You are on page 1of 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/251427439

Coupled simulations of fluvial erosion and mass wasting for cohesive river
banks

Article  in  Journal of Geophysical Research Atmospheres · August 2007


DOI: 10.1029/2006JF000722

CITATIONS READS
110 270

3 authors:

Stephen E. Darby Massimo Rinaldi


University of Southampton Universidad de Sevilla
49 PUBLICATIONS   2,252 CITATIONS    230 PUBLICATIONS   4,068 CITATIONS   

SEE PROFILE SEE PROFILE

Stefano Dapporto
Regione Toscana
22 PUBLICATIONS   465 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Geomorphic Change Detection View project

REFORM - REstoring rivers FOR effective catchment Management View project

All content following this page was uploaded by Massimo Rinaldi on 21 May 2014.

The user has requested enhancement of the downloaded file.


JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, F03022, doi:10.1029/2006JF000722, 2007
Click
Here
for
Full
Article

Coupled simulations of fluvial erosion and mass wasting


for cohesive river banks
Stephen E. Darby,1 Massimo Rinaldi,2 and Stefano Dapporto3
Received 15 November 2006; revised 19 April 2007; accepted 25 May 2007; published 25 August 2007.

[1] The erosion of sediment from riverbanks affects a range of physical and ecological
issues. Bank retreat often involves combinations of fluvial erosion and mass wasting, and
in recent years, bank retreat models have been developed that combine hydraulic erosion
and limit equilibrium stability models. In related work, finite element seepage analyses
have also been used to account for the influence of pore water pressure in controlling the
onset of mass wasting. This paper builds on these previous studies by developing a
simulation modeling approach in which the hydraulic erosion, finite element seepage, and
limit equilibrium stability models are, for the first time, fully coupled. Application of the
model is demonstrated by undertaking simulations of a single flow event at a single study
site for scenarios where (1) there is no fluvial erosion and the bank geometry profile
remains constant throughout, (2) there is no fluvial erosion but the bank profile is
deformed by simulated mass wasting, and (3) the bank profile is allowed to freely deform
in response to both simulated fluvial erosion and mass wasting. The results are limited in
scope to the specific conditions encountered at the study site, but they nevertheless
demonstrate the significant role that fluvial erosion plays in steepening the bank profile or
creating overhangs, thereby triggering mass wasting. However, feedbacks between the
various processes also lead to unexpected outcomes. Specifically, fluvial erosion also
affects bank stability indirectly, as deformation of the bank profile alters the hydraulic
gradients driving infiltration into the bank, thereby modulating the evolution of the pore
water pressure field. Consequently, the frequency, magnitude, and mode of bank erosion
events in the fully coupled scenario differ from the two scenarios in which not all the
relevant bank process interactions are included.
Citation: Darby, S. E., M. Rinaldi, and S. Dapporto (2007), Coupled simulations of fluvial erosion and mass wasting for cohesive
river banks, J. Geophys. Res., 112, F03022, doi:10.1029/2006JF000722.

1. Introduction range associated with low-energy UK rivers [Walling et


al., 1998, 1999] and the upper end corresponding to incised
[2] The erosion of sediment from riverbanks is a key channel systems [Wasson et al., 1998; Simon and Darby,
factor affecting a range of physical and ecological issues in 2002; Simon and Rinaldi, 2006]. With such a significant
the fluvial environment. These include the establishment of fraction of bank-derived material within the alluvial sedi-
river and floodplain morphology and their associated habitats mentary system, knowledge of the rates, patterns and
[Thorne and Lewin, 1979; Darby and Thorne, 1996; Millar, controls on bank erosion events is necessary for a complete
2000; Goodson et al., 2002; Eaton et al., 2004], turbidity understanding of the fluvial sediment transport regime.
problems [Bull, 1997; Green et al., 1999; U.S. Environmental [3] Given the importance of bank-derived sediment
Protection Agency, 2000], as well as nutrient and contami- within the fluvial sediment transfer system, it is not sur-
nant dynamics [Marron, 1992; Reneau et al., 2004]. prising that bank erosion has been the subject of much
Indeed, the contribution of bank-derived materials to fluvial research. A particular focus of recent work has been the
sediment budgets may be higher than previously thought. application of slope-stability modeling to quantify the role
Bank-derived sediments typically contribute between about and relative influence of the factors that control mass wast-
37% and 80% of the total suspended sediment yield ing. Such studies have focused on the respective roles of bank
emanating from catchments, with the lower end of this shape [Osman and Thorne, 1988; Darby et al., 2000],
riparian vegetation [Abernethy and Rutherfurd, 2000; Simon
1
School of Geography, University of Southampton, Southampton, UK. and Collison, 2002], as well as bank and channel hydrology
2
Department of Civil Engineering, University of Florence, Florence, [Rinaldi and Casagli, 1999; Casagli et al., 1999; Simon et al.,
Italy. 1999; Rinaldi et al., 2004]. Although fewer studies have been
3
Regione Toscana, Sistema Regionale di Protezione Civile, Florence,
Italy. concerned with the process of fluvial erosion (i.e., the
removal of bank sediments by the direct action of the flow),
Copyright 2007 by the American Geophysical Union. notable recent contributions have started to address the issue
0148-0227/07/2006JF000722$09.00

F03022 1 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

influence of dynamically interacting processes on the ero-


sion of bank sediments. This is achieved by coupling a
hydraulic erosion model with a finite element seepage
analysis and limit equilibrium stability methods to address
transient mass wasting triggered by bank profile deformation
and/or variations in bank pore water pressures [Dapporto,
2003; Dapporto and Rinaldi, 2003]. The new method
requires the finite element mesh used in the seepage
analysis to be adapted to the deforming bank profile,
thereby allowing for the possibility of feedbacks to occur
between simulated bank stability, fluvial erosion and pore
water pressure fields. Having outlined the simulations, we
go on to analyze the results to explore the consequences of
some of these key feedbacks.

2. Methods
[6] In this paper we seek to develop insight into the
dynamics of bank erosion by developing simulations that
couple a fluvial erosion model with a finite element seepage
analysis and limit equilibrium methods for evaluating bank
stability. The approach involves applying these three sub-
models at each of a series of discrete time steps throughout a
flow event hydrograph (Figure 1). This study therefore
builds on previous models [Darby and Thorne, 1996;
Langendoen, 2000; Simon et al., 2006] which consider
the combined effects of fluvial erosion and mass wasting,
but differs in that a finite element seepage analysis is also
included to account for the influence of pore water pressure
Figure 1. Logic diagram showing the computational in controlling the onset of mass wasting. Some previous
sequence used to couple the finite element seepage, bank studies [Rinaldi and Casagli, 1999; Casagli et al., 1999;
stability, and fluvial erosion analyses. Simon et al., 1999; Rinaldi et al., 2004] have analyzed this
latter effect, but only in the context of river banks that are
of quantifying entrainment thresholds and process rates not deformed by fluvial erosion. To our knowledge, ours is
[Lawler et al., 1997a; Simon et al., 2000; Dapporto, 2001; therefore the first study to fully couple all three relevant
Lawler, 2004, 2005]. process submodels, though one study [Simon et al., 2006]
[4] Nevertheless, predictions of bank erosion rates remain has combined measured pore water pressure data with
poor [Darby et al., 1998; Mosselman, 1998], with two key fluvial erosion and bank stability models. This section
issues standing out as likely limiting factors. First, existing details the submodels and their parameterization, focusing
studies utilize data collected at relatively coarse timescales, on how the finite element seepage and bank stability
understandably so given the problems of working during analyses are applied in a form that can accommodate the
floods, but preevent versus postevent studies cannot resolve effects of fluvial erosion.
process thresholds, timing and rates [Lawler, 2004, 2005]. 2.1. Study Site Description
Second, previous research tends to view individual processes [7] The study site investigated herein, the Sieve River, is
in isolation, but bank retreat is the net result of interacting located within the mountain belt of the Northern Apennines
processes [Wolman, 1959; Thorne, 1982; Lawler et al., (Figure 2a), and has been described in previous papers [e.g.,
1997b]. Although the adoption of approaches that break Rinaldi et al., 2004]. The Sieve River (840 km2) is the main
systems down into manageable components is a logical first tributary of the Arno (Tuscany, Central Italy) and at the
step, interactions and feedbacks between processes in non- study site presents a single-thread, sinuous pattern, with bed
linear systems can lead to outcomes that are not always sediments predominantly composed of gravel, and a chan-
predictable a priori. In short, viewing bank processes in nel gradient of 0.0030. Mean daily discharge at the study
isolation is unrealistic, introducing the possibility that the reach, which is located close to the basin outlet (830 km2),
conclusions derived from such studies may be biased. This is 15 m3 s1, while the 2-year return period peak flow (Q2)
limitation is starting to be addressed by modeling, with a is 410 m3 s1. The eroding banks have an average height
particular focus on coupling fluvial erosion and mass ranging between 5.00 and 5.50 m and are layered, with the
wasting models [Darby and Thorne, 1996; Darby et al., bank material stratigraphy (defined by grain size analyses
1998; Langendoen, 2000; Simon et al., 2006]. However, as combined with the results of one static and six dynamic
yet no study has coupled fluvial erosion, seepage and penetration tests) arranged as shown in Figure 2b. Thus the
stability submodels in a fully integrated analysis. bank in question comprises a cohesive upper part overlying
[5] To address this gap, we herein develop a new simu- a gravel toe as is commonly the case in upland and
lation approach that seeks to couple hydraulic and geotech- piedmont zones in Europe and elsewhere, but it is distinct
nical simulations in a way that facilitates analysis of the

2 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 2. Sieve River study reach showing (a) location of the reach within the Sieve basin (area 1 is
lacustrine, fluvio-lacustrine, and alluvial deposits of the Upper Pliocene and Quaternary; area 2 is
arenaceous-marly and calcaleous flysch formations of the Cretaceous and Miocene) and (b) arrangement
of the bank materials observed within the study reach.
from the fine-grained bank settings that are more usually extent that the parameter values can be estimated accurately.
associated with lowland environments and which have been [10] For granular (noncohesive) sediments, critical shear
the subject of related research [e.g., Simon et al., 1999, stress can be estimated by applying the same methods that
2000; Simon and Collison, 2002]. are used to predict the entrainment of bed material, albeit
[8] Of numerous cross sections surveyed along the reach, with modifications to take into account the effect of bank
a representative section was selected for further analysis by angle on the downslope component of the particle weight
choosing the profile with the rate of bank retreat that was [Lane, 1955] and the influence of packing and cementing
closest to the mean rate of bank retreat observed along the [e.g., Millar and Quick, 1993; Millar, 2000]. However,
reach as a whole. On the basis of the Q2 discharge values there are no theoretical or empirical methods to determine
referred to above, the reach has a rather high value of stream kd for granular sediments. Theoretical determination of
power per unit bed area (350 W/m2). Mean rates of bank critical shear stress for cohesive materials is even more
erosion for most flows are nevertheless only moderate complex, given that it is widely recognized that it depends
(0.40 m/yr during February 1996 to February 2000), since on several factors, including (amongst others) clay and
the study site is located in an almost straight reach, with organic content, and the composition of interstitial fluids
fluvial erosion, and shear and cantilever-type failures all [Arulanandan et al., 1980; Grissinger, 1982]. However,
contributing to the observed retreat. The availability of data, recent studies have deployed in situ jet-testing devices [e.g.,
the presence of a range of interacting bank processes, and its Hanson, 1990; Hanson and Simon, 2001] to obtain direct
location within a straight reach (which simplifies the re- measurements of both bank erodibility parameters [e.g.,
quired hydraulic computations) makes the Sieve study site Dapporto, 2001]. Bearing these difficulties in mind, we
ideal for this research. estimated the values of the erodibility parameters (kd and t c)
in equation (1) as follows:
2.2. Fluvial Erosion Model [11] 1. Loose gravel. At the Sieve River study site a wedge
[9] Fluvial bank erosion rates can be quantified using an of loose gravel is located at the bank toe prior to a flow
excess shear stress formula such as [Partheniades, 1965; event. We estimated the critical shear stress (t c = 27.9 Pa, see
Arulanandan et al., 1980] Table 1) for this material using Lane [1955]. We assumed
that after this critical shear stress was exceeded, the material
e ¼ kd ðt b  t c Þa ð1Þ was removed from the toe exposing the basal packed gravel
(layer 2) to the direct action of the flow.
where e (m/s) is the fluvial bank erosion rate per unit time and [12] 2. Packed gravel. For this material, it is not possible to
unit bank area, t b (Pa) is the boundary shear stress applied by apply the Lane [1955] equation because it is restricted to
the flow, kd (m3/Ns) and t c (Pa) are erodibility parameters conditions (not met for this layer) in which the bank angle is
(erodibility coefficient, kd, and critical shear stress, t c) and a less than the friction angle of the material. We therefore
(dimensionless) is an empirically derived exponent, gener- estimated the critical shear stress (t c = 32.5 Pa, see Table 1)
ally assumed to equal 1.0. Although equation (1) appears using [Millar, 2000]
simple, in practice it is necessary to define the erodibility
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
parameters and boundary shear stress. These are all highly
tc sin2 q
variable, helping to explain why observed rates of fluvial ¼ 0:048 tan f0 1 2 0 ð2Þ
erosion range over several orders of magnitude [Hooke, g ðs  1ÞD50bank sin f
1980]. Consequently fluvial erosion is predictable only to the

3 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Table 1. Geotechnical and Hydraulic Characteristics of the Bank Materialsa


Sediment Layers
Parameter 1 2 3 Notes
Effective cohesion, c0, kPa n/a n/a 2.0 Data based on a single triaxial test
Effective friction angle, f0, deg n/a n/a 35 Data based on a single triaxial test
Matric suction angle, fb, deg n/a n/a 15 – 35 See text for explanation
Unit weight, g, kN/m3 n/a n/a 15.1 – 19.3 Data based on a single soil sample removed from
layer 3 during dry (low value) and saturated
(high value) conditions, respectively
Porosity, n, % 35.0 35.0 39.5 Data based on single samples removed from
each layer
Saturated hydraulic conductivity, ksat, m/s 6.0  104 1.0  104 1.4  106 ± 7.8  107 Values for layer 3 are the mean and standard
deviation (quoted range) of 9 Amoozemeter tests
and 3 open single ring infiltrometer tests. See text
for details concerning layers 1 and 2
Critical shear stress, t c, Pa 27.9 32.5 1.8 ± 0.7 Values for layer 3 are the mean and standard
deviation (quoted range) of 4 jet tests
6
3
Erodibility coefficient, kd, m /Ns n/a 1.3  10 5.4  106 ± 2.6  106 Values for layer 3 are the mean and standard
deviation (quoted range) of 4 jet tests
a
The positions and thicknesses of the sediment layers 1, 2, and 3 are illustrated in Figure 3; n/a means not applicable.

where D50bank (0.0187 m) is the median grain size of the This procedure is clearly an idealization of the actual near-
bank material, f0 (82°) is an equivalent friction angle of bank shear stress distribution, but was used simply to
material, estimated from the steepest angle that the bank demonstrate the methodological approach required to com-
forms at the bankfull waterline [Millar, 2000], q (70°) is bine the fluvial erosion and mass-wasting process models.
the bank angle, g is the unit weight of water, assumed here Having obtained the near-bank shear stress, the bank profile
to be 9810 N/m3, and s is the specific gravity of the was deformed by fluvial shear erosion estimated using
sediment, assumed here to be 2.65. The erodibility coef- equation (1).
ficient (kd = 1.3  106, see Table 1) was then determined
via model calibration (i.e., by forcing best agreement 2.3. Modeling Saturated and Unsaturated Flow
between calculated and measured bank toe retreat). [15] Changes in pore water content and pressures are
[13] 3. Cohesive portion of the bank. For these materials, recognized as one of the most important factors controlling
both erodibility parameters were estimated by averaging the the onset and timing of bank collapse [Thorne, 1982;
results from a series of four jet tests (Table 1). As is commonly Springer et al., 1985]. Pore water has at least four main
the case [e.g., Hanson and Simon, 2001; Dapporto, 2001] effects: (1) reducing shear strength (under conditions of
error estimates (Table 1) indicate a substantial degree (±39% positive pore water pressure), (2) increasing the unit weight
for tc and ±48% for kd) of natural variability. However, since of the bank material, (3) providing an additional destabiliz-
the primary purpose of this paper is to develop and demon- ing force due to the presence of water in tension cracks, and
strate the fully coupled approach, for reasons of simplicity we (4) providing additional (stabilizing or destabilizing) seep-
ignore this source of uncertainty and conduct a single set of age forces. A crucial point when accounting for pore water
simulations using only the mean parameter values. pressures is their transient character, driven as they are by
[14] To characterize the near-bank shear stresses exerted dynamic hydrological variables (rainfall, as well as the
on the river banks during the simulated event we adopted varying level of the water in the river). The actual mecha-
the following approach. It is important to note that shear nisms and timing of bank failures induced by pore water
stress and therefore lateral erosion was computed for each pressure effects are therefore difficult to predict if their
time step of the discretized hydrographs, following the temporal changes are not accounted for [Rinaldi and Casagli,
procedure shown in Figures 1 and 3a. Boundary shear stress 1999; Casagli et al., 1999; Simon et al., 2000, 2006]. For
was initially estimated using t mean = gRS, where t mean = this reason, bank stability response during flow events
mean boundary shear stress (Pa), g = unit weight of the requires knowledge of the dynamics of saturated and
water (N/m3), R = hydraulic radius (m), and S = energy unsaturated seepage flows. Herein we follow previous
slope (m/m). To transform t mean to the near-bank shear studies [e.g., Dapporto et al., 2001; Dapporto, 2003;
stress, we selected a function describing the distribution of Rinaldi et al., 2001, 2004] that have modeled pore water
boundary shear stress around the wetted perimeter. Numer- pressure distributions in riverbanks using the software
ous experimental studies have attempted to derive such a SEEP/W [Geo-Slope International, 2001a]. The software
function, but most have been conducted in trapezoidal performs a two-dimensional, finite element seepage analysis
flumes with gentle bank slopes. In this study we therefore using the governing equations of motion (Darcy’s law) and
used a function (Figure 3a) obtained from laboratory flume mass conservation, the latter expressed here in a form
experiments [Leutheusser, 1963] within straight, rectangu- extended to unsaturated conditions [Richards, 1931;
lar, channels with vertical banks, which provides a closer Fredlund and Rahardjo, 1993] as
analogy to the near-vertical bank encountered at the Sieve    
River. The Leutheusser distribution function (Figure 3a) @ @H @ @H @q
kx þ ky þQ¼ ð3Þ
was applied at each of 32 computational nodes spaced apart @x @x @y @y @t
up the bank profile at a uniform vertical distance of 0.15 m.

4 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 3. Overview of the coupled bank erosion analysis for the Sieve River study site: (a) method of
estimating fluvial erosion; (b) geometry of the finite element seepage analysis, indicating the different
types of assigned boundary conditions (bank profile and bank top boundaries are represented by solid and
open circles, respectively; all the other boundary nodes are zero flux boundaries) and sediment horizons;
and (c) slide- and cantilever-failure mass wasting analyses applied to the upper cohesive part of the river
bank (sediment layer 3). Terms are defined as follows: C*, effective cohesion; CT, total cohesion; f0,
effective friction angle; fb, friction angle in terms of matric suction; g, bulk unit weight; L, length of the
failure surface for the cantilever failure block; A, area of the layer i of the cantilever failure block; t c,
critical shear stress to entrain the bank material; kd, erodibility coefficient; LE, lateral erosion increment
simulated in the time step; e, bank erosion rate; Dt, time step; t b, shear stress exerted on the river bank;
t mean, mean boundary shear stress.

5 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 4. Bank material characteristics at the Sieve River study site showing (a) grain size distributions
of each layer (1, 2, and 3, as illustrated in Figure 3) of bank sediment, based on bulk samples collected
from each layer; (b) soil water characteristic curves; and (c) hydraulic conductivity functions for each of
the bank sediments. Terms are defined as follows: d, grain size; ua, pore air pressure; u, pore water
pressure; q, volumetric water content; k, hydraulic conductivity.

where H = total head (m), kx = hydraulic conductivity in resulting range. As noted above, model results are sensitive
the x direction (m/s), ky = hydraulic conductivity in the to the k-u relation, but we have confidence in the parame-
y direction (m/s), Q = unit flux passing in or out of an terization as the range of modeled k-u curves was found to
elementary cube (in this case an elementary square, given be small and sensitivity tests indicate that the results
that the equation is in two dimensions) (m2/m2s), q = reported below are not materially affected.
volumetric water content (m3/m3), and t = time (s).Critical [17] Having defined the bank sediment properties, finite
to the success of the modeling is accurate parameterization of element seepage analyses were performed by discretizing
the hydraulic and physical properties of the bank sediments. It the river banks into a series of finite elements, and assigning
is also necessary to appropriately set the initial and boundary k-u and q-u functions to each defined layer of bank material.
conditions, and to discretize (Figure 3b) equation (3). Each of For this study we used a mesh comprising 1600 quadrilateral
these aspects is now considered. and triangular finite elements (Figure 3b). This mesh is
[16] Parameterization of the hydraulic and physical prop- considerably finer in resolution than the 517 element mesh
erties of the bank sediments (Figure 4 and Table 1) used at the Sieve River study site in an earlier study [Rinaldi
primarily involves defining the relations between hydraulic et al., 2004] in which a constant bank profile was assumed,
conductivity (k) and pore water pressure (u), and between but the increased computational requirements are not prob-
volumetric moisture content (q) and pore water pressure for lematic with modern desktop computers. We return to the
each layer of sediment. Accurate parameterization of these significance of the mesh resolution in the next section.
relations, referred to here as the conductivity (k-u) and soil [18] Regarding the boundary conditions employed along
characteristic (q-u) curves, respectively, is important be- the borders of the finite element grid, for bank profile nodes
cause simulated pore pressures are highly sensitive to these (indicated as closed circles on Figure 3b) a total head versus
hydraulic functions [Rinaldi et al., 2004]. The soil charac- time function was used to represent the flow hydrograph of
teristic (q-u) curves (Figure 4b) were constructed via direct the simulated event, the latter being defined using 15-min
measurements of pore water pressure (by deployment of resolution time series data obtained from pressure trans-
tensiometer arrays) and water content values (by removal of ducers installed in the channel adjacent to the bank. To
soil samples for standard laboratory analysis) over a sea- simulate the effects of infiltrating rainfall, a rainfall intensity
sonal cycle. The k-u curves (Figure 4c) were estimated versus time function was used along the bank top nodes
following procedures described in Rinaldi et al. [2004]. In (indicated as open circles on Figure 3b), using 15-min
summary, a range of empirical relationships [Green and resolution time series data obtained from an automated rain
Corey, 1971; Van Genuchten, 1980; Fredlund et al., 1994] gauge installed at the study site. Zero flux boundary con-
was used to model possible forms of the k-u relations, based ditions were assigned along the remaining vertical and
on measured grain size distributions as determined from horizontal boundaries of the finite element grid. This
bulk samples extracted from each horizon (Figure 4a). For approach follows Rinaldi et al. [2004], who argued that
the cohesive layer 3 that is central to the stability analyses the zero flux assumption has little affect on simulated pore
undertaken herein, the resulting k-u relations were then pressures in the area of interest, namely the cohesive upper
constrained (by displacing the curves vertically) to match part of the bank. Initial conditions were defined using the
measured values of saturated hydraulic conductivity groundwater level measured before the start of the flow
(Table 1) obtained in situ using a series of 9 Amoozemeter event using piezometers installed in the banks (Figure 3b).
[Amoozegar, 1989] and 3 open single ring infiltrometer [19] The seepage analysis is performed by discretizing a
[Daniel, 1987] tests. Consistent with the rationale adopted specific flow event hydrograph into a series of time steps. For
in section 2.2, we undertook a single set of simulations this study we selected a single peak flow event (Q = 790 m3/s),
based on the median k-u relations (Figure 4c) from the that occurred during 18–20 November 1999. This event

6 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 5. Flow hydrograph and rainfall time series for the investigated flow event (18 – 20 November
1999) on the Sieve River.

represents a large magnitude (16-year RI) flood, providing a status as a bank profile, so that the associated boundary
case study wherein high rates of fluvial erosion induce mass condition does not change, but it necessitates manual
failures. The flow hydrograph and rainfall inputs for this remeshing at the end of each time step. Furthermore, as
period were discretized into 25 time steps, with shorter time finite elements in the seepage model cannot be destroyed,
steps during phases of rapidly varying flow (Figure 5). this scheme cannot be applied if the increment of simulated
fluvial erosion is greater than the thickness of the boundary
2.4. Mesh Adaptation cell.
[20] The seepage analysis described above is identical to [22] 2. For the second case (FE  CW), the change in grid
previous studies [Dapporto et al., 2001; Dapporto, 2003; geometry is simulated artificially by adjusting (again, man-
Rinaldi et al., 2001, 2004] that maintain a constant bank ually) the hydraulic conductivity and volumetric water
profile. However, in this study bank profiles are deformed content of eroded cell(s) to replicate the conductivity and
to accord with fluvial erosion and/or mass wasting (see saturated state of subaqueous in-channel cells (Figure 6c).
below) simulated at the end of each discrete time step. This was achieved by assigning k-u and q-u functions
Accordingly, special attention must be paid to the manner in appropriate for a (hypothetical) coarse sediment, so that
which the finite element mesh is adapted to the new bank the behavior of the cells is that of a very permeable material
geometry. Finite element mesh adaptation was achieved in (in practical terms, the water level within this material is
this study using two manual procedures (Figure 6): equal to the river stage). Elements exposed by fluvial
[21] 1. If the magnitude of the simulated fluvial erosion erosion are then updated with new bank profile (total head
(FE) is less than the width of the boundary cell (CW), the versus time) boundary conditions. This scheme contrasts
boundary node is shifted horizontally inward by an amount with the first in that the status and associated boundary
equal to the simulated fluvial erosion (Figure 6b). This conditions of the adjusted nodes must be updated manually,
scheme has the advantage that the shifted node retains its

Figure 6. Illustration of the two possible schemes used to adapt the finite element mesh to an increment
(FE) of simulated fluvial erosion: (a) unmodified near-bank cells prior to fluvial erosion; (b) scheme 1, in
which the bank profile nodes (indicated by the large solid circles) are displaced inward by an amount
equal to the simulated fluvial erosion; and (c) scheme 2, in which the geometry of the original mesh
remains unchanged but the physical properties of the ‘‘eroded’’ cells are updated and boundary conditions
are assigned to the newly exposed bank profile nodes (indicated by the large solid circles).

7 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

albeit without the need to remesh at the end of each time cantilever failure mechanisms defined by Thorne and Tovey
step. [1981], namely the shear-type failure, primarily for reasons
[23] That both schemes involve manually updating the of simplicity (the other two mechanisms, toppling and
mesh and/or boundary conditions is a clear disadvantage in tensile failures, respectively, are more complex to model
that the procedures require significant investments of oper- for the unsaturated conditions of interest here). This is not a
ator time. However, manual remeshing is a necessary major limitation as shear-type failures appear to be the most
consequence of employing SEEP/W, as users cannot access common observed at the study site, though it is admittedly
the source code and implement customized, automated, sometimes difficult to discriminate between shear versus
mesh generation algorithms. Which of the two meshing toppling failures in the field. What is clear is that tensile
schemes is adopted depends on the relative scales of the failures are rarely observed. For shear-type cantilever fail-
boundary cells and simulated fluvial erosion in a discrete ures (Figure 3c) the factor of safety can be written as
time step. Thus coarse (CW  FE) elements on the bank P
profile cause the first scheme (Figure 6b) to be adopted, Li CTi
whereas the second scheme (Figure 6c) is associated with i
Fs ¼ P ð5Þ
fine resolution elements and/or high fluvial erosion rates g i Ai
(CW < FE). The second scheme is advantageous in the
sense that manually updating the element properties and where Li is the vertical length (layer i) of the cantilever
boundary conditions is much simpler than manually block (m), CTi is the total cohesion (layer i) of the
remeshing, but it requires the use of much higher resolution cantilever block (kPa), g i is the unit weight (layer i) of
finite element grids than those used in past studies. This the cantilever block (kN/m3), and Ai is the cross-sectional
constraint can, to some extent, be relaxed in the first area (layer i) of the cantilever block (m2). Note that the total
scheme, but it is more labor intensive than the second in cohesion is the sum of the effective and apparent cohesion,
that manual updating of the mesh is required at the end of the latter being calculated from the third term on the right
each time step during which fluvial erosion and/or mass hand side of equation (4) according to the pore water
wasting is predicted to occur. Note that in the fully coupled pressure distribution in the cantilever block.
simulations reported below (scenario 3 in section 3, below) [27] Both stability analyses were performed for the cohe-
the bank profile is deformed after 15 of the 25 time steps. sive part of the bank that is subject to mass failure (Figure 3c).
For this reason we found it significantly easier to utilize the A single set of sediment samples were removed from this
high-resolution mesh as this enabled us to manually update horizon to estimate the soil density and shear strength
the boundary conditions (scheme 2) rather than the mesh (Table 1) used in the stability analyses. The latter was
itself, making the task manageable. defined using a triaxial shear test to estimate the effective
cohesion and effective friction angle components of shear
2.5. Riverbank Stability Analyses strength. Unlike the finite element seepage analysis, the
[24] The pore water pressure distributions obtained by the bank stability model requires no computational grid and in
seepage analysis in each time step are used as input data for each time step computations were simply performed with a
the bank stability analyses, thus providing one of the two bank profile shape updated in accordance with the pattern of
key factors that force a transient response in bank stability fluvial erosion simulated in that time step, together with the
through the flow event. The second forcing factor is, of specified geotechnical properties of the soils and the pore
course, the deformation of the bank profile by fluvial water pressure field imported from the SEEP/W output in
erosion, details of which were provided in subsection 2.2. the corresponding time step. It should be noted that the bank
[25] Two specific mechanisms of bank failure were mod- stability analysis also includes the stabilizing hydrostatic
eled in this study (Figure 3c). First, the software SLOPE/W confining pressures exerted by the water in the river
[Geo-Slope International, 2001b] was used for the applica- channel. However, the variation of unit weight with chang-
tion of limit equilibrium methods to determine stability with ing soil moisture content during an event, and the variability
respect to slide-type failures. The factor of safety was, in all of the parameter fb, also require consideration. For each
cases reported herein, computed by the Morgenstern-Price time step and sediment layer of the simulation, the unit
method, with the Mohr-Coloumb criterion in terms of weight was updated using
effective stress used for portions of the bank with positive
pore water pressures, while the Fredlund et al. [1978] g ¼ g d þ 9:81q ð6Þ
criterion was used for parts of the bank with negative pore
water pressures, the latter being expressed as where g is the unit weight of the soil during the time step
(kN/m3), g d is the unit weight of soil under completely dry
t ¼ c0 þ ðs  ua Þ tan f0 þ ðua  uÞ tan fb ð4Þ conditions (kN/m3), and q is the volumetric water content
(m3/m3) during the time step, which is estimated from the
where t = shear strength (kPa), c0 = effective cohesion soil water characteristic curve using the simulated pore
(kPa), s = normal stress (kPa), ua = pore air pressure (kPa), water pressure in that time step. In contrast, a function
f0 = effective friction angle (°), u = pore water pressure relating fb to the matric suction was defined on the basis of
(kPa) and fb = angle (°) expressing the rate of increase in the soil water characteristic curves, which has the effect of
strength relative to the matric suction (ua  u). varying fb from a minimum value of 15° to a maximum
[26] Second, the analysis of Thorne and Tovey [1981] was value equal to f0 (35°) when the soil approaches saturation.
used to determine bank stability with respect to cantilever- This was achieved by dividing the failure envelope into a
type failures. We analyzed only one of the three possible series of linear segments with varying fb angles and

8 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 7. Simulated fluvial erosion for the investigated flow event (18 – 20 November 1999) on the
Sieve River.

different intercepts (c*), each segment corresponding to a gravel and the wedge of loose sediment at the bank toe is
range of matric suction [Fredlund and Rahardjo, 1993]. therefore removed. Subsequently, the packed gravel at the
[28] Both stability analyses were performed at the end of toe is directly exposed to fluvial action and, by time step 7
each time step, with the most likely failure mechanism (t = 8.5 hours), boundary shear stress exceeds the critical
being discriminated by the method that provides the lowest value for incipient motion of this material. At the same
simulated factor of safety. For time steps in which the time, the river stage just exceeds the contact between the
simulated minimum factor of safety was less than the gravel and the upper cohesive part of the bank, and the
critical value of 1, the bank is unstable and the bank profile boundary shear stress is sufficient to trigger erosion of
was updated in accordance with the geometry of the the cohesive sediment. Consequently, the fluvial erosion
simulated failure (Figure 1). In doing so it was assumed rate of both layers rapidly increases, in phase with in-
that failed material is completely and instantaneously re- creasing stage (Figure 7). The rate of fluvial erosion of the
moved by the flow. This is reasonable for the Sieve study basal packed gravel is higher than for the cohesive layer,
site because field observations indicate that failed debris promoting the generation of an overhanging bank profile,
stored at the bank toe has very low effective cohesion and a and reaches a maximum during time step 12 (t = 12 hours),
high degree of saturation, so the material is quickly disag- the peak flow stage. Most (about 60%) of the simulated
gregated and removed by the flow. fluvial erosion occurs during the peak flow phase, between
time steps 12 and 14 (t = 12 to 15 hours). As stage recedes,
3. Results fluvial erosion of the upper cohesive part is terminated by
t = 17.5 hours, with fluvial erosion of the packed gravel at the
[29] In this section we present simulations of bank toe continuing until t = 20.5 hours. Overall, the flow event
response during and after the flow event of 18– 20 November induces a mean retreat along the bank profile of 0.90 m.
1999. Simulations are undertaken for three scenarios:
[30] 1. There is no fluvial erosion and the bank profile 3.2. Pore Water Pressure
remains constant throughout. This scenario emulates the [35] Changes in riverbank pore water pressures are usually
approach adopted in many previous studies. considered to be controlled by rainfall and river stage
[31] 2. There is no fluvial erosion but the bank profile is variations. However, as shown here bank profile deformation
deformed by simulated mass wasting. due to fluvial erosion and mass wasting may also play a role
[32] 3. The bank profile is allowed to freely deform in in the transient evolution of bank pore water pressures. For
response to both simulated fluvial erosion and mass wasting. reasons of clarity, in this section we focus attention on the
[33] Comparing simulation results from these scenarios differences between the two ‘‘extreme’’ scenarios 1 (where
enables the effects of feedbacks between fluvial erosion, the bank profile is constant throughout the simulation), and 3
mass wasting and the evolving pore water pressure field to (where the bank is deformed by both fluvial erosion and mass
be isolated. wasting).
[36] The pore water pressures simulated within the cohe-
3.1. Simulated Fluvial Erosion sive layer of the Sieve river bank are presented in Figure 8.
[34] Simulated fluvial erosion (Figure 7), which is In both scenarios, at the beginning of the simulation (from
common to scenarios 2 and 3, commences at time step 5 time step 0 to 6, t = 0 to 8 hours) changes in soil moisture
(t = 7.5 hours), when the boundary shear stress exceeds content are caused by rainfall infiltration into the soil,
the critical conditions for incipient motion of the loose leading to an initial reduction of matric suction. The wetting

9 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 8. Simulated bank pore water pressure distributions for selected time steps of the Sieve River
simulation for scenarios 1 (no bank deformation), 2 (bank profile deformed only by simulated mass
wasting), and 3 (bank profile deformed by both simulated fluvial erosion and mass wasting). Locations
and associated factor of safety values for the simulated slide and cantilever failure surfaces are also
illustrated, as are the positions of the groundwater surface. For scenarios 2 and 3, the initial bank profile is
also indicated.

front is partially saturated, with a mean suction of about equilibrium with the river stage during the whole drawdown
2 kPa (Figure 8, time step 6). phase (Figure 8, time steps 17 and 20).
[37] As the flow stage rises (from time steps 6 to 12, [39] To highlight differences in pore water pressures
t = 8 to 12 hours) and eventually exceeds the contact arising between the two scenarios we extracted time series
between the packed gravel at the toe and the cohesive part of pore water pressure along the sliding surface (excluding
of the bank, a steep wetting front develops at the contact the part occupied by the tension crack) and computed the
between the river and the bank profile, inducing a rise in the integral of the simulated values (shaded areas in the sketch
groundwater level. From this moment the pore water of Figure 9). As a proxy for the stress exerted by pore
pressure distribution is characterized by a minimum in the water along the sliding surface, this index (Pw) has
central part of the cohesive layer and by higher values in the significance in the assessment of bank stability conditions.
upper right and lower regions, due to the infiltration of both Figure 9a shows the trend of integral values computed
rainfall and river flow induced wetting fronts. However, only on the saturated portion of the sliding surface (Pw(+),
subsequently the two scenarios diverge (see Figure 8, time highlighted as the dark grey area in the sketch) for both
steps 17 and 20). In scenario 3, the river flow induced scenarios, while time series of the integral value across the
wetting front and bank retreat have comparable rates of entire sliding surfaces (Pw, sum of the light and dark grey
motion: As a consequence, the bank profile retreats into areas) are shown in Figure 9b.
relatively dry areas of the bank, inducing lower pore water [40] In interpreting these time series it should be noted
pressures along the bank profile. In terms of the effect of that in both scenarios the geometry of the sliding surface
this on bank stability, the destabilizing effect of bank pore varies throughout the simulation. However, positive pore
water pressure is reduced in scenario 3; however, as water pressures develop in a limited part of the bank close to
described in the following section, bank steepening due to the river, so changes in sliding surface geometry only have
fluvial erosion outweighs this effect and still promotes a net minor effects on the computation of the Pw(+) values.
decrease in bank stability. Values of Pw(+) in scenario 1 are always higher than the
[38] During the hydrograph’s falling limb the groundwater values computed in scenario 3 (Figure 9a). The importance
level and pore water pressure distribution lower progres- of Pw(+) as a measure of pore water pressure effects on bank
sively, with notable differences between the two scenarios. stability is apparent in the computation of the factor of
In scenario 1, the groundwater level initially recedes at a safety, where the ‘‘weight’’ of positive pore water pressures
slower rate than the hydrograph, reaching a new equilibrium is higher than that of negative pressures, the former being
with river stage only by time step 20 (t = 18.5 hours). In multiplied by tanf0 (0.70 in this study, see Table 1), the
contrast, in scenario 3 the groundwater level remains in latter by tanfb (0.47 to 0.70 in this study, see Table 1).

10 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 9. Evolution of average pore water pressure values integrated along the failure surface in the
Sieve River simulation for scenarios 1 (no bank deformation) and 3 (bank profile deformed by fluvial
erosion and mass wasting): (a) pore water pressure integral along the saturated portion of the failure
surface and (b) pore water pressure integral along the entire failure surface (excluding tension cracks).

[41] Pore water pressure integrals computed along the have modeled bank stability variations without considering
whole failure surface (Pw) exhibit a less regular trend the effects of fluvial erosion.
(Figure 9b), the curve for the deforming bank profile [43] In contrast, factor of safety with respect to cantilever
(scenario 3) alternating in position above and below the failure declines on the hydrograph’s rising limb, falling
curve for the constant bank profile (scenario 1). During the from 2.23 initially to a value denoting instability (0.99) at
initial drawdown phase, the former is characterized by the event peak (t = 12 hours). This is caused by a rapid rise
lower Pw values; this is not surprising considering the in pore water pressure along the failure surface, due to the
differences in pore water pressure distributions already cantilever failure surface’s location closer to the advancing
described. However, at time step 17 (t = 16.5 hours) Pw wetting front. After the hydrograph peak the flow recedes,
suddenly increases and approximates the value of the but pore pressures remain elevated. Hence there is a period
constant bank profile line. This apparently abnormal trend in which the downward trend in stability continues until the
is due to a change in the geometry of the computed sliding factor of safety minimizes (0.51) during t = 13 to 15.5 hours.
surface, which has a shorter length and a deeper tension The factor of safety subsequently recovers, but stability is
crack line than in the previous time steps. not regained until t = 29 hours, when the rate of flow
recession slows sufficiently for the slowly equilibrating pore
3.3. Bank Stability and Sediment Entrainment pressure field to restabilize the bank (see Figures 8 and 9).
[42] Time series of factor of safety with respect to the [44] Scenario 1 is thus characterized by phases of bank
slide and cantilever failure mechanisms, and associated instability at the peak of the event and on the falling limb of
volumes of sediment entrained by each process, are shown the hydrograph for cantilever and slide-type failures, re-
in Figure 10. Factor of safety values in scenario 1 are purely spectively. Although the results for this scenario are biased
hypothetical because the two failure mechanisms are mod- in the sense that the bank profile shape is not updated in
eled independently (so there is no consequence if one of the accordance with these predicted failures, they show how
factor of safety curves falls below the critical value of stability fluctuations for each failure mechanism are driven
unity), but they present a reference for comparison with by the evolving pore water and hydrostatic confining
the other scenarios. In scenario 1 (Figure 10a) the trend of pressure fields. In particular, the distinctive response of
factor of safety with respect to slide failure follows the each failure type is conditioned by the positions of the
hydrograph shape, due to the stabilizing influence of the respective failure surfaces relative to the advancing wetting
hydrostatic confining pressure exerted by water in the river fronts. The shallow-seated cantilever failure surface is
[see also Simon et al., 1999, Figure 6.6b, p.137]. As flow located close to the advancing wetting fronts, so bank
stage increases, the factor of safety rises from 1.25 initially response is more sensitive to pore water pressure. In
to a maximum of 5.31 at the event peak, prior to falling on contrast the deep-seated slide failure surface is relatively
the recession limb. The factor of safety (0.98) at the end of distant from the wetting front, causing a lag in the failure
the event (t = 40 hours) is lower than at the start because response to pore water pressure changes induced by varia-
of the decreased extent of negative, and increased extent of tions in river stage.
positive, pore water pressures, caused by the infiltration of [45] Factor of safety time series in scenario 2 (bank
rainfall- and river flow – induced wetting fronts into the profile deformed by mass wasting only) are identical to
bank. The onset of instability with respect to slide failure those simulated in scenario 1 until time step 12 (t = 12 hours),
occurs on the falling limb of the hydrograph (t = 16.5 hours), when a cantilever failure is triggered (Figure 10c). In this
similar to other studies [Casagli et al., 1999; Simon et al., scenario the cantilever failure results in the collapse and
1999; Simon and Collison, 2002; Rinaldi et al., 2004] that removal of the overhanging upper, cohesive, part of the
bank. The absence of fluvial erosion, however, prevents the

11 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Figure 10. Simulated (left) bank stability and (right) sediment entrainment responses for the 18 –
20 November 1999 flow event on the Sieve River: (a, b) scenario 1 with the bank profile constant
throughout the simulation; (c, d) scenario 2 with the bank profile deformed by mass wasting; and (e, f )
scenario 3 with the bank profile deformed by both fluvial erosion and mass wasting. The event
hydrograph is also shown (solid lines). The dotted horizontal lines indicate the critical factor of safety
value of unity (Fs < 1 implies bank collapse), with the arrows indicating the onset of simulated failure
episodes. Note that factor of safety data for the cantilever failures are plotted only for those points in time
when a cantilevered (overhanging) bank profile is actually present. Also note that since the bank
geometry in scenario 1 is not updated, either in response to mass wasting or fluvial erosion, there is no
corresponding sediment entrainment time series (see inset in Figure 10b).

overhang from reforming, so that (unlike scenario 1) the of safety at the peak of the flow event is 9.99 in scenario 2,
time series of factor of safety with respect to cantilever compared to a value of 5.31 in scenario 1). This increased
failure is terminated at this point. In this scenario the change stability with respect to the slide failure in scenario 2 versus 1
in bank geometry also has an effect on the simulated factor is maintained throughout the recession limb of the hydro-
of safety with respect to the slide failure mechanism, graph. Indeed, unlike scenario 1 the factor of safety with
resulting in a substantial increase in stability (e.g., the factor respect to the slide failure mechanism does not fall below

12 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

the critical value of unity, a minimum value of 1.25 instead coupled analysis (scenario 3, Figure 10f) is seen to provide
being attained by t = 18.5 hours. modified frequencies, timings and modes of erosion relative
[46] In the fully coupled scenario 3 (Figure 10e), the to scenario 2 (Figure 10d). The significance of this point is
factor of safety with respect to cantilever failure is initially now discussed.
(t 8 hours) identical to that in scenarios 1 and 2, but
subsequently declines more rapidly, falling below the 4. Concluding Discussion
critical value at t = 10.5 hours, before the corresponding
failure at the peak of the hydrograph (t = 12 hours) in [49] Our results (Figure 10) indicate that when bank
scenarios 1 and 2 (Figures 10a and 10c). Since the simulated response models include feedbacks between fluvial ero-
pore water pressures in scenario 3 are lower than in sion, bank pore water pressure and mass wasting (scenario 3),
scenarios 1 (see Figure 9a) and 2, this variation must be predictions of the modes (cantilever versus slide failures),
caused by the rapid increase in bank shear stress and fluvial frequencies, magnitudes, and timings of bank erosion epi-
erosion that is predicted to accompany the rise in stage (see sodes are distinct from predictions made by neglecting such
Figure 7). Subsequently, the bank remains unstable with effects (scenarios 1 and 2). It must be emphasized that the
respect to cantilever failures until t = 16.5 hours as multiple scope of our simulations are, of course, limited to the
small-scale cantilevers form and fail. Recovery to a stable sedimentary conditions encountered at the Sieve River bank
condition (with respect to cantilever failure) also occurs study site, which comprises a cohesive upper layer overly-
earlier than in scenario 1, a consequence of updating the ing a gravel toe. This arrangement is common in upland and
bank profile shape. Thus in scenario 3, multiple cantilever piedmont zones in Europe and elsewhere, but is distinct
failures are generated by the repeated collapse and refor- from the fine-grained bank settings that are more usually
mation (by fluvial erosion) of overhanging bank profiles. associated with lowland environments. The scope of the
[47] It is not possible, however, to compare the cantilever present work is further restricted by our focus on a single,
factor of safety time series for the different scenarios exemplar, event hydrograph. This means that our key
throughout the entire recessional limb because the analysis findings might best be presented as tentative hypotheses
is terminated at t = 16.5 hours by the onset of a slide failure that require testing in a wider range of environments.
that reshapes the bank profile, removing the potential for Nevertheless, if our findings are more widely transferable
fluvial erosion to generate overhangs. Indeed, the onset of a then there are a number of implications with respect to
major slide failure in this scenario is distinct from scenario 2 understanding the mechanisms by which bank materials
(where fluvial erosion is absent, Figure 10c). It seems that may be contributed to an alluvial stream, and in this section
fluvial erosion, which is predicted to occur throughout the we discuss these issues.
flow event in scenario 3, deforms the bank profile suffi- [50] Notwithstanding the preceding caveats, it can be
ciently to significantly reduce the factor of safety with noted that some of our findings are consistent with previous
respect to the slide failure mechanism (e.g., a value of research [Rinaldi and Casagli, 1999; Casagli et al., 1999;
4.08 at the peak of the hydrograph, compared to 5.31 and Simon et al., 1999; Simon and Collison, 2002; Rinaldi et al.,
9.99 in scenarios 1 and 2, respectively). This reduction is 2004] in that hydrologically driven variations in bank pore
‘‘significant’’ in the sense that by t = 16.5 hours, only water and hydrostatic confining pressures are key controls
4.5 hours after the peak of the event, the combination of on the transient response of mass wasting. Our simulations
reduced confining pressure, elevated pore pressures, and also highlight the significant role that fluvial erosion can
fluvial erosion of the bank profile is sufficient to initiate a have as a triggering mechanism for mass failure [Darby and
major slide failure (Figure 10e), where no such failure Thorne, 1996; Langendoen, 2000; Simon et al., 2006], with
occurred previously (Figures 10a and 10c). This is very significant additional volumes of mass wasted sediment
similar to results published by Simon et al. [2006] for the contributed as a result of a bank profile being destabilized
Upper Truckee River, California, in which measured pore by fluvial erosion (Figure 10). However, the simulation data
water pressure distributions were combined with simulated also reveal a previously undocumented effect wherein the
bank toe erosion and bank stability analyses. evolving pore water pressure field in the near-stream region
[48] Differences in the unit volumes of bank material of the bank is modulated by deformation of the bank profile.
entrained in scenarios 2 and 3 are summarized in the Specifically, (positive) pore water pressure values in the
annotated insets in the figures (Figures 10d and 10f). These near-stream region are decreased below what they would be
show that the combined effects of fluvial erosion and mass in the absence of bank deformation because bank failure
wasting in scenario 3 provide a total eroded sediment and/or fluvial erosion expose interior regions with higher
volume (11.65 m3/m) some 33 times greater than the values of matric suction. The magnitude of this effect is, in
scenario 2 simulations in which fluvial erosion is absent. these simulations, insufficient to cause a delay in the onset
Clearly, most (9.18 m3/m) of this additional material is of mass failure.
eroded directly by hydraulic action, but significant addi- [51] The limited number of simulations in this study
tional (by a factor of about 7) volumes are also contributed makes it difficult to discriminate quantitatively the condi-
by mass wasting, of which about two thirds and one third tions under which this modulating effect may be a more
are attributable to the slide and cantilever failure mecha- significant control affecting the onset of bank instability.
nisms, respectively. That high rates of fluvial erosion cause However, in circumstances where fluvial erosion
an increase in the magnitude of bank erosion, both directly (1) removes bank materials that are saturated or at least
and, via its influence on mass wasting, indirectly is not partially saturated and (2) proceeds at a rate greater than
surprising. However, the interaction of pore water pressure, the advance of the wetting front induced by the presence of
fluvial erosion and mass failure dynamics in the fully flow within the channel, the effect is to remove regions of the

13 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

bank face with relatively high pore pressures, ‘‘replacing’’ would be distinct from riverbanks represented using an
them with drier regions of reduced (or negative) pore uncoupled approach (Figures 10a and 10c). However,
water pressure within the bank interior. In terms of bank whether the additional material delivered to the toe by
environments where such conditions might be favored, fluvially triggered bank failures can be supplied to the in-
flashy stream environments and/or banks composed of channel sediment transfer system will again depend on the
erodible materials of low hydraulic conductivity are impor- basal residence time of that material. Thus a fruitful avenue
tant. In such cases the bank face would be prone to periods for future research would be to quantify these residence
of wetting that are sufficiently short in duration to maintain times. If this could be achieved it would provide the means
a relatively dry bank interior. However, even in such cases if to explain how some rivers are able to both erode and
the effect is to be significant (in the sense of being transmit bank sediments effectively enough to produce the
sufficiently large to influence mass wasting; that is, the high sediment yields described in the introduction to this
factor of safety is modified in such a way that its value paper. We suggest that the development of conceptual
crosses the critical value of unity) an additional necessary sediment delivery models, which focus on characterizing
condition is still required, namely that antecedent stability transient bank response (especially the phasing and mech-
conditions must be marginal (i.e., Fs
1). These constraints anism of erosion episodes in relation to the capacity of the
suggest that the circumstances under which fluvial erosion flow to remove eroded material), would represent an ad-
acts to delay the onset of mass wasting may be restricted to vance toward this goal.
very specific bank settings or times.
[52] What is clear is that our fully coupled simulations [55] Acknowledgments. This research was supported by the Royal
predict transient mass wasting responses that differ from Society (Joint Project Grant 15077). We would also like to thank Rob
Millar and Andrew Simon for their constructive reviews.
those studies that suggest that mass failure is essentially a
quasi-catastrophic event that occurs on the falling limb of References
event hydrographs [e.g., Thorne, 1982; Huang, 1983; Abam, T. K. S. (1997), Aspects of alluvial river bank recession: Some
Abam, 1997; Darby et al., 1998]. In this study mass wasting examples from the Niger delta, Environ. Geol., 31, 211 – 220.
instead occurs as a series of failure episodes, with progres- Abernethy, B., and I. D. Rutherfurd (2000), The effect of riparian tree roots
sive fluvial erosion undermining the bank, modulating the on the mass-stability of riverbanks, Earth Surf. Processes Landforms, 25,
921 – 937.
evolving pore water pressure field, and triggering failures Amoozegar, A. (1989), A compact constant-head permeameter for measur-
prior to, at, and subsequent to event peaks. What is not clear ing saturated hydraulic conductivity of the vadose zone, J. Soil Sci. Soc.
from this preliminary investigation is whether this is gen- Am., 53, 1356 – 1361.
Arulanandan, K., E. Gillogley, and R. Tully (1980), Development of a
eral, or whether our findings are specific to the hydrologic quantitative method to predict critical shear stress and rate of erosion
and sedimentary setting of this study. Nevertheless, a focus of natural undisturbed cohesive soils, Rep. GL-80-5, U.S. Army Corps
on the timing of mass wasting events, and specifically their of Eng., Waterways Exp. Station, Vicksburg, Miss.
Bull, L. J. (1997), Magnitude and variation in the contribution of bank
phasing in relation to the event hydrograph, is a generically erosion to the suspended sediment load of the River Severn, UK, Earth
important aspect of bank erosion dynamics as it has a direct Surf. Processes Landforms, 22, 1109 – 1124.
bearing on the capacity of the flow to remove the coarse Carson, M. A., and M. J. Kirkby (1972), Hillslope Form and Process,
fraction of the failed debris, with the potential for removal Cambridge Univ. Press, Cambridge, U. K.
Casagli, N., M. Rinaldi, A. Gargini, and A. Curini (1999), Pore water
significantly decreased if failure occurs late on the recession pressure and streambank stability: Results from a monitoring site on
limb [Rinaldi and Darby, 2007]. the Sieve River, Italy, Earth Surf. Processes Landforms, 24, 1095 – 1114.
[53] The broader significance of these findings in relation Daniel, D. E. (1987), Hydraulic conductivity tests for clay liners, in
Geotechnical and Geohydrological Aspects of Waste Management, edited
to the delivery of bank-derived material to the alluvial by A. Van Zyl, pp. 825 – 844, Lewis, Chelsea, Minn.
sedimentary system can be explained with reference to the Dapporto, S. (2001), Non-vertical jet testing of cohesive streambank toe
concept of basal endpoint control [Carson and Kirkby, material, Sch. of Geogr., Univ. of Nottingham, Nottingham, U. K.
Dapporto, S. (2003), Processi di erosione e meccanismi di instabilità di
1972; Thorne, 1982]. The basal endpoint concept empha- sponde fluviali: Monitoraggio, modellazione e analisi, Ph.D. thesis, Univ.
sizes that the residence time of bank-derived material stored of Florence, Florence, Italy.
in the basal zone is a critical factor controlling long-term Dapporto, S., and M. Rinaldi (2003), Modelling of river bank retreat by
bank retreat rates. Neglecting the complicating effects of combining fluvial erosion, seepage and mass failure, Geophys. Res.
Abstr., 5, 03,425.
variations in the caliber of the failed bank material [Wood et Dapporto, S., M. Rinaldi, and N. Casagli (2001), Mechanisms of failure and
al., 2001], basal storage of eroded sediment is favored for pore water pressure conditions: Analysis of a riverbank along the Arno
riverbanks prone to large-scale slide failure(s), that deliver River (central Italy), Eng. Geol., 61, 221 – 242.
Darby, S. E., and C. R. Thorne (1996), Numerical simulation of bed
large volumes of sediment, timed during the recession limb topography and channel widening in straight sand-bed rivers. I: Model
of the hydrograph (when the erosivity of the flow is development, J. Hydraul. Eng., 122, 184 – 193.
declining). In contrast, bank environments favoring multiple Darby, S. E., et al. (1998), River width adjustment II: Modeling. Final
small-scale cantilever failures occurring before or at the report of the ASCE Task Committee on Hydraulics, Bank Mechanics
and Modeling of River Width Adjustment, J. Hydraul. Eng., 124,
peak of the event would likely substantially reduce the 903 – 917.
residence time of eroded bank material stored at the toe of Darby, S. E., D. Gessler, and C. R. Thorne (2000), A computer program for
the bank. stability analysis of steep, cohesive riverbanks, Earth Surf. Processes
Landforms, 25, 175 – 190.
[54] It can also be noted that fluvial erosion has the effect of Eaton, B. C., M. Church, and R. G. Millar (2004), Rational regime model of
increasing the volume of sediment derived from mass wast- alluvial channel morphology and response, Earth Surf. Processes Land-
ing (Figure 10), so it can be expected that the basal endpoint forms, 29, 511 – 529.
Fredlund, D. G., and H. Rahardjo (1993), Soil Mechanics for Unsaturated
status of riverbanks analyzed in this way (Figures 10e and Soils, John Wiley, New York.
10f as well as studies by Darby and Thorne [1996], Fredlund, D. G., N. R. Morgenstern, and R. A. Widger (1978), The shear
Langendoen [2000], Simon et al. [2006], amongst others) strength of unsaturated soils, Can. Geotech. J., 15, 312 – 321.

14 of 15
F03022 DARBY ET AL.: COUPLED BANK EROSION SIMULATIONS F03022

Fredlund, D. G., A. Xing, and S. Huang (1994), Predicting the permeability the River Restoration, Dev. Earth Surf. Processes, vol. 11, edited by H.
function for unsaturated soils using the soil-water characteristic curve, Habersack, H. Piégay, and M. Rinaldi, pp. 217 – 243, Elsevier, Dordrecht,
Can. Geotech. J., 31, 533 – 546. Netherlands.
Geo-Slope International (2001a), SEEP/W for finite element seepage ana- Rinaldi, M., S. Dapporto, and N. Casagli (2001), Monitoring and modeling
lysis, version 5, user manual, Calgary, Alberta, Canada. of unsaturated flow and mechanisms of riverbank failure in gravel bed
Geo-Slope International (2001b), SLOPE/W for slope stability analysis, rivers, in Gravel Bed Rivers 2000 [CD-ROM], special publication, edited
version 5, user manual, Calgary, Alberta, Canada. by T. Noland and C. R. Thorne, N. Z., Hydrol. Soc., Wellington.
Goodson, J. M., A. M. Gurnell, P. G. Angold, and I. P. Morrissey (2002), Rinaldi, M., N. Casagli, S. Dapporto, and A. Gargini (2004), Monitoring
Riparian seed banks along the lower River Dove UK, Geomorphology, and modeling of pore water pressure changes and riverbank stability
47, 45 – 60. during flow events, Earth Surf. Processes Landforms, 29, 237 – 254.
Green, R. E., and J. C. Corey (1971), Calculation of hydraulic conductivity: Simon, A., and A. J. Collison (2002), Quantifying the mechanical and
A further evaluation of some predictive methods, Proc. Soil Sci. Soc. hydrological effects of vegetation on streambank stability, Earth Surf.
Am., 35, 3 – 8. Processes Landforms, 27, 527 – 546.
Green, T. R., S. G. Beavis, C. R. Dietrich, and A. J. Jakeman (1999), Simon, A., and S. E. Darby (2002), Effectiveness of grade-control struc-
Relating stream bank erosion to in-stream transport of suspended sedi- tures in reducing erosion along incised river channels: The case of Ho-
ment, Hydrol. Processes, 13, 777 – 787. tophia Creek, Mississippi, Geomorphology, 42, 229 – 254.
Grissinger, E. H. (1982), Bank erosion of cohesive materials, in Gravel-Bed Simon, A., and M. Rinaldi (2006), Disturbance, stream incision, and chan-
Rivers, edited by R. D. Hey et al., pp. 273 – 287, John Wiley, Chichester, nel evolution: The roles of excess transport capacity and boundary ma-
U. K. terials in controlling channel response, Geomorphology, 79, 361 – 383.
Hanson, G. J. (1990), Surface erodibility of earthen channels at high stres- Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen (1999), Stream-
ses. part II—Developing an in situ testing device, Trans. ASAE, 33(1), bank mechanics and the role of bank and near-bank processes in incised
132 – 137. channels, in Incised River Channels, edited by S. E. Darby and A. Simon,
Hanson, G. J., and A. Simon (2001), Erodibility of coesive streambeds in pp. 123 – 152, John Wiley, Chichester, U. K.
the loess area of the midwestern USA, Hydrol. Processes, 15, 23 – 38. Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen (2000), Bank and
Hooke, J. M. (1980), Magnitude and distribution of rates of river bank near-bank processes in an incised channel, Geomorphology, 35, 193 –
erosion, Earth Surf. Processes Landforms, 5, 143 – 157. 218.
Huang, Y. H. (1983), Stability Analysis of Earth Slopes, Van Nostrand Simon, A., N. L. Pollen, and E. J. Langendoen (2006), Influence of two
Reinhold, New York. woody riparian species on critical conditions for streambank stability:
Lane, E. W. (1955), Design of stable channels, Trans. Am. Soc. Civ. Eng., Upper Truckee River, California, J. Am. Water Resour. Assoc., 41, 99 –
120, 1234 – 1260. 113.
Langendoen, E. J. (2000), CONCEPTS—Conservational Channel Evolu- Springer, F. M., C. R. Ullrich, and D. J. Hagerty (1985), An analysis of
tion and Pollutant Transport System, Res. Rep. 16, Natl. Sediment. Lab., streambank stability, J. Geotech. Eng., 111(5), 624 – 640.
U.S. Dep. of Agric., Oxford, Miss. Thorne, C. R. (1982), Processes and mechanisms of river bank erosion, in
Lawler, D. M. (2004), The importance of high-resolution monitoring in Gravel-Bed Rivers, edited by R. D. Hey et al., pp. 227 – 271, John Wiley,
erosion and deposition dynamics studies: Examples from estuarine and Chichester, U. K.
fluvial systems, Geomorphology, 64, 1 – 23. Thorne, C. R., and J. Lewin (1979), Bank processes, bed-material move-
Lawler, D. M. (2005), Defining the moment of erosion: The principle of ment and planform development in a meandering river, in Adjustment of
thermal consonance timing, Earth Surf. Processes Landforms, 30, 1597 – the Fluvial System, edited by D. D. Rhodes and G. P. Williams, pp. 117 –
1615. 137, Kendall Hunt, Dubuque, Iowa.
Lawler, D. M., L. Bull, and N. M. Harris (1997a), Bank erosion events and Thorne, C. R., and N. K. Tovey (1981), Stability of composite river banks,
processes in the Upper Severn, Hydrol. Earth Syst. Sci., 1, 523 – 534. Earth Surf. Processes Landforms, 6, 469 – 484.
Lawler, D. M., C. R. Thorne, and J. M. Hooke (1997b), Bank erosion and U.S. Environmental Protection Agency (2000), Quality of the Nation’s
instability, in Applied Fluvial Geomorphology for River Engineering and Rivers and Streams, Washington, D. C.
Management, edited by C. R. Thorne et al., pp. 137 – 172, John Wiley, Van Genuchten, M. T. (1980), A closed-form equation for predicting the
Chichester, U. K. hydraulic conductivity of unsaturated soils, J. Soil. Sci. Soc. Am., 44,
Leutheusser, H. J. (1963), Turbulent flow in rectangular ducts, J. Hydraul. 892 – 898.
Div. Am. Soc. Civ. Eng., 89, 1 – 19. Walling, D. E., P. N. Owens, and G. J. L. Leeks (1998), The role of channel
Marron, D. C. (1992), Floodplain storage of mine tailings in the Belle and floodplain storage in the suspended sediment budget of the River
Fourche River, Earth Surf. Processes Landforms, 17, 675 – 685. Ouse, Yorkshire, UK, Geomorphology, 22, 225 – 242.
Millar, R. G. (2000), Influence of bank vegetation on alluvial channel Walling, D. E., P. N. Owens, and G. J. L. Leeks (1999), Fingerprinting
patterns, Water Resour. Res., 36, 1109 – 1118. suspended sediment sources in the catchment of the River Ouse, Yorkshire,
Millar, R. G., and M. C. Quick (1993), Effect of bank stability on geometry UK, Hydrol. Processes, 13, 955 – 975.
of gravel rivers, J. Hydraul. Eng., 119, 1343 – 1363. Wasson, R. J., R. K. Mazari, B. Starr, and G. Clifton (1998), The recent
Mosselman, E. (1998), Morphological modeling of rivers with erodible history of erosion and sedimentation on the Southern Tablelands of SE
banks, Hydrol. Processes, 12, 1357 – 1370. Australia: Sediment flux dominated by channel incision, Geomorphology,
Osman, A. M., and C. R. Thorne (1988), Riverbank stability analysis. 24, 291 – 308.
I: Theory, J. Hydraul. Eng., 114, 134 – 150. Wolman, M. G. (1959), Factors influencing the erosion of cohesive river
Partheniades, E. (1965), Erosion and deposition of cohesive soils, J. Hydraul. banks, Am. J. Sci., 257, 204 – 216.
Div. Am. Soc. Civ. Eng., 91, 105 – 139. Wood, A., A. Simon, P. W. Downs, and C. R. Thorne (2001), Bank-toe
Reneau, S. L., P. G. Drakos, D. Katzman, D. V. Malmon, E. V. McDonald, processes in incised channels: The role of apparent cohesion in the en-
and R. T. Ryti (2004), Geomorphic controls on contaminant distribution trainment of failed bank materials, Hydrol. Processes, 15, 39 – 61.
along an ephemeral stream, Earth Surf. Processes Landforms, 29, 1209 –
1223. 
Richards, L. A. (1931), Capillary conduction of liquid through porous S. Dapporto, Regione Toscana, Sistema Regionale di Protezione Civile,
medium, J. Phys., 1, 318 – 333. Via Val di Pesa 3, Firenze I-50126, Italy. (stefano.dapporto@regione.
Rinaldi, M., and N. Casagli (1999), Stability of streambanks formed in toscana.it)
partially saturated soils and effects of negative pore water pressures: S. E. Darby, School of Geography, University of Southampton,
The Sieve River (Italy), Geomorphology, 26, 253 – 277. Highfield, Southampton SO17 1BJ, UK. (s.e.darby@soton.ac.uk)
Rinaldi, M., and S. E. Darby (2007), Modelling river bank erosion pro- M. Rinaldi, Department of Civil Engineering, University of Florence, Via
cesses and mass failure mechanisms: Progress towards fully-coupled S. Marta 3, Firenze I-50139, Italy. (mrinaldi@dicea.unifi.it)
simulations, in Gravel-Bed Rivers 6—From Process Understanding to

15 of 15

You might also like