusted to 5.39 and 4.85 m (5.15 m in the reference
case), respectively.
The computed wave overtopping characteristics for these scenarios are given in
Table 3. The computed results of grass erosion after 6 h in comparison with the reference
case are shown in Fig. 11.

123
440 Nat Hazards (2012) 63:417–447

Fig. 11 Effect of seaward slope steepness and wave period on grass erosion, average overtopping discharge
q = 50 l/s/m

Interestingly, under the same average overtopping rate and the same steepness of the
outer slope, longer waves induce less grass erosion, whilst more severe erosion is expected
if the outer dike slope is more gentle. At first, this sounds contradictory. However, a closer
inspection of Fig. 11 and the wave overtopping characteristics given in Table 3 illustrates
that the extent of grass erosion is clearly proportional to the total overtopping duration at
the dike crest (expressed through Fcd, see Tuan et al. 2006). In the three considered cases,
the time fraction that the bed shear stress induced by wave overtopping is in excess of the
critical shear stress prevails, the overtopping duration thus plays an important role.
Because wave overtopping events are highly sensitive to the change of the water level or
the crest freeboard (see Tuan and Oumeraci 2010), the above trends of grass erosion may
be explained as follows.
In comparison with the reference case, with longer waves, the design dike height must
be increased (equivalent to a decrease of the storm surge level in the computation) in order
to maintain the same average discharge. Therefore, although longer waves may induce a
number of overtopping events of larger volume, the number of overtopping events and the
total overtopping duration, however, become smaller. In contrast, for the same average
overtopping discharge, a reduction in the design dike height that would result from a more
gentle outer dike slope under the same wave conditions (equivalent to an increase in the
storm surge level in the computation) will increase the total overtopping time.

4.2 Effect of inner dike slope steepness

The geometry of the inner dike slope modifies wave overtopping on the slope significantly.
Grass erosion is therefore a function of the slope geometry. In this case, we investigate the

123
Nat Hazards (2012) 63:417–447 441

Fig. 12 Grass erosion on a gentle inner slope of 1/4

effect of the slope steepness on the extent of grass erosion. The inner slope gradient used
here is 1/4 instead of 1/3 in the reference case, and the same dug spot of 1 m 9 5 cm is
schematized. All other conditions (crest height, outer slope steepness, and wave parame-
ters) are kept the same as the reference case.
The wave overtopping flow velocity is smaller as the slope is more gentle. In this case,
the model computed the maximum velocity (occur on the slope during the same wave as in
the reference case) umax is 5.7 m/s, which is smaller than that in the reference case
(umax = 6.2 m/s). The predicted maximum erosion depth after 6 h is about 5 cm (see
Fig. 12), which is only half of that compared to the reference case with the slope of 1/3.

4.3 Grass erosion on ideally uniform grass slopes

Irrespective of any (artificial or natural) weak spots in the inner slope, ideally uniform
grass slopes are used to investigate locations on slope that are most vulnerable to erosion
due to wave overtopping.
It is convenient to make use of the dike geometry and wave conditions of the reference
case (see Table 2). The inner slope is ideally uniform without any weak spots. Therefore,
in order to observe erosion on the slope, either much more severe wave overtopping is
required or the strength of the entire grass slope has to be reduced. These two scenarios of
erosion are addressed in the following.
In the first case, that is keeping the same strength (moderate-quality grass) of the grass
slope, the overtopping rate is raised from q = 50 l/s/m (surge level at 5.15 m) to
q = 112 l/s/m, which corresponds to an increase of the surge level to 5.40 m.

123
442 Nat Hazards (2012) 63:417–447

In the second case, that is keeping the same overtopping rate q = 50 l/s/m, the strength
of the entire grass slope is assumed to reduce to that at 5 cm below the surface (equivalent
to the grass strength at the surface of the damaged spot in the reference case).
The model predictions of the grass erosion profiles for these two cases are shown,
respectively, in Fig. 13a and b, showing that the maximum superficial erosion depth after
6.0 h occurs at the end of the inner slope is about 10 cm in both cases. As quantitative data
of grass and soil characteristics are unavailable for model computation, measured grass
erosion profiles from a field experiment of a weakly growth grass slope in Vietnam (WRU-
CE 2009) are also given in Fig. 13c for qualitative comparison of the main erosion features.
It appears in general that the erosion depth is increasingly larger in the direction down the
slope. At the end of the slope (where the slope makes a transition with the ground), a deep
scour hole is present, especially for the case of weak (thin grass) slope as shown in Fig. 13b.
This noticeable feature predicted by the model is in agreement with observations from field
experiments as shown in Fig. 13c and those addressed in Sect. 2.3 (Fig. 7). This result
confirms the necessity and applicability of the assumption about the analogy with the flow
structure in turbulent wall jets of the wave overtopping flow on grass slopes.
It is noted that the bed on the inland side is kept dry in all the above simulations. In
reality, the scour at the end of the inner slope may not occur when the inland water level is
sufficiently high. This is confirmed by both field experiments in Vietnam (WRU-CE 2009)
and the present model simulations.

4.4 Effect of root decay parameter

The root density decay parameter b is an important parameter describing the variation of
the grass strength in depth. A larger b value means a stronger rate of root decay, which
results in a smaller grass working depth or a weaker grass cover (see Sect. 2.2.1 and also
Fig. 8). It is noted that all computations so far have been performed with root decay
parameter b = 0.12 derived from the moderate grass category by VTV (2007). To
investigate the effect of b on the extent of erosion, the value b = 0.30 (according to the
data by Stanczak 2008) is used for the computation of grass erosion for the reference case.
The model results shown in Fig. 14 indicate that, in comparison with the reference case,
grass erosion with b = 0.3 accelerates very fast as the grass strength reduces rapidly over
depth. As a result, a much larger extent of grass erosion is expected as predicted by the
model. After 2.5 h, the scour already reaches the grass base (at depth of 30 cm), whilst in
the reference case, only half of this value is reached after 6 h (see Fig. 9).

4.5 Breach initiation time

In the breaching of grassed sea-dikes, the phase of breach development is defined to start as
soon as the sand core is exposed to the flow attack. The preceding period, which causes a
breach to develop, is commonly addressed as the breach initiation phase. The time scale for
breach initiation is significantly larger than that for the breach development (after the
breach initiation phase up to the final breach). It usually takes many hours for a breach to
initiate, whilst just half an hour to several hours for complete breaching.
Given the data that have been validated against the field experiments in previous sections,
it is now possible to anticipate the time required for a breach to fully initiate, that is until the
core is exposed. In this study, we limit ourselves to the prediction for the reference case only,
that is, a moderate-quality grass dike slope with a natural weak spot (1 m 9 5 cm) in the inner
dike slope subject to wave overtopping with the mean rate q = 50 l/s/m.

123
Nat Hazards (2012) 63:417–447 443

Fig. 13 Erosion of ideally uniform grass slopes: (a) q = 112 l/s/m on moderate-quality grass, (b) q = 50 l/
s/m on reduced-strength grass, (c) measured erosion profiles of a weakly growth grass slope after 6 h of
successive testing with q = 40–70 l/s/m (WRU-CE 2009)

If breach initiation can be subdivided into two successive processes, grass turf erosion
and (bare) clay cover erosion, the breach initiation time is thus determined as the sum of
the erosion times of these two processes. It is observed from field experiments and also
from the present model simulations that during the process of grass erosion, the erosion
channel expands more rapidly in length than in depth. Hence, it is plausible to assume that
the grass turf fails as a whole as soon as the erosion depth has reached the maximum
working depth.

123
444 Nat Hazards (2012) 63:417–447

Fig. 14 Effect of root density decay parameter b on grass erosion

It can be deduced from the previous computation for the reference case (see Sect. 3) that
the failure time for the grass turf (working depth up to 40 cm) is about 7.0 h. The failure
time for the clay cover also depends on its thickness, the model is therefore applied to
simulate erosion of a bare clay slope (also moderate-quality clay, see Table 2) under the
same overtopping rate q = 50 l/s/m. The model predictions of clay erosion are shown in
Fig. 15 together with two practical thicknesses of the clay cover (1.0 and 2.0 m). In this
simulation, the grass turf was assumed to remain on the crest of the dike as usually
observed in the field, which slowed down the upstream expansion of the erosion channel. It
follows from the computation that, on average, every 1.0 m in thickness of the clay cover
can withstand about 2 h.
On the whole, we can estimate that the breach initiation time in this particular case is
about 9 and 11 h for grass slope dikes with the clay cover thickness of 1.0 and 2.0 m,
respectively.
It should be stressed that the breach initiation time estimated herein is based on the
cause by superficial erosion and therefore should be considered as an upper boundary. In
reality, other phenomena like flow concentration and mass failure mechanisms may occur
simultaneously and thus shorten the breach initiation time considerably.

5 Conclusions

This paper presents a numerical model of grass erosion on the landward sea-dike slope
induced by wave overtopping. To better describe grass erodibility, the concept of depth-
dependent critical velocity for grass erosion is introduced, in which the effect of root

123
Nat Hazards (2012) 63:417–447 445

Fig. 15 Erosion of bare clay cover, q = 50 l/s/m

reinforcement in the subsoil is implemented with an improved approach for determining


the root cohesion. Besides the root area ratio, the strength of a grass cover can now be
characterized with several other new parameters relating to the root characteristics such as
the root density decay parameter and the root mobilized strength coefficient. Wave
overtopping on grass slopes under prototype conditions is highly aerated and turbulent. For
modelling grass erosion under this extraordinary flow conditions, the bed shear stress based
on the depth-averaged parameters as applied for ordinary open-channel flows appears to be
underestimated. A refinement of the computed depth-averaged wave overtopping flow field
is therefore necessary for a proper determination of the bed shear stress. To this end, the
present model assumes an analogy between the flow structure in wave overtopping on grass
slopes and that in turbulent wall jets, whereby the bed shear stress can be determined in
connection with the flow turbulence.
The developed model has been validated against the data of grass erosion from field
wave overtopping experiments conducted recently in the Netherlands. In general, the
extent of grass erosion is fairly well predicted by the model. Grass erosion initiated from a
naturally weak or damaged spot in the slope can effectively be modelled. Moreover, the
formation of a scour hole at the toe of the landward slope as commonly observed from field
experiments is successfully simulated.
A series of numerical experiments are carried out to study the influence of several key
parameters on grass erosion such as steepness of outer and inner slopes, wave period, and
root density decay parameter. It is shown that, though under the same wave overtopping
rate, grass erosion is still strongly affected by these parameters. Also, if the superficial
erosion processes of grass turf and clay cover are assumed responsible for breach initiation
of grassed sea-dikes, the model can be applied to tentatively anticipate the breach initiation
time.
As the model performance has been examined with one dataset of Dutch dike grassland
only, more quantitative data of grass erosion of other grass species and soil conditions are
needed for model validation and improvement.

123
446 Nat Hazards (2012) 63:417–447

In general, more fundamental studies on the underlying physical principles are neces-
sary for the future development of predictive tools for grassed sea-dike erosion due to wave
overtopping, which in essence are the erodibility of compact and root-permeated soils and
the detailed flow structure of bubbly turbulent wave overtopping on grassed slopes. A
model extension to two dimensions (2DH) is also necessary to resolve the issue of flow
concentration.

Acknowledgments The study was funded by the Alexander von Humboldt (AvH) research foundation in
collaboration with LWI-Braunschweig Technical University, Germany. Sincerely thanks are due to WRU-
CE project for sharing data of field wave overtopping experiments in Vietnam.

References

Akkerman GJ, Gerven Van KAJ, Schaap HA, van der Meer JW (2007) Wave overtopping erosion tests at
Groningen sea dyke. Report of ComCoast workpackage 3: development of alternative overtopping-
resistant sea defences, phase 3, Rijkswaterstaat, Delft, 99 pp
CEM (2002) Coastal Engineering manual. US Army Corps of Engineers, Engineer Manual 1110-2-1100,
Washington, DC, USA
Chanson H (2006) Air bubble entrainment in hydraulic jumps. Similitude and scale effects. Report CH57/05,
Department of civil engineering–University of Queensland, Australia, 117 pp
Chanson H, Brattberg T (2000) Experimental study of the air-water shear flow in a hydraulic jump. Int J
Multiphase Flow 26:583–607
Cheng H, Yang X, Liu A, Fu H, Wan M (2003) A study on the performance and mechanism of soil-
reinforcement by herb root system. In: Proceedings of 3nd international conference vetiver and
exhibition, Guangzhou, China, 7 pp
De Baets S, Poesen J, Reubens B, Wemans K, De Baerdemaeker J, Muys B (2008) Root tensile strength and
root distribution of typical Mediterranean plant species and their contribution to soil shear strength.
Plant Soil 305:207–226
Dean RG, Rosati JD, Walton TL, Edge BL (2010) Erosional equivalences of levees: steady and intermittent
wave overtopping. Ocean Eng 37:04–113
EurOtop (2007) Wave overtopping of sea defences and related structures: assessment manual. Environment
Agency UK/Expertise Netwerk Waterkeren NL/Kuratorium fur Forschung im Kusteningenieurswesen,
DE
Geisenhainer P, Oumeraci H (2008). Seadike breach initiation and development: large-scale experiments in
GWK. Report Floodsite T06-08-12
Hager WH (1992) Energy dissipators and hydraulic jumps. Kluwer Academic, Water Science and Tech-
nology Library, Dordrecht, the Netherlands, 288 pp
Hewlett HWM, Boorman LA, Bramley ME (1987) Design of reinforced grass waterways. Construction
industry research and information association, Rep. No. 116, London
Hoffmans GJCM, Verheij HJ (1997) Scour manual. Balkema, Rotterdam 205 pp
Hoffmans G, Akkerman GJ, Verheij HJ, Van Hoven A, Van der Meer JW (2008) The erodibility of grassed
inner dike slopes against wave overtopping. In: Proceedings of 31st international conference coastal
engineering, ASCE, Hamburg, Germany
Kruse GAM (1994a) Initial report on tests of grass banks (original title in Dutch: Meetverslag grastalud-
proeven). Soil Mechanics Delft, report CO-334430/17, Delft
Kruse GAM (1994b) First analysis of Deltagoot tests on a grass bank (original title in Dutch: Eerste analyse
van Deltagootproeven op een grastalud). Soil Mechanics Delft, report CO-334430/25, Delft
Mirtskhoulava TSE (1991) Scouring by flowing water of cohesive and non-cohesive beds. Hydraulic Res
29(3):341–354
Ohtsu IO, Yasuda Y, Awazu S (1990) Free and submerged hydraulic jumps in rectangular channels. Report
of Research Institution of Science and Technology, No. 35, Nihon University, Japan
Operstein V, Frydman S (2000) The influence of vegetation on soil strength. Ground Improv 4:81–89
Partheniades E (1965) Erosion and deposition of cohesive soils. J of Hydr Div ASCE 91(HY1)
Pollen N, Simon A (2005) Estimating the mechanical effects of riparian vegetation on stream bank stability
using a fiber bundle model. Water Resour Res 41:1–11
Ribberink JS (1998) Bed-load transport for steady flows and unsteady oscillatory flows. Coast Eng 34:59–82

123
Nat Hazards (2012) 63:417–447 447

Robinson KM (1996) Gully erosion and headcut advance. Doctoral dissertation, Oklahoma State University,
Stillwater, Oklahoma, USA
Schüttrumpf H, Oumeraci H (2004) Learning from seadike failures. PIANC Bulletin No. 117, pp. 47–60
Smith GM (1993) Grass dikes–grass erosion, residual strength and wave overtopping (original title in Dutch:
Grasdijken: graserosie, reststerkte en golfoverslag). Preliminary report on large-scale study H1565,
Delft Hydraulics, Delft
Sprangers JTCM (1999) Vegetation dynamics and erosion resistance of seadyke grassland. Doctoral dis-
sertation, Wageningen Agriculture University, Wageningen, 167 pp
Stanczak G (2008) Breaching of sea dikes initiated from the seaside by breaking wave impacts. Doctoral
dissertation, Leichtweiss-Institute, Technical University of Braunschweig, Germany
Tosi M (2007) Root tensile strength relationships and their slope stability implications of three shrub species
in the Northern Apennines (Italy). Geomorphology 87:268–283
Tuan TQ (2007) Seasonal breaching of coastal barriers. Doctoral dissertation, Delft University of Tech-
nology, Delft, the Netherlands, 179 pp
Tuan TQ, Oumeraci H (2010) A numerical model of wave overtopping on sea-dikes. Coast Eng
57(8):757–772
Tuan TQ, Verhagen HJ, Visser PJ, Stive MJF (2006) Wave overwash at low-crested beach barriers. Coast
Eng J 48(4):371–393
Van den Bos W (2006) Erodibility of a grass revetment during wave overtopping (in Dutch). MSc. thesis,
Delft University of Technology, Delft, the Netherlands
Van der Meer JW (2008) Erosion strength of inner slopes of dikes against wave overtopping: preliminary
conclusions after two years of testing with the Wave Overtopping Simulator. ComCoast, Rijkswaterstaat-
Dienst Weg- en Waterbouwkunde, Rijkswaterstaat-Waterdienst, Projectbureau Zeeweringen
Van der Meer JW, Bernardini P, Snijders W, Regeling HJ (2006) The wave overtopping simulator. In:
Proceedings of 30th international conference coastal engineering, ASCE, San Diego
Van der Meer JW, Verheij HJ, Lindenberg J, Van Hoven A, Hoffmans GJCM (2007a) Wave overtopping
and strenght of inner slopes of dykes (in Dutch: Golfoverslag en sterkte binnentaluds van dijken).
Rapport Predictiespoor SBW, Infram, WL|Delft Hydraulics, Geodelft, report 05i028
Van der Meer JW, Bernardini P, Steendam GJ, Akkerman GJ, Hoffmans GJCM (2007b) The wave over-
topping simulator in action. Proceedings of Coastal Structures, Venice, Italy
Van der Meer JW, Hardeman B, Steendam GJ, Schüttrumpf H, Verheij H (2010) Flow depths and velocities
at crest and inner dike slope of a dike, in theory and with wave simulator. In: Proceedings of 32nd
international conference coastal engineering, ASCE, Shanghai, China
Verheij HJ (1995) Investigation of the strength of a grass cover upon river dikes (in Dutch). Report Q1878,
Deltares, Delft
VTV (2007) The safety of the primary defences in the Netherlands (in Dutch: Voorschrift Toetsen op
Veiligheid Primaire Waterkeringen). DGW/WV/2007/1047, Ministerie van Verkeer en Waterstaat,
472 pp
Winterwerp JC, Van Kesteren WGM (2004) Introduction to the physics of cohesive sediment in the marine
environment. Elsevier, Amsterdam
WRU-CE (2009) Field wave overtopping tests on grassed sea-dikes at Nam Dinh province–Vietnam (in
Vietnamese). WRU-CE project report, Water Resources University
Wu TH, McKinnell WP III, Swanston DN (1979) Strength of tree roots and landslides on Prince of Wales
Island, Alaska. Can J Geotech Res 16(l):19–33
Young MJ (2005) Wave overtopping and grass cover layer failure on the inner slope of dikes. Master thesis,
UNESCO-IHE, Delft

123

You might also like