2021 RG 000761

Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

REVIEW ARTICLE A Review on Bank Retreat: Mechanisms, Observations, and

10.1029/2021RG000761
Modeling
Key Points:
Kun Zhao1,2 , Giovanni Coco2 , Zheng Gong1,3 , Stephen E. Darby4 , Stefano Lanzoni5 ,
• B ank retreat is a key control on river
Fan Xu6 , Kaili Zhang1,2 , and Ian Townend7
and estuary morphodynamics
• Multi-variable predictors of bank 1
State Key Laboratory of Hydrology-Water Resources and Hydraulic Engineering, Hohai University, Nanjing, China, 2School
retreat from the literature are
systematically reviewed of Environment, University of Auckland, Auckland, New Zealand, 3Jiangsu Key Laboratory of Coast Ocean Resources
• Numerical approaches to bank retreat Development and Environment Security, Hohai University, Nanjing, China, 4School of Geography and Environmental
are critically appraised, resulting in a Sciences, University of Southampton, Southampton, UK, 5Department of Civil, Environmental and Architectural
series of recommendations
Engineering, University of Padua, Padua, Italy, 6State Key Laboratory of Estuarine and Coastal Research, East China Normal
University, Shanghai, China, 7Ocean and Earth Sciences, University of Southampton, Southampton, UK
Correspondence to:
Z. Gong,
[email protected] Abstract Bank retreat plays a fundamental role in fluvial and estuarine dynamics. It affects the
cross-sectional evolution of channels, provides a source of sediment, and modulates the diversity of habitats.
Citation: Understanding and predicting the geomorphological response of fluvial/tidal channels to external driving forces
Zhao, K., Coco, G., Gong, Z., Darby, underpins the robust management of water courses and the protection of wetlands. Here, we review bank retreat
S. E., Lanzoni, S., Xu, F., et al. (2022). with respect to mechanisms, observations, and modeling, covering both rivers and (previously neglected) tidal
A review on bank retreat: Mechanisms,
channels. Our review encompasses both experimental and in situ observations of failure mechanisms and bank
observations, and modeling. Reviews of
Geophysics, 60, e2021RG000761. https:// retreat rates, modeling approaches and numerical methods to simulate bank erosion. We identify that external
doi.org/10.1029/2021RG000761 forces, despite their distinct characteristics, may have similar effects on bank stability in both river and tidal
channels, leading to the same failure mode. We review existing data and empirical functions for bank retreat
Received 18 NOV 2021
rate across a range of spatial and temporal scales, and highlight the necessity to account for both hydraulic
Accepted 13 MAY 2022
and geotechnical controls. Based on time scale considerations, we propose a new hierarchy of modeling styles
that accounts for bank retreat, leading to clear recommendations for enhancing existing modeling approaches.
Author Contributions:
Finally, we discuss systematically the feedbacks between bank retreat and morphodynamics, and suggest that to
Conceptualization: Kun Zhao,
Giovanni Coco move this agenda forward will require a better understanding of multifactor-driven bank retreat across a range
Data curation: Kun Zhao, Kaili Zhang of temporal scales, with particular attention to the differences (and similarities) between riverine and estuarine
Funding acquisition: Zheng Gong
environments, and the role of feedbacks exerted by the collapsed bank soil.
Supervision: Giovanni Coco,
Zheng Gong
Validation: Fan Xu Plain Language Summary As one of the most persistent and dominant processes that shape rivers,
Writing – review & editing: Kun Zhao, estuaries, and coastal regions, bank retreat is of great relevance to the management of water courses and the
Giovanni Coco, Stephen E. Darby,
Stefano Lanzoni, Fan Xu, Ian Townend
protection of wetlands. Here, we review bank retreat processes with respect to both rivers and tidal channels,
including observations of erosion/failure processes and bank retreat rate, as well as methods proposed for
numerical modeling. We highlight that different hydrodynamic forces associated with fluvial and estuarine
settings may nevertheless have similar effects on bank stability. Using data from the literature we show that
observed bank retreat rates are not only controlled by flow strength and soil composition (e.g., discharge,
vegetation cover, and critical shear stress for erosion), but are also related to other factors, such as channel width
and bank geometry. We summarize recent developments in bank retreat modeling, discuss their advantages
and challenges, and propose some recommendations for their future use. Overall, this review deepens our
understanding of bank erosion/collapse processes, and how they shape both fluvial and estuarine landscapes.

1. Introduction
Bank retreat results from the complex of processes causing the detachment of bank soil under the actions of
hydraulic and gravitational forces, subaerial processes, bioturbation, as well as resisting actions that depend on
soil properties, and which are possibly mediated by vegetation cover. In general, the processes which cause bank
© 2022. The Authors.
This is an open access article under retreat can be divided in two main classes: flow-driven bank erosion and gravity-driven bank collapse (Simon
the terms of the Creative Commons et al., 2000; Thorne & Tovey, 1981). Bank retreat is of fundamental importance to fluvial, estuarine, and coastal
Attribution License, which permits use, dynamics, and encompass a wide range of spatial and temporal scales, with important physical, ecological, and
distribution and reproduction in any
medium, provided the original work is socio-economic repercussions (Figure 1). Bank retreat drives the cross-sectional evolution of channels (Gong
properly cited. et al., 2018; Thorne et al., 1998a; van der Wegen et al., 2008), facilitates the initiation and development of channel

ZHAO ET AL. 1 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 1. Bank retreat events and their impacts across a range of temporal and spatial scales. Aerial photos are with permission from Google Earth, and the image of
Venice Lagoon is taken by Milan Radisics on 4 August 2018.

meandering (Ikeda et al., 1981; Marani et al., 2002; Seminara, 2006), alters floodplain morphology (Beechie
et al., 2006; Han & Brierley, 2020), and affects the dynamics of free and forced bars forming within the channel bed
(Blondeaux & Seminara, 1985; Solari et al., 2002; van Dijk et al., 2012). Bank retreat has also been suggested as a
major source of sediment load, thereby significantly altering sediment transfer dynamics. This latter phenomenon
occurs at a global scale as shown by studies conducted in New Zealand (Griffiths, 1979; Watson & Basher, 2006),
Europe (Duró et al., 2020; J. W. Poesen et al., 1996), the Midwestern (Simon et al., 2000) and southern United
States (Simon & Darby, 2002), Canada (Nanson & Hickin, 1986), Russia (Dong, Nittrouer, et al., 2019), China
(Xia et al., 2016; Yao et al., 2011), Egypt (Abate et al., 2015), and many other countries.

From an ecological perspective, bank retreat modulates the diversity of species and vegetation units (Piégay
et al., 1997), provides sediment supply to create habitats on a floodplain and over channel bars (Florsheim

ZHAO ET AL. 2 of 51
Reviews of Geophysics 10.1029/2021RG000761

et al., 2008), affects nutrient and contaminant dynamics (Marron, 1992; Reneau et al., 2004), induces salt-marsh
loss (Deegan et al., 2012) and so modulates the global carbon cycle (Kirwan & Mudd, 2012). With respect to
socio-economic effects, bank retreat is responsible for farmland and wetland loss (Qin et al., 2018; Turner, 1990),
population displacement (Best, 2019), reservoir siltation (Ben Slimane et al., 2016), damage to riparian infra-
structure and hydraulic structures (C. R. Hackney et al., 2020; Hooke, 1979), and creates pathways for pollutant
transport (Castillo & Gómez, 2016).

As pointed out above, bank retreat is in general caused by flow-driven particle-by-particle erosion (flow-driven
bank erosion) and gravity-driven mass failure (bank collapse). Flow-driven bank erosion consists of the removal
of bank materials under the direct action of water flow (near-bank channel flow, over-bank flow, and seepage
flow), and can be enhanced by subaerial processes. Bank collapse, in contrast, occurs when the forces that tend
to move soil downslope (soil weight, seepage forces, and excess pore water pressure) exceed the resisting forces
of the bank (cohesion imparted by the soil and root matrix, matric suction, and hydrostatic pressure head) (Fox
& Wilson, 2010; Langendoen & Simon, 2008; Rinaldi & Darby, 2007; Simon et al., 2000; Thorne, 1982). Unlike
the continuous, progressive, nature of flow-driven bank erosion, bank collapse is episodic and is associated with
various modes of failure, including shear, tensile, and toppling, that depend on bank soil properties, bank height,
near-bank water depth and the presence or absence of vegetation (Cancienne & Fox, 2008; Nardi et al., 2012;
Patsinghasanee et al., 2018; Thorne et al., 1998a; Zhao et al., 2020).

A large body of literature has been produced on bank dynamics in fluvial environments. Considerably less atten-
tion has been devoted to tidal environments, where research has to date focused mainly on the retreat of salt marsh
borders rather than tidal channel banks. Existing reviews on the subject tend to focus on specific aspects, rather
than delineating a general framework which addresses comprehensively the mechanisms that cause bank retreat
and synthesizes observations, as well as modeling frameworks describing the commonly observed bank failure
mechanisms. It is thus timely to provide a thorough review that highlights recent developments about the under-
standing and the prediction of bank retreat and point out future research needs that deserve attention to improve
our fundamental knowledge of bank retreat processes.

On the basis of field and laboratory observations, we specifically focus on a unifying description of failure
mechanisms leading to bank retreat, and refer the reader to Piégay et al. (2005) and Florsheim et al. (2008) for
prior comprehensive reviews of the geomorphic and ecological functions of bank retreat. The reader is also
referred to Couper (2004), who reviewed how temporal and spatial scales are treated in bank retreat research,
and discussed the linkages between these scales. We do not consider in situ measurement instrumentation, and
refer readers interested in this aspect to Lawler (1993a). Since our focus is bank retreat, we only review the effect
of subsurface flow on bank erosion/collapse and refer the reader to Fox and Wilson (2010) and Bernatek-Jakiel
and Poesen (2018) for the specific aspects of seepage flows. Existing review articles of bank retreat mainly focus
on fluvial environments (Castro-Bolinaga & Fox, 2018; Chassiot et al., 2020; Thorne et al., 1998a), while tidal
systems are not discussed by any of these works. The present contribution is intended to fill this gap. We note
that a large body of literature concerning numerical models have also been reviewed, for instance by Thorne
et al. (1998b) and Rinaldi and Darby (2007). Still, numerical methods advance rapidly and an assessment of
recent developments is now timely.

The present contribution aims at setting the various processes causing bank retreat in both fluvial and tidal
environments within a rational and comprehensive framework which integrates experimental observations and
numerical modeling. Given the similarity with failure mechanisms observed in many other physical systems, indi-
cating linkages in a broader context, we have considered also the literature concerning cliff retreat in salt-marsh
and marine-ice landscapes, slope stability, gully and tidal channel headcuts, and the stability of reservoir banks.
Overall, we pursue a holistic view, covering observations of failure mechanisms, empirical predictive functions,
as well as mathematical and numerical modeling.

This review is structured as follows. Section 2 defines a general classification of bank retreat and provides a
detailed description of the various types of failure mechanisms. Section 3 reviews the mechanisms of bank
collapse, as observed mainly from laboratory experiments. We distinguish bank collapse with respect to different
driving forces: subaerial processes, bank surface flow (including near-bank channel flow and overbank flow),
seepage flow, fluctuations in soil pore-water pressure, and waves. We discuss also the role of vegetation, biolog-
ical disturbances, and hydrostatic pressure head. Section 4 reviews the empirical relations developed to predict

ZHAO ET AL. 3 of 51
Reviews of Geophysics 10.1029/2021RG000761

bank retreat rate in terms of hydraulic and geometric factors such as discharge, precipitation, and bank height.
The comparison of the existing relations highlights the necessity to account for both hydraulic and geotechnical
factors. Section 5 reviews the approaches adopted to model bank erosion/collapse and proposes a hierarchical
classification of models. The advantages and limitations of each method are pointed out, and some modeling
recommendations are provided. Section 6 reviews the feedbacks between bank retreat and morphodynamics,
showing how bank retreat affects morphodynamic evolution and analyzing the importance of its integration into
morphodynamic models. Finally, Section 7 discusses the open questions, and provides recommendations for
future research.

2. Classification of Bank Retreat


We consider in detail the main aspects characterizing the two classes of processes which cause bank retreat:
flow-driven bank erosion and bank collapse. The latter can be further classified using additional criteria. In
terms of failure mode, Thorne and Tovey (1981) identified three types of cantilever failure: shear, beam, and
tensile. Based on the shape of the failure surface, Simon et al. (2000) distinguished between planar and rotational
failure types. Nardi et al. (2012) and Zhao et al. (2020) extended the above classification on the basis of experi-
mental observations of non-cohesive riverbanks. However, many misinterpretations still exist partly because of
the different terminologies often used to describe the same processes. For instance, the term beam failure and
toppling failure are applied to describe the same failure mode (Bendoni et al., 2014; Nardi et al., 2012; Samadi
et al., 2013), and pop-out failure is also called tensile failure (Fox & Felice, 2014). When investigating bank
failure mechanisms along Arno River, Dapporto et al. (2003) used the term slab-type failure to include both
mechanisms of toppling and planar failures, and some new terms are also introduced to distinguish different bank
collapse processes such as soil fall, alcove-type failure, and shallow slide failure. Therefore, it is necessary to start
by defining the classification of bank erosion that we will adopt in this review (Figure 2). In this contribution, we
use the failure mode to distinguish each type of bank collapse.

2.1. Flow-Driven Bank Erosion


2.1.1. Surface Flow Erosion

This type of erosion is quite common in sand- and silt-composed banks, and is also termed as fluvial erosion
(Darby et al., 2007; Thorne & Tovey, 1981). Here, we use the general term surface flow erosion to include both
fluvial and estuarine settings (Fagherazzi et al., 2004; Gong et al., 2018). Surface flow erosion can be defined
as the removal of bank material by the water flowing near to the bank of a channel (near-bank channel flow), or
over the bank surface itself (overbank flow). This type of erosion is commonly enhanced by subaerial processes
(Rinaldi & Darby, 2007). The rate of erosion depends on near-bank and/or overbank hydrodynamics and bank soil
erodibility. It is usually evaluated using an excess shear stress formula of the type (Partheniades, 1965):

(1)
𝐸𝐸𝑙𝑙 = 𝐾𝐾𝑙𝑙 (𝜏𝜏𝑏𝑏 − 𝜏𝜏𝑐𝑐 )

where Εl is the unit length erosion rate (m/s), Kl is the volumetric erosion coefficient for surface flow erosion
(m 3/N/s), τb is the boundary shear stress exerted by the flowing water (Pa), and τc is the critical shear stress for
sediment erosion (Pa). Previous studies show a high degree of variability in τc, with values ranging from 0.001
to 2 Pa. Typically, the value of τc depends on the type and texture of sediment composing the bank, the presence
of vegetation, etc. In particular, the two extreme values indicated for τc roughly correspond to very-erodible and
slightly-erodible material, respectively (Hanson & Simon, 2001; Midgley et al., 2012; Simon & Thomas, 2002).
The coefficient Kl can be assumed to depend on τc and estimated through the relation:

(2)
𝐾𝐾𝑙𝑙 = 𝑎𝑎 ⋅ 𝜏𝜏𝑐𝑐−0.5

where a is a regression coefficient, taking the value of 0.2 according to Hanson and Simon (2001) and 0.1 accord-
ing to Simon and Collison (2002). Nevertheless, a large variability is expected for a since, to our knowledge,
Equation 2 has been calibrated on the basis of two relatively limited databases.

ZHAO ET AL. 4 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 2. Sketch of typical bank configurations showing controlling factors in (a) fluvial and (b) tidal environments. Classification of bank retreat mechanisms and of
the associated controlling factors: bank collapse induced by (c) seepage erosion and (d) water level changes; (e–g) cantilever failure resulting from (e) tensile, (f) shear,
and (g) toppling mechanisms; (h) pop-out failure; (i) sliding failure with planar or rotational failure plane; and (j) soil creep. Here, σc and σt are the compressive stress
and tensile strength, respectively, along the failure plane.

2.1.2. Seepage Erosion


Seepage erosion results from the entrainment of soil particles by subsurface flow (T. Dunne, 1990). An exit
point/region on the bank surface is necessary for the soil particles to be dislodged from the bank interior (Wilson
et al., 2007). Seepage erosion is commonly observed in layered streambanks (Figure 2c) where high infiltra-
tion rates cause the development of perched water tables between layers of varying hydraulic conductivity (Fox
et al., 2007). Note that seepage erosion is often termed as subsurface flow erosion, sapping, piping, and internal

ZHAO ET AL. 5 of 51
Reviews of Geophysics 10.1029/2021RG000761

Table 1
Empirical Relations for Estimating Bank Retreat Rate Obtained From Laboratory Experiments
Coefficient Units

Type Ks (or b) a R 2
Left Right References
Induced by seepage
Dimensional form
𝐴𝐴   𝐴𝐴𝑚𝑚 = 𝑏𝑏𝑏𝑏𝑎𝑎𝑠𝑠 0.79 1.25 0.92 kg/s cm 3/s Fox et al. (2006)a
0.29 2.2 0.91 kg/s cm /s
3
Akay et al. (2018)b
0.08 2.3 0.91 With fibrous protection
  Em = Ks(i − ic) a 0.04 1.2 0.54 kg/s (−) Chu-Agor et al. (2009)
  tb = b ⋅ i a
81.87 −1.43 0.91 min (−) Karmaker and Dutta (2013)
150.57 −1.36 0.95
249.28 −1.42 0.93
2.08 −2.46 min (−) Masoodi et al. (2018)
Dimensionless form
𝐴𝐴   𝐴𝐴𝐴𝐴
𝑠𝑠 = Ks(𝐴𝐴𝑠𝑠 )
∗ ∗ a
584 1.04 0.86 (−) (−) Fox et al. (2006)
90 4 (−) (−) Akay et al. (2018)
25 5.4 With fibrous protection
𝐴𝐴   𝐴𝐴𝐴𝐴
𝑠𝑠 = K

𝐴𝐴s(𝐴𝐴𝑠𝑠∗ − 𝐴𝐴𝑐𝑐𝑐𝑐
∗ a
) 584 1.04 (−) (−) Fox et al. (2007)
Induced by lateral flow
  El = be at Related to initial width, flow 0.88 cm/min min Qin et al. (2018)
rate, and slope m/s s Wells et al. (2013)
a
An average of the coefficients in Fox et al. (2006) is provided. bThe coefficient in Akay et al. (2018) has been recalculated
to ensure unit consistency. Variables are summarized in the list of Symbols.

erosion (Fox & Wilson, 2010; Wilson et al., 2013). Based on laboratory experiments, the rate of seepage erosion
is typically quantified by an excess shear stress equation of the form (Fox et al., 2007):

(3) ∗ 𝑎𝑎
𝑞𝑞𝑠𝑠∗ = 𝐾𝐾𝑠𝑠 (𝜏𝜏𝑠𝑠∗ − 𝜏𝜏𝑐𝑐𝑐𝑐 )

where 𝐴𝐴𝑠𝑠∗ is the dimensionless sediment𝐴𝐴 flux (= 𝑞𝑞𝑠𝑠 ∕ (Δ − 1)𝑔𝑔𝑔𝑔3 , with Δ the ratio of bank soil density to water

𝐴𝐴
density, g the acceleration due to gravity (m/s 2), and D the grain size of the bank material (mm)),
𝐴𝐴 𝐴𝐴𝑠𝑠∗ is the dimen-
𝐶𝐶 ′′ 𝑈𝑈𝑑𝑑
sionless shear stress induced by the seepage𝐴𝐴flow (= (𝑠𝑠−1)𝐾𝐾 2
, with 𝐴𝐴2′′ an empirical parameter depending on
𝐴𝐴
sat 𝜆𝜆𝑏𝑏
sediment packing, Ud Darcy's velocity (m/s), Ksat saturated hydraulic conductivity (m/s), and λb porosity of the
bank material),
𝐴𝐴 ∗
𝐴𝐴𝑐𝑐𝑐𝑐 is the corresponding critical value for seepage erosion, Ks is the seepage erodibility coefficient,
and a is an exponent. Table 1 reports the values suggested in literature for the empirical parameters Ks and a.

2.2. Bank Collapse


2.2.1. Tensile Failure

This type of failure occurs when the tensile stress induced by the weight of the lower part of a cantilever block
exceeds the critical tensile strength of the bank soil (Thorne & Tovey, 1981). It is characterized by the presence
of tension cracks on the portion of bank that is going to fail (Figure 2e). The detachment may also occur along a
horizontal or arched surface under the action of a slight rotational component (Nardi et al., 2012). The collapse
of bank material leads to the formation of an alcove-shaped surface, and therefore tensile failure is sometimes
termed as alcove-type failure (Dapporto et al., 2003). Tensile failure is commonly observed when the upper
bank is composed of cohesive layers or covered by vegetation (J. E. Pizzuto, 1984), the roots of which impart

ZHAO ET AL. 6 of 51
Reviews of Geophysics 10.1029/2021RG000761

considerable resistance to tensile stresses. In many studies, tensile failure has also been defined using the terms
undercutting (Wilson et al., 2007), soil fail, and shallow slide failure (Dapporto et al., 2003).

2.2.2. Shear Failure


Shear failure occurs when the driving shear force, FD, acting on a potential failure surface overcomes the resist-
ing shear force, FR. The analysis of this kind of failure is typically performed on the basis of a safety factor, FS,
defined as (Osman & Thorne, 1988):

(4)
𝐹𝐹𝑆𝑆 = 𝐹𝐹𝑅𝑅 ∕𝐹𝐹𝐷𝐷

Shear failure occurs when FS is less than 1. The resisting forces are calculated by the sum of shear stress, τ
(kPa), acting along the potential failure surface. This overall stress can be expressed by the relation (Fredlund &
Rahardjo, 1993):

(5)
[ ] [ ]
𝜏𝜏 = 𝑐𝑐 ′ + (𝜎𝜎 − 𝑢𝑢𝑎𝑎 ) tan 𝜙𝜙′ + (𝑢𝑢𝑎𝑎 − 𝑢𝑢𝑤𝑤 ) tan 𝜙𝜙𝑏𝑏

where c′ is the effective soil cohesion (kPa), σ is the total normal stress (kPa), ua is the pore-air pressure (kPa),
the difference σ − ua is the net (effective) normal stress, and ϕ′ is the effective internal friction angle (°). The first
term in square brackets on the right-hand side of Equation 5 represents the shear stress for saturated soils. The
second term accounts for the additional effects which arise when the soil is partially saturated. Here, uw is the
pore-water pressure (kPa), (ua − uw) is the matric suction in unsaturated soil (kPa), and ϕ b is the angle expressing
the rate of increase in strength relative to the matric suction (°). The value of ϕ b is generally between 10° and 20°,
and approaches ϕ′ at saturation (Fredlund & Rahardjo, 1993).

Shear failure can be categorized into cantilever shear failures (Figure 2f) and sliding failure (Figure 2i). For
a cantilever profile, shear failure is determined by shear acting along a vertical, or inclined, surface, thereby
detaching the cantilever block as the crack propagates down from the bank top. This kind of failure is restricted
to sandy soils of low cohesion or silty soils of high water-content (e.g., at bank-full stage or when the tidal plain
is inundated, see Figure 1 in Zhao et al. (2020)) and to areas where the vegetation cover is sparse (Thorne &
Tovey, 1981). For high banks, sliding failures are more common than other types of failure because shear stress
increases with bank height faster than soil strength (Terzaghi, 1951). Sliding failures can be further subdivided
into planar and rotational failures (Figure 2i), depending on the shape of the failure surface (Simon et al., 2000).

2.2.3. Toppling Failure

Toppling failure is characterized by a rotational component of movement (Figure 2g). It is also called beam or
slab failure and is a very common mechanism in both fluvial and tidal environments (Allen, 1989; Francalanci
et al., 2013; Nardi et al., 2012; Samadi et al., 2013). Toppling failure is accompanied by one or several deep
tension cracks on the bank top, and eventually occurs when the moment along the failure plane overcomes the
resistance provided by soil cohesion and/or vegetation roots (Van Eerdt, 1985). Contrary to cantilever shear fail-
ure, toppling failure is driven by excess moment rather than excess shear stress (see Figures 2f and 2g).

2.2.4. Pop-Out Failure

Pop-out failure (also called tension failure, see e.g., Fox and Felice (2014)) occurs when seepage forces are
greater than soil resistance, or shear/tension strength decreases owing to an increased soil pore-water pressure
(Figure 2h). Unlike seepage erosion characterized by particle entrainment and mobilization, pop-out failure
consists of a block failure with the formation of tension cracks (Chu-Agor et al., 2008).

2.2.5. Erosion and Failure Resulting From Loss of Matric Suction

This type of processes occurs due to a loosening of the weak links between particles (especially sand) when the
pores are saturated by water and the weight of bank soil increases (Nardi et al., 2012). This kind of failure differs
from tensile failure because the bank soil is close to saturation. It can be regarded as the consequence of tensile
failures and occurs at the endpoint of the commonly observed arched shape close to the water surface.

2.2.6. Soil Creep

Soil creep is a gravity-induced viscous-like slow deformation which produces a net downslope transport of bank
soils (Figure 2j) and is commonly observed in salt-marsh banks (Mariotti et al., 2016, 2019). Although the

ZHAO ET AL. 7 of 51
Reviews of Geophysics 10.1029/2021RG000761

behavior is similar to shear failure, soil creep occurs at a much slower rate of deformation (around 20–50 mm/year,
see Mariotti et al. (2019)).

3. Mechanisms Leading to Bank Retreat


Bank collapse can be triggered by subaerial process, surface flows which establish in fluvial and tidal environ-
ments, seepage flow, wind and ship waves, as well as by variations in soil strength (i.e., soil cohesion) resulting
from water level changes, rainfall impacts and infiltration, and evaporation. Other factors such as vegetation root
reinforcement and hydrostatic pressure head, in contrast, favor bank stability and so might affect the bank collapse
mode. Based on experimental and in situ observations, here, we review mechanisms leading to bank collapse with
respect to subaerial process, surface flow, seepage flow, fluctuations in soil pore-water pressure head, and waves,
paying specific attention to the role of vegetation, biological disturbances, and hydrostatic pressure head.

3.1. Subaerial Process

Subaerial processes are commonly associated with the slow-motions driven by local climate variations that
directly deliver bank soil to the channel, or which act as preparatory processes which weaken the face of the bank
prior to surface flow erosion (Couper & Maddock, 2001). Their contribution to bank retreat has been analyzed
by introducing the concept of “process zonation”, that is, identifying the dominant factor among near-bank
hydrodynamics, bank soil properties, vegetation cover, and local climate change (e.g., temperature and rainfall)
(Abernethy & Rutherfurd, 1998; Henshaw et al., 2013; Lawler, 1992; Lawler et al., 1999).

In the middle and lower reaches of a river (horizontal zonation), fluvial erosion and bank collapse events domi-
nate bank retreat, while subaerial erosion serves as a preparatory process reducing the resistance of bank soils
by up to 80% of their original strength (Kimiaghalam et al., 2015; Wynn et al., 2008). In contrast, bank retreat
is controlled by subaerial processes in headwater streams or where hydrodynamic power is insufficient to cause
soil erosion without prior weakening of bank materials (Prosser et al., 2000; Veihe et al., 2011). For example,
it has been demonstrated that 85% of the annual retreat of coastal cliffs in a cold temperate climate may occur
due to subaerial erosion (Bernatchez & Dubois, 2008). Similarly, direct measurements carried out in bedrock
rivers show that the lateral bank retreat can be substantially enhanced by the higher erodibility of rock exposed to
continuous wetting-drying cycles (Montgomery, 2004). Subaerial erosion typically affects the upper part of the
bank while the lower part is subject to fluvial erosion. This gives rise to a vertical zonation of the controls of bank
erosion which eventually determine the bank morphology depending on the soil characteristics (e.g., the silt-clay
content (Couper, 2003; J. Pizzuto, 2009; Veihe et al., 2011)).

Overall, subaerial processes may play a fundamental role in bank retreat. Repeated wetting-drying cycles (leading
to associated desiccation) and freeze-thaw sequences are widely recognized as major subaerial processes in both
fluvial and estuarine systems (Bernatchez & Dubois, 2008; Couper & Maddock, 2001). Wetting of bank soils
is generally the result of prolonged high flows, exchanges between surface water and groundwater (due to water
level variations), and infiltration of precipitation (Simon et al., 2000; Xin et al., 2022). This process affects bank
retreat by increasing soil pore-water pressure with a consequent weakening of the linkage between soil particles
(e.g., matric suction mentioned by Nardi et al. (2012)) which eventually makes the bank soils more susceptible to
surface flow erosion (Zhao et al., 2020). We refer the reader to Section 3.4 for more details about this mechanism.
On the other hand, the reduction in soil-moisture content associated with drying may lead to soil desiccation with
consequent fissuring and exfoliation of the bank surface (Figure 3a). The strength of the bank is therefore weak-
ened, especially for cohesive banks where soil aggregates are separated by desiccation cracks (Couper, 2003;
Thorne & Tovey, 1981). For example, in the St. Lawrence River, desiccation followed by a rapid rewetting was
found to increase bank retreat rate by several orders of magnitude (Gaskin et al., 2003).

Freezing periods cause an expansion of water within soil voids which weakens the linkage between soil particles,
while the subsequent thawing action directly induces subaerial erosion (Yumoto et al., 2006). Short-duration of
freeze-thaw successions facilities the growth of needle-ice on the bank face (Figure 3b), while annual freeze-thaw
cycles may lead to the development of deep cracking and, consequently, soil mass failure (Kimiaghalam
et al., 2015; Lawler, 1993b). In general, freeze-thaw cycles play an important role in subaerial erosion in regions
where the bank is subject to deep seasonal frost (Chassiot et al., 2020). Although the number of freeze-thaw
cycles is likely the major factor for subaerial erosion (Kimiaghalam et al., 2015; Wynn et al., 2008), several

ZHAO ET AL. 8 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 3. (a) Soil desiccation on a muddy bank. The wet (left) and dried (right) mud areas are distinguished by the white
dashed line (adapted from Zeng et al. (2022)). (b) Typical needle ice formation under freeze-thaw cycles (adapted from C. Li
et al. (2018)).

studies have also highlighted that silt-clay content (Couper, 2003), soil water content (Ferrick & Gatto, 2005), and
vegetation cover (Prosser et al., 2000; Wynn & Mostaghimi, 2006) matter as well. Banks with a higher soil mois-
ture and silt-clay content are more affected by freeze-thaw cycles, while vegetation cover dramatically reduces
the frequency of these cycles.

3.2. Surface Flow


3.2.1. Near-Bank Channel Flow

Near-bank channel flow is generally responsible for the removal of soil particles from the bank. For high enough
flow velocities, this removal may undermine the bank toe, eventually leading to the formation of a bank profile
with cantilever-shape (Thorne & Tovey, 1981). Samadi et al. (2011) carried out several flume experiments using
soil blocks of silt and silt-clay, respectively, to mimic the actual bank configuration caused by artificially repro-
duced undermining. The observed failure process included: (a) tensile failure in the lower part of the bank; (b)
tension cracks on the bank top; and (c) toppling failure forming a vertical cracked interface (Figure 4a). The flume
experiments also suggested that the pattern of cantilever failure is dominated by toppling failure rather than the
simple shear-type mechanism considered by Darby et al. (2007) and Rinaldi et al. (2008).

To overcome the drawbacks induced by an artificially reproduced undermining, several downscaled physical
experiments have been conducted in recent years under conditions of uniform and steady flow. Patsinghasanee
et al. (2018) used two types of cohesive materials with different silt-clay contents to investigate the process of
bank collapse under similar hydraulic conditions. The observed failure sequence was similar to that reported by
Samadi et al. (2011). The tension cracks on the bank top seemed to develop only when the cantilever was close to
failure. A slight increase in silt-clay content was found to have negligible effect on bank failure patterns, but could
significantly shorten the time taken to collapse. A number of studies have been also carried out to investigate the
role of impulsive flow due to dam-breaks (Cantelli et al., 2004; Zech et al., 2008), near-bank bed evolution (Yu
et al., 2015), bank height (Patsinghasanee et al., 2017), channel bed slope (Qin et al., 2018), and near-bank turbu-
lence (Das et al., 2019; Roy et al., 2019). Even though the focus of Qin et al. (2018) was gully erosion, this type
of process can be regarded as a downscaled case of bank erosion relevant also for tidal environments (e.g., for
channel head migration). It is however worthwhile to note that downscaled physical models are useful to obtain
general information on bank failure patterns (D. M. Wood, 2014), but entail uncertainties about the interpretation
of the collected data which restrict their use in formal quantitative analysis. For example, tension cracks on the
surface of the bank (Figure 4a), and the consequent tensile failures documented by Samadi et al. (2011) are not
observed in many downscaled experiments. Besides, only a few of the existing downscaled physical experiments
attempted to capture variations in bank soil parameters (e.g., soil stress and pore-water pressure) occurring during
bank collapse.

Recently, Zhao et al. (2020) set up a laboratory experiment with limited scaling effects to investigate the role
of bank height, Hb, and near-bank water depth, Hw, on bank collapse. Results showed that the failure patterns
correlate to the ratio Hw/Hb (i.e., the relative water depth). For relatively small values of Hw/Hb (less than 0.5,

ZHAO ET AL. 9 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 4. (a) Typical overhanging failures in response to artificial undermining (adapted from Samadi et al. (2013)). Bank height in subplot (a) was 0.8 m. (b and c)
Typical erosion and failure processes experienced by vertical banks in response to near-bank channel flow, for a water depth of 0.15 m, and bank height of (b) 0.6 m and
(c) 0.2 m (adapted from Zhao et al. (2020)).

Figure 4b), the decrease in matric suction led to the occurrence of cracks and consequent tensile failures. Several
deep tension cracks were observed on the bank top, followed immediately by toppling failure. For values of Hw/
Hb close to 1 (Figure 4c), tension cracks were first observed on the bank top. A large (compared to the stress
exerted by soil weight) hydrostatic pressure head prevented the occurrence of toppling failure. Shear failures thus
occurred along a vertical (or inclined) surface separating the cantilever block from the bank top. Extending the
experimental analysis to a broader range of values of the ratio Hw/Hb, K. Zhang et al. (2021) further pointed out
the need to include water content variation when predicting shear or toppling failure. Their experimental frame-
work provides a new way to predict failure patterns by combining hydraulic and geotechnical control factors.

Natural rivers usually convey bedload and/or suspended load downstream. The transported grains, in turn, stresses
the channel bed/bank, possibly affecting bank retreat. Bedload saltation produces an abrasion of the exposed bank
surface. The development of an alluvial cover due to bed aggradation alters the cross-sectional distribution of
shear stresses affecting indirectly bank retreat. Field and laboratory evidence of abrasion-induced bank retreat is
plentiful in bedrock rivers (see e.g., the shallow depressions in Figure 5a), where the channel bed and banks often
consist of rocks covered by a thin layer of alluvium (Carter & Anderson, 2006; Fuller et al., 2016; Hartshorn
et al., 2002). The direct bumping and abrasion by saltating grains can increase channel width by up to an order of
magnitude as compared to channels with negligible bedload (Baynes et al., 2020; Fuller et al., 2016). On the other
hand, bank retreat of bedrock channels can be indirectly affected by the increased bed roughness which charac-
terizes the formation of an alluvial cover. Experimental observations have shown that this increase in roughness
enhances the deflection of saltating bedload particles toward the bank surface (Figure 5b and 5c), causing rates
of lateral erosion to grow by as much as a factor of 7 (Fuller et al., 2016).

The effects of alluvial cover on bank retreat depend also on the channel planform (e.g., meandering, braiding,
and pinch-swell undulation), which strongly reflects on flow and sediment routing (Carter & Anderson, 2006;
Mishra et al., 2018). In braided gravel-bed channels, lateral incision is weakly sensitive to the rate of sediment

ZHAO ET AL. 10 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 5. (a) Field evidence of bedrock wall erosion by sediment impacts. (b and c) Laboratory evidence showing the role of
bed roughness on bedrock wall erosion (adapted from Fuller et al. (2016)).

transport (Bufe et al., 2016, 2019). Conversely, in sharp bedrock bends an increased sediment supply accelerates
bank retreat and shifts its location upstream (Mishra et al., 2018). In bedrock channels, the development of an
alluvial cover alters the cross-sectional distribution of shear stress, ultimately leading to a transition from vertical
incision of the rocky bed to lateral expansion due to bank retreat (Turowski et al., 2008). As a result, a substantial
sediment supply not only leads to the formation of a significant alluvial cover but is commonly accompanied by
a fast lateral bedrock erosion (Finnegan et al., 2007; T. Li et al., 2020).

In alluvial meandering rivers, high sediment loads favor a rapid accretion of the point bar at the inner bank,
which forces erosion at the outer bank (bar push effects) and consequently enhances the annual migration rate
(Constantine et al., 2014; Donovan et al., 2021; E. Eke et al., 2014). On the other hand, nearly clear water condi-
tions associated with a scarce sediment supply facilitate a progressive incision of the channel bed, which causes
outer bank instability and ultimately speeds up meander migration. For example, the sharp reduction in sediment
supply (around 80%) caused by upstream damming was responsible for the increased bank retreat rate observed in
the Jingjing Reach of the Changjiang River over the last 20 years (Xia, Zong, Zhang, et al., 2014; Xia et al., 2017).
Overall, sediment load plays a vital role in the bank retreat process, and more
research is needed to quantify its contribution to bank retreat rate.

In cold environments, the failure by near-bank channel flow may be also


affected by soil deterioration (i.e., freeze-thaw cycles) and the formation of
ice blocks, which has been proven to be important especially in coastal and
estuarine settings (Bernatchez & Dubois, 2008). The collapsed ice blocks may
be swept away at high tide, thus further enhancing bank erosion by attack-
ing the bank surface and by undercutting bank materials, while during ebb
tides they add friction to the system (Black et al., 2018). We refer the reader
interested in erosion processes in cold environments to Chassiot et al. (2020),
who systematically evaluated and summarized the effects of ice on bank
retreat. Another mechanism resulting in the formation of a cantilever is found
in marine ice cliffs, where the cantilever is caused by the action of warmer
deep water (DeConto & Pollard, 2016). Regardless of the peculiarities of
cold environments, bank collapse processes are still detected whereby the
cliff evolution is related to cliff height and water depth (Bassis et al., 2021).
3.2.2. Overbank Flow

Overbank flow occurs over the bank surface, as commonly observed in gullies
Figure 6. Headcut retreat for: (a) gully and (b and c) tidal channel head. In and tidal channels, and controls headcut retreat in these environments. Gully
panel (a), the length and width of the drainage area are 5 and 2 m, respectively, headcuts, characterized by near-vertical steps, form often in dry-hot valleys
and the initial headcut height is 0.5 m. In panel (b), the circled bird footprint is as concentrated overland flow initiates erosion on a gully bed composed of
about 0.04 m. In panel (c), the width and length of the flume are 0.4 and 2 m, a hard-upper layer and a soft-lower layer (A. Chen et al., 2013; Rengers &
respectively, and the time between each snapshot is 6 hr (adapted from Dong,
Xiong, et al. (2019) and Kleinhans et al. (2009)).
Tucker, 2014). When passing over the overhanging block (Figure 6a), the

ZHAO ET AL. 11 of 51
Reviews of Geophysics 10.1029/2021RG000761

overbank flow either moves along the sidewall scouring the overhanging profile (on-wall runoff) or directly falls
into the channel forming a plunge pool (off-wall runoff). With the development of scouring erosion in the lower
headcut layer, the overhanging block gradually becomes larger, ultimately triggering different types of mass
failure such as tensile, sliding, and toppling (A. Chen et al., 2015). In the presence of a vegetation cover, tensile
failure occurs in the middle face of the headcut below the root zone, leaving behind an overhanging block. Failure
takes place when block weight exceeds vegetation root strength (Rengers & Tucker, 2014).

Experimental studies have addressed the development and migration of gully headcuts with emphasis on differ-
ent factors such as bed slope (Bennett, 1999), flow discharge (Bennett et al., 2000), and soil texture (Wells
et al., 2009). Several studies focus on mass failure rather than plunge pool erosion. For example, Stein and
LaTray (2002) set up a specific gully bed where a relatively erosive base soil is overlaid by a relatively hard
soil layer. For this geometry, mass failures were observed as a result of undercutting in the plunge pool immedi-
ately downstream. Experimental and in situ studies have also been conducted to investigate the failure process
of headcuts in response to on-wall runoff. A. Chen et al. (2013) performed flow scouring experiments with a
set-up involving a lower sandy layer and an upper clay layer. Results showed that the collapse of overhanging
layers was dominated by the development of the lower scour holes. The occurrence of upper cracks significantly
accelerated the collapse process. Compared to off-wall runoff, on-wall runoff played a more important role in
the development of scour holes, and headcut collapse was found to depend on runoff duration rather than runoff
intensity. This implies that variations in soil strength, rather than cantilever development, are responsible for mass
failure, since soil cohesion decreases dramatically with the increase in soil moisture (Rajaram & Erbach, 1999).
Similar findings emerged from the in situ observations of Rengers and Tucker (2015), who pointed out that high
water content rather than hydraulic scour was the main cause for mass failure. Other factors such as summer flash
floods, winter snowmelt, prolonged summer dry periods, and drying-rewetting cycles can also be correlated with
headcut collapse (Dong, Xiong, et al., 2019; Rengers & Tucker, 2014). For instance, Dong, Xiong, et al. (2019)
conducted 11 in situ flow scouring experiments with different discharges. Results showed that the variation in
discharge had little effect on mass failure frequency, whereas drying-rewetting cycles accounted for 64% of the
observed mass failure. We refer the reader to J. Poesen et al. (2003), Valentin et al. (2005), and Castillo and
Gómez (2016) for more information on gully headcut erosion.

Similar to gully headcuts, but with a much smaller spatial scale, a fast retreat of tidal channel heads is commonly
observed on tidal flats. Symonds and Collins (2007) reported an annual headcut retreat rate of up to 400 m/year in
response to a managed coastal realignment. Through laboratory experiments and in situ observations, Kleinhans
et al. (2009) found that the retreat of tidal channel heads was induced not only by gradual erosion of the chan-
nel bed, but also by repeated cantilever or sliding failures at the headcut border typical of cohesive sediments
(Figures 6b and 6c). They also concluded that the failure was mainly related to soil weakness induced by waves,
rain, or excess pore-water pressure. Other studies have been carried out to investigate the role of tidal channel
headcuts on, for example, the evolution of tidal flats (Ni et al., 2014), tidal channels (Xu et al., 2019), and tidal
networks (Geng et al., 2019; Kleinhans et al., 2012; Xu et al., 2017).

3.3. Seepage Flow

Bank collapse related to seepage flow can be attributed directly to seepage erosion and resultant tensile failure
(direct influence), or indirectly to the effects of seepage flow on soil properties (indirect influence) (Fox &
Wilson, 2010). Since the indirect influence is mainly related to soil pore-water pressure, we refer the reader
to Section 3.4 for more details. To mimic bank collapse in response to seepage erosion, Fox et al. (2006) and
Wilson et al. (2007) performed lysimeter experiments using reconstructed banks packed with three different soil
layers (Figure 7). They suggested that increasing bank height, or hydraulic head, leads to more bank retreat (i.e.,
a decrease in bank stability), and alters the associated pattern of bank collapse. For large bank heights and high
hydraulic heads, they observed a sequence of seepage erosion at the bank toe, tensile failure in the middle of the
bank, and toppling failure. For small bank heights and low hydraulic heads, tensile failure was absent and bank
erosion was sometimes characterized by seepage erosion (i.e., no toppling failure occurred).

A series of studies were also carried out to investigate how seepage flow erosion is affected by bank slope
(Fox et al., 2007), soil density (soil texture) (Chu-Agor et al., 2008; Fox & Felice, 2014), bank stratification
(Lindow et al., 2009), root reinforcement (Akay et al., 2018; Cancienne & Fox, 2008), and soil chemical prop-
erties (Masoodi et al., 2017, 2019). Cancienne and Fox (2008) found that, under the same hydraulic conditions,

ZHAO ET AL. 12 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 7. Erosion and failure processes of river banks in response to seepage: (a) seepage erosion; (b) tensile failure in the
middle of the bank; (c) tension cracks on the bank top; and (d) toppling failure. Bank height is 0.8 m, and seepage flow within
the bank body is from right to left (adapted from Fox et al. (2006)).

seepage erosion and undercutting patterns exhibit remarkable differences in vegetated and unvegetated contexts.
Unlike their unvegetated counterparts, vegetated bank blocks were characterized by the absence of unimodal
undercutting (see Figure 2 in Chu-Agor et al. (2008)) and the formation of shallow multimodal cuts along the
entire width of the bank face. To evaluate and predict the development of seepage erosion followed by toppling
failure or the triggering of pop-out failure, Fox and Felice (2014) proposed a dimensionless seepage mechanism
number based on the ratio between resistive cohesion forces and destabilizing forces. This number reasonably
predicted the type of observed seepage failures, and is suitable for streambanks or hillslopes experiencing steady
seepage forces. On the basis of in situ observations, Masoodi et al. (2019) found a high correlation coefficient
(R 2 = 0.81) between soil dispersion (see Glossary) and the volume of the undercutting cavity caused by seepage.
At the same time, many studies aimed at developing empirical sediment transport functions relating seepage
erosion rates to controlling factors (Chu-Agor et al., 2008, 2009; Fox & Felice, 2014; Fox et al., 2006, 2007;
Karmaker & Dutta, 2013; Wilson et al., 2007). For example, on the basis of 71 lysimeter experiments, Karmaker
and Dutta (2013) suggested that the seepage flow gradient and the vertical stratigraphy have a dominant effect on
bank stability. We refer the reader to Section 4.1 for more details about these empirical relations.

3.4. Fluctuations in Soil Pore-Water Pressure

The bank soil tends to be unsaturated above the water surface and saturated below the water surface. In tidal
settings, due to the periodical rising and falling of the water level, this difference is less evident (e.g., when
the water level in tidal channels is just below the elevation of the bank top during the early stage of the ebb
tide, see Figure 1 in Zhao et al. (2020)). For unsaturated soils, increasing soil pore-water pressure implies a
smaller matric suction and hence reduced soil strength. For saturated soils, the matric suction is zero and the
soil shear strength ultimately depends on the effective normal stress. The transition from saturated to unsatu-
rated soil (and vice versa), clearly indicated by soil pore-water pressure, is mainly driven by variations in water
level, and often leads to bank collapse. For example, Nardi et al. (2012) investigated the basic processes and
possible factors influencing the instability of relatively coarse (sandy-gravel) river banks in response to water

ZHAO ET AL. 13 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 8. Erosion and failure processes of river banks in response to water level (a and b) rising and (c) falling. (a) Coarse sandy-gravel bank of height 0.7 m; (b) fine
sand bank of height 0.5 m; and (c) compacted sand bank of height 0.47 m (adapted from Nardi et al. (2012), Arai et al. (2018), and Chen, Hsieh, and Yang (2017)).

level changes (Figure 8a). Contrary to cohesive banks, for which bank failure is commonly observed during the
falling water stage (saturated-unsaturated transition), they suggested that the reduction of apparent cohesion (e.g.,
matric suction) during the rising water stage (unsaturated-saturated transition) caused the instability of the banks.
This finding indicates an inherent difference in bank failure mechanisms between sandy-gravel and sand-clay
banks, arising from the distinct response of soil strength to very high water-content. In contrast to clay soils, silt
is prone to losing almost all strength (i.e., static liquefaction) when approaching a saturated state (Yamamuro &
Lade, 1998). Using finer cohesionless uniform sand, Arai et al. (2018) observed more cantilevers and, conse-
quently, toppling failures as a result of the stronger apparent cohesion and tensile strength (Figure 8b). They also
stated that dry granular flow, as reported by Nardi et al. (2012), was absent, emphasizing the role of soil grain
size on bank collapse patterns.

For cohesive banks and decreasing water levels, Francalanci et al. (2013) observed that excess pore-water pres-
sure and decreasing hydrostatic pressure head could trigger tensile failures in the middle part of banks mimick-
ing marsh cliffs. In tidal environments, these two pressures tend however to counterbalance at high tides. C.
Chen, Hsieh, and Yang (2017) and Khatun et al. (2019) both stated that a rapid drawdown was a major cause
for sand-compacted bank instability (Figure 8c). However, the main cause of bank failure with respect to water
level drawdown is still unclear. Two destabilizing effects are in fact triggered during a falling water stage (Simon
et al., 2000). On one hand, the drawdown of water level implies a progressive reduction of the stabilizing action
exerted by the hydrostatic pressure head. On the other hand, positive pore-water pressure leads to a reduction
in soil shear strength. Applying stress-strain analysis and taking into account varying hydrostatic pressure head,

ZHAO ET AL. 14 of 51
Reviews of Geophysics 10.1029/2021RG000761

Gong et al. (2018) concluded that bank retreat rate is partly related to the rate of water level change. In particu-
lar, the maximum bank retreat was found to occur in conjunction with the maximum rate of decrease in hydro-
static pressure head. Deng et al. (2018) demonstrated that pore-water pressure, resulting from the delay between
groundwater level and river stage, was mainly responsible for shear-type bank collapse. For banks with relatively
low soil permeability, the initial groundwater level may also play an important role in controlling bank stability.
Anyhow, bank instability caused by a falling water stage might in general be related to bank geometry, stratigra-
phy, and soil properties such as soil density and permeability (Pollen-Bankhead & Simon, 2010). Further research
is needed to fully unravel the failure mechanism with respect to water level drawdown.

Other factors such as evapotranspiration and infiltration (induced by vegetation, seepage and rainfall) can cause
fluctuations in soil pore-water pressure and thus are also likely to affect bank stability. Vegetation affects soil
pore-water pressure either by extracting soil moisture via evaporation, or by intercepting rainfall that would
otherwise infiltrate into the bank. Both processes potentially reduce positive pore-water pressure and instead
facilitate the development of matric suction. For instance, Simon and Collison (2002) and Pollen-Bankhead
and Simon (2010) pointed out that the variation in matric suction due to evapotranspiration might provide more
benefits to riverbank stability than the mechanical effects associated with vegetative root reinforcement. Given
the timescale, evapotranspiration may be relevant in fluvial environments, while it likely plays only a minor
role in tidal settings. Indeed, the period during which the salt marsh surface is inundated could be too short
for evapotranspiration to result in clear reductions in matric suction. Nevertheless, Dacey and Howes (1984)
observed that marsh plants might lower the water table and reduce pore pressure very quickly. Their findings
identified a new mechanism for salt-marsh bank stability whereby the generation of excess pore-water pressure
and removal of hydrostatic pressure during the ebb tide are counteracted by a quick increase in matric suction.
Wynn and Mostaghimi (2006) suggested that vegetation covers facilitate stream bank stability not only by inter-
cepting rainfall, but also by reducing subaerial processes such as soil desiccation and freeze-thaw cycles, which
make the banks more vulnerable to surface flow erosion (Chassiot et al., 2020). However, as pointed out by
Durocher (1990), canopy interception and stemflow can concentrate rainfall locally around the plant roots, deter-
mining local highs of pore-water pressure which weaken bank stability.

Seepage flow has also been suggested to affect soil pore-water pressure and, consequently, bank stability. Fox
et al. (2006) observed a sharp increase in soil pore-water pressure owing to the transition from unsaturated to satu-
rated conditions caused by seepage, followed immediately by several bank failures. Lindow et al. (2009) found
that variations in soil pore-water pressure consequent to seepage depend on the initial bank slope. For banks with
a gentle slope, a relatively small increase in soil pore-water pressure is sufficient to trigger bank collapse (Fox
& Wilson, 2010). In tidal settings, seepage flow due to diurnal changes of water level is responsible for periodic
transitions from unsaturated to saturated conditions and vice versa. In the presence of high water levels (e.g.,
inundating tides and large river flow), water overtops and enters through the creek bank, making at least part of
the bank soil fully saturated. This process is possibly enhanced by a higher permeability of the upper soil layer,
as a result of macropores created by plant root and organism burrowing (Harvey et al., 2019; Xin et al., 2022). In
contrast, groundwater gradually seeps out from the bank during falling tides and, together with high evapotranspi-
ration (due to some combination of sun, wind and temperature), can lead to complete drying or, even desiccation
of bank material (Derksen Hooijberg et al., 2019; McKew et al., 2011). The drying-wetting of tidal channel banks
and marsh edges during tidal cycles thus plays a key role on retreat processes in tidal environments.

Finally, infiltration of rainfall may lead to a loss of matric suction and, therefore, may reduce bank soil strength
and bank stability (Simon et al., 2000). Using a prototype model, Okura et al. (2002) showed that the generation
of excess pore-water pressure as a result of continuous rainfall infiltration resulted in bank instability and land-
slide fluidization. For tidal environments, Mariotti et al. (2019) observed that creep movement of a salt-marsh
bank was more accentuated during rainfall events. Therefore, they suggested that this behavior is likely governed
by effective stresses and their dependence on pore water pressure. L. Z. Wu et al. (2017) investigated the role
of rainfall intensity on the development of pore-water pressure. Results showed that increasing rainfall intensity
facilitated the development of high soil pore-water pressure, possibly producing gravity-driven landslides. A
review of slope stability analysis under rainfall-driven infiltration, with a focus on conceptual models, analytical
analysis, and numerical modeling, has also been undertaken by L. L. Zhang et al. (2011).

ZHAO ET AL. 15 of 51
Reviews of Geophysics 10.1029/2021RG000761

3.5. Waves

Bank instability is also caused by the attack by wind waves and ship wakes in rivers, salt marsh channels, and
navigable waterways. The associated mechanisms of bank collapse can be attributed to (a) terrace erosion adja-
cent to the bank toe, (b) removal of soil particles from the bank (wave erosion), (c) mechanical fatigue of the bank
soil, and (d) changes in soil pore-water pressure. On the basis of in situ experiments, Nanson et al. (1994) found
that for small rivers, boat-generated waves may dominate bank erosion as compared to surface flow erosion and
seepage erosion. Duró et al. (2019) reported that, for a relatively constant water level, boat-induced waves caused
the formation of a mildly-sloping terrace adjacent to the bank toe, which progressively dominated the bank
erosion process. Emergent vegetation however plays a significant role in reducing wave erosion by attenuating the
wave impact (Coops et al., 1996) and reinforcing the soil through the root system (Gabel et al., 2017).

Sunamura (1982) designed a laboratory experiment to investigate wave erosion mechanisms at the base of a beach
cliff. He found that turbulence due to wave runup destabilized the sand and generated a shear stress on the cliff
face, leading to cliff erosion/collapse. To quantify the relative importance of vessel-generated versus wind waves
to salt marsh cliff retreat, Houser (2010) carried out a field study in the Savannah River. He found that locally
generated wind waves were largely responsible for the observed cliff retreat. In the case of soil-geotextile filtra-
tion, Faure et al. (2010) observed that erosion of the middle part of revetments was induced by the up-and-down
drag force exerted by waves along the bank. More recently, studies have been carried out by Ji et al. (2017, 2019)
to investigate the role of bank profile morphology on the stability of reservoir banks. Under the action of wave
erosion, a concave cavity was found to occur near the water surface, leading to tension cracks on the bank top
and consequent toppling or sliding failures. Using the width of the collapsed portion of the bank as a metric,
they concluded that a convex shape was the most unfavorable condition for bank stability, followed by a concave
shape, straight shape, and a multi-stepped shape. Contrary to surface flow erosion, wave erosion occurs only
around the water surface and so results in a large underwater shoal, which in turn provides resistance to bank
collapse. As a result, this kind of failure mechanism is episodic and essentially inactive until the submerged bank
toe is removed by other processes such as surface flow erosion or artificial dredging.

Apart from direct erosion under wave attack, bank soil may also be rendered unstable as a result of mechanical
fatigue in response to wave loading (Coops et al., 1996; Hooke, 1979). Ginsberg and Perillo (1990) reported that
the continuous impinging due to locally-generated waves may lead to mechanical fatigue, significantly reducing
channel bank stability. To gain more insight into the effect of dynamic wave loads, Bendoni et al. (2014) devel-
oped a theoretical model to account for the instantaneous action of waves rather than averaging the waves over
long time intervals. Contrary to a simple static model, the dynamic response can account for the elastic potential
energy and inertial effects, resulting in higher stress and consequently predicting more rapid bank failure. Using
laboratory experiments, they also concluded that water inside the tension cracks and low water levels in front of
the bank were the two most favorable conditions to trigger bank collapse.

High excess pore-water pressures inside the bank, triggered by cyclic wave loadings, can also be an important
cause for upper bank collapse (Faure et al., 2010). Francalanci et al. (2013) carried out a prototype laboratory
experiment to reproduce salt marsh cliff retreat under the combined action of tides and wind waves. The succes-
sive development of tension cracks and upper bank deformation, tensile failure in the middle of the front surface,
and toppling failure was the typical succession of the observed failure processes (Figure 9a). This kind of failure
sequence is similar to that experienced by banks in response to surface flow (Patsinghasanee et al., 2018; Samadi
et al., 2011), seepage flow (Fox et al., 2006; Wilson et al., 2007), or water level changes (Arai et al., 2018; Nardi
et al., 2012). Overall, this shows that external forces operating in different environmental contexts may have
similar effects on bank stability, leading to similar failure mode(s). Although in the experiments of Francalanci
et al. (2013) tidal cycles were the main reason for crack formation and upper bank deformation, the authors
suggested that wind waves are likely to provide an additional mechanism accelerating the occurrence of bank
collapse. Any overtopping waves are in fact likely to induce water infiltration within the tension cracks on the
bank top and promote the occurrence of extra pore-water pressures inside the bank. They also pointed out that
the presence of vegetation roots made some notable differences (Figure 9). For instance, tension cracks on vege-
tated banks were generally smaller and narrower, and measured pore-water pressures at low tides were generally
lower (10% decrease) than when vegetation was absent, both factors contributing to delay the occurrence of bank
collapse.

ZHAO ET AL. 16 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 9. Erosion and failure processes of marsh cliffs in response to tidal currents and wind waves, (a) without and (b) with vegetation. The onward time from the
beginning of Experiments 4.1 and 8.1 are reported on each photo, showing the occurrence of mass failure, while other photos display the final geometry of bank surface
during each experiment. Bank height was 0.6 m (adapted from Francalanci et al. (2013)).

The combined influences of waves and vegetation roots on bank stability has been highlighted by many other
studies (Mariotti et al., 2019; Y. Chen & Collins, 2007; Y. Chen et al., 2011). Contrary to coarse roots with high
resistance to soil creep, Y. Chen et al. (2019) found that fine roots with high resistance were more effective in
stabilizing marsh banks in response to flow- and wave-induced erosion. However, Feagin et al. (2009) argued that
soil type, rather than salt marsh plants, is the primary variable affecting the erosion rate of the salt marsh edge,
thus challenging the common perspective that salt marsh plants prevent wave-induced erosion. This conclusion is
also consistent with the work of Bendoni et al. (2016), who stated that vegetation roots prevent particle-by-par-
ticle erosion on the upper layers of the bank and, hence, facilitate the formation of cantilever profiles, which in
turn, may lead to more frequent mass failure events.

Biological disturbances matter as well. Biofilms provide a positive or negative effect on sediment stability depend-
ing on the driving forces (e.g., steady flow vs. waves (X. Chen et al., 2021; X. D. Chen, Zhang, et al., 2017)), while
a negative effect is provided by shells and crabs (Bortolus & Iribarne, 1999; Quaresma et al., 2007; Thompson
& Amos, 2002). A comprehensive review is provided by Harvey et al. (2019), who summarized the mechanisms
of bank instability induced by burrows as: (a) altering bank geometry, (b) inducing subsurface flow (seepage),
and (c) facilitating erosion at bank surface and burrow entrances. Overall, more research is needed to establish
general parameterizations that can describe and quantitatively predict the influence of vegetation and biological
disturbances on salt-marsh stability.

4. Bank Retreat Rate


On the basis of laboratory experiments and in situ observations, many empirical relations have been developed to
relate bank retreat rate to controlling factors such as flow discharge, near-bank flow velocity and bank geometry.
A summary of typical empirical functions to predict bank retreat rate, and their regression coefficients, is shown

ZHAO ET AL. 17 of 51
Reviews of Geophysics 10.1029/2021RG000761

Table 2
Empirical Correlations for Estimating Bank Retreat Rate Obtained From Field Observations
Coefficient Units

Type K (or b) a Qc/c/Pc d R 2


Left Right Time scalea References
(1) Induced by seepage
𝐴𝐴 𝐴𝐴𝑠𝑠 = 𝑏𝑏𝑏𝑏𝑎𝑎𝑠𝑠 3.7 2.12 0.69 g/L L/day Event Wilson et al. (2007)
Em = Ks(Qs − Qc) a
1,700 1 0.2 0.89 g/min L/min Event Midgley et al. (2013)
Vs = b ∗ CEI + d 99.156 1 −39627 0.77 cm 3 (−) (−) Masoodi et al. (2017)
(2) Induced by near-bank channel flow
E𝐴𝐴
l = b ∗ 𝐴𝐴𝑑𝑑
𝑎𝑎
2.45 0.45 0.4 m/year km 2 Short Hooke (1980)
0.05 0.44 0.67 m/year km 2
(−) Van De Wiel (2003)
0.012 0.4 0.64 m/year km 2 Long De Rose and Basher (2011)
E𝐴𝐴
l = b ∗ 𝐴𝐴𝑐𝑐
𝑎𝑎
6.66E−9 0.86 0.75 m/s m/s Short J. E. Pizzuto and Meckelnburg (1989)
E𝐴𝐴
l = b ∗ 𝐴𝐴𝑓𝑓 + c
𝑎𝑎
6.07 1 4.53 0.94 mm/year (−) Short Lawler (1986)
El = b ∗ Ql a
0.0016 0.6 m/year m /s
3
(−) Rutherfurd (2000)
Ea = b ∗ e (Ql /a) + c 0.6 472.299 0.636 0.98 km 2/year m 3/s Long Yao et al. (2011)
Ev = b ∗ Hub + c a
22.88 1 −3.93 0.73 m /year
3
m Seasonal Z. Zhang et al. (2019)
(3) Induced by overbank flow
E𝐴𝐴
l = b ∗ 𝐴𝐴𝑑𝑑
𝑎𝑎
5.1 0.5 0.62 m/year km 2 Medium Seginer (1966) (Bror-Hayil)
6 0.5 0.84 Ruhama
2.1 0.5 0.85 Tkuma
0.01 0.23 0.39 m/year m2 Short Vandekerckhove, Poesen, et al. (2001)
Ea𝐴𝐴= b ∗ 𝐴𝐴𝑎𝑎𝑑𝑑 0.4 0.59 0.77 m 2/year m2 Long Burkard and Kostaschuk (1997)
Vg𝐴𝐴= b ∗ 𝐴𝐴𝑎𝑎𝑑𝑑 1.71 0.6 0.65 m 3
m2 Medium Vandekerckhove et al. (2000)
Ev𝐴𝐴= b ∗ 𝐴𝐴𝑎𝑎𝑑𝑑 0.02 0.57 0.93 m /year
3
m 2
Medium Vandekerckhove, Muys, et al. (2001)
0.04 0.38 0.39 m 3/year m2 Short Vandekerckhove, Poesen, et al. (2001)
0.069 0.38 0.51 m 3/year m2 Medium Vandekerckhove et al. (2003)
Ev𝐴𝐴= b ∗ 𝐴𝐴𝑠𝑠𝑎𝑎 5.56E−3 2.31 0.67 m /year
3
mm Medium Capra et al. (2009)
El = b ∗ S + c a
4.85 1 30.64 0.8 m/s (−) Medium Samani et al. (2010)
El 𝐴𝐴
= b ∗ (𝐴𝐴𝑑𝑑 ∗ Ps) a 6.466E−9 1.424 0.89 (−) (−) Long Rieke-Zapp and Nichols (2011)
Ea = b ∗ [(Φ60Ai) 0.24S] a 0.154 3.2588 0.62 m 2/year m2 Medium Z. Li et al. (2015)
E𝐴𝐴
l=b∗ 𝐴𝐴𝑠𝑠2 + c ∗ Ps + d 7E−4 2 −0.06 1.11 0.94 m mm Short Dong, Wu, et al. (2019)
(4) Induced by waves
El = Kw(Pw − Pc) a 0.35 1.1 0.8 m/year kW/m Medium Schwimmer (2001)
Ea = Kw(Pw − Pc) a
0.036 1 0.89 m /year
2
kW/m Medium Marani et al. (2011)
0.098 0.75 m 2/year kW/m Seasonal Bendoni et al. (2016)
0.117 0.73 m 2/year kW/m Seasonal
0.413 0.77 m 2/year kW/m Seasonal
0.33 0.54 m /year
2
kW/m Seasonal
a
“Event” timescales denote time intervals of one or several flood events, “seasonal” timescales indicate time intervals of less than one year, “short” timescales indicate
time intervals of 1–5 years, “medium” timescales correspond to time intervals of 5–50 years, and “long” timescales denote a time intervals of more than 50 years.
Variables are summarized in Symbols.

in Tables 1 and 2. Following Lawler (1993a), in this review timescales are defined thus: “Event” timescales
denote time intervals of one or several flood events, “seasonal” timescales indicate time intervals of less than one
year, “short” timescales indicate time intervals of 1–5 years, “medium” timescales correspond to time intervals
of 5–50 years, and “long” timescales denote a time intervals of more than 50 years.

ZHAO ET AL. 18 of 51
Reviews of Geophysics 10.1029/2021RG000761

4.1. Hydraulic-Based Empirical Relations

There are numerous empirical relations that express the bank retreat rate as a function of some hydraulic param-
eters. These relations are developed to account for the average bank retreat (including both erosion and collapse)
over the measured period. The seepage-induced bank retreat rate E, for both laboratory- and field-scale settings, is
commonly estimated in the dimensional form by an excess discharge (E = Ks(Qs − Qc) a) or gradient (E = Ks(i − ic) a)
formulation. Here, Ks is seepage erodibility coefficient, Qs is seepage discharge, Qc is critical seepage discharge
for bank erosion, i is hydraulic gradient driving seepage flow, and ic is critical hydraulic gradient for seepage
erosion. Note that for non-cohesive soils, the excess discharge relation is reduced to a power law correlation
𝐴𝐴 (𝐴𝐴 = 𝑏𝑏𝑏𝑏𝑎𝑎𝑠𝑠, see Table 1) (Akay et al., 2018; Chu-Agor et al., 2009; Fox et al., 2007; Howard & McLane, 1988;
Midgley et al., 2013). Given that the direct measurement of E is sometimes complex and time-consuming, espe-
cially in the field, other variables, such as seepage sediment concentration and seepage-induced cavity volume,
have been used to indirectly evaluate the retreat rate in excess discharge formulas (Masoodi et al., 2017; Wilson
et al., 2007). In particular, Masoodi et al. (2017) suggested a linear correlation between seepage-induced cavity
volume and soil chemical properties. This suggestion has implications for estuarine contexts, where seepage
erosion might be more complicated as a result of the elusive dynamics of salinity (Hua et al., 2019).

Contrary to excess discharge formulas (see Fox et al. (2006) in Table 1), gradient-type relations directly link the
retreat rate to boundary conditions (hydraulic head) and bank geometry without requiring additional input data
(e.g., seepage velocity). This facilitates data analysis and inter-comparison between experiments, and provides
more robust predictions. Since the regression parameters in the above dimensional-form relations are highly
affected by hydraulic settings and soil properties, a more generic dimensionless formula 𝐴𝐴 (𝐴𝐴𝑠𝑠∗𝐴𝐴= K𝐴𝐴
s(𝐴𝐴𝑠𝑠 − 𝐴𝐴𝑐𝑐𝑐𝑐) was
∗ ∗

proposed by Fox et al. (2007), 𝐴𝐴where 𝐴𝐴


𝐴𝐴𝑠𝑠∗ and 𝐴𝐴𝑠𝑠∗ are the dimensionless sediment flux and shear stress, respectively,
induced by seepage flow (see Section 2.1 for more details). The proposed dimensionless formula is not restricted
to specific sites, and instead can be used for a range of contexts.

For retreat driven by surface flow, power law correlations are usually assumed. The associated controlling factors
are distinct for near-bank channel flow and overbank flow conditions (Table 2). For near-bank channel flow,
previous studies mainly focused on the effects of peak and annual mean flow discharges (Rutherfurd, 2000;
Yao et al., 2011), near-bank flow velocity (J. E. Pizzuto & Meckelnburg, 1989), frost (Lawler, 1986), bend
curvature (Lagasse et al., 2004; Nanson & Hickin, 1983), and freeze-thaw cycles (J. Pizzuto, 2009). For over-
bank flow more attention has been paid to the consequences of precipitation (Capra et al., 2009; Dong, Wu,
et al., 2019; Rieke-Zapp & Nichols, 2011), channel slope gradient (Samani et al., 2010), and vegetation cover
(Z. Li et al., 2015). A more general controlling factor, the drainage area considered as a surrogate of the flow
discharge, is also used for both near-bank channel flow and overbank flow (Figure 10). Contrary to overbank flow,
bank retreat induced by near-bank channel flow is more sensitive to temporal scale, as indicated by an evident gap
in the observed retreat rate between short- and long-term time scales (e.g., river linear retreat rate, indicated by
red dots and yellow squares in Figure 10). Therefore, the drainage area can be taken as a rough index for hydro-
dynamic intensity discriminating between different spatial and temporal scales (Burkard & Kostaschuk, 1997;
Hooke, 1980; Seginer, 1966; Vandekerckhove, Poesen, et al., 2001; Vandekerckhove et al., 2000, 2003). More
recently, machine learning algorithms based on large datasets, have been applied to develop predictors of gully
erosion induced by overbank flow (Amiri et al., 2019; Arabameri et al., 2019; Rahmati et al., 2017). Since the
factors dominating overbank flow are relatively easily collected (e.g., precipitation, drainage area, and soil char-
acteristics), machine learning is likely to improve our predictive skill of gully erosion rates.

Employing laboratory-scale data, Wells et al. (2013) proposed an exponential function to describe changes in
channel width over time, with the exponent determined by channel slope and flow discharge (Qin et al., 2018).
Clearly, this relation is strictly related to bank retreat and, therefore, can be used to derive a bank retreat predictor.
Note that, since the cross-section widening rate was found to decrease gradually with time (due to the increasing
channel width), the associated bank retreat relation is deemed to provide more realistic predictions than when
assuming a constant bank retreat rate over time.

As for tidal systems, the retreat rate of salt-marsh cliffs has usually been evaluated using a power law of the
type E = Kw(Pw − Pc) a, where Pw represents the mean wave power calculated over a representative period, Pc is
a threshold value for wave-induced retreat, Kw is the erodibility coefficient for wave erosion, and a is an empir-
ical exponent (here a nonlinear correlation is considered to obtain a more general form) (Bendoni et al., 2016;

ZHAO ET AL. 19 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 10. Correlation between drainage area and the linear and volumetric retreat rates, showing the variability of different
predictors for bank retreat rate. Dots represent observations collected over short time scales (1–5 years). Open and filled
symbols refer to data collected over medium (5–50 years) or long (>50 years) time scales, respectively. Colors are used to
distinguish different references.

Marani et al., 2011; Mariotti & Fagherazzi, 2010; Schwimmer, 2001). When accounting for their results in the
context of previous studies, Bendoni et al. (2016) found that at short temporal scales (months), the marsh cliff
retreat rate is much higher than the one obtained over much larger time intervals (decades), a result consistent
with the tendency observed in fluvial systems (Hooke, 1980). This finding can be partly explained by a dynamic
equilibrium theory (Zhou et al., 2017) whereby bank dynamic is not only associated with a monotonic retreat,
but is possibly characterized by a periodic cycle of erosion and accretion (e.g., the catch-up behavior reported in
meander migration (Mason & Mohrig, 2019; Nanson & Hickin, 1983; Zhao et al., 2021)). In other words, when
considering the bank retreat rate over relatively long-time periods, other dynamics such as subaerial processes
prior to bank erosion/collapse (e.g., freeze-thaw and drying-wetting cycles) (Chassiot et al., 2020), decomposi-
tion of collapsed bank soil (C. Hackney et al., 2015; Fagherazzi et al., 2004; K. Zhang et al., 2021), and bank
accretion (Asahi et al., 2013) can all lead to a lower than expected bank retreat rate. The issue associated with the
choice of a suitable temporal scale thus turns out to be crucial, since it determines the robustness and accuracy
of the developed empirical relations given the period of time to be considered for predicting the bank dynamic
(Hooke, 1980; J. Pizzuto et al., 2010).

In general, the performance of hydraulic-based empirical relations depends on the choice of the representative
discharge (e.g., mean or peak discharge). In the case of banks composed of sand and silt, which constantly
undergo erosion, the annual mean discharge is the best choice. It in fact summarizes the overall information
concerning the relevant hydrologic processes. In contrast, the peak discharge is more suitable for estimating the
retreat rate of cohesive and bedrock banks, for which bank erosion is only active during high flows. In these cases,
using the mean discharge likely leads to an overestimation of the overall bank retreat. Anyhow, when using a
constant formative discharge, an intermittency factor is needed to account for the effect of temporal variations in
hydrological forcings. This factor is defined as the fraction of time the channel is actually experiencing effective
erosive conditions (Paola et al., 1992; Parker et al., 1998). However, very few of the empirical relations discussed
above have been developed accounting explicitly for the intermittency of the formative discharge. This limitation
is possibly one of the factors leading to the large scatter characterizing the various empirical coefficients reported
in Table 2. More efforts are needed to provide appropriate criteria for the application of the hydraulic-based
empirical relations developed so far for estimating bank retreat.

4.2. Empirical Relations Accounting Directly for Bank Collapse

Hydraulic-based empirical relations commonly fail when including mass failure events at short temporal scales
(Bendoni et al., 2016). Indeed, contrary to flow-induced bank erosion, which is mainly related to flow velocity

ZHAO ET AL. 20 of 51
Reviews of Geophysics 10.1029/2021RG000761

and soil properties (e.g., critical shear stress for bank erosion), the scale and frequency of bank collapse depends
on different factors such as bank geometry, soil properties, near-bank hydrodynamics and biological disturbances
(C. Chen, Hsieh, & Yang, 2017; Fox et al., 2006; Nardi et al., 2012; Samadi et al., 2013; Sanders et al., 2021).
Numerous empirical relations have thus been developed to predict bank retreat rate accounting explicitly for bank
collapse.

The effect of bank geometry on bank collapse and the consequent bank retreat rate is commonly estimated in
the form of an excess bank height or bank slope (Jang & Shimizu, 2005; Mosselman, 1995). For field-scale
applications (e.g., the Yellow River investigated by Z. Zhang et al. (2019) and Liu et al. (2021)), the relation
is reduced to a simple linear correlation, without any threshold. In fact, bank collapse occurs when the driving
force overcomes the resisting forces regardless of bank height. Bank height seems only to affect the scale and
frequency of bank collapse (see Figure 4 in Section 3.2). Soil properties and near-bank hydrodynamics also affect
bank stability, and therefore alter bank retreat rates. Although exhibiting a relative weak correlation, Xia, Zong,
Zhang, et al. (2014) suggested that a decrease in clay content or an increase in drawdown rate generally favors
bank collapse and, hence, accelerates bank retreat rate. The effect of the former factor can be partly explained
by an increased thickness of the cohesive layer in the upper bank implying the formation of heavier overhanging
soil blocks, while the effect of the latter is attributed to the fact that a rapid drawdown of water level generates
excess pore-water pressure combined with a loss of hydrostatic pressure, both of which favor bank instability. As
for biological disturbances, Sanders et al. (2021) found that burrow metrics (e.g., burrow density) have a strong
positive linear correlation with bank retreat rate and the area of collapsed bank. Given that salt-marsh channels
are generally covered by halophytic vegetation with a strong root matrix, biological disturbances turn out to be
crucial to bank stability and, hence, bank collapse (Harvey et al. (2019) and see Section 3.5). Other factors, such
as the critical length of overhanging soil blocks and time to collapse, have also been used to indirectly evaluate
collapse-induced retreat rates. For event-scale cohesive overhang failures, J. Pizzuto (2009) proposed a critical
value of the overhang length depending linearly from overhang height. In the case of seepage-induced bank
collapse, several studies suggested a correlation between time to collapse and the seepage gradient (Karmaker
& Dutta, 2013; Masoodi et al., 2018). Although providing some useful information, the above correlations only
account for the individual effects of hydraulic and geotechnical conditions, without an integrated representation
of hydraulic and geotechnical control factors.

The seminal in situ work by Hickin and Nanson (1984) suggests that, in fluvial contexts, the linear migration
rate of channel banks, El, can be evaluated by a combination of stream power, ω, bank height, Hb, channel width,
wc, and bank soil resistance, γb, proposing the linear relation El = aω/γbwc/Hb. Subsequently, several studies
have used the ratio between bank height and near-bank water depth, Hb/Hw, as a proxy for bank stability when
seeking empirical laws for bank/cliff retreat rate. Dapporto et al. (2003) suggested that the critical value of
Hb/Hw triggering bank collapse can be estimated based on the peak river stage. This value is taken to surro-
gate the complex coupling between stress-strain and seepage processes. For tidal settings, Marani et al. (2011)
attempted to build a correlation between volumetric retreat rate of salt marsh borders, Ev, mean wave power,
Pw, and the ratio Hb/Hw, proposing a linear relation of the type Ev = Pw ⋅ Hb/Hw. However, the heterogeneity of
marshes (Houttuijn Bloemendaal et al., 2021) and the relatively large time intervals over which the data were
averaged (decades), smooth out the effects that geotechnical factors exert on cliff retreat rates. As a result, Marani
et al. (2011) do not find a clear correlation between Ev/Pw and the ratio Hb/Hw. To separate the effects of bank
collapse on bank retreat rate, Zhao et al. (2020) defined a dimensionless linear retreat rate, rl, quantifying the
erosion controlled by near-bank channel flow. It reads
𝐸𝐸 𝑤𝑤
𝑟𝑟𝑙𝑙 = 𝑙𝑙 ⋅ 𝑡𝑡
(6)
𝑈𝑈𝑐𝑐 𝑤𝑤𝑐𝑐

where El is the linear retreat rate (m/s), Uc is the near-bank flow velocity (m/s), wc is the channel width (m), and wt
is the width of the overhanging bank material (m). The ratio wt/wc accounts for the protective effect of collapsed
bank soil (slump blocks) on the bank retreat rate. Figure 11a shows the normalized retreat rate defined by Equa-
tion 6 as a function of the ratio Hb/Hw and the best fit line obtained from the laboratory data available in literature
(Braudrick et al., 2009; Patsinghasanee et al., 2017; Qin et al., 2018; Shu et al., 2019; van Dijk et al., 2012;
Vargas Luna et al., 2019; Wells et al., 2013; Zhao et al., 2020) (further details on the variables used in Figure 11

ZHAO ET AL. 21 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 11. (a) Correlation between the normalized linear bank retreat rate, rl, and the ratio Hb /Hw. The various quantities have been determined from laboratory
experiments available in literature and are defined as follows: Uc, near-bank flow velocity; wc, channel width; wt, the width of the overhanging bank material; Hb, bank
height; and Hw, near-bank water depth. (b and c) Comparison of linear bank retreat rates predicted by empirical relations (Equations 7a and 7b) (blue triangles) and
those formulated by Z. Zhang et al. (2019) (red circles), Rutherfurd (2000) and Yao et al. (2011) (red squares), J. E. Pizzuto and Meckelnburg (1989) (red crosses), and
the excess shear stress Equation 1 (red plus). Previous empirical relations are based on (a) geotechnical factors and (b) hydraulic factors. Subplot (a) is adapted from
Zhao et al. (2020).

are provided in Data Availability Statement). The normalized retreat rate rl exhibits two distinct behaviors for
small (<7.5) and large (>10) values of the ratio Hb/Hw, namely:

(7a)
( 2 )
𝑟𝑟𝑙𝑙 = 0.00003 (𝐻𝐻𝑏𝑏 ∕𝐻𝐻𝑤𝑤 − 1) 1 < 𝐻𝐻𝑏𝑏 ∕𝐻𝐻𝑤𝑤 < 7.5 𝑅𝑅 = 0.83

(7b)
𝑟𝑟𝑙𝑙 = 0.0097(𝐻𝐻𝑏𝑏 ∕𝐻𝐻𝑤𝑤 )−2.05 𝐻𝐻𝑏𝑏 ∕𝐻𝐻𝑤𝑤 > 10
( 2 )
𝑅𝑅 = 0.94

ZHAO ET AL. 22 of 51
Reviews of Geophysics 10.1029/2021RG000761

In the former case (Equation 7a) rl increases linearly with Hb/Hw. Conversely, in the latter case rl decreases
following a power law. These two trends can be explained by noting that the ratio Hb/Hw can be taken as a meas-
ure of the degree of bank stability (Zhao et al., 2020). In general, an increase in Hb/Hw leads to a decreased bank
stability enhancing bank collapse and, hence, bank retreat rate. However, for large values of Hb/Hw, the near-bank
flow velocity and water depth are small. The bank erosion processes thus weaken leading to a reduction of the
frequency of bank collapse and, hence, to the progressively decreasing bank retreat rate shown in Figure 11a.

Figures 11b and 11c report the comparison between the empirical relations (Equations 7a and 7b) and a number
of predictors previously proposed in literature accounting separately for hydraulic and geotechnical control
factors in terms of bank height (Z. Zhang et al., 2019), discharge (Rutherfurd, 2000; Yao et al., 2011), flow
velocity (J. E. Pizzuto & Meckelnburg, 1989) and bed shear stress (see Equation 1 and Table 2 for the consid-
ered formulas). Various observations emerge from this comparison. First, it is evident that both hydraulic (flow
velocity and discharge) and geotechnical factors (bank height) must be accounted for to obtain robust predictions
of bank retreat, at least at laboratory-scale. In particular, geotechnical factors (controlling bank collapse) exert
a relative stronger influence on bank retreat rate at the laboratory scale, when compared to hydraulic factors.
Second, a discrepancy between the importance of geotechnical factor in the field and laboratory contexts is
apparent when noting that an increased bank height decelerates bank retreat rate in the field while accelerates it
in the laboratory (Hickin & Nanson, 1984; Zhao et al., 2020). This implies that the inclusion of bank collapse
may be far more complex than previously thought, since collapsed bank soil affects the interplay between hydrau-
lic and geotechnical factors, as conjectured by the concept of basal endpoint control (Carson & Kirkby, 1972;
Thorne & Tovey, 1981) and discussed by C. Hackney et al. (2015). Third, even though the variables required for
empirical relations (e.g., those in Table 2) are commonly available, we highlight that empirical approaches have
been usually pursued without a clear physical basis (except for Marani et al. (2011) who adopted dimensional
analysis to derive their relations). Most of these empirical relations were instead obtained by direct fitting of
a relatively limited amount of data and they also lack systematic validation. Hence, the coefficients appearing
in the various relations are strictly valid only for those specific sites where the measurements were originally
collected. For instance, the bank erodibility coefficients in Equation 1 can vary by several orders of magnitude
(Gong et al., 2018; Parker et al., 2011), and should be regarded as calibration parameters (Crosato, 2007; Rinaldi
& Darby, 2007). Overall, the empirical relations are useful for contexts where monitoring of hydrodynamics is
impossible (e.g., ephemeral gully) or situations where the key mechanisms are still elusive. Empirical relations
are also meaningful when modeling long-term morphodynamic evolution. They in fact provide an acceptable
time averaged description of the effects of some complicated processes such as secondary flow and sequences of
bank collapse events (see Section 5.2). Although some parameterized relations have been proposed to account for
subaerial processes (J. Pizzuto, 2009; Wynn et al., 2008), additional advances are still needed to obtain a more
realistic representation of the role which these important processes play on bank erosion and collapse. Finally, we
notice that very few empirical bank retreat relations have been developed to include multiple erosion mechanisms
in estuaries and tidal channels (e.g., coupled effects of seepage and near-bank channel flow during ebb tides).
Efforts are strongly needed to improve our understanding and, hence, the predictive capability of the complex
feedbacks acting in estuarine and coastal systems.

5. A Hierarchy of Models for Bank Retreat


Studies carried out to model bank retreat have followed two distinct paths: a hydraulic approach and a geotechni-
cal approach (Rinaldi & Darby, 2007). The hydraulic approach, based on strong simplifications of bank collapse,
relies on some empirical parameterization for bank retreat processes, usually by means of surrogates of the
shear stress that the near-bank flow exerts on the bank. This approach has been used to describe the evolution
of rivers, estuaries and tidal channels over a range of time scales (Bogoni et al., 2017; Duan & Julien, 2005;
Ikeda et al., 1981; Jang & Shimizu, 2005; J. E. Pizzuto, 1990; Jia et al., 2010; Lanzoni & Seminara, 2006;
Lopez Dubon & Lanzoni, 2019; Nagata et al., 2000; Parker et al., 2011; van der Wegen et al., 2008; van Dijk
et al., 2019). In contrast, the geotechnical approach focuses on the transient process of bank collapse that controls
the intermittent evolution of channel cross-sections or salt marsh borders (Bendoni et al., 2014; Gong et al., 2018;
Istanbulluoglu et al., 2005; Kleinhans et al., 2009; Langendoen & Simon, 2008; Osman & Thorne, 1988; Samadi
et al., 2013; Thorne & Tovey, 1981; Van Eerdt, 1985). The contribution to bank collapse has been investigated
numerically by focusing on different factors such as bank height (Zhao et al., 2019), soil properties (Simon

ZHAO ET AL. 23 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 12. Diagram showing the hierarchy of models used to simulate bank retreat. The color bar represents the degree of simplification used to treat bank collapse.
Note that the literature listed in this figure aims to provide some typical examples and is not meant to be comprehensive.

et al., 2000), pore-water pressure (Darby et al., 2007; Darby & Thorne, 1996a), and vegetation roots (Krzeminska
et al., 2019; Pollen Bankhead & Simon, 2009; T. H. Wu et al., 1979). Although the classification into a hydraulic
approach and a geotechnical approach proposed by Rinaldi and Darby (2007) is helpful to distinguish between
models, it fails to indicate whether a model describes the interplay between flow-driven bank erosion and bank
collapse or treats only one of the two mechanisms. We thus propose to categorize the existing models of bank
retreat into a hierarchy of models: purely hydraulic models, not accounting explicitly for bank collapse; models
based on empirical relations that parameterize bank collapse; models considering the static equilibrium of bank
soil through the limit equilibrium method (LEM); and models based on bank soil stress-strain deformations. A
summary of these four typical modeling approaches is presented in Figure 12 and Table 3.

5.1. Hydraulic Models

This type of models describes bank retreat from a hydraulic perspective, without any representation of bank
collapse. Hence, hydraulic models should not be applied to contexts where bank collapse controls bank retreat,
such as in the case of steep and cohesive riverbanks or when assessing the effects of the recession of flood hydro-
graphs and the short-term evolution of tidal creeks on muddy flats.

Many hydraulic models have been developed to simulate bank retreat associated with the long-term evolution of
meandering channels using surrogates (i.e., excess velocity, water depth) of the shear stress exerted by the chan-
nel flow on the bank (Ikeda et al., 1981; Kitanidis & Kennedy, 1984; Langendoen et al., 2016; Odgaard, 1989).
The most common model was introduced by Hasegawa (1977) and Ikeda et al. (1981). In this approach (hereafter
denoted as HIPS) the bank retreat rate is taken to be proportional to the curvature-induced difference between
the near-bank and the cross-sectionally averaged velocity. Since the thickness of the near-bank boundary layer is
commonly uncertain (Parker et al., 2011), the definition of the near-bank velocity in the HIPS equation is some-
times arbitrary. The HIPS equation provides a simple and direct way to simulate channel migration, and it has

ZHAO ET AL. 24 of 51
Reviews of Geophysics 10.1029/2021RG000761

Table 3
Summary of Typical Modeling Approaches Used for Computing Bank Retreat
Mechanism of bank retreat

Model Flow-driven bank erosion Bank collapse Highlights References


Hydraulic model/Author
HIPS Excess velocity equation No Easy to couple with morphodynamic models Hasegawa (1977); Ikeda
of long-term evolution of meandering et al. (1981)
channels
Delft3D Near-bank sediment flux No Lateral bank retreat is replaced with a van der Wegen
vertical decrease in bank height, which et al. (2008)
is easy to couple with other processes,
for example, tides, waves, and sediment
dynamics
Duró et al. Wave-induced shear stress No Account for the shear stress distribution Duró et al. (2020)
induced by ship waves
Parameterized bank collapse model/Author
Hasegawa and Mosselman No Critical bank height Bank collapse promotes bank retreat by Hasegawa (1989);
collapse events, or alternatively, prevents Mosselman (1998)
bank retreat by collapsed bank soil
Duan and Julien Excess shear stress equation Parameterized A vertical bank profile is maintained since Duan and Julien (2005)
the upper bank retreat rate due to
collapse is assumed to keep up with the
basal erosion rate
Nays2D Integration of the sediment Critical repose angle Outer bank erosion and inner bank Jang and Shimizu (2005);
continuity equation deposition processes are separated and Parker et al. (2011)
the effects of slump blocks are taken into
account
Bank collapse occurs when the local bank Nagata et al. (2000)
slope exceeds the angle of repose for
bank material
Mariotti et al. No Critical bank slope A linear relation linking soil creep with bed Mariotti et al. (2016)
slope and soil diffusivity rate
van Dijk et al. No Local slope Bank collapse is parameterized on the basis van Dijk et al. (2019)
of field data, relating collapse frequency
to the local slope angle and bed elevation
Limit equilibrium method/Author (slide failure)
BSTEM Excess shear stress equation Planar failure Matric suction, positive pore-water pressure, Simon et al. (2000);
hydrostatic pressure, and vegetation roots Simon and
are coupled for layered cohesive banks Collison (2002)
New algorithms are derived to account for Langendoen and
the effect of tension cracks on planar Simon (2008)
failure
The near-bank groundwater table is assumed Midgley et al. (2012)
to change instantly or gradually in
response to water level change in the
channel
Coupled with a process-based Motta et al. (2014); Lai
morphodynamic model accounting et al. (2015)
for meander migration and planform
evolution
Rinaldi and Darby Excess shear stress equation Sliding failure Fluvial erosion, finite element seepage Rinaldi et al. (2004);
analysis and bank stability analysis are Darby et al. (2007)
fully coupled
Near-bank bed deformation is coupled Deng et al. (2018)

ZHAO ET AL. 25 of 51
Reviews of Geophysics 10.1029/2021RG000761

Table 3
Continued
Mechanism of bank retreat

Model Flow-driven bank erosion Bank collapse Highlights References


Chu-Agor et al. No Pop-out failure New algorithms are derived for pop-out Chu-Agor et al. (2008)
failure along a failure plane parallel and
perpendicular respectively to the bank
face
Thorne and Tovey No Cantilever failure First method specific for cantilever failure Thorne and Tovey (1981)
Van Eerdt No Toppling failure A triangular distribution of both tensile and Van Eerdt (1985)
compressive stresses is derived along the
failure plane
Xia et al. No Toppling failure The effect of tension cracks on the bank top Xia, Zong, Deng,
is included and a constant ratio between et al. (2014)
tensile strength to compressive stress is
applied
Bendoni et al. No Toppling failure A dynamic wave-induced load is accounted Bendoni et al. (2014)
for
Patsinghasanee et al. Excess shear stress equation Cantilever failure Cantilever stability analysis is coupled with Patsinghasanee
fluvial erosion and bedload sediment et al. (2017)
transport
Stress-strain analysis/Author
Samadi et al. No Stress-strain analysis An elastic-plastic stress-strain model is Samadi et al. (2013)
applied to investigate toppling failure
due to undermining
Gong and Zhao Excess shear stress equation Stress-strain analysis Stress-strain analysis is coupled with lateral Gong et al. (2018); Zhao
flow erosion, sediment dynamics, and et al. (2019, 2021)
river meandering

been used in numerous studies (see e.g., among many others, Camporeale et al., 2007; Frascati & Lanzoni, 2009;
Lanzoni & Seminara, 2006; Parker et al., 2011).

Another common approach pursued for alluvial rivers and estuarine environments is the so called dry-cell erosion
(DCE) method, incorporated in models solving the full set of shallow water equations (e.g., Delft3D, Lesser
et al. (2004)). It allows the redistribution of an erosion flux from a wet cell to the adjacent dry cells. As a result,
the lateral bank retreat is replaced with a vertical decrease in bank height (van der Wegen et al., 2008; Zhao
et al., 2019). Because of its convenience, especially for complicated bank alignments, the DCE has been exten-
sively applied to the simulation of large-scale contexts such as estuaries and tidal networks (Guo et al., 2021; Xu
et al., 2017).

For bedrock rivers evolving over millennial timescales, the inclusion of a lateral erosion law into long-term land-
scape evolution models (e.g., cellular models) is a challenge (Lague, 2014). Several models have been proposed at
either river-scale (Hancock & Anderson, 2002; Inoue et al., 2021; Malatesta et al., 2017; Murray & Paola, 1994),
or catchment-scale (Coulthard et al., 2013). On the basis of empirical data, Howard and Knutson (1984) related
bank retreat rate to the local and upstream-integrated curvature. This relation was subsequently adopted in cellu-
lar models to drive lateral erosion and consequent meandering (Coulthard & Wiel, 2006), and to investigate
how meander migration is affected by bedrock lithology (Limaye & Lamb, 2014). Nevertheless, a physics-based
parametrization of bank retreat accounting for the combined action of vertical and lateral incisions in mixed
bedrock-alluvial meandering channels is still missing. On the other hand, studies have also been carried out to
deal with the broader catchment topography where bedrock rivers are embedded (Carretier et al., 2016; Langston
& Tucker, 2018). In these studies, bank retreat is simply related to local erosion flux, or elevation difference
between nodes representing bank and channel, respectively. This highly simplified approach is a first step toward
a fully understanding of bedrock valleys evolution over long time periods and large space scales.

ZHAO ET AL. 26 of 51
Reviews of Geophysics 10.1029/2021RG000761

5.2. Parameterized Bank Collapse Models

This type of models represents the processes of bank collapse using bank geometry (e.g., bank height and
slope) or other empirical parameterizations (see Section 4). On the basis of an excess bank height formula,
Mosselman (1998) suggested that bank retreat rate is proportional to bank height on the short- or medium-term
timescales typical of meandering river dynamics. Alternatively, Hasegawa (1989) assumed that bank retreat is
inversely proportional to bank height. This assumption is based on the concept of basal endpoint control, set out
by Carson and Kirkby (1972) and Thorne (1982), whereby bank retreat rate is determined by the balance between
the supply of slump blocks by bank collapse and their removal by fluvial (grain-by-grain) erosion.

Using bank slope, two alternative approaches have been proposed, depending on the potential failure mechanism.
For planar bank failure and non-cohesive bank material, J. E. Pizzuto (1990) and Nagata et al. (2000) derived
a model in which failure of the upper bank occurs as the bank slope exceeds the angle of repose, as a result
of lower basal erosion (Figure 13a). This approach was improved by Parker et al. (2011), who introduced the
sediment continuity equation for the near-bank region (Figure 13b). It was applied by Jang and Shimizu (2005)
and Dulal et al. (2010) to simulate flume experiments of meander dynamics through the Nays2D software. This
model solves numerically the two-dimensional shallow water equations coupled with the Exner sediment balance
equation and relates bank retreat rate directly to the near-bank sediment flux (Shimizu et al., 2019). The charac-
teristic timescale of bank processes is thus of the same order of magnitude as for bed processes. This modeling
approach performs well in terms of meander planform and width variations, when compared to channel evolution
in both experimental and natural contexts (Asahi et al., 2013; E. C. Eke et al., 2014). In the case of vertical banks
composed of cohesive sediment, Duan and Julien (2005) assumed that the upper bank retreat due to bank collapse
keeps up with basal erosion, maintaining a vertical bank profile during the retreat. However, this assumption is
arbitrary, since recent laboratory experiments have shown that the location of tension cracks on the bank top is
always beyond the endpoint of the cantilever (i.e., closer to the landward boundary). To obtain a more realistic
representation of bank collapse in meandering rivers, Zhao et al. (2021) suggested that the flow-induced bank
erosion rate can be amplified by the contribution of bank collapse to bank retreat, Cbc. This contribution, evalu-
ated on the basis of experimental observations, was found to depend on the ratio of bank height to near-bank water
depth, Hb/Hw. The contribution of bank collapse to bank retreat is thus taken as continuous, and its average effect
is linked to flow-driven bank erosion.

To evaluate the performance of existing approaches to bank retreat, Stecca et al. (2017) set up a framework in
which three modeling steps are introduced: bank identification, bank retreat simulation, and bank updating. A set
of bank retreat models were then constructed by choosing different options for each step. This approach quanti-
fied how the cross-sectional evolution of rivers can be affected by each of the steps, and therefore can guide the
practical application of existing bank retreat models. A similar strategy was adopted by J. Pizzuto (2009) who
simulated the evolution of bank profiles by assembling several regression equations including lower hydraulic
erosion, upper subaerial erosion, and the volume of the failed overhanging block. This research can be regarded as
a first attempt to develop a detailed numerical model using empirical methods. More efforts are needed to extend
this framework to the entire river reach and to decadal timescales.

More recently, several empirical parameterizations, validated by field data, have been introduced to describe
bank retreat in tidal environments. For instance, a shoal margin collapse predictor was introduced by van Dijk
et al. (2019), relating collapse frequency to the local maximum bank slope angle. Several authors have assumed
a linear relation linking soil creep with slope and soil diffusivity of salt marsh borders (Kirwan & Murray, 2007;
Larsen et al., 2007; Mariotti et al., 2016). This approach has been used to study sediment exchanges between tidal
flats and creeks. Nevertheless, soil creep is a complicated process related to temperature, soil water content, and
vegetation roots, and more research is needed to further validate the simple linear relationship so far employed.

Overall, parameterized bank collapse models can be used to account for collapse-induced retreat rate in long-term
and large-scale simulations. However, because of the strong simplifications employed in their derivation, they
should be used in contexts where bank stability is mainly related to bank geometry (i.e., homogeneous bank soils
without other external forces), or where detailed information of bank materials and external forces are unavaila-
ble. Moreover, since models of this type fail to account for the intermittency of bank collapse, they should only
be applied to short-term contexts with extreme caution.

ZHAO ET AL. 27 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 13. Schematization of the various bank erosion/collapse models proposed in literature: (a) bank collapse of non-cohesive sediment (Nagata et al., 2000); (b)
fluvial erosion based on sediment conservation (Parker et al., 2011); (c and d) shear and tensile failure (Thorne & Tovey, 1981); (e) toppling failure (Van Eerdt, 1985);
(f) toppling failure under the action of waves (Bendoni et al., 2014); (g) Bank Stability and Toe Erosion Model (Simon et al., 2000); (h) pop-out failure (Chu-Agor
et al., 2008); and (i) stress-strain analysis of a tidal channel bank (adapted from Gong et al. (2018)). We refer the reader to the list of Symbols for the description of the
parameters appearing in the various sketches.

5.3. Limit Equilibrium Methods

From a geotechnical perspective, bank collapse occurs when the driving forces (or moments), FD, acting on a
potential failure surface overcome the resisting forces (or moments), FR. As a result, a factor of safety, Fs = FR/
FD, is commonly applied to estimate bank stability. To calculate Fs, a potential failure surface is assumed and one
or more equations of static equilibrium (equilibrium of forces or moments) are used to calculate the forces along
the incipient failure surface (Duncan et al., 2014). This procedure is termed the LEM. It accounts not only for
bank geometry, but also for soil properties and some relevant external forces, such as hydrostatic pressure head,

ZHAO ET AL. 28 of 51
Reviews of Geophysics 10.1029/2021RG000761

soil pore-water pressure, and reinforcement provided by vegetation roots. Although a number of failure mecha-
nisms have been identified by Thorne (1982), applications of the LEM have concentrated mostly on cantilever
and sliding failures. Given the large body of literature concerning LEM (see e.g., the reviews by Rinaldi and
Darby (2007) and Klavon et al. (2017)), below we only summarize some recent advances.

For cantilever failures (Figures 13c–13e), recent research has focused on the influence of heterogeneous bank
material and tension cracks, as well as on developing a more realistic distribution of stress along the failure
surface. On one hand, the potential cantilever block has been divided into horizontal slices, in order to account
for the effect of heterogeneous bank material and upper cracks (Langendoen & Simon, 2008). On the other hand,
the uniform stress assumption put forward by Thorne and Tovey (1981) has been gradually replaced with a trian-
gular distribution of both tensile and compressive stresses along the failure plane (Figure 13e). This approach,
first proposed by Van Eerdt (1985) for tidal environments, was later on widely employed to investigate the rela-
tion between the cantilever lengths under tensile (lt) and compressive stress (lc) (Arai et al., 2018; Micheli &
Kirchner, 2002; Patsinghasanee et al., 2017; Xia, Zong, Deng, et al., 2014). With reference to salt marsh borders,
Bendoni et al. (2014) developed a theoretical model to evaluate the effects of dynamic loads on cantilever stability
(Figure 13f). Bank failure was assumed to occur when the tensile strength of the bank material is exceeded at least
on one point of the failure surface, rather than using an average value for the whole failure surface.

For sliding failures, one of the most advanced and commonly used tools is the Bank Stability and Toe Erosion
Model (BSTEM; Figure 13g) first proposed by Simon et al. (2000). This modeling approach accounts for the
effects of positive pore water pressure and hydrostatic pressure (Darby & Thorne, 1996a), matric suction in
unsaturated soils (Casagli et al., 1999), layered bank materials (Simon et al., 2000), vegetation roots (Simon
& Collison, 2002), and upper tension cracks (Langendoen & Simon, 2008). In particular, the procedure intro-
duced by Langendoen and Simon (2008) discretizes the bank soil above the potential failure surface into several
vertical slices to deal with complicated bank geometry, upper tension cracks and external actions such as those
due to vegetation roots. BSTEM is, however, limited to gentle failure plane angles as a result of computational
issues (Lai et al., 2015). BSTEM has also been used in tidal environments to investigate the stability of tidal
channel headcuts (Kleinhans et al., 2009). Despite of its comprehensiveness, the main limitation of the BSTEM
approach is that the near-bank groundwater table is assumed to be horizontal and constant (except for Midgley
et al. (2012)). Hence, soil pore-water pressure variations and seepage forces are completely neglected and the
model fails to predict failures due to seepage erosion unless the near-bank groundwater flow is separately consid-
ered (Lindow et al., 2009; Wilson et al., 2007).

The transient character of pore-water pressure has been accounted for through a finite element seepage analysis
of the saturated and unsaturated flow that establishes inside a bank during a single flood (Rinaldi et al., 2004).
The distribution of pore-water pressure can then be used as input data for bank stability analysis by LEM. This
approach has been further extended by Darby et al. (2007) and Deng et al. (2018) to account for fluvial erosion
and near-bank bed deformation. With respect to seepage force effects, Chu-Agor et al. (2008) computed the safety
factor Fs of cohesive slopes subject to pop-out failures (Figure 13h). They derived new equations for Fs along
failure planes parallel and perpendicular to the bank face. We refer the reader to Fox and Wilson (2010) and
Rinaldi and Nardi (2013) for more details on modeling the interactions between seepage flow and bank collapse.

The LEM evaluates bank stability considering the effects of both soil properties and external forces and, most
importantly, provides a simple and direct stability index (Fs). This method is appropriate for engineering projects
of bank protection structures such as revetments, breakwaters and seawalls. It is anyhow of great importance to
integrate LEMs and hydraulic models to better describe bank retreat processes in the presence of floodplain heter-
ogeneity (e.g., composite riverbanks) or external forces such as pore-water pressure, seepage forces and waves.
We therefore suggest using LEMs in conjunction with hydraulic models for short- or medium-term predictions
when the effects of floodplain heterogeneity (Bogoni et al., 2017) and external forces are non-negligible.

Finally, it is worthwhile to note that, although the LEM provides a good agreement with field observations
in terms of bank retreat rate and bank line evolution (Daly et al., 2015; Lai et al., 2015; Midgley et al., 2012;
Patsinghasanee et al., 2017), it also has some limitations (Duncan et al., 2014; Gong et al., 2018; Rinaldi &
Darby, 2007). The fundamental one is that bank materials delimited by the potential failure surface are assumed
not to be subject to deformation. Besides, additional assumptions are commonly required for toppling failure. For

ZHAO ET AL. 29 of 51
Reviews of Geophysics 10.1029/2021RG000761

instance, a relation between the compressive (lc) and tensile (lt) lengths along the failure plane is required (Arai
et al., 2018; Patsinghasanee et al., 2017; Xia, Zong, Deng, et al., 2014).

5.4. Stress-Strain Analysis

Stress-strain analysis has been developed to account for any deformation of bank material during bank collapse
and has been applied to both fluvial and tidal contexts (Gong et al., 2018; K. Zhang et al., 2021; Masoodi
et al., 2019; Samadi et al., 2013; Zhao et al., 2019, 2021). In the case of fluvial environments Samadi et al. (2013)
employed an elastic-plastic constitutive model (using the SIGMA/W software, see https://www.geoslope.com)
to investigate the development of stress and strain under the action of undermining. They found that the location
of the maximum tensile stress on the bank top, where tension cracks occur, was always beyond the endpoint
of the cantilever (i.e., closer to the landward boundary, see Figure 4a). This finding challenges the commonly
adopted assumption that toppling failure occurs along the endpoint of the cantilever (Figure 13e) (Micheli &
Kirchner, 2002; Van Eerdt, 1985). It also agrees with laboratory results indicating that failures are in general char-
acterized by a greater retreat distance in the upper portion of the bank as compared to the lower undermining part
(K. Zhang et al., 2021; Samadi et al., 2011). For tidal environments, Gong et al. (2018) proposed a process-based
bank retreat model coupling the stress-deformation analysis with tide-induced bank erosion (Figure 13i). The
Mohr-Coulomb failure criterion was applied to evaluate the state of soil elements after tide-driven bank erosion
and, hence, to account for possible bank collapse. According to the state of soil elements (stable or subject to
tensile and shear failure), three stages of the failure process are identified, namely, shear failure at the bank toe
(Stage I), tensile failure on the bank top (Stage II), and sectional cracking from the bank top to the toe (Stage III).
Note that both shear and tensile failures refer to soil element failure, rather than to overall failure of the cantilever
block described in Section 2.2. This type of analysis was later extended by Zhao et al. (2019), who considered
the effects of suspended sediment transport. Their results highlighted a negative effect of bank height on bank
stability, in good agreement with experimental observations (Zhao et al., 2020).

In summary, the stress-strain analysis should be applied when bank deformation and the consequent stress
concentrations due to the removal of bank soil (e.g., by near-bank channel flow, seepage flow and waves) are
non-negligible, such as the evaluation of overhanging stability. Since this approach captures many details in terms
of stress-strain behavior, it can be used for calibration of parameterized bank collapse models or to provide robust
assumptions for LEMs. For example, the modeling of soil creep can take advantage of stress-strain analysis. The
information provided by stress-strain analysis can in fact be employed to express soil diffusion as a function of
effective shear stresses and thus of pore-water pressure. Overall, stress-strain analyses should be used for short-
or medium-term predictions in the presence of complex external forces and therefore of soil deformation/stress
states that cannot be easily treated through simplified approaches. It could also be used for long-term predictions
in order to provide calibration and assumptions for more simplified models.

6. Feedbacks Between Bank Retreat and Morphodynamics


6.1. Effects of Bank Retreat on Morphodynamics

Bank retreat plays a key role on the morphodynamic evolution of natural rivers over a wide range of scales
(Figure 1). In general, bank retreat controls the equilibrium river width by eventually reducing the shear stress
on the banks to a critical threshold value (Francalanci et al., 2020; K. B. J. Dunne & Jerolmack, 2020). From
the perspective of cross-sectional evolution, bank retreat is responsible for the instantaneous adjustment of the
bank line (C. Hackney et al., 2015; Darby et al., 2007), which modulates near-bank hydrodynamics and, together
with the transverse transport of sediment driven by secondary flow circulations which establish in meander-
ing bends (Bolla Pittaluga & Seminara, 2011), gradually shift the thalweg from one bank to the other (Fryirs
& Brierley, 2012; Parker et al., 2011; Stecca et al., 2017; Yuan et al., 2021). In various settings, bank-derived
materials have been reported as the dominant factor of sediment budgets (Kronvang et al., 2013; Trimble, 2009).
The collapsed sediments not only significantly affect the local morphodynamics but also feed the river reach
downstream of the collapsed bank. For rivers in the loess area more than 80% of the total suspended sediment
can be ascribed to bank erosion and collapse (Simon et al., 2000). In addition, the transit of bank-derived sedi-
ment from a given catchment to the ocean contributes to the global carbon cycle (Galy et al., 2015; Golombek
et al., 2021; Repasch et al., 2022). For instance, lateral erosion of rivers cutting through floodplains releases

ZHAO ET AL. 30 of 51
Reviews of Geophysics 10.1029/2021RG000761

additional organic carbon fluxes to downstream depositional sinks (e.g., tidal flats and deltas). Hence, rivers with
high channel mobility can enhance CO2 drawdown (Repasch et al., 2021).

At the medium-term timescale (decades), bank retreat contributes significantly to river morphodynamics (Simon
et al., 2000; Thorne et al., 1998a). In meandering rivers, the interaction between outer bank retreat and inner bank
accretion determines channel widening or narrowing (Asahi et al., 2013; Parker et al., 2011; Zhao et al., 2021;
Zolezzi et al., 2012). For instance, high-resolution field observations have documented the continuous interplay
between bank-pull and bar-push mechanisms, whereby river bends widen and narrow in discrete steps while main-
taining a statistically constant mean channel width (Lopez Dubon & Lanzoni, 2019; Mason & Mohrig, 2019). In
the context of braided rivers, divergent flow due to braid bars facilitates channel widening, while the stabilizing
action of vegetation counteracts bank erosion and reduces bar and channel dynamics (Bristow & Best, 1993;
Schuurman et al., 2013). Laboratory experiments showed that bank strength provided by vegetation to a cohe-
sionless material is the necessary ingredient to narrow and deepen channels favoring, the transition from braiding
to meandering pattern (Braudrick et al., 2009; Tal & Paola, 2007; Van Dijk et al., 2013). In other words, an
increase in bank stability and, hence, a reduction of bank retreat rate controls the transition between braiding and
meandering rivers (Fredsøe, 1978; Gibling & Davies, 2012; Howard, 2009; Ielpi et al., 2022).

Over timescales of hundreds to thousands of years, bank retreat plays a fundamental role in determining flood-
plain heterogeneity, potentially forming strath terraces, and controlling the overall landscape evolution. Scroll
bars and oxbow lakes consequent to river bend cutoffs are responsible for spatial heterogeneity in floodplain
erodibility which, in turn, affects the long-term planform evolution of meandering rivers (Bogoni et al., 2017;
Güneralp & Rhoads, 2011). The interaction between lateral bank retreat and vertical bed incision controls valley
morphology. Remarkable examples are the spectacular lithological structures created by highly sinuous mean-
ders in deep slot canyons and the stepped strath terraces forming within broader mountain valleys (Hancock &
Anderson, 2002; Limaye & Lamb, 2014; Venditti et al., 2014). Generally, erosion-resistant rocks and high verti-
cal incision rates favor the formation of narrow canyons, while wide strath terraces tend to develop in relatively
weak sedimentary rocks with high bank retreat rates (Brocard & Van der Beek, 2006; Limaye & Lamb, 2014).
Field observations indicate a strong lithologic influence on strath terrace formation. Indeed, the asymmetry in
erodibility between submerged and emerged rocks promotes lateral widening rather than vertical incision of
bedrock rivers (Montgomery, 2004). In contrast, the intermittent collapse of the valley margins due to lateral
undercutting can facilitate vertical incision over lateral widening. The formation of talus piles from slump blocks,
in fact, shields the valley slopes and potentially prevents terrace formation (Malatesta et al., 2017).

The evolution of open coasts, tidal inlet and estuary systems is also related to bank retreat processes (Anthony
et al., 2010; Francalanci et al., 2013; Guo et al., 2021; Hughes, 2012; Zhang et al., 2015). Tidal meanders are
one of the most active geomorphic units of unvegetated intertidal mudflats. They migrate frequently owing
to outer bank retreat and, hence, can cause damage to coastal infrastructure such as seawalls (Figure 14) and
offshore wind structures. The planimetric shape of tidal channels is generally funneled in order to accommodate
a seaward increasing tidal prism (Gong et al., 2018; Lanzoni & D'Alpaos, 2015; Marani et al., 2002; van der
Wegen et al., 2008), while the bank profile is typically characterized by convex or concave shapes depending on
the mechanism dominating bank retreat (Zhao et al., 2019). For small tidal channels composed of highly cohesive
sediment, meander dynamics are very slow. The high thresholds for erosion typical of muddy sediment and the
limited size of the tidal prism leads to a situation whereby bank retreat typically occurs only in very sharp bends
(Kleinhans et al., 2009). Erosion of cohesive sediment in tidal settings is also affected by the possible presence
of cohesive extracellular polymeric substances (EPS) generated by microorganisms abundant on intertidal flats
(Figure 14c). Although EPS are widely regarded as bed “stabilizers,” enhancing sediment strength (Flemming &
Wuertz, 2019), recent flume experiments show that under wave actions, the inclusion of EPS may induce higher
mobility of the sediment, liquefying an otherwise stable bed (X. Chen et al., 2021). As a result, how EPS affects
bank stability and, consequently, tidal channel dynamics remains unclear.

Halophytic vegetation has a stabilizing effect on channel banks (Y. Chen et al., 2019). This effect, coupled with the
competing feedbacks produced by vegetation encroachment on salt marsh surfaces (Sgarabotto et al., 2021), leads
to small width-to-depth ratios. Moreover, the lateral migration rate can be reduced by several orders of magnitude
as compared to unvegetated mudflats (Finotello et al., 2018; Gabet, 1998). Salt marsh channels are thus often
thought of as being relatively stable landscape features (D'Alpaos et al., 2007; Kearney & Fagherazzi, 2016).
However, when normalized by local channel width, the observed migration rates of tidal and fluvial meanders

ZHAO ET AL. 31 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 14. (a) Short-term evolution of tidal meanders showing a fast migration of their centerlines toward seawalls (the centerlines are obtained using the waterline
method (Kang et al., 2017)). (b) Bank collapse of tidal channels on unvegetated intertidal mudflats (radial sand ridges in the southern Yellow Sea, China; image taken
on May 2019). (c) Evidence of abundant microorganisms on tidal channel banks (Jiangsu coast, China; image taken by K. Zhao on July 2021).

are shown to be quite similar (Finotello et al., 2018). Hence, in the long-term, the role of bank retreat in salt
marsh channels is non-negligible. It is also worthwhile to note that existing numerical models fail to reproduce
the formation of the highly curved branches that are typically observed in salt marsh channels, partly due to the
oversimplification or even exclusion of bank retreat process (Fagherazzi et al., 2012; Geng et al., 2021; Kirwan
& Murray, 2007; Temmerman et al., 2007). Over large spatial scales, salt marsh stability or deterioration is
controlled by the mutual interaction between vertical sea level rise and sediment supply, as well as the lateral
cliff retreat and vegetation colonization (Bendoni et al., 2016; Feagin et al., 2009; Francalanci et al., 2013).
Although the adaption to sea level rise can make salt marshes reach equilibrium in the vertical direction (Kirwan
et al., 2010), they may be inherently unstable in the horizontal direction (Fagherazzi et al., 2013), leaving unan-
swered the question of whether a salt marsh can really survive to future rates of sea level rises.

6.2. Integration Between Morphodynamic Models and Bank Retreat Processes

At the beginning of the 1990s, critical issues concerning the integration between morphodynamic and bank
retreat models were discussed by the ASCE Task Committee (Thorne et al., 1998b). This discussion included the
importance of: (a) the accurate prediction of the boundary shear stress distribution in the near-bank region, (b)
the simulation of the corresponding sediment fluxes over the entire channel width, (c) the calculation of the rate
of flow-induced bank erosion, and resultant bank profile deformation, (d) the evaluation of bank stability and
consequent adjustment of bank line, and (e) the exchanges of sediment (e.g., slump blocks) between the banks and
the bed material. Over the last 30 years, researchers have addressed the above issues with particular emphasis on
the deformation of the bank profile and its feedbacks on near-bank morphodynamics (Klavon et al., 2017; Rinaldi
& Darby, 2007; Rinaldi & Nardi, 2013).

The bank profile caused by erosion of bank material and/or consequent bank collapse has in general been
computed using near-bank bed shear stresses, provided by depth-averaged hydrodynamic models, either line-
arized (E. Eke et al., 2014; Motta et al., 2014; Zhao et al., 2021) or fully numerical (Darby & Thorne, 1996b;

ZHAO ET AL. 32 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 15. Illustration of the different approaches adopted for coupling hydrodynamic and bank retreat models (adapted from Zhao et al. (2019)). Panels (a–c) show the
procedure of the first method discussed in the text, while (a, b, d, e) display the procedure of the second method.

Langendoen et al., 2016; Rinaldi et al., 2008). However, vertical variations in shear stress are neglected by these
models and, hence, the submerged part of the bank is inevitably taken to have a rectangular shape. As a result,
this modeling approach is suitable only for steep banks controlled by cantilever failure (e.g., tidal creek banks on
muddy flats (Gong et al., 2018)). For inclined riverbanks composed of multiple stratigraphic layers, additional
procedures are needed to simulate the distribution of shear stresses along the wetted bank (Langendoen, 2000;
Motta et al., 2014). For example, the shear stress acting on each layer of the bank can be calculated by scaling
the shear stress at the bank toe by the hydraulic radius of the layer (see Figure 2 in Lai et al. (2015)). When part
of the bank is eroded by the flow, the computational domain (described through either structured or unstructured
grids) must be adjusted to adapt to the new bank geometry (Darby et al., 2007; Gong et al., 2018; Patsinghasanee
et al., 2017; Zhao et al., 2019).
Three main approaches have been put forward to couple together hydrodynamic and bank retreat models
(Figure 15). The first method records the accumulated retreat distance (Figures 15a–15c), and the hydrodynamic
mesh is updated when the cumulative retreat distance exceeds the transverse grid size (Darby et al., 2002; Deng
et al., 2019; Gong et al., 2018; Jia et al., 2010). The channel width thus varies intermittently, and instantaneous
feedbacks on the flow field are not accounted for. A possible solution to this shortcoming is the use of the so-called
immersed boundary method, whereby the bank line is followed with a cut-cell approach (Canestrelli et al., 2016;
Mittal & Iaccarino, 2005). The second method replaces bank retreat with a decrease in bank height (Abderrezzak
et al., 2016; Rousseau et al., 2017; Stecca et al., 2017; van der Wegen et al., 2008; Zhao et al., 2019). Once the
bank height decreases to a threshold value, mesh nodes representing the bank top are transformed to mesh nodes
representing the channel bed (e.g., point I in Figure 15d and 15e). Although bank retreat is not simulated by a
horizontal widening, this method accounts for a progressive reduction in near-bank flow velocity. Contrary to the
above fixed-mesh approaches, the third method re-generates boundary-fitted curvilinear or unstructured grids,
based on the simulated bank retreat distance (Asahi et al., 2013; Langendoen et al., 2016). The mesh nodes on
the bank top are shifted, and an additional criterion is applied to smooth the bank line. Since the grid lines do not
always align with the bank top, this method performs better in terms of river width adjustment (Lai et al., 2015).

Despite the remarkable advances performed in coupling bank retreat process with hydraulic modeling, some chal-
lenges deserve further attention. First, the distribution of near-bank shear stresses under complex bank/channel
bed topography remains unclear. This complexity is partly the result of bank undermining, which forms a concave

ZHAO ET AL. 33 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 16. (a) Deformation of bank profile as a result of undermining (Chongming Dongtan wetland, China; image taken by
K. Zhao on June 2016). (b) Illustration of dramatic riverbank retreat and resultant slump blocks on the Colville River delta in
northern Alaska (adapted from Walker (2013)).

cavity (Figure 16a). Also, the initiation of gullies, streams, and tidal creeks is affected by the configuration of
the bank profile and the related shear stress distribution. Second, for medium- and long-term simulations little
attention has been paid to the representation of bank retreat as an intermittent process. Since the estimation of
bank stability is complex and computationally costly, suitable parametrizations able to account for intermittent
bank collapse events are needed. The third issue concerns the interactions between groundwater fluctuations
and bank retreat. Notwithstanding much progress achieved in recent years (see Sections 5.3 and 5.4), the need
for a continuous adaptation of the computational mesh limits the application of LEMs and stress-strain analyses
to a broader context. The last issue is associated with the fate of slump blocks (Figure 16b), as discussed below
(Section 7.3).

7. Open Questions and Future Research Needs


The studies reviewed in the previous sections have produced remarkable advances in our understanding of bank
retreat in terms of failure mechanisms, development of empirical predictive relations, and numerical modeling.
However, many essential questions remain wholly or partially unanswered. In this section, we discuss three chal-
lenges that deserve attention in order to improve our fundamental understanding of bank retreat processes. The
first challenge is to unravel the complex mechanism of multifactor-driven bank erosion in both fluvial and tidal
environments, with particular attention to the interactions between hydrodynamics, soil properties, and biological
processes. The second challenge is to clarify whether the assessment of bank retreat in fluvial and tidal envi-
ronments needs somewhat ad hoc approaches, given the different temporal and spatial scales operating in these
different settings. The third issue is to assess the role that collapsed bank soil has on channel morphodynamics,
eventually affecting bank retreat, hydrodynamics, and sediment dynamics.

7.1. Multifactor-Driven Bank Retreat

There is ample experimental evidence (at both the field and laboratory scales) that bank retreat is driven by
the interactions between hydrodynamics and geotechnical factors, biological processes, and sediment dynamics
(Figure 17). However, much of the reviewed literature, with some notable exceptions, focuses on the effects of a
single factor and deals with fluvial environments, with their distinctive boundary conditions. Studies that account
for two or more factors, especially in the context of tidal environments, are therefore needed to improve our
understanding of bank retreat across a broad range of real-world systems.

The consequences of temporal variations in soil pore-water pressure on the physical properties of bank soil surely
merit attention. Changes in soil pore-water pressure are commonly associated with floods, periodic tides, overtop-
ping waves, seepage erosion, and drying-wetting cycles. Although quantitative observations have been conducted
for soil pore-water pressures (Midgley et al., 2013; Nardi et al., 2012), we still do not fully understand how
changes in pore-water pressures may influence bank stability through bank soil properties. Samadi et al. (2013)
replaced fluvial erosion processes with artificial undermining, thus separating the effects of soil properties. The
results of these controlled laboratory experiments underlined the importance of bank soil properties on bank
collapse mode, and indicated that tensile failure is absent when soil cohesion is low. The role of soil pore-water

ZHAO ET AL. 34 of 51
Reviews of Geophysics 10.1029/2021RG000761

Figure 17. Sketch of the multiple influences that drive bank retreat.

pressure has been long neglected in numerical modeling. Only a few studies have been conducted to explore the
interactions with other factors (Darby et al., 2007; Deng et al., 2018; K. Zhang et al., 2021). For instance, Darby
et al. (2007) considered the transient variation in pore-water pressure in response to flood events. They found
that the deformation of the bank profile due to fluvial erosion can itself alter the hydraulic head driving infiltra-
tion into the bank, and hence the distribution of pore-water pressure. These studies are as yet limited to fluvial
erosion. Other erosion processes typical of tidal environment, such as recursive water level oscillations, subsur-
face flow and wind waves, have been shown to affect soil pore-water pressure (Cao et al., 2012; Fox et al., 2006;
Francalanci et al., 2013; Xin et al., 2022) and therefore call for their implementation in numerical models.

Surprisingly, the role of vegetation on bank stability has received relatively little attention (Camporeale et al., 2013).
While qualitative comparisons have been performed through laboratory experiments carried out with and with-
out vegetation (Cancienne & Fox, 2008; Francalanci et al., 2013), a quantitative and systematic analysis is still
lacking. Moreover, previous laboratory experiments have been designed to maintain both geometrical (e.g., bank

ZHAO ET AL. 35 of 51
Reviews of Geophysics 10.1029/2021RG000761

height) and material (e.g., soil cohesion) scaling, following the suggestions of D. M. Wood (2014). Neglecting
the reinforcement ensured by vegetation roots might explain the field-laboratory discrepancy whereby the domi-
nant failure mechanism in downscaled laboratory experiments is toppling failure (Patsinghasanee et al., 2017;
Samadi et al., 2013; Zhao et al., 2020), while shear-type failures are most commonly observed in natural rivers
(Langendoen & Simon, 2008; Simon et al., 2000). Vegetation roots counteract the development of tension cracks
(Francalanci et al., 2013) typical of toppling failures (Patsinghasanee et al., 2015; Samadi et al., 2013; Zhao
et al., 2020). Clearly, maintaining geometrical, material and vegetation (e.g., roots density per unit width and
diameter) scaling in laboratory experiments is not a trivial task. More efforts are thus needed to unravel what role
vegetation roots play in bank stability, and how vegetation interacts with hydrodynamic forces, finding a reason-
able balance between the requirements for geometrical, material, and vegetation scaling.

In numerical models, the role of vegetation is commonly parameterized in terms of root cohesion (i.e., mechan-
ical effects). Hydrological effects (e.g., rainfall interception and soil moisture extraction) have received little
attention, as pointed out by Simon and Collison (2002). The mechanical effects so far treated are limited to the
root enhancement of shear strength. Little attention has been paid to tensile strength which, as noted above, was
proven to be important for toppling failure (Van Eerdt, 1985; Xia, Zong, Deng, et al., 2014). The cyclic fluctu-
ations of pore-water pressure, as a result of tides and waves, are also strongly influenced by vegetation cover,
and hence may affect the stability of salt marshes borders. The model developed by Xin et al. (2013), coupling
subsurface flow and plant growth, led to the identification of three characteristic zones for plant growth along salt
marsh creeks. These zones are determined by the combined influence of spring-neap tides and evapotranspira-
tion. This review thus supports the need to improve the understanding of bank stability with respect to vegetation,
particularly for environments such as salt marshes where the toppling mechanism dominates, and where tides and
wind waves control subsurface flow and, hence, pore-water pressure variations.

Further evidence for the role of multiple stressors (surface and seepage flow, ship waves) on bank retreat has
been provided by field measurements. For instance, Rengers and Tucker (2014) observed the following cycle for
multifactor-driven bank retreat. A pop-out failure driven by seepage flow leaved an overhanging block, followed
by toppling failure. The collapsed soil was then eroded by near-bank channel flow. Based on detailed monitoring
of upper-bank erosion, Duró et al. (2019) stated that floods were not necessary for the basal clean-out of failed
material, since ship waves during the dry season acted to disaggregate and remove the collapsed bank soil. More
research is however needed to improve the accuracy of the prediction of multifactor-driven bank retreat. For
long-term timescales empirical relations should be developed which integrate factors like water and sediment
discharges, wave power, and rainfall. An attempt in this sense has been pursued to predict the vertical incision
rate of bedrock rivers under various processes (Lague, 2014). Other factors such as soil chemical properties
(e.g., soil dispersion), biological perturbators (e.g., biofilms and crabs) are also likely to play a major role in
multifactor-driven bank retreat.

7.2. Similarities and Differences Between Fluvial and Tidal Environments

As reviewed in Section 3, there are some similarities between fluvial and tidal environments, such as the domi-
nant driving forces and failure modes associated with bank stability. Notwithstanding these similarities, that for
example, reflect in comparable migration rates of tidal and fluvial meanders when normalized to the channel
width (Finotello et al., 2018), the external forces imposed on river banks exhibit somewhat distinct character-
istics, arising from the differences in the drivers of flow motion. While hydrodynamic of rivers is governed by
topographic gradients, in the case of tidal landscapes water surface gradients drive the flow (Coco et al., 2013).
Here, we discuss various aspects that are clearly distinct between fluvial and tidal environments, thus requiring
specific attention to improve our understanding of bank retreat process.

Basically, the time-scale typical of water level variations within tidal systems (hours) is much faster than its
fluvial counterpart (months). The hydrograph of the pore-water pressure variation is also quite different when
dealing with tides and waves (Xin et al., 2011, 2016). In the case of waves (generated either by winds or by
boats), the frequency at which bank soil is subject to saturation and unsaturation cycles grows to a large extent as
compared to tidal forcing. On the other hand, Xin et al. (2011) found that the simulated groundwater dynamics
exhibited significant flow asymmetry over tidal cycles, and the timescale of pore-water circulation decreased
landward by orders of magnitude. Their work underlines the hydrological complexity of intertidal marshes, and
shows the importance of subsurface flow dynamics over a range of spatial scales. Since bank retreat rate is related

ZHAO ET AL. 36 of 51
Reviews of Geophysics 10.1029/2021RG000761

to the ratio Hb/Hw (Figure 11a in Section 4.2), for tidal systems the periodical variation in water level results in
an elusive curve of bank retreat rate. Also, tidal channel banks potentially experience two periodical erosion
mechanisms over one tidal cycle: coupled erosion by seepage and near-bank channel flow during ebb tides or
alternatively, by overbank flow at flood or ebb peak. As a result, previous models or empirical relations devel-
oped for riverine morphodynamics (e.g., BSTEM) might be unsuitable for tidal environments, due to their strong
simplifications or disregard of some key processes.

Another important distinction is that tidal channel banks experience periodical changes in flow direction. Accord-
ing to bend instability theory, meander bends migrate either upstream or downstream, depending on the phase
lag between the bend apex and the peak flow location (Lanzoni & Seminara, 2006). In tidal environments, the
flow field and consequently the location characterized by the maximum bank erosion rate differ between flood
and ebb tides (Solari et al., 2002). The key factor controlling the phase of the point bar pattern relative to channel
curvature is thus the flood- or ebb-dominant character of the basic flow (Tambroni et al., 2017). On the other
hand, bidirectional flows result in bank soil heterogeneity mainly in the cross-shore direction (due to sediment
sorting (Zhou et al., 2015, 2016)), rather than in the vertical direction as commonly observed in riverbanks (e.g.,
composite bank (Samadi et al., 2013; Xia, Zong, Deng, et al., 2014)). These processes, for example, can explain
why tidal meanders are in general less morphologically complex and display more spatially homogeneous charac-
teristics when compared to fluvial meanders (Finotello et al., 2020). Other processes, such as salinity dynamics,
affect bank stability through soil dispersion (Masoodi et al., 2019). More efforts are therefore needed to investi-
gate bank retreat resulting from the above processes, and ultimately to unravel the intrinsic differences between
fluvial and tidal systems.

7.3. Collapsed Bank Soil

Collapsed bank soil (also referred to as slump blocks) is the product of bank collapse. When such material depos-
its at the base of the bank, it modulates bank erosion, affects the flow field, and alters the channel topography. Part
of the collapsed bank soil is likely to protect the bank from direct erosion (A. L. Wood et al., 2001). Fagherazzi
et al. (2004) observed that collapsed bank soil in salt-marsh creeks was able to persist for several years, and was
responsible for the channel erosion paradox (Gabet, 1998), whereby marsh creeks are likely to migrate laterally
at a quite slow rate despite the widespread occurrence of bank collapse. Midgley et al. (2013) reported that
seepage erosion became temporarily restricted after bank collapse, since the collapsed bank soil blocked flow
pathways and limited particle mobilization. From a morphodynamic perspective, collapsed bank soil abruptly
alters the local topography and the flow resistance of the near-bank channel bed (K. Zhang et al., 2021), possibly
generating complex turbulence at the bank toe and affecting the cross-sectional evolution of channels. Based
on three-dimensional flow field observations, C. Hackney et al. (2015) stated that collapsed bank soil may
also deflect flow onto the bank, thus enhancing bank erosion rate. In addition, collapsed bank soil is a source
of sediment that may alter downstream channel morphology. Large failures are likely to cause the temporary
constriction of the cross-sectional area of the flow, especially for small rivers covered by trees. This narrowing
induces velocity gradients along the adjacent cross sections, and might be a reason for the inception of bar-pool
patterns along channels (Duró et al., 2016). Another distinct phenomenon is detected in cold environments,
where collapsed bank soil is often surrounded by ice blocks, and therefore can be transported more downstream
by means of drifting ice (Black et al., 2018). Collapsed bank soil has also been proven to act as a bridge between
near-bank hydrodynamics and morphodynamics (e.g., through the concept of basal endpoint control mentioned in
Section 4.2), and therefore plays an important role in the bank retreat cycle (Figure 17). More research is needed
to understand how collapsed bank soil, bank erosion/collapse, hydrodynamics, and sediment dynamics interact
to form the overall architecture and morphology of fluvial and tidal settings, depending also on the size of the
considered channel and the intensity of the flow therein.

Recently, the fate of collapsed bank soil has been investigated through numerical modeling (Asahi et al., 2013;
Darby et al., 2002; Deng et al., 2019; E. Eke et al., 2014; Lai et al., 2015; Langendoen, 2000; Parker et al., 2011;
Zhao et al., 2019). Parker et al. (2011) proposed an armoring coefficient, as a multiplier of sediment flux, to
account for the armoring effect of collapsed bank soil. While accounting for the protection effects, this method
neglects the consequences on local topography and hydrodynamics. The role of apparent cohesion in the removal
of collapsed bank soil was also underlined by A. L. Wood et al. (2001), who suggested that not only block size,
but also the vertical distribution of apparent cohesion should be accounted for to enable improved estimation of

ZHAO ET AL. 37 of 51
Reviews of Geophysics 10.1029/2021RG000761

the entrainment of the collapsed bank soil. An attempt to account globally for all these effects in a probabilistic
approach has been recently put forward by Lopez Dubon and Lanzoni (2019). More research is clearly needed to
better describe the feedbacks between bank erosion/collapse and the collapsed soil.

8. Summary and Conclusions


We have presented a comprehensive review of bank retreat dynamics with respect to mechanisms, observations,
and modeling, covering both rivers and tidal channels. Our review includes the commonly observed failure mech-
anisms and other factors that can cause bank retreat, synthesizes laboratory and in situ observations of bank
retreat rate, and discusses the numerical methods used to simulate bank retreat.

On the basis of observational data, we have reviewed mechanisms of bank collapse with respect to subaerial
processes, surface flow (including near-bank channel flow and overbank flow), seepage flow, fluctuations in soil
pore-water pressure and waves. Specific attention has been paid to the role of vegetation, biological disturbances,
and hydrostatic pressure head. We have provided evidence demonstrating that the various types of external forces,
despite their distinct characteristics, may have similar effects on bank stability, leading to the same failure mode:
typical erosion and tensile failure in the middle and lower part of the bank, followed by toppling failure.

Existing empirical relations for predicting bank retreat exhibit a large scatter over spatial and temporal scales, a
consequence of the inherent complexity of the interactions and feedbacks between the mechanisms controlling
bank retreat. When compared with the hydraulic-based or geotechnical-based empirical predictors, the relations
that integrate hydraulic and geotechnical impacts are more accurate, thus highlighting the necessity to account for
both hydraulic and geotechnical parameters.

Based on the way in which bank collapse is accounted for, we categorize existing modeling approaches into a
hierarchy: purely hydraulic (in which only sediment erosion is accounted for), parameterized bank collapse, static
equilibrium, and stress-strain analysis. We also discuss the advantages and challenges of these methods, and
provide some model recommendations in terms of real-world applications.

Overall, this review recognizes the role of bank retreat on the overall architecture and morphology of rivers and
tidal channels. We also propose three research directions (multifactor-driven bank retreat, discrepancy between
fluvial and tidal environments, and the role of collapsed bank soil) that we believe are critical to advance the
future understanding and prediction of morphodynamics in these systems.

Symbols
a, b, c, d Empirical regression coefficient
Achunk Volume of the collapsed bank soil per unit length
Ad Drainage area (km 2)
Φ60Ai Effective drainage area (m 2)
Bc Width of cantilever (m)
Bb Width of marsh cliff block (m)
c′ Effective soil cohesion (kPa)
𝐴𝐴 𝐴𝐴2′′  Empirical parameter depending on sediment packing
cr Cohesion provided by vegetation roots (kPa)
CEI Cavity Erosion Index
Cs Concentration of sediment mobilized by seepage sediment (g/L)
D Grain size of bank material (mm)
Df Days with air frost (day)
Em Mass erosion rate for bank materials (kg/s)
El Linear erosion rate for bank materials (m/year)
Ev Volumetric erosion rate for bank materials (m 3/year)
Ea Areal erosion rate for bank materials (m 2/year)
Fs Safety factor
FR Resisting forces used to calculate safety factor
FD Destabilizing forces used to calculate safety factor

ZHAO ET AL. 38 of 51
Reviews of Geophysics 10.1029/2021RG000761

Fw Hydrodynamic thrust due to water waves


g Acceleration due to gravity (m/s 2)
Hb Bank height (m)
Hc Height of the cantilever (m)
Hp Thickness of the cantilever subject to tensile failure (m)
Ht Depth of tension crack on the bank top (m)
Hub Upper bank height (m)
Hw Near-bank water depth (m)
i Hydraulic gradient driving seepage flow
ic Critical hydraulic gradient for seepage erosion
Kl Volumetric erodibility coefficient for surface flow erosion (m 3/N/s)
Kw Erodibility coefficient for wave erosion
Ks Seepage erodibility coefficient
Ksat Saturated hydraulic conductivity (m/s)
lt Length of cantilever under tensile stress (m)
lc Length of cantilever under compressive stress (m)
P Hydrostatic pressure per unit width (kN/m)
Pc Threshold value for wave-induced retreat
Pic Hydrostatic pressure inside the crack (kN/m)
Pw Mean wave power calculated over a representative period (W/m)
Ps Precipitation intensity (mm)
Qs Seepage discharge (L/s)
Ql Channel flow discharge (L/s)
Qc Critical seepage discharge for bank erosion (L/s)
𝐴𝐴 𝐴𝐴𝑠𝑠∗  Dimensionless seepage-induced sediment flux
qb Volumetric sediment transport rate per unit width in the transverse direction (m 2/s)
rl Dimensionless normalized bank retreat rate in response to surface flow
S Slope gradient
t Time (s)
tb Elapsed time of first collapse (min)
U Hydrostatic-uplift force per unit width (kN/m)
Uc Near-bank flow velocity (m/s)
Ud Darcy velocity (m/s)
uw Pore water pressure (kPa)
ua Pore air pressure (kPa)
Vs Seepage-induced cavity volume (cm 3)
Vg Eroded bank gully volume (m 3)
W Soil weight per unit width (kN/m)
wt Overhanging block width (m)
wc Channel width (m)
zb, zl Dimensions of the failure soil block (m)
α Bank angle (°)
β Failure-plane angle (°)
η Near-bank bed elevation (m)
θ Oscillation angle around an equilibrium configuration (°)
ε Bank retreat or accretion rate (m/s)
Δ Ratio of bank soil density to water density
σ Total normal stress (kPa)
σc Compressive stress along the failure plane (kPa)
σt Tensile strength along the failure plane (kPa)
τ Shear strength of bank materials (kPa)
𝐴𝐴 𝐴𝐴𝑠𝑠∗  Dimensionless shear stress induced by seepage
τb Boundary shear stress applied by the near-bank flow (Pa)
τc Critical shear stress for bank erosion (Pa)

ZHAO ET AL. 39 of 51
Reviews of Geophysics 10.1029/2021RG000761

𝐴𝐴 ∗
𝐴𝐴𝑐𝑐𝑐𝑐  Dimensionless critical shear stress for seepage erosion
λ Direction of the seepage vector (°)
λb Porosity of the bank material
φ′ Effective internal friction angle (°)
φb Angle expressing the rate of increase in strength relative to the matric suction (°)
ω Stream power per unit bed area
γb Bank soil resistance

Glossary
Apparent cohesion the cohesion of grains caused by surface tension in the surrounding
pore water.
Angle of repose the steepest angle of descent or dip relative to the horizontal plane
without slumping.
Bank stratification bank composed of heterogeneous layers.
Cohesive a collection of sediment particles that cohere, or stick together, largely
due to electrochemical forces.
Drainage area the land area where precipitation falls off into river basins and gullies,
usually identified by the line along the highest topographic eleva-
tion (in fluvial settings) or the zero flux divide (in tidal settings) and
connecting its ends to the outer section of the basin/gullies.
Dry granular flow a kind of bank collapse consisting of an avalanche of granular, loose
sediment, creating a fan-shaped debris accumulation close to the angle
of repose.
Effective internal friction angle internal friction angle in the context of saturated soils.
Effective soil cohesion soil cohesion in the context of saturated soils.
Elastic-plastic model a description of the relation between stress and strain.
Elastic potential energy the energy stored as a result of applying a force to deform an elastic
object.
Evapotranspiration sum of evaporation and plant transpiration.
Evaporation the process by which water changes from liquid to a gas or vapor.
Fluidization  a process similar to liquefaction whereby a granular material is
converted from a static solid-like state to a dynamic fluid-like state.
Gully erosion the removal of soil along drainage lines by surface water runoff.
Headcuts an erosional feature characterized by an abrupt vertical drop. Also
called “knickpoint.”
Hydraulic conductivity the ability of the material to transmit fluid through pore spaces and
fractures in the presence of an applied hydraulic gradient.
Hydraulic head liquid pressure above a vertical datum.
Infiltration the process by which water on the ground surface enters the soil.
Inner bank the bank with the smallest radius of curvature around a bend, commonly
characterized by bank accretion.
Internal friction angle measure of the ability of a unit of a soil to withstand a shear stress,
defined by the angle between the applied shear stress and the normal
effective stress at which shear failure occurs.
Matric suction the pressure that a dry soil exerts on the surrounding soils to equalize
everywhere the moisture content.
Mechanical fatigue weakening of bank soils caused by cyclic loading such as waves.
Mohr-Coulomb failure criterion a mathematical model describing the state of soil units.
Normal stress the stress perpendicular to a specific plane.
Outer bank the bank with the largest radius of curvature around a bend, commonly
characterized by bank erosion.
Perched water table an aquifer that occurs above the regional water table, in the vadose
zone.

ZHAO ET AL. 40 of 51
Reviews of Geophysics 10.1029/2021RG000761

Plunge pool erosion bed erosion immediately downstream of headcuts.


Root cohesion soil cohesion provided by vegetation roots.
Shear strength the resistance of a material to breaking under shear.
Shear stress the component of stress coplanar with a material cross section.
Soil cohesion soil shear strength that is independent of interparticle friction.
Soil desiccation soil free from all moisture.
Soil diffusivity a coefficient to describe soil creep.
Soil dispersion soils become vulnerable as a result of hydration of sodium ions between
clays.
Soil permeability the property of soils to transmit water and air.
Static liquefaction the sudden loss of strength when loose soil, typically granular materials
such as sand or silt, is loaded and cannot drain.
Stemflow the flow of intercepted water down the trunk or stem of a plant.
Subaerial process weathering and mass movement process.
Tensile strength the resistance of a material to breaking under tension.
Tensile stress the stress caused by pulling the material.
Tension crack cracks induced by tensile force.
Undercutting bank erosion occurring at the lower part of the bank.
Weathering the process that changes solid rock into sediments.

Data Availability Statement


All data shown in the figures of this review is available on GitHub website (https://github.com/zk1357/Review-
on-bank-retreat), and is extracted from the following references: Seginer (1966), Hooke (1980), Vandekerckhove,
Muys, et al. (2001), Vandekerckhove, Poesen, et al. (2001), Vandekerckhove et al. (2003), Braudrick et al. (2009),
De Rose and Basher (2011), van Dijk et al. (2012), Wells et al. (2013), Patsinghasanee et al. (2017), Qin
et al. (2018), Vargas Luna et al. (2019), Shu et al. (2019), and Zhao et al. (2020).

Acknowledgments References
This research was supported by the
National Natural Science Foundation of Abate, M., Nyssen, J., Steenhuis, T. S., Moges, M. M., Tilahun, S. A., Enku, T., & Adgo, E. (2015). Morphological changes of Gumara River
China (51925905), the China Postdoctoral channel over 50 years, upper Blue Nile basin, Ethiopia. Journal of Hydrology, 525, 152–164. https://doi.org/10.1016/j.jhydrol.2015.03.044
Science Foundation (2021M701050), Abderrezzak, K. E. K., Moran, A. D., Tassi, P., Ata, R., & Hervouet, J. (2016). Modelling river bank erosion using a 2D depth-averaged numer-
the Fundamental Research Funds for the ical model of flow and non-cohesive, non-uniform sediment transport. Advances in Water Resources, 93, 75–88. https://doi.org/10.1016/j.
Central Universities (B220202079), and advwatres.2015.11.004
the National Natural Science Founda- Abernethy, B., & Rutherfurd, I. D. (1998). Where along a river's length will vegetation most effectively stabilise stream banks? Geomorphology,
tion of China (51879095). The authors 23(1), 55–75. https://doi.org/10.1016/S0169-555X(97)00089-5
acknowledge valuable comments from Akay, O., Özer, A. T., Fox, G. A., & Wilson, G. V. (2018). Application of fibrous streambank protection against groundwater seepage erosion.
Sebastien Carretier, Simona Francalanci, Journal of Hydrology, 565, 27–38. https://doi.org/10.1016/j.jhydrol.2018.08.010
and one anonymous reviewer and the Allen, J. (1989). Evolution of salt-marsh cliffs in muddy and sandy systems: A qualitative comparison of British west-coast estuaries. Earth
managing editors, which led to significant Surface Processes and Landforms, 14(1), 85–92. https://doi.org/10.1002/esp.3290140108
improvement of the paper. Special thanks Amiri, M., Pourghasemi, H. R., Ghanbarian, G. A., & Afzali, S. F. (2019). Assessment of the importance of gully erosion effective factors
are given to Qingyun Duan and Pei Xin, using Boruta algorithm and its spatial modeling and mapping using three machine learning algorithms. Geoderma, 340, 55–69. https://doi.
who provide suggestions for the structure org/10.1016/j.geoderma.2018.12.042
and revision of this review. Additional Anthony, E. J., Gardel, A., Gratiot, N., Proisy, C., Allison, M. A., Dolique, F., & Fromard, F. (2010). The Amazon-influenced muddy coast
thanks go to Yanyan Kang for her help on of South America: A review of mud-bank–shoreline interactions. Earth-Science Reviews, 103(3–4), 99–121. https://doi.org/10.1016/j.
Figure 14. earscirev.2010.09.008
Arabameri, A., Yamani, M., Pradhan, B., Melesse, A., Shirani, K., & Bui, D. T. (2019). Novel ensembles of COPRAS multi-criteria decision-
making with logistic regression, boosted regression tree, and random forest for spatial prediction of gully erosion susceptibility. The Science of
the Total Environment, 688, 903–916. https://doi.org/10.1016/j.scitotenv.2019.06.205
Arai, R., Ota, K., Sato, T., & Toyoda, Y. (2018). Experimental investigation on cohesionless sandy bank failure resulting from water level rising.
International Journal of Sediment Research, 33(1), 47–56. https://doi.org/10.1016/j.ijsrc.2017.08.002
Asahi, K., Shimizu, Y., Nelson, J., & Parker, G. (2013). Numerical simulation of river meandering with self-evolving banks. Journal of Geophys-
ical Research: Earth Surface, 118(4), 2208–2229. https://doi.org/10.1002/jgrf.20150
Bassis, J. N., Berg, B., Crawford, A. J., & Benn, D. I. (2021). Transition to marine ice cliff instability controlled by ice thickness gradients and
velocity. Science, 372(6548), 1342–1344. https://doi.org/10.1126/science.abf6271
Baynes, E. R., Lague, D., Steer, P., Bonnet, S., & Illien, L. (2020). Sediment flux-driven channel geometry adjustment of bedrock and mixed
gravel-bedrock rivers. Earth Surface Processes and Landforms, 45(14), 3714–3731. https://doi.org/10.1002/esp.4996
Beechie, T. J., Liermann, M., Pollock, M. M., Baker, S., & Davies, J. (2006). Channel pattern and river-floodplain dynamics in forested mountain
river systems. Geomorphology, 78(1–2), 124–141. https://doi.org/10.1016/j.geomorph.2006.01.030

ZHAO ET AL. 41 of 51
Reviews of Geophysics 10.1029/2021RG000761

Bendoni, M., Francalanci, S., Cappietti, L., & Solari, L. (2014). On salt marshes retreat: Experiments and modeling toppling failures induced by
wind waves. Journal of Geophysical Research: Earth Surface, 119(3), 603–620. https://doi.org/10.1002/2013JF002967
Bendoni, M., Mel, R., Solari, L., Lanzoni, S., Francalanci, S., & Oumeraci, H. (2016). Insights into lateral marsh retreat mechanism through
localized field measurements. Water Resources Research, 52(2), 1446–1464. https://doi.org/10.1002/2015WR017966
Bennett, S. J. (1999). Effect of slope on the growth and migration of headcuts in rills. Geomorphology, 30(3), 273–290. https://doi.org/10.1016/
S0169-555X(99)00035-5
Bennett, S. J., Alonso, C. V., Prasad, S. N., & Römkens, M. J. (2000). Experiments on headcut growth and migration in concentrated flows typical
of upland areas. Water Resources Research, 36(7), 1911–1922. https://doi.org/10.1029/2000WR900067
Ben Slimane, A., Raclot, D., Evrard, O., Sanaa, M., Lefevre, I., & Le Bissonnais, Y. (2016). Relative contribution of rill/interrill and gully/
channel erosion to small reservoir siltation in Mediterranean environments. Land Degradation & Development, 27(3), 785–797.
https://doi.org/10.1002/ldr.2387
Bernatchez, P., & Dubois, J. M. (2008). Seasonal quantification of coastal processes and cliff erosion on fine sediment shorelines in a cold
temperate climate, north shore of the St. Lawrence maritime estuary, Québec. Journal of Coastal Research, 24(10024), 169–180. https://doi.
org/10.2112/04-0419.1
Bernatek-Jakiel, A., & Poesen, J. (2018). Subsurface erosion by soil piping: Significance and research needs. Earth-Science Reviews, 185,
1107–1128. https://doi.org/10.1016/j.earscirev.2018.08.006
Best, J. (2019). Anthropogenic stresses on the world’s big rivers. Nature Geoscience, 12(1), 7–21. https://doi.org/10.1038/s41561-018-0262-x
Black, C., Hill, P. S., & DeGelleke, L. (2018). Formation, collapse and composition of ice banks in a macrotidal channel of the Bay of Fundy.
Cold Regions Science and Technology, 155, 29–36. https://doi.org/10.1016/j.coldregions.2018.06.012
Blondeaux, P., & Seminara, G. (1985). A unified bar–bend theory of river meanders. Journal of Fluid Mechanics, 157, 449–470. https://doi.
org/10.1017/S0022112085002440
Bogoni, M., Putti, M., & Lanzoni, S. (2017). Modeling meander morphodynamics over self-formed heterogeneous floodplains. Water Resources
Research, 53(6), 5137–5157. https://doi.org/10.1002/2017WR020726
Bolla Pittaluga, M., & Seminara, G. (2011). Nonlinearity and unsteadiness in river meandering: A review of progress in theory and modelling.
Earth Surface Processes and Landforms, 36(1), 20–38. https://doi.org/10.1002/esp.2089
Bortolus, A., & Iribarne, O. (1999). Effects of the SW Atlantic burrowing crab Chasmagnathus granulata on a Spartina salt marsh. Marine Ecol-
ogy Progress Series, 178, 79–88. https://doi.org/10.3354/meps178079
Braudrick, C. A., Dietrich, W. E., Leverich, G. T., & Sklar, L. S. (2009). Experimental evidence for the conditions necessary to sustain meandering
in coarse-bedded rivers. Proceedings of the National Academy of Sciences, 106(40), 16936–16941. https://doi.org/10.1073/pnas.0909417106
Bristow, C. S., & Best, J. L. (1993). Braided rivers: Perspectives and problems. Geological society, London, special publications, 75(1), 1–11.
https://doi.org/10.1144/gsl.sp.1993.075.01.01
Brocard, G. Y., & Van der Beek, P. A. (2006). Influence of incision rate, rock strength, and bedload supply on bedrock river gradients and
valley-flat widths: Field-based evidence and calibrations from western Alpine rivers (southeast France), S. D. Willett et al. Special Papers –
Geological Society of America, 398, 101–126.
Bufe, A., Paola, C., & Burbank, D. W. (2016). Fluvial bevelling of topography controlled by lateral channel mobility and uplift rate. Nature
Geoscience, 9(9), 706–710. https://doi.org/10.1038/ngeo2773
Bufe, A., Turowski, J. M., Burbank, D. W., Paola, C., Wickert, A. D., & Tofelde, S. (2019). Controls on the lateral channel-migration rate
of braided channel systems in coarse non-cohesive sediment. Earth Surface Processes and Landforms, 44(14), 2823–2836. https://doi.
org/10.1002/esp.4710
Burkard, M. B., & Kostaschuk, R. A. (1997). Patterns and controls of gully growth along the shoreline of Lake Huron, Earth Surface
Processes and landforms. The Journal of the British Geomorphological Group, 22(10), 901–911. https://doi.org/10.1002/(SICI)1096-
9837(199710)22:10<901::AIDESP743>3.0.CO;2-O
Camporeale, C., Perona, P., Porporato, A., & Ridolfi, L. (2007). Hierarchy of models for meandering rivers and related morphodynamic processes.
Reviews of Geophysics, 45(1), RG1001. https://doi.org/10.1029/2005RG000185
Camporeale, C., Perucca, E., Ridolfi, L., & Gurnell, A. M. (2013). Modeling the interactions between river morphodynamics and riparian vege-
tation. Reviews of Geophysics, 51(3), 379–414. https://doi.org/10.1002/rog.20014
Cancienne, R. M., & Fox, G. A. (2008). Laboratory experiments on three-dimensional seepage erosion undercutting of vegetated banks. Ameri-
can Society of Agricultural and Biological Engineers. https://doi.org/10.13031/2013.25035
Canestrelli, A., Spruyt, A., Jagers, B., Slingerland, R., & Borsboom, M. (2016). A mass-conservative staggered immersed boundary model for
solving the shallow water equations on complex geometries. International Journal for Numerical Methods in Fluids, 81(3), 151–177. https://
doi.org/10.1002/fld.4180
Cantelli, A., Paola, C., & Parker, G. (2004). Experiments on upstream-migrating erosional narrowing and widening of an incisional channel
caused by dam removal. Water Resources Research, 40(3), W03304. https://doi.org/10.1029/2003WR002940
Cao, M., Xin, P., Jin, G., & Li, L. (2012). A field study on groundwater dynamics in a salt marsh-Chongming Dongtan wetland. Ecological
Engineering, 40, 61–69. https://doi.org/10.1016/j.ecoleng.2011.12.018
Capra, A., Porto, P., & Scicolone, B. (2009). Relationships between rainfall characteristics and ephemeral gully erosion in a cultivated catchment
in Sicily (Italy). Soil and Tillage Research, 105(1), 77–87. https://doi.org/10.1016/j.still.2009.05.009
Carretier, S., Martinod, P., Reich, M., & Godderis, Y. (2016). Modelling sediment clasts transport during landscape evolution. Earth Surface
Dynamics, 4(1), 237–251. https://doi.org/10.5194/esurf-4-237-2016
Carson, M. A., & Kirkby, M. J. (1972). Hillslope form and process. Cambridge University Press.
Carter, C. L., & Anderson, R. S. (2006). Fluvial erosion of physically modeled abrasion-dominated slot canyons. Geomorphology, 81(1–2),
89–113. https://doi.org/10.1016/j.geomorph.2006.04.006
Casagli, N., Rinaldi, M., Gargini, A., Curini, A., Thorne, C. R., Billi, P., & Rinaldi, M. (1999). Pore water pressure and streambank stability:
Results from a monitoring site on the Sieve River, Italy. Earth Surface Processes and Landforms, 24(12), 1095–1114. https://doi.org/10.1002/
(SICI)1096-9837(199911)24:12<1095::AIDESP37>3.0.CO;2-F
Castillo, C., & Gómez, J. A. (2016). A century of gully erosion research: Urgency, complexity and study approaches. Earth-Science Reviews, 160,
300–319. https://doi.org/10.1016/j.earscirev.2016.07.009
Castro-Bolinaga, C. F., & Fox, G. A. (2018). Streambank erosion: Advances in monitoring, modeling and management. Water, 10(10), 1346.
https://doi.org/10.3390/w10101346
Chassiot, L., Lajeunesse, P., & Bernier, J. F. (2020). Riverbank erosion in cold environments: Review and outlook. Earth-Science Reviews, 207,
103231. https://doi.org/10.1016/j.earscirev.2020.103231

ZHAO ET AL. 42 of 51
Reviews of Geophysics 10.1029/2021RG000761

Chen, A., Zhang, D., Peng, H., Fan, J., Xiong, D., & Liu, G. (2013). Experimental study on the development of collapse of overhanging layers of
gully in Yuanmou Valley, China. Catena, 109, 177–185. https://doi.org/10.1016/j.catena.2013.04.002
Chen, A., Zhang, D., Yan, B., Lei, B., & Liu, G. (2015). Main types of soil mass failure and characteristics of their impact factors in the Yuanmou
Valley, China. Catena, 125, 82–90. https://doi.org/10.1016/j.catena.2014.10.011
Chen, C., Hsieh, T., & Yang, J. (2017). Investigating effect of water level variation and surface tension crack on riverbank stability. Journal of
Hydro-Environment Research, 15, 41–53. https://doi.org/10.1016/j.jher.2017.02.002
Chen, X., Zhang, C., Townend, I. H., Paterson, D. M., Gong, Z., Jiang, Q., et al. (2021). Biological cohesion as the architect of bed movement
under wave action. Geophysical Research Letters, 48(5), e2020G–e92137G. https://doi.org/10.1029/2020gl092137
Chen, X. D., Zhang, C. K., Paterson, D. M., Thompson, C., Townend, I. H., Gong, Z., et al. (2017). Hindered erosion: The biological mediation
of noncohesive sediment behavior. Water Resources Research, 53(6), 4787–4801. https://doi.org/10.1002/2016WR020105
Chen, Y., & Collins, M. B. (2007). The influence of root systems on the geomorphology of a tidal creek: Exbury Marsh. Southern England.
Chen, Y., Collins, M. B., & Thompson, C. E. (2011). Creek enlargement in a low-energy degrading saltmarsh in southern England. Earth Surface
Processes and Landforms, 36(6), 767–778. https://doi.org/10.1002/esp.2104
Chen, Y., Thompson, C., & Collins, M. (2019). Controls on creek margin stability by the root systems of saltmarsh vegetation. Beaulieu Estuary,
Southern England, Anthropocene Coasts, 2(1), 21–38. https://doi.org/10.1139/anc-2018-0005
Chu-Agor, M. L., Fox, G. A., Cancienne, R. M., & Wilson, G. V. (2008). Seepage caused tension failures and erosion undercutting of hillslopes.
Journal of Hydrology, 359(3), 247–259. https://doi.org/10.1016/j.jhydrol.2008.07.005
Chu-Agor, M. L., Fox, G. A., & Wilson, G. V. (2009). Empirical sediment transport function predicting seepage erosion undercutting for cohesive
bank failure prediction. Journal of Hydrology, 377(1), 155–164. https://doi.org/10.1016/j.jhydrol.2009.08.020
Coco, G., Zhou, Z., van Maanen, B., Olabarrieta, M., Tinoco, R., & Townend, I. (2013). Morphodynamics of tidal networks: Advances and
challenges. Marine Geology, 346, 1–16. https://doi.org/10.1016/j.margeo.2013.08.005
Constantine, J. A., Dunne, T., Ahmed, J., Legleiter, C., & Lazarus, E. D. (2014). Sediment supply as a driver of river meandering and floodplain
evolution in the Amazon Basin. Nature Geoscience, 7(12), 899–903. https://doi.org/10.1038/ngeo2282
Coops, H., Geilen, N., Verheij, H. J., Boeters, R., & van der Velde, G. (1996). Interactions between waves, bank erosion and emergent vegetation:
An experimental study in a wave tank. Aquatic Botany, 53(3–4), 187–198. https://doi.org/10.1016/0304-3770(96)01027-3
Coulthard, T. J., Neal, J. C., Bates, P. D., Ramirez, J., de Almeida, G. A., & Hancock, G. R. (2013). Integrating the LISFLOOD-FP 2D hydro-
dynamic model with the CAESAR model: Implications for modelling landscape evolution. Earth Surface Processes and Landforms, 38(15),
1897–1906. https://doi.org/10.1002/esp.3478
Coulthard, T. J., & Wiel, M. J. V. D. (2006). A cellular model of river meandering. Earth Surface Processes and Landforms, 31(1), 123–132.
https://doi.org/10.1002/esp.1315
Couper, P. (2003). Effects of silt–clay content on the susceptibility of river banks to subaerial erosion. Geomorphology, 56(1–2), 95–108. https://
doi.org/10.1016/s0169-555x(03)00048-5
Couper, P. R. (2004). Space and time in river bank erosion research: A review. Area, 36(4), 387–403. https://doi.org/10.1111/j.0004-0894.2004.00239.x
Couper, P. R., & Maddock, I. P. (2001). Subaerial river bank erosion processes and their interaction with other bank erosion mechanisms on the
River Arrow, Warwickshire, UK. Earth Surface Processes and Landforms, 26(6), 631–646. https://doi.org/10.1002/esp.212
Crosato, A. (2007). Effects of smoothing and regridding in numerical meander migration models. Water Resources Research, 43(1), W01401.
https://doi.org/10.1029/2006wr005087
Dacey, J. W., & Howes, B. L. (1984). Water uptake by roots controls water table movement and sediment oxidation in short Spartina marsh.
Science, 224(4648), 487–489. https://doi.org/10.1126/science.224.4648.487
D'Alpaos, A., Lanzoni, S., Marani, M., & Rinaldo, A. (2007). Landscape evolution in tidal embayments: Modeling the interplay of erosion, sedi-
mentation, and vegetation dynamics. Journal of Geophysical Research, 112(F1), F01008. https://doi.org/10.1029/2006jf000537
Daly, E. R., Miller, R. B., & Fox, G. A. (2015). Modeling streambank erosion and failure along protected and unprotected composite streambanks.
Advances in Water Resources, 81, 114–127. https://doi.org/10.1016/j.advwatres.2015.01.004
Dapporto, S., Rinaldi, M., Casagli, N., & Vannocci, P. (2003). Mechanisms of riverbank failure along the Arno River, Central Italy, Earth Surface
Processes and landforms. The Journal of the British Geomorphological Research Group, 28(12), 1303–1323. https://doi.org/10.1002/esp.550
Darby, S. E., Alabyan, A. M., & Van de Wiel, M. J. (2002). Numerical simulation of bank erosion and channel migration in meandering rivers.
Water Resources Research, 38(9), 1–2. https://doi.org/10.1029/2001wr000602
Darby, S. E., Rinaldi, M., & Dapporto, S. (2007). Coupled simulations of fluvial erosion and mass wasting for cohesive river banks. Journal of
Geophysical Research, 112(F3), F03022. https://doi.org/10.1029/2006JF000722
Darby, S. E., & Thorne, C. R. (1996a). Development and testing of riverbank-stability analysis. Journal of Hydraulic Engineering, 122(8),
443–454. https://doi.org/10.1061/(ASCE)0733-9429(1996)122:8(443)
Darby, S. E., & Thorne, C. R. (1996b). Numerical simulation of widening and bed deformation of straight sand-bed rivers. I: Model development.
Journal of Hydraulic Engineering, 122(4), 184–193. https://doi.org/10.1061/(asce)0733-9429(1996)122:4(184)
Das, V. K., Roy, S., Barman, K., Debnath, K., Chaudhuri, S., & Mazumder, B. S. (2019). Investigations on undercutting processes of cohesive
river bank. Engineering Geology, 252, 110–124. https://doi.org/10.1016/j.enggeo.2019.03.004
DeConto, R. M., & Pollard, D. (2016). Contribution of Antarctica to past and future sea-level rise. Nature, 531(7596), 591–597. https://doi.
org/10.1038/nature17145
Deegan, L. A., Johnson, D. S., Warren, R. S., Peterson, B. J., Fleeger, J. W., Fagherazzi, S., & Wollheim, W. M. (2012). Coastal eutrophication as
a driver of salt marsh loss. Nature, 490(7420), 388–392. https://doi.org/10.1038/nature11533
Deng, S., Xia, J., & Zhou, M. (2019). Coupled two-dimensional modeling of bed evolution and bank erosion in the Upper JingJiang Reach of
middle Yangtze River. Geomorphology, 344, 10–24. https://doi.org/10.1016/j.geomorph.2019.07.010
Deng, S., Xia, J., Zhou, M., Li, J., & Zhu, Y. (2018). Coupled modeling of bank retreat processes in the Upper Jingjiang Reach, China. Earth
Surface Processes and Landforms, 43(14), 2863–2875. https://doi.org/10.1002/esp.4439
Derksen Hooijberg, M., Angelini, C., Hoogveld, J. R., Lamers, L. P., Borst, A., Smolders, A., et al. (2019). Repetitive desiccation events weaken
a salt marsh mutualism. Journal of Ecology, 107(5), 2415–2426. https://doi.org/10.1111/1365-2745.13178
De Rose, R. C., & Basher, L. R. (2011). Measurement of river bank and cliff erosion from sequential LIDAR and historical aerial photography.
Geomorphology, 126(1–2), 132–147. https://doi.org/10.1016/j.geomorph.2010.10.037
Dong, T. Y., Nittrouer, J. A., Czapiga, M. J., Ma, H., McElroy, B., Il'Icheva, E., et al. (2019). Roles of bank material in setting bankfull hydraulic
geometry as informed by the Selenga River delta, Russia. Water Resources Research, 55(1), 827–846. https://doi.org/10.1029/2017WR021985
Dong, Y., Wu, Y., Qin, W., Guo, Q., Yin, Z., & Duan, X. (2019). The gully erosion rates in the black soil region of northeastern China: Induced
by different processes and indicated by different indexes. Catena, 182, 104146. https://doi.org/10.1016/j.catena.2019.104146

ZHAO ET AL. 43 of 51
Reviews of Geophysics 10.1029/2021RG000761

Dong, Y., Xiong, D., Su, Z., Duan, X., Lu, X., Zhang, S., & Yuan, Y. (2019). The influences of mass failure on the erosion and hydraulic
processes of gully headcuts based on an in situ scouring experiment in dry-hot valley of China. Catena, 176, 14–25. https://doi.org/10.1016/j.
catena.2019.01.004
Donovan, M., Belmont, P., & Sylvester, Z. (2021). Evaluating the relationship between meander-bend curvature, sediment supply, and migration
rates. Journal of Geophysical Research: Earth Surface, 126(3), e2020J–e6058J. https://doi.org/10.1029/2020jf006058
Duan, J. G., & Julien, P. Y. (2005). Numerical simulation of the inception of channel meandering. Earth Surface Processes and Landforms: The
Journal of the British Geomorphological Research Group, 30(9), 1093–1110. https://doi.org/10.1002/esp.1264
Dulal, K. P., Kobayashi, K., Shimizu, Y., & Parker, G. (2010). Numerical computation of free meandering channels with the application of slump
blocks on the outer bends. Journal of Hydro-Environment Research, 3(4), 239–246. https://doi.org/10.1016/j.jher.2009.10.012
Duncan, J. M., Wright, S. G., & Brandon, T. L. (2014). Soil strength and slope stability. John Wiley & Sons.
Dunne, K. B. J., & Jerolmack, D. J. (2020). What sets river width? Science Advances, 6(41), eabc1505. https://doi.org/10.1126/sciadv.abc1505
Dunne, T. (1990). Hydrology, mechanics, and geomorphic implications of erosion by subsurface flow. In C. G. Higgins & D. R. Coates (Eds.),
Groundwater geomorphology: The role of subsurface water in Earth-surface processes and landforms (pp. 1–28). Geological Society of
America Special Paper. 252.
Duró, G., Crosato, A., Kleinhans, M. G., Roelvink, D., & Uijttewaal, W. (2020). Bank erosion processes in regulated navigable rivers. Journal of
Geophysical Research: Earth Surface, 125(7), e2019J–e5441J. https://doi.org/10.1029/2019JF005441
Duró, G., Crosato, A., Kleinhans, M. G., Winkels, T. G., Woolderink, H. A., & Uijttewaal, W. S. (2019). Distinct patterns of bank erosion in a
navigable regulated river. Earth Surface Processes and Landforms, 45(2), 361–374. https://doi.org/10.1002/esp.4736
Duró, G., Crosato, A., & Tassi, P. (2016). Numerical study on river bar response to spatial variations of channel width. Advances in Water
Resources, 93, 21–38. https://doi.org/10.1016/j.advwatres.2015.10.003
Durocher, M. G. (1990). Monitoring spatial variability of forest interception. Hydrological Processes, 4(3), 215–229. https://doi.org/10.1002/
hyp.3360040303
Eke, E., Parker, G., & Shimizu, Y. (2014). Numerical modeling of erosional and depositional bank processes in migrating river bends with
self-formed width: Morphodynamics of bar push and bank pull. Journal of Geophysical Research: Earth Surface, 119(7), 1455–1483. https://
doi.org/10.1002/2013JF003020
Eke, E. C., Czapiga, M. J., Viparelli, E., Shimizu, Y., Imran, J., Sun, T., & Parker, G. (2014). Coevolution of width and sinuosity in meandering
rivers. Journal of Fluid Mechanics, 760, 127–174. https://doi.org/10.1017/jfm.2014.556
Fagherazzi, S., Gabet, E. J., & Furbish, D. J. (2004). The effect of bidirectional flow on tidal channel planforms. Earth Surface Processes and
Landforms, 29(3), 295–309. https://doi.org/10.1002/esp.1016
Fagherazzi, S., Kirwan, M. L., Mudd, S. M., Guntenspergen, G. R., Temmerman, S., D'Alpaos, A., et al. (2012). Numerical models of salt marsh
evolution: Ecological, geomorphic, and climatic factors. Reviews of Geophysics, 50(1), RG1002. https://doi.org/10.1029/2011rg000359
Fagherazzi, S., Mariotti, G., Wiberg, P. L., & McGlathery, K. J. (2013). Marsh collapse does not require sea level rise. Oceanography, 26(3),
70–77. https://doi.org/10.5670/oceanog.2013.47
Faure, Y., Ho, C. C., Chen, R., Le Lay, M., & Blaza, J. (2010). A wave flume experiment for studying erosion mechanism of revetments using
geotextiles. Geotextiles and Geomembranes, 28(4), 360–373. https://doi.org/10.1016/j.geotexmem.2009.11.002
Feagin, R. A., Lozada-Bernard, S. M., Ravens, T. M., Möller, I., Yeager, K. M., & Baird, A. H. (2009). Does vegetation prevent wave erosion
of salt marsh edges? Proceedings of the National Academy of Sciences, 106(25), 10109–10113. https://doi.org/10.1073/pnas.0901297106
Ferrick, M. G., & Gatto, L. W. (2005). Quantifying the effect of a freeze–thaw cycle on soil erosion: Laboratory experiments, Earth Surface
Processes and landforms. The Journal of the British Geomorphological Research Group, 30(10), 1305–1326. https://doi.org/10.1002/esp.1209
Finnegan, N. J., Sklar, L. S., & Fuller, T. K. (2007). Interplay of sediment supply, river incision, and channel morphology revealed by the transient
evolution of an experimental bedrock channel. Journal of Geophysical Research, 112(F3), F03S11. https://doi.org/10.1029/2006jf000569
Finotello, A., D’Alpaos, A., Bogoni, M., Ghinassi, M., & Lanzoni, S. (2020). Remotely-sensed planform morphologies reveal fluvial and tidal
nature of meandering channels. Scientific Reports, 10, 1–13. https://doi.org/10.1038/s41598-019-56992-w
Finotello, A., Lanzoni, S., Ghinassi, M., Marani, M., Rinaldo, A., & D’Alpaos, A. (2018). Field migration rates of tidal meanders recapitulate
fluvial morphodynamics. Proceedings of the National Academy of Sciences, 115(7), 1463–1468. https://doi.org/10.1073/pnas.1711330115
Flemming, H., & Wuertz, S. (2019). Bacteria and archaea on Earth and their abundance in biofilms. Nature Reviews Microbiology, 17(4),
247–260. https://doi.org/10.1038/s41579-019-0158-9
Florsheim, J. L., Mount, J. F., & Chin, A. (2008). Bank erosion as a desirable attribute of rivers. BioScience, 58(6), 519–529. https://doi.
org/10.1641/B580608
Fox, G. A., Chu-Agor, M. L. M., & Wilson, G. V. (2007). Erosion of noncohesive sediment by ground water seepage: Lysimeter experiments and
stability modeling. Soil Science Society of America Journal, 71(6), 1822–1830. https://doi.org/10.2136/sssaj2007.0090
Fox, G. A., & Felice, R. G. (2014). Bank undercutting and tension failure by groundwater seepage: Predicting failure mechanisms. Earth Surface
Processes and Landforms, 39(6), 758–765. https://doi.org/10.1002/esp.3481
Fox, G. A., & Wilson, G. V. (2010). The role of subsurface flow in hillslope and stream bank erosion: A review. Soil Science Society of America
Journal, 74(3), 717–733. https://doi.org/10.2136/sssaj2009.0319
Fox, G. A., Wilson, G. V., Periketi, R. K., & Cullum, R. F. (2006). Sediment transport model for seepage erosion of streambank sediment. Journal
of Hydrologic Engineering, 11(6), 603–611. https://doi.org/10.1061/(ASCE)1084-0699(2006)11:6(603)
Francalanci, S., Bendoni, M., Rinaldi, M., & Solari, L. (2013). Ecomorphodynamic evolution of salt marshes: Experimental observations of bank
retreat processes. Geomorphology, 195(Supplement C), 53–65. https://doi.org/10.1016/j.geomorph.2013.04.026
Francalanci, S., Lanzoni, S., Solari, L., & Papanicolaou, A. N. (2020). Equilibrium cross section of river channels with cohesive erodible banks.
Journal of Geophysical Research: Earth Surface, 125(1), 1–20. https://doi.org/10.1029/2019JF005286
Frascati, A., & Lanzoni, S. (2009). Morphodynamic regime and long-term evolution of meandering rivers. Journal of Geophysical Research,
114(2), F02002. https://doi.org/10.1029/2008JF001101
Fredlund, D. G., & Rahardjo, H. (1993). Soil mechanics for unsaturated soils. John Wiley & Sons.
Fredsøe, J. (1978). Meandering and braiding of rivers. Journal of Fluid Mechanics, 84(4), 609–624. https://doi.org/10.1017/s0022112078000373
Fryirs, K. A., & Brierley, G. J. (2012). Geomorphic analysis of river systems: An approach to reading the landscape. John Wiley & Sons.
Fuller, T. K., Gran, K. B., Sklar, L. S., & Paola, C. (2016). Lateral erosion in an experimental bedrock channel: The influence of bed roughness
on erosion by bed load impacts. Journal of Geophysical Research: Earth Surface, 121(5), 1084–1105. https://doi.org/10.1002/2015jf003728
Gabel, F., Lorenz, S., & Stoll, S. (2017). Effects of ship-induced waves on aquatic ecosystems. The Science of the Total Environment, 601,
926–939. https://doi.org/10.1016/j.scitotenv.2017.05.206
Gabet, E. J. (1998). Lateral migration and bank erosion in a saltmarsh tidal channel in San Francisco Bay, California. Estuaries and Coasts, 21(4),
745–753. https://doi.org/10.2307/1353278

ZHAO ET AL. 44 of 51
Reviews of Geophysics 10.1029/2021RG000761

Galy, V., Peucker-Ehrenbrink, B., & Eglinton, T. (2015). Global carbon export from the terrestrial biosphere controlled by erosion. Nature,
521(7551), 204–207. https://doi.org/10.1038/nature14400
Gaskin, S. J., Pieterse, J., Shafie, A. A., & Lepage, S. (2003). Erosion of undisturbed clay samples from the banks of the St. Lawrence River.
Canadian Journal of Civil Engineering, 30(3), 585–595. https://doi.org/10.1139/l03-008
Geng, L., D'Alpaos, A., Sgarabotto, A., Gong, Z., & Lanzoni, S. (2021). Intertwined eco-morphodynamic evolution of Salt marshes and emerging
tidal channel networks. Water Resources Research, 57(11), e2021W–e30840W. https://doi.org/10.1029/2021wr030840
Geng, L., Gong, Z., Zhou, Z., Lanzoni, S., & D'Alpaos, A. (2019). Assessing the relative contributions of the flood tide and the ebb tide to tidal
channel network dynamics. Earth Surface Processes and Landforms, 45(1), 237–250. https://doi.org/10.1002/esp.4727
Gibling, M. R., & Davies, N. S. (2012). Palaeozoic landscapes shaped by plant evolution. Nature Geoscience, 5(2), 99–105. https://doi.org/
10.1038/ngeo1376
Ginsberg, S. S., & Perillo, G. M. E. (1990). Channel bank recession in the Bahía Blanca estuary, Argentina. Journal of Coastal Research, 6(4),
999–1009.
Golombek, N. Y., Scheingross, J. S., Repasch, M. N., Hovius, N., Menges, J., Sachse, D., et al. (2021). Fluvial organic carbon composition regu-
lated by seasonal variability in lowland river migration and water discharge. Geophysical Research Letters, 48(24), e2021G–e93416G. https://
doi.org/10.1029/2021gl093416
Gong, Z., Zhao, K., Zhang, C., Dai, W., Coco, G., & Zhou, Z. (2018). The role of bank collapse on tidal creek ontogeny: A novel process-based
model for bank retreat. Geomorphology, 311, 13–26. https://doi.org/10.1016/j.geomorph.2018.03.016
Griffiths, G. A. (1979). Recent sedimentation history of the Waimakariri River, New Zealand. Journal of Hydrology, 18(1), 6–28.
Güneralp, İ., & Rhoads, B. L. (2011). Influence of floodplain erosional heterogeneity on planform complexity of meandering rivers. Geophysical
Research Letters, 38(14), L14401. https://doi.org/10.1029/2011gl048134
Guo, L., Xu, F., Van Der Wegen, M., Townend, I., Wang, Z. B., & He, Q. (2021). Morphodynamic adaptation of a tidal basin to centennial
sea-level rise: The importance of lateral expansion. Continental Shelf Research, 226, 104494. https://doi.org/10.1016/j.csr.2021.104494
Hackney, C., Best, J., Leyland, J., Darby, S. E., Parsons, D., Aalto, R., & Nicholas, A. (2015). Modulation of outer bank erosion by slump blocks:
Disentangling the protective and destructive role of failed material on the three-dimensional flow structure. Geophysical Research Letters,
42(24), 10–663. https://doi.org/10.1002/2015GL066481
Hackney, C. R., Darby, S. E., Parsons, D. R., Leyland, J., Best, J. L., Aalto, R., et al. (2020). River bank instability from unsustainable sand mining
in the lower Mekong River. Nature Sustainability, 3(3), 1–9. https://doi.org/10.1038/s41893-019-0455-3
Han, M., & Brierley, G. (2020). Channel geomorphology and riparian vegetation interactions along four anabranching reaches of the Upper
Yellow River. Progress in Physical Geography: Earth and Environment, 44(6), 898–922. https://doi.org/10.1177/0309133320938768
Hancock, G. S., & Anderson, R. S. (2002). Numerical modeling of fluvial strath-terrace formation in response to oscillating climate. The Geolog-
ical Society of America Bulletin, 114(9), 1131–1142. https://doi.org/10.1130/0016-7606(2002)114<1131:nmofst>2.0.co;2
Hanson, G. J., & Simon, A. (2001). Erodibility of cohesive streambeds in the loess area of the midwestern USA. Hydrological Processes, 15(1),
23–38. https://doi.org/10.1002/hyp.149
Hartshorn, K., Hovius, N., Dade, W. B., & Slingerland, R. L. (2002). Climate-driven bedrock incision in an active mountain belt. Science,
297(5589), 2036–2038. https://doi.org/10.1126/science.1075078
Harvey, G. L., Henshaw, A. J., Brasington, J., & England, J. (2019). Burrowing invasive species: An unquantified erosion risk at the aquatic-
terrestrial Interface. Reviews of Geophysics, 57(3), 1018–1036. https://doi.org/10.1029/2018RG000635
Hasegawa, K. (1977). Computer simulation of the gradual migration of meandering channels. Proceedings of the Hokkaido Branch, Japan Society
of Civil Engineering (pp. 197–202).
Hasegawa, K. (1989). Universal bank erosion coefficient for meandering rivers. Journal of Hydraulic Engineering, 115(6), 744–765. https://doi.
org/10.1061/(ASCE)0733-9429(1989)115:6(744)
Henshaw, A. J., Thorne, C. R., & Clifford, N. J. (2013). Identifying causes and controls of river bank erosion in a British upland catchment.
Catena, 100, 107–119. https://doi.org/10.1016/j.catena.2012.07.015
Hickin, E. J., & Nanson, G. C. (1984). Lateral migration rates of river bends. Journal of Hydraulic Engineering, 110(11), 1557–1567. https://doi.
org/10.1061/(ASCE)0733-9429(1984)110:11(1557)
Hooke, J. M. (1979). An analysis of the processes of river bank erosion. Journal of Hydrology, 42(1), 39–62. https://doi.org/10.1016/0022-
1694(79)90005-2
Hooke, J. M. (1980). Magnitude and distribution of rates of river bank erosion. Earth Surface Processes, 5(2), 143–157. https://doi.org/10.1002/
esp.3760050205
Houser, C. (2010). Relative importance of vessel-generated and wind waves to salt marsh erosion in a restricted fetch environment. Journal of
Coastal Research, 262, 230–240. https://doi.org/10.2112/08-1084.1
Houttuijn Bloemendaal, L. J., FitzGerald, D. M., Hughes, Z. J., Novak, A. B., & Phippen, P. (2021). What controls marsh edge erosion? Geomor-
phology, 386, 107745. https://doi.org/10.1016/j.geomorph.2021.107745
Howard, A. D. (2009). How to make a meandering river. Proceedings of the National Academy of Sciences, 106(41), 17245–17246. https://doi.
org/10.1073/pnas.0910005106
Howard, A. D., & Knutson, T. R. (1984). Sufficient conditions for river meandering: A simulation approach. Water Resources Research, 20(11),
1659–1667. https://doi.org/10.1029/wr020i011p01659
Howard, A. D., & McLane, C. F., III. (1988). Erosion of cohesionless sediment by groundwater seepage. Water Resources Research, 24(10),
1659–1674. https://doi.org/10.1029/WR024i010p01659
Hua, X., Huang, H., Wang, Y., Lan, Y., Zhao, K., & Chen, D. (2019). Abnormal ETM in the North Passage of the Changjiang River estuary:
Observations in the wet and dry seasons of 2016, estuarine. Coastal and Shelf Science, 227, 106334. https://doi.org/10.1016/j.ecss.2019.106334
Hughes, Z. J. (2012). Tidal channels on tidal flats and marshes. In Principles of tidal sedimentology (pp. 269–300). Springer.
Ielpi, A., Lapôtre, M. G. A., Gibling, M. R., & Boyce, C. K. (2022). The impact of vegetation on meandering rivers. Nature Reviews Earth &
Environment, 3(3), 165–178. https://doi.org/10.1038/s43017-021-00249-6
Ikeda, S., Parker, G., & Sawai, K. (1981). Bend theory of river meanders. Part 1. Linear development. Journal of Fluid Mechanics, 112, 363–377.
https://doi.org/10.1017/S0022112081000451
Inoue, T., Mishra, J., & Parker, G. (2021). Numerical simulations of meanders migrating laterally as they incise into bedrock. Journal of Geophys-
ical Research: Earth Surface, 126(5), e2020J–e5645J. https://doi.org/10.1029/2020jf005645
Istanbulluoglu, E., Bras, R. L., Flores Cervantes, H., & Tucker, G. E. (2005). Implications of bank failures and fluvial erosion for gully develop-
ment: Field observations and modeling. Journal of Geophysical Research, 110(F1), F01014. https://doi.org/10.1029/2004JF000145
Jang, C., & Shimizu, Y. (2005). Numerical simulation of relatively wide, shallow channels with erodible banks. Journal of Hydraulic Engineer-
ing, 131(7), 565–575. https://doi.org/10.1061/(ASCE)0733-9429(2005)131:7(565)

ZHAO ET AL. 45 of 51
Reviews of Geophysics 10.1029/2021RG000761

Ji, F., Liu, C., Shi, Y., Feng, W., & Wang, D. (2019). Characteristics and parameters of bank collapse in coarse-grained-material reservoirs based
on back analysis and long sequence monitoring. Geomorphology, 333, 92–104. https://doi.org/10.1016/j.geomorph.2019.02.018
Ji, F., Shi, Y., Zhou, H., Liu, H., & Liao, Y. (2017). Experimental research on the effect of slope morphology on bank collapse in mountain reser-
voir. Natural Hazards, 86(1), 165–181. https://doi.org/10.1007/s11069-016-2679-0
Jia, D., Shao, X., Wang, H., & Zhou, G. (2010). Three-dimensional modeling of bank erosion and morphological changes in the Shishou bend of
the middle Yangtze River. Advances in Water Resources, 33(3), 348–360. https://doi.org/10.1016/j.advwatres.2010.01.002
Kang, Y., Ding, X., Xu, F., Zhang, C., & Ge, X. (2017). Topographic mapping on large-scale tidal flats with an iterative approach on the waterline
method, Estuarine. Coastal and Shelf Science, 190, 11–22. https://doi.org/10.1016/j.ecss.2017.03.024
Karmaker, T., & Dutta, S. (2013). Modeling seepage erosion and bank retreat in a composite river bank. Journal of Hydrology, 476, 178–187.
https://doi.org/10.1016/j.jhydrol.2012.10.032
Kearney, W. S., & Fagherazzi, S. (2016). Salt marsh vegetation promotes efficient tidal channel networks. Nature Communications, 7(1), 1–7.
https://doi.org/10.1038/ncomms12287
Khatun, S., Ghosh, A., & Sen, D. (2019). An experimental investigation on effect of drawdown rate and drawdown ratios on stability of cohesion-
less river bank and evaluation of factor of safety by total strength reduction method. International Journal of River Basin Management, 17(3),
289–299. https://doi.org/10.1080/15715124.2018.1498856
Kimiaghalam, N., Goharrokhi, M., Clark, S. P., & Ahmari, H. (2015). A comprehensive fluvial geomorphology study of riverbank erosion on the
Red River in Winnipeg, Manitoba, Canada. Journal of Hydrology, 529, 1488–1498. https://doi.org/10.1016/j.jhydrol.2015.08.033
Kirwan, M. L., Guntenspergen, G. R., D'Alpaos, A., Morris, J. T., Mudd, S. M., & Temmerman, S. (2010). Limits on the adaptability of coastal
marshes to rising sea level. Geophysical Research Letters, 37(23). https://doi.org/10.1029/2010gl045489
Kirwan, M. L., & Mudd, S. M. (2012). Response of salt-marsh carbon accumulation to climate change. Nature, 489(7417), 550–553. https://doi.
org/10.1038/nature11440
Kirwan, M. L., & Murray, A. B. (2007). A coupled geomorphic and ecological model of tidal marsh evolution. Proceedings of the National
Academy of Sciences, 104(15), 6118–6122. https://doi.org/10.1073/pnas.0700958104
Kitanidis, P. K., & Kennedy, J. F. (1984). Secondary current and river-meander formation. Journal of Fluid Mechanics, 144, 217–229. https://
doi.org/10.1017/S0022112084001580
Klavon, K., Fox, G., Guertault, L., Langendoen, E., Enlow, H., Miller, R., & Khanal, A. (2017). Evaluating a process-based model for use in
streambank stabilization: Insights on the Bank Stability and Toe Erosion Model (BSTEM). Earth Surface Processes and Landforms, 42(1),
191–213. https://doi.org/10.1002/esp.4073
Kleinhans, M. G., Schuurman, F., Bakx, W., & Markies, H. (2009). Meandering channel dynamics in highly cohesive sediment on an intertidal
mud flat in the Westerschelde estuary, The Netherlands. Geomorphology, 105(3), 261–276. https://doi.org/10.1016/j.geomorph.2008.10.005
Kleinhans, M. G., van der Vegt, M., van Scheltinga, R. T., Baar, A. W., & Markies, H. (2012). Turning the tide: Experimental creation of tidal
channel networks and ebb deltas. Netherlands Journal of Geosciences, 91(3), 311–323. https://doi.org/10.1017/S0016774600000469
Kronvang, B., Andersen, H. E., Larsen, S. E., & Audet, J. (2013). Importance of bank erosion for sediment input, storage and export at the catch-
ment scale. Journal of Soils and Sediments, 13(1), 230–241. https://doi.org/10.1007/s11368-012-0597-7
Krzeminska, D., Kerkhof, T., Skaalsveen, K., & Stolte, J. (2019). Effect of riparian vegetation on stream bank stability in small agricultural
catchments. Catena, 172, 87–96. https://doi.org/10.1016/j.catena.2018.08.014
Lagasse, P. F., Zevenbergen, L. W., Spitz, W. J., & Thorne, C. R. (2004). Methodology for predicting channel migration. National Cooperative
Highway Research Program, Web-Only Document, 67, 162.
Lague, D. (2014). The stream power river incision model: Evidence, theory and beyond. Earth Surface Processes and Landforms, 39(1), 38–61.
https://doi.org/10.1002/esp.3462
Lai, Y. G., Thomas, R. E., Ozeren, Y., Simon, A., Greimann, B. P., & Wu, K. (2015). Modeling of multilayer cohesive bank erosion with a coupled
bank stability and mobile-bed model. Geomorphology, 243, 116–129. https://doi.org/10.1016/j.geomorph.2014.07.017
Langendoen, E. J. (2000). Concepts: Conservational channel evolution and pollutant transport system. USDA-ARS National Sedimentation
Laboratory.
Langendoen, E. J., Mendoza, A., Abad, J. D., Tassi, P., Wang, D., Ata, R., et al. (2016). Improved numerical modeling of morphodynamics of
rivers with steep banks. Advances in Water Resources, 93, 4–14. https://doi.org/10.1016/j.advwatres.2015.04.002
Langendoen, E. J., & Simon, A. (2008). Modeling the evolution of incised streams. II: Streambank erosion. Journal of Hydraulic Engineering,
134(7), 905–915. https://doi.org/10.1061/(ASCE)0733-9429(2008)134:7(905)
Langston, A. L., & Tucker, G. E. (2018). Developing and exploring a theory for the lateral erosion of bedrock channels for use in landscape
evolution models. Earth Surface Dynamics, 6(1), 1–27. https://doi.org/10.5194/esurf-6-1-2018
Lanzoni, S., & D'Alpaos, A. (2015). On funneling of tidal channels. Journal of Geophysical Research: Earth Surface, 120(3), 433–452. https://
doi.org/10.1002/2014jf003203
Lanzoni, S., & Seminara, G. (2006). On the nature of meander instability. Journal of Geophysical Research, 111(F4), F04006. https://doi.
org/10.1029/2005JF000416
Larsen, L. G., Harvey, J. W., & Crimaldi, J. P. (2007). A delicate balance: Ecohydrological feedbacks governing landscape morphology in a lotic
peatland. Ecological Monographs, 77(4), 591–614. https://doi.org/10.1890/06-1267.1
Lawler, D. M. (1986). River bank erosion and the influence of frost: A statistical examination (pp. 227–242). Transactions of the Institute of
British Geographers. https://doi.org/10.2307/622008
Lawler, D. M. (1992). Process dominance in bank erosion systems. Paper presented at Lowland Floodplain Rivers Geomorphological Perspectives.
Lawler, D. M. (1993a). The measurement of river bank erosion and lateral channel change: A review. Earth Surface Processes and Landforms,
18(9), 777–821. https://doi.org/10.1002/esp.3290180905
Lawler, D. M. (1993b). Needle ice processes and sediment mobilization on river banks: The River Ilston, West Glamorgan, UK. Journal of
Hydrology, 150(1), 81–114. https://doi.org/10.1016/0022-1694(93)90157-5
Lawler, D. M., Grove, J. R., Couperthwaite, J. S., & Leeks, G. (1999). Downstream change in river bank erosion rates in the Swale–Ouse system, north-
ern England. Hydrological Processes, 13(7), 977–992. https://doi.org/10.1002/(sici)1099-1085(199905)13:7<977::aid-hyp785>3.0.co;2-5
Lesser, G. R., Roelvink, J. A., van Kester, J. A. T. M., & Stelling, G. S. (2004). Development and validation of a three-dimensional morphological
model. Coastal Engineering, 51(8–9), 883–915. https://doi.org/10.1016/j.coastaleng.2004.07.014
Li, C., Holden, J., & Grayson, R. (2018). Effects of needle ice on peat erosion processes during overland flow events. Journal of Geophysical
Research: Earth Surface, 123(9), 2107–2122. https://doi.org/10.1029/2017jf004508
Li, T., Fuller, T. K., Sklar, L. S., Gran, K. B., & Venditti, J. G. (2020). A mechanistic model for lateral erosion of bedrock channel banks by
bedload particle impacts. Journal of Geophysical Research: Earth Surface, 125(6), e2019J–e5509J. https://doi.org/10.1029/2019jf005509

ZHAO ET AL. 46 of 51
Reviews of Geophysics 10.1029/2021RG000761

Li, Z., Zhang, Y., Zhu, Q., He, Y., & Yao, W. (2015). Assessment of bank gully development and vegetation coverage on the Chinese Loess
Plateau. Geomorphology, 228, 462–469. https://doi.org/10.1016/j.geomorph.2014.10.005
Limaye, A. B., & Lamb, M. P. (2014). Numerical simulations of bedrock valley evolution by meandering rivers with variable bank material.
Journal of Geophysical Research: Earth Surface, 119(4), 927–950. https://doi.org/10.1002/2013jf002997
Lindow, N., Fox, G. A., & Evans, R. O. (2009). Seepage erosion in layered stream bank material. Earth Surface Processes and Landforms, 12(34),
1693–1701. https://doi.org/10.1002/esp.1874
Liu, C., Liu, A., He, Y., & Chen, Y. (2021). Migration rate of river bends estimated by tree ring analysis for a meandering river in the source region
of the Yellow River. International Journal of Sediment Research, 36(5), 593–601. https://doi.org/10.1016/j.ijsrc.2021.04.001
Lopez Dubon, S., & Lanzoni, S. (2019). Meandering evolution and width variations: A physics-statistics-based modeling approach. Water
Resources Research, 55(1), 76–94. https://doi.org/10.1029/2018WR023639
Malatesta, L. C., Prancevic, J. P., & Avouac, J. P. (2017). Autogenic entrenchment patterns and terraces due to coupling with lateral erosion in
incising alluvial channels. Journal of Geophysical Research: Earth Surface, 122(1), 335–355. https://doi.org/10.1002/2015jf003797
Marani, M., D’Alpaos, A., Lanzoni, S., & Santalucia, M. (2011). Understanding and predicting wave erosion of marsh edges. Geophysical
Research Letters, 38(21), L21401. https://doi.org/10.1029/2011gl048995
Marani, M., Lanzoni, S., Zandolin, D., Seminara, G., & Rinaldo, A. (2002). Tidal meanders. Water Resources Research, 38(11), 7-1–7-14. https://
doi.org/10.1029/2001WR000404
Mariotti, G., & Fagherazzi, S. (2010). A numerical model for the coupled long-term evolution of salt marshes and tidal flats. Journal of Geophys-
ical Research, 115(F1), F01004. https://doi.org/10.1029/2009jf001326
Mariotti, G., Kearney, W. S., & Fagherazzi, S. (2016). Soil creep in salt marshes. Geology, 44(6), 459–462. https://doi.org/10.1130/G37708.1
Mariotti, G., Kearney, W. S., & Fagherazzi, S. (2019). Soil creep in a mesotidal salt marsh channel bank: Fast, seasonal, and water table mediated.
Geomorphology, 334, 126–137. https://doi.org/10.1016/j.geomorph.2019.03.001
Marron, D. C. (1992). Floodplain storage of mine tailings in the Belle Fourche river system: A sediment budget approach. Earth Surface Processes
and Landforms, 17(7), 675–685. https://doi.org/10.1002/esp.3290170704
Mason, J., & Mohrig, D. (2019). Differential bank migration and the maintenance of channel width in meandering river bends. Geology, 47(12),
1136–1140. https://doi.org/10.1130/G46651.1
Masoodi, A., Majdzadeh Tabatabai, M. R., Noorzad, A., & Samadi, A. (2017). Effects of soil physico-chemical properties on stream bank erosion
induced by seepage in northeastern Iran. Hydrological Sciences Journal, 62(16), 2597–2613. https://doi.org/10.1080/02626667.2017.1403030
Masoodi, A., Majdzadeh Tabatabai, M. R., Noorzad, A., & Samadi, A. (2019). Riverbank stability under the influence of soil dispersion phenom-
enon. Journal of Hydrologic Engineering, 24(3), 5019001. https://doi.org/10.1061/(ASCE)HE.1943-5584.0001756
Masoodi, A., Noorzad, A., Majdzadeh Tabatabai, M. R., & Samadi, A. (2018). Application of short-range photogrammetry for monitoring
seepage erosion of riverbank by laboratory experiments. Journal of Hydrology, 558, 380–391. https://doi.org/10.1016/j.jhydrol.2018.01.051
McKew, B. A., Taylor, J. D., McGenity, T. J., & Underwood, G. J. (2011). Resistance and resilience of benthic biofilm communities from a
temperate saltmarsh to desiccation and rewetting. The ISME Journal, 5(1), 30–41. https://doi.org/10.1038/ismej.2010.91
Micheli, E. R., & Kirchner, J. W. (2002). Effects of wet meadow riparian vegetation on streambank erosion. 2. Measurements of vegetated bank
strength and consequences for failure mechanics. Earth Surface Processes and Landforms, 27(7), 687–697. https://doi.org/10.1002/esp.340
Midgley, T. L., Fox, G. A., & Heeren, D. M. (2012). Evaluation of the Bank Stability and Toe Erosion Model (BSTEM) for predicting lateral
retreat on composite streambanks. Geomorphology, 145–146, 107–114. https://doi.org/10.1016/j.geomorph.2011.12.044
Midgley, T. L., Fox, G. A., Wilson, G. V., Heeren, D. M., Langendoen, E. J., & Simon, A. (2013). Seepage-induced streambank erosion and
instability: In situ constant-head experiments. Journal of Hydrologic Engineering, 18(10), 1200–1210. https://doi.org/10.1061/(ASCE)
HE.1943-5584.0000685
Mishra, J., Inoue, T., Shimizu, Y., Sumner, T., & Nelson, J. M. (2018). Consequences of abrading bed load on vertical and lateral bedrock erosion in
a curved experimental channel. Journal of Geophysical Research: Earth Surface, 123(12), 3147–3161. https://doi.org/10.1029/2017jf004387
Mittal, R., & Iaccarino, G. (2005). Immersed boundary methods. Annual Review of Fluid Mechanics, 37(1), 239–261. https://doi.org/10.1146/
annurev.fluid.37.061903.175743
Montgomery, D. R. (2004). Observations on the role of lithology in strath terrace formation and bedrock channel width. American Journal of
Science, 304(5), 454–476. https://doi.org/10.2475/ajs.304.5.454
Mosselman, E. (1995). A review of mathematical models of river planform changes. Earth Surface Processes and Landforms, 20(7), 661–670.
https://doi.org/10.1002/esp.3290200708
Mosselman, E. (1998). Morphological modelling of rivers with erodible banks. Hydrological Processes, 12(8), 1357–1370. https://doi.org/
10.1002/(SICI)1099-1085(19980630)12:8<1357::AID-HYP619>3.0.CO;2-7
Motta, D., Langendoen, E. J., Abad, J. D., & García, M. H. (2014). Modification of meander migration by bank failures. Journal of Geophysical
Research: Earth Surface, 119(5), 1026–1042. https://doi.org/10.1002/2013JF002952
Murray, A. B., & Paola, C. (1994). A cellular model of braided rivers. Nature, 371(6492), 54–57. https://doi.org/10.1038/371054a0
Nagata, N., Hosoda, T., & Muramoto, Y. (2000). Numerical analysis of river channel processes with bank erosion. Journal of Hydraulic Engi-
neering, 126(4), 243–252. https://doi.org/10.1061/(ASCE)0733-9429(2000)126:4(243)
Nanson, G. C., & Hickin, E. J. (1983). Channel migration and incision on the Beatton River. Journal of Hydraulic Engineering, 109(3), 327–337.
https://doi.org/10.1061/(ASCE)0733-9429(1983)109:3(327)
Nanson, G. C., & Hickin, E. J. (1986). A statistical analysis of bank erosion and channel migration in western Canada. The Geological Society of
America Bulletin, 97(4), 497–504. https://doi.org/10.1130/0016-7606(1986)97<497:ASAOBE>2.0.CO;2
Nanson, G. C., Von Krusenstierna, A., Bryant, E. A., & Renilson, M. R. (1994). Experimental measurements of river-bank erosion caused by
boat-generated waves on the Gordon River, Tasmania. Regulated Rivers: Research & Management, 9(1), 1–14. https://doi.org/10.1002/rrr.
3450090102
Nardi, L., Rinaldi, M., & Solari, L. (2012). An experimental investigation on mass failures occurring in a riverbank composed of sandy gravel.
Geomorphology, 163–164, 56–69. https://doi.org/10.1016/j.geomorph.2011.08.006
Ni, W., Wang, Y. P., Symonds, A. M., & Collins, M. B. (2014). Intertidal flat development in response to controlled embankment retreat: Freiston
Shore, The Wash, UK. Marine Geology, 355, 260–273. https://doi.org/10.1016/j.margeo.2014.06.001
Odgaard, A. J. (1989). River-meander model. I: Development. Journal of Hydraulic Engineering, 115(11), 1433–1450. https://doi.org/10.1061/
(ASCE)0733-9429(1989)115:11(1433)
Okura, Y., Kitahara, H., Ochiai, H., Sammori, T., & Kawanami, A. (2002). Landslide fluidization process by flume experiments. Engineering
Geology, 66(1–2), 65–78. https://doi.org/10.1016/S0013-7952(02)00032-7
Osman, A. M., & Thorne, C. R. (1988). Riverbank stability analysis. I: Theory. Journal of Hydraulic Engineering, 114(2), 134–150. https://doi.
org/10.1061/(ASCE)0733-9429(1988)114:2(134)

ZHAO ET AL. 47 of 51
Reviews of Geophysics 10.1029/2021RG000761

Paola, C., Heller, P. L., & Angevine, C. L. (1992). The large-scale dynamics of grain-size variation in alluvial basins, 1: Theory. Basin Research,
4(2), 73–90. https://doi.org/10.1111/j.1365-2117.1992.tb00145.x
Parker, G., Paola, C., Whipple, K. X., & Mohrig, D. (1998). Alluvial fans formed by channelized fluvial and sheet flow. I: Theory. Journal of
Hydraulic Engineering, 124(10), 985–995. https://doi.org/10.1061/(asce)0733-9429(1998)124:10(985)
Parker, G., Shimizu, Y., Wilkerson, G. V., Eke, E. C., Abad, J. D., Lauer, J. W., et al. (2011). A new framework for modeling the migration of
meandering rivers. Earth Surface Processes and Landforms, 36(1), 70–86. https://doi.org/10.1002/esp.2113
Partheniades, E. (1965). Erosion and deposition of cohesive soils. Journal of the Hydraulics Division, 91(1), 105–139. https://doi.org/10.1061/
jyceaj.0001165
Patsinghasanee, S., Kimura, I., Shimizu, Y., & Nabi, M. (2015). Cantilever failure investigations for cohesive riverbanks. Proceedings of the
Institution of Civil Engineers-Water Management (pp. 93–108). Thomas Telford Ltd. https://doi.org/10.1680/jwama.15.00033
Patsinghasanee, S., Kimura, I., Shimizu, Y., & Nabi, M. (2018). Experiments and modelling of cantilever failures for cohesive riverbanks. Journal
of Hydraulic Research, 56(1), 76–95. https://doi.org/10.1080/00221686.2017.1300194
Patsinghasanee, S., Kimura, I., Shimizu, Y., Nabi, M., & Chub-Uppakarn, T. (2017). Coupled studies of fluvial erosion and cantilever failure for
cohesive riverbanks: Case studies in the experimental flumes and U-Tapao River. Journal of Hydro-Environment Research, 16(Supplement
C), 13–26. https://doi.org/10.1016/j.jher.2017.04.002
Piégay, H., Cuaz, M., Javelle, E., & Mandier, P. (1997). Bank erosion management based on geomorphological, ecological and economic criteria
on the Galaure River, France, Regulated Rivers. Research Management, 13(5), 433–448. https://doi.org/10.1002/(SICI)1099-1646(199709/
10)13:5<433::AID-RRR467>3.0.CO;2-L
Piégay, H., Darby, S. E., Mosselman, E., & Surian, N. (2005). A review of techniques available for delimiting the erodible river corridor: A
sustainable approach to managing bank erosion. River Research and Applications, 21(7), 773–789. https://doi.org/10.1002/rra.881
Pizzuto, J. (2009). An empirical model of event scale cohesive bank profile evolution. Earth Surface Processes and Landforms, 34(9), 1234–
1244. https://doi.org/10.1002/esp.1808
Pizzuto, J., O'Neal, M., & Stotts, S. (2010). On the retreat of forested, cohesive riverbanks. Geomorphology, 116(3–4), 341–352. https://doi.
org/10.1016/j.geomorph.2009.11.008
Pizzuto, J. E. (1984). Bank erodibility of shallow sandbed streams. Earth Surface Processes and Landforms, 9(2), 113–124. https://doi.org/
10.1002/esp.3290090203
Pizzuto, J. E. (1990). Numerical simulation of gravel river widening. Water Resources Research, 26(9), 1971–1980. https://doi.org/10.1029/
WR026i009p01971
Pizzuto, J. E., & Meckelnburg, T. S. (1989). Evaluation of a linear bank erosion equation. Water Resources Research, 25(5), 1005–1013. https://
doi.org/10.1029/WR025i005p01005
Poesen, J., Nachtergaele, J., Verstraeten, G., & Valentin, C. (2003). Gully erosion and environmental change: Importance and research needs.
Catena, 50(2–4), 91–133. https://doi.org/10.1016/S0341-8162(02)00143-1
Poesen, J. W., Vandaele, K., & Van Wesemael, B. (1996). Contribution of gully erosion to sediment production on cultivated lands and range-
lands. IAHS Publications-Series of Proceedings and Reports-Intern Assoc Hydrological Sciences (Vol. 236, pp. 251–266).
Pollen Bankhead, N., & Simon, A. (2009). Enhanced application of root-reinforcement algorithms for bank-stability modeling. Earth Surface
Processes and Landforms, 34(4), 471–480. https://doi.org/10.1002/esp.1690
Pollen-Bankhead, N., & Simon, A. (2010). Hydrologic and hydraulic effects of riparian root networks on streambank stability: Is mechanical
root-reinforcement the whole story? Geomorphology, 116(3–4), 353–362. https://doi.org/10.1016/j.geomorph.2009.11.013
Prosser, I. P., Hughes, A. O., & Rutherfurd, I. D. (2000). Bank erosion of an incised upland channel by subaerial processes: Tasmania, Australia,
Earth Surface Processes and Landforms. The Journal of the British Geomorphological Research Group, 25(10), 1085–1101. https://doi.
org/10.1002/1096-9837(200009)25:10<1085::aid-esp118>3.0.co;2-k
Qin, C., Zheng, F., Wells, R. R., Xu, X., Wang, B., & Zhong, K. (2018). A laboratory study of channel sidewall expansion in upland concentrated
flows. Soil and Tillage Research, 178, 22–31. https://doi.org/10.1016/j.still.2017.12.008
Quaresma, V. D. S., Amos, C. L., & Bastos, A. C. (2007). The influence of articulated and disarticulated cockle shells on the erosion of a cohesive
bed. Journal of Coastal Research, 236, 1443–1451. https://doi.org/10.2112/05-0449.1
Rahmati, O., Tahmasebipour, N., Haghizadeh, A., Pourghasemi, H. R., & Feizizadeh, B. (2017). Evaluation of different machine learning models
for predicting and mapping the susceptibility of gully erosion. Geomorphology, 298, 118–137. https://doi.org/10.1016/j.geomorph.2017.09.006
Rajaram, G., & Erbach, D. C. (1999). Effect of wetting and drying on soil physical properties. Journal of Terramechanics, 36(1), 39–49. https://
doi.org/10.1016/S0022-4898(98)00030-5
Reneau, S. L., Drakos, P. G., Katzman, D., Malmon, D. V., McDonald, E. V., & Ryti, R. T. (2004). Geomorphic controls on contaminant distri-
bution along an ephemeral stream. Earth Surface Processes and Landforms: The Journal of the British Geomorphological Research Group,
29(10), 1209–1223. https://doi.org/10.1002/esp.1085
Rengers, F. K., & Tucker, G. E. (2014). Analysis and modeling of gully headcut dynamics, North American high plains. Journal of Geophysical
Research: Earth Surface, 119(5), 983–1003. https://doi.org/10.1002/2013JF002962
Rengers, F. K., & Tucker, G. E. (2015). The evolution of gully headcut morphology: A case study using terrestrial laser scanning and hydrological
monitoring. Earth Surface Processes and Landforms, 40(10), 1304–1317. https://doi.org/10.1002/esp.3721
Repasch, M., Scheingross, J. S., Hovius, N., Lupker, M., Wittmann, H., Haghipour, N., et al. (2021). Fluvial organic carbon cycling regulated by
sediment transit time and mineral protection. Nature Geoscience, 14(11), 842–848. https://doi.org/10.1038/s41561-021-00845-7
Repasch, M., Scheingross, J. S., Hovius, N., Vieth-Hillebrand, A., Mueller, C. W., Höschen, C., et al. (2022). River organic carbon fluxes
modulated by hydrodynamic sorting of particulate organic matter. Geophysical Research Letters, 49(3), e2021G–e96343G. https://doi.
org/10.1029/2021gl096343
Rieke-Zapp, D. H., & Nichols, M. H. (2011). Headcut retreat in a semiarid watershed in the southwestern United States since 1935. Catena, 87(1),
1–10. https://doi.org/10.1016/j.catena.2011.04.002
Rinaldi, M., Casagli, N., Dapporto, S., & Gargini, A. (2004). Monitoring and modelling of pore water pressure changes and riverbank stability
during flow events. Earth Surface Processes and Landforms, 29(2), 237–254. https://doi.org/10.1002/esp.1042
Rinaldi, M., & Darby, S. E. (2007). 9 Modelling river-bank-erosion processes and mass failure mechanisms: Progress towards fully coupled
simulations. Developments in Earth Surface Processes, 11, 213–239. https://doi.org/10.1016/S0928-2025(07)11126-3
Rinaldi, M., Mengoni, B., Luppi, L., Darby, S. E., & Mosselman, E. (2008). Numerical simulation of hydrodynamics and bank erosion in a river
bend. Water Resources Research, 44(9), W09428. https://doi.org/10.1029/2008WR007008
Rinaldi, M., & Nardi, L. (2013). Modeling interactions between Riverbank Hydrology and mass failures. Journal of Hydrologic Engineering,
10(18), 1231–1240. https://doi.org/10.1061/(ASCE)HE.1943-5584.0000716

ZHAO ET AL. 48 of 51
Reviews of Geophysics 10.1029/2021RG000761

Rousseau, Y. Y., Van de Wiel, M. J., & Biron, P. M. (2017). Simulating bank erosion over an extended natural sinuous river reach using a
universal slope stability algorithm coupled with a morphodynamic model. Geomorphology, 295, 690–704. https://doi.org/10.1016/j.
geomorph.2017.08.008
Roy, S., Barman, K., Das, V. K., Debnath, K., & Mazumder, B. S. (2019). Experimental investigation of undercut mechanisms of river bank
erosion based on 3D turbulence characteristics. Environmental Processes, 7, 1–26. https://doi.org/10.1007/s40710-019-00417-3
Rutherfurd, I. D. (2000). Some human impacts on Australian stream channel morphology. In S. Brizga & B. Finlayson, (Eds.), River management:
The Australasian experience (pp. 11–49). John Wiley and Sons.
Samadi, A., Amiri-Tokaldany, E., Davoudi, M. H., & Darby, S. E. (2013). Experimental and numerical investigation of the stability of overhang-
ing riverbanks. Geomorphology, 184, 1–19. https://doi.org/10.1016/j.geomorph.2012.03.033
Samadi, A., Davoudi, M. H., & Amiri-Tokaldany, E. (2011). Experimental study of cantilever failure in the upper part of cohesive riverbanks.
Research Journal of Environmental Sciences, 5(5), 444–460. https://doi.org/10.3923/rjes.2011.444.460
Samani, A. N., Ahmadi, H., Mohammadi, A., Ghoddousi, J., Salajegheh, A., Boggs, G., & Pishyar, R. (2010). Factors controlling gully advance-
ment and models evaluation (Hableh Rood Basin, Iran). Water Resources Management, 24(8), 1531–1549. https://doi.org/10.1007/s11269-
009-9512-4
Sanders, H., Rice, S. P., & Wood, P. J. (2021). Signal crayfish burrowing, bank retreat and sediment supply to rivers: A biophysical sediment
budget. Earth Surface Processes and Landforms, 46(4), 837–852. https://doi.org/10.1002/esp.5070
Schuurman, F., Marra, W. A., & Kleinhans, M. G. (2013). Physics-based modeling of large braided sand-bed rivers: Bar pattern formation,
dynamics, and sensitivity. Journal of Geophysical Research: Earth Surface, 118(4), 2509–2527. https://doi.org/10.1002/2013jf002896
Schwimmer, R. A. (2001). Rates and processes of marsh shoreline erosion in Rehoboth Bay, Delaware, USA. Journal of Coastal Research,
672–683.
Seginer, I. (1966). Gully development and sediment yield. Journal of Hydrology, 4, 236–253. https://doi.org/10.1016/0022-1694(66)90082-5
Seminara, G. (2006). Meanders. Journal of Fluid Mechanics, 554, 271–297. https://doi.org/10.1017/S0022112006008925
Sgarabotto, A., D’Alpaos, A., & Lanzoni, S. (2021). Effects of vegetation, sediment supply and sea level rise on the morphodynamic evolution of
tidal channels. Water Resources Research, 57(7), e2020WR028577. https://doi.org/10.1029/2020wr028577
Shimizu, Y., Nelson, J., Arnez Ferrel, K., Asahi, K., Giri, S., Inoue, T., et al. (2019). Advances in computational morphodynamics using the Inter-
national River Interface Cooperative (iRIC) software. Earth Surface Processes and Landforms, 45(1), 11–37. https://doi.org/10.1002/esp.4653
Shu, A., Duan, G., Rubinato, M., Tian, L., Wang, M., & Wang, S. (2019). An experimental study on mechanisms for sediment transformation due
to riverbank collapse. Water, 11(3), 529. https://doi.org/10.3390/w11030529
Simon, A., & Collison, A. J. C. (2002). Quantifying the mechanical and hydrologic effects of riparian vegetation on streambank stability. Earth
Surface Processes and Landforms, 27(5), 527–546. https://doi.org/10.1002/esp.325
Simon, A., Curini, A., Darby, S. E., & Langendoen, E. J. (2000). Bank and near-bank processes in an incised channel. Geomorphology, 35(3–4),
193–217. https://doi.org/10.1016/S0169-555X(00)00036-2
Simon, A., & Darby, S. (2002). Effectiveness of grade-control structures in reducing erosion along incised river channels: The case of Hotophia
Creek, Mississippi. Geomorphology, 42(3–4), 229–254. https://doi.org/10.1016/S0169-555X(01)00088-5
Simon, A., & Thomas, R. E. (2002). Processes and forms of an unstable alluvial system with resistant, cohesive streambeds. Earth Surface
Processes and Landforms: The Journal of the British Geomorphological Research Group, 27(7), 699–718. https://doi.org/10.1002/esp.347
Solari, L., Seminara, G., Lanzoni, S., Marani, M., & Rinaldo, A. (2002). Sand bars in tidal channels Part 2. Tidal meanders. Journal of Fluid
Mechanics, 451, 203–238. https://doi.org/10.1017/S0022112001006565
Stecca, G., Measures, R., & Hicks, D. M. (2017). A framework for the analysis of noncohesive bank erosion algorithms in morphodynamic
modeling. Water Resources Research, 53(8), 6663–6686. https://doi.org/10.1002/2017WR020756
Stein, O. R., & LaTray, D. A. (2002). Experiments and modeling of head cut migration in stratified soils. Water Resources Research, 38(12),
20–21. https://doi.org/10.1029/2001WR001166
Sunamura, T. (1982). A wave tank experiment on the erosional mechanism at a cliff base. Earth Surface Processes and Landforms, 7(4), 333–343.
https://doi.org/10.1002/esp.3290070405
Symonds, A. M., & Collins, M. B. (2007). The establishment and degeneration of a temporary creek system in response to managed coastal
realignment: The Wash, UK. Earth Surface Processes and Landforms: The Journal of the British Geomorphological Research Group, 32(12),
1783–1796. https://doi.org/10.1002/esp.1495
Tal, M., & Paola, C. (2007). Dynamic single-thread channels maintained by the interaction of flow and vegetation. Geology, 35(4), 347–350.
https://doi.org/10.1130/g23260a.1
Tambroni, N., Luchi, R., & Seminara, G. (2017). Can tide dominance be inferred from the point bar pattern of tidal meandering channels? Journal
of Geophysical Research: Earth Surface, 122(2), 492–512. https://doi.org/10.1002/2016JF004139
Temmerman, S., Bouma, T. J., Van de Koppel, J., Van der Wal, D., De Vries, M. B., & Herman, P. (2007). Vegetation causes channel erosion in
a tidal landscape. Geology, 35(7), 631–634. https://doi.org/10.1130/g23502a.1
Terzaghi, K. (1951). Theoretical soil mechanics. Chapman and Hall, Limited.
Thompson, C., & Amos, C. L. (2002). The impact of mobile disarticulated shells ofCerastoderma edulis on the abrasion of a cohesive substrate.
Estuaries, 25(2), 204–214. https://doi.org/10.1007/BF02691308
Thorne, C. R. (1982). Processes and mechanisms for river bank erosion. In R. D. Hey, J. C. Bathurst & C. R. Thorne (Eds.), Gravel-bed rivers:
Fluvial Processes, Engineering, and Management (pp. 227–259). John Wiley & Sons.
Thorne, C. R., Alonso, C., Borah, D., Darby, S., Diplas, P., Julien, P., et al. (1998a). River width adjustment. I: Processes and mechanisms. Journal
of Hydraulic Engineering, 124(9), 881–902. https://doi.org/10.1061/(ASCE)0733-9429(1998)124:9(881)
Thorne, C. R., Alonso, C., Borah, D., Darby, S., Diplas, P., Julien, P., et al. (1998b). River width adjustment. II: Modeling. Journal of Hydraulic
Engineering, 124(9), 903–917. https://doi.org/10.1061/(ASCE)0733-9429(1998)124:9(903)
Thorne, C. R., & Tovey, N. K. (1981). Stability of composite river banks. Earth Surface Processes and Landforms, 5(6), 469–484. https://doi.
org/10.1002/esp.3290060507
Trimble, S. W. (2009). Fluvial processes, morphology and sediment budgets in the Coon Creek Basin, WI, USA, 1975–1993. Geomorphology,
108(1–2), 8–23. https://doi.org/10.1016/j.geomorph.2006.11.015
Turner, R. E. (1990). Landscape development and coastal wetland losses in the northern Gulf of Mexico. American Zoologist, 30(1), 89–105.
https://doi.org/10.1093/icb/30.1.89
Turowski, J. M., Hovius, N., Meng Long, H., Lague, D., & Men Chiang, C. (2008). Distribution of erosion across bedrock channels. Earth Surface
Processes and Landforms: The Journal of the British Geomorphological Research Group, 33(3), 353–363. https://doi.org/10.1002/esp.1559
Valentin, C., Poesen, J., & Li, Y. (2005). Gully erosion: Impacts, factors and control. Catena, 63(2–3), 132–153. https://doi.org/10.1016/j.
catena.2005.06.001

ZHAO ET AL. 49 of 51
Reviews of Geophysics 10.1029/2021RG000761

Vandekerckhove, L., Muys, B., Poesen, J., De Weerdt, B., & Coppé, N. (2001). A method for dendrochronological assessment of medium-term
gully erosion rates. Catena, 45(2), 123–161. https://doi.org/10.1016/S0341-8162(01)00142-4
Vandekerckhove, L., Poesen, J., & Govers, G. (2003). Medium-term gully headcut retreat rates in Southeast Spain determined from aerial photo-
graphs and ground measurements. Catena, 50(2–4), 329–352. https://doi.org/10.1016/S0341-8162(02)00132-7
Vandekerckhove, L., Poesen, J., Wijdenes, D. O., & Gyssels, G. (2001). Short-term bank gully retreat rates in Mediterranean environments.
Catena, 44(2), 133–161. https://doi.org/10.1016/S0341-8162(00)00152-1
Vandekerckhove, L., Poesen, J., Wijdenes, D. O., Gyssels, G., Beuselinck, L., & De Luna, E. (2000). Characteristics and controlling factors of
bank gullies in two semi-arid Mediterranean environments. Geomorphology, 33(1–2), 37–58. https://doi.org/10.1016/S0169-555X(99)00109-9
Van De Wiel, M. J. (2003). Numerical modelling of channel adjustment in alluvial meandering rivers with riparian vegetation. University of
Southampton.
van der Wegen, M., Wang, Z. B., Savenije, H. H. G., & Roelvink, J. A. (2008). Long-term morphodynamic evolution and energy dissipation in a
coastal plain, tidal embayment. Journal of Geophysical Research, 113(F3), F03001. https://doi.org/10.1029/2007JF000898
van Dijk, W. M., Hiatt, M. R., van der Werf, J. J., & Kleinhans, M. G. (2019). Effects of shoal margin collapses on the morphodynamics of a sandy
estuary. Journal of Geophysical Research: Earth Surface, 124(1), 195–215. https://doi.org/10.1029/2018JF004763
van Dijk, W. M., Lageweg, W. I., & Kleinhans, M. G. (2012). Experimental meandering river with chute cutoffs. Journal of Geophysical Research,
117(F3), F03023. https://doi.org/10.1029/2011JF002314
Van Dijk, W. M., Teske, R., Van de Lageweg, W. I., & Kleinhans, M. G. (2013). Effects of vegetation distribution on experimental river channel
dynamics. Water Resources Research, 49(11), 7558–7574. https://doi.org/10.1002/2013wr013574
Van Eerdt, M. M. (1985). Salt marsh cliff stability in the Oosterschelde. Earth Surface Processes and Landforms, 10(2), 95–106. https://doi.
org/10.1002/esp.3290100203
Vargas Luna, A., Duró, G., Crosato, A., & Uijttewaal, W. (2019). Morphological adaptation of river channels to vegetation establishment: A
laboratory study. Journal of Geophysical Research: Earth Surface, 124(7), 1981–1995. https://doi.org/10.1029/2018JF004878
Veihe, A., Jensen, N. H., Schiøtz, I. G., & Nielsen, S. L. (2011). Magnitude and processes of bank erosion at a small stream in Denmark. Hydro-
logical Processes, 25(10), 1597–1613. https://doi.org/10.1002/hyp.7921
Venditti, J. G., Rennie, C. D., Bomhof, J., Bradley, R. W., Little, M., & Church, M. (2014). Flow in bedrock canyons. Nature, 513(7519), 534–537.
https://doi.org/10.1038/nature13779
Walker, H. J. (2013). A picture is Worth 934 words: Riverbank erosion and Block Collapse in an Arctic delta. Focus on Geography, 56(3),
114–115. https://doi.org/10.1111/foge.12019
Watson, A. J., & Basher, L. R. (2006). Stream bank erosion: A review of processes of bank failure, measurement and assessment techniques, and
modelling approaches. A report prepared for stakeholders of the Motueka Integrated Catchment Management Programme and the Raglan Fine
Sediment Study. Landcare Research.
Wells, R. R., Alonso, C. V., & Bennett, S. J. (2009). Morphodynamics of headcut development and soil erosion in upland concentrated flows. Soil
Science Society of America Journal, 73(2), 521–530. https://doi.org/10.2136/sssaj2008.0007
Wells, R. R., Momm, H. G., Rigby, J. R., Bennett, S. J., Bingner, R. L., & Dabney, S. M. (2013). An empirical investigation of gully widening
rates in upland concentrated flows. Catena, 101, 114–121. https://doi.org/10.1016/j.catena.2012.10.004
Wilson, G. V., Nieber, J. L., Sidle, R. C., & Fox, G. A. (2013). Internal erosion during soil pipeflow: State of the science for experimental and
numerical analysis. Transactions of the ASABE, 56(2), 465–478. https://doi.org/10.13031/2013.42667
Wilson, G. V., Periketi, R. K., Fox, G. A., Dabney, S. M., Shields, F. D., & Cullum, R. F. (2007). Soil properties controlling seepage erosion
contributions to streambank failure. Earth Surface Processes and Landforms: The Journal of the British Geomorphological Research Group,
32(3), 447–459. https://doi.org/10.1002/esp.1405
Wood, A. L., Simon, A., Downs, P. W., & Thorne, C. R. (2001). Bank-toe processes in incised channels: The role of apparent cohesion in the
entrainment of failed bank materials. Hydrological Processes, 15(1), 39–61. https://doi.org/10.1002/hyp.151
Wood, D. M. (2014). Geotechnical modelling. CRC press.
Wu, L. Z., Zhou, Y., Sun, P., Shi, J. S., Liu, G. G., & Bai, L. Y. (2017). Laboratory characterization of rainfall-induced loess slope failure. Catena,
150, 1–8. https://doi.org/10.1016/j.catena.2016.11.002
Wu, T. H., McKinnell, W. P., III, & Swanston, D. N. (1979). Strength of tree roots and landslides on prince of Wales Island, Alaska. Canadian
Geotechnical Journal, 16(1), 19–33. https://doi.org/10.1139/t79-003
Wynn, T. M., Henderson, M. B., & Vaughan, D. H. (2008). Changes in streambank erodibility and critical shear stress due to subaerial processes
along a headwater stream, southwestern Virginia, USA. Geomorphology, 97(3–4), 260–273. https://doi.org/10.1016/j.geomorph.2007.08.010
Wynn, T. M., & Mostaghimi, S. (2006). Effects of riparian vegetation on stream bank subaerial processes in southwestern Virginia, USA. Earth
Surface Processes and Landforms: The Journal of the British Geomorphological Research Group, 31(4), 399–413. https://doi.org/10.1002/
esp.1252
Xia, J., Deng, S., Lu, J., Xu, Q., Zong, Q., & Tan, G. (2016). Dynamic channel adjustments in the Jingjiang Reach of the middle Yangtze River.
Scientific Reports, 6(1), 22802. https://doi.org/10.1038/srep22802
Xia, J., Deng, S., Zhou, M., Lu, J., & Xu, Q. (2017). Geomorphic response of the Jingjiang Reach to the Three Gorges Project operation. Earth
Surface Processes and Landforms, 42(6), 866–876. https://doi.org/10.1002/esp.4043
Xia, J., Zong, Q., Deng, S., Xu, Q., & Lu, J. (2014). Seasonal variations in composite riverbank stability in the Lower Jingjiang Reach, China.
Journal of Hydrology, 519(Part D), 3664–3673. https://doi.org/10.1016/j.jhydrol.2014.10.061
Xia, J., Zong, Q., Zhang, Y., Xu, Q., & Li, X. (2014). Prediction of recent bank retreat processes at typical sections in the Jingjiang Reach. Science
China Technological Sciences, 57(8), 1490–1499. https://doi.org/10.1007/s11431-014-5597-y
Xin, P., Dan, H., Zhou, T., Lu, C., Kong, J., & Li, L. (2016). An analytical solution for predicting the transient seepage from a subsurface drainage
system. Advances in Water Resources, 91, 1–10. https://doi.org/10.1016/j.advwatres.2016.03.006
Xin, P., Kong, J., Li, L., & Barry, D. A. (2013). Modelling of groundwater–vegetation interactions in a tidal marsh. Advances in Water Resources,
57, 52–68. https://doi.org/10.1016/j.advwatres.2013.04.005
Xin, P., Wilson, A., Shen, C., Ge, Z., Moffett, K. B., Santos, I. R., et al. (2022). Surface water and groundwater interactions in salt marshes and their
impact on plant ecology and coastal biogeochemistry. Reviews of Geophysics, 60(1), e2021R–e2740R. https://doi.org/10.1029/2021rg000740
Xin, P., Yuan, L. R., Li, L., & Barry, D. A. (2011). Tidally driven multiscale pore water flow in a creek-marsh system. Water Resources Research,
47(7), W07534. https://doi.org/10.1029/2010wr010110
Xu, F., Coco, G., Tao, J., Zhou, Z., Zhang, C., Lanzoni, S., & D'Alpaos, A. (2019). On the morphodynamic equilibrium of a short tidal channel.
Journal of Geophysical Research: Earth Surface, 124(2), 639–665. https://doi.org/10.1029/2018JF004952
Xu, F., Coco, G., Zhou, Z., Tao, J., & Zhang, C. (2017). A numerical study of equilibrium states in tidal network morphodynamics. Ocean
Dynamics, 12(67), 1593–1607. https://doi.org/10.1007/s10236-017-1101-0

ZHAO ET AL. 50 of 51
Reviews of Geophysics 10.1029/2021RG000761

Yamamuro, J. A., & Lade, P. V. (1998). Steady-state concepts and static liquefaction of silty sands. Journal of Geotechnical and Geoenvironmen-
tal Engineering, 124(9), 868–877. https://doi.org/10.1061/(ASCE)1090-0241(1998)124:9(868)
Yao, Z., Ta, W., Jia, X., & Xiao, J. (2011). Bank erosion and accretion along the Ningxia-Inner Mongolia reaches of the Yellow River from 1958
to 2008. Geomorphology, 127(1–2), 99–106. https://doi.org/10.1016/j.geomorph.2010.12.010
Yu, M., Wei, H., & Wu, S. (2015). Experimental study on the bank erosion and interaction with near-bank bed evolution due to fluvial hydraulic
force. International Journal of Sediment Research, 30(1), 81–89. https://doi.org/10.1016/s1001-6279(15)60009-9
Yuan, S., Tang, H., Li, K., Xu, L., Xiao, Y., Gualtieri, C., et al. (2021). Hydrodynamics, sediment transport and morphological features at
the confluence between the Yangtze River and the Poyang Lake. Water Resources Research, 57(3), e2020W–e28284W. https://doi.
org/10.1029/2020wr028284
Yumoto, M., Ogata, T., Matsuoka, N., & Matsumoto, E. (2006). Riverbank freeze-thaw erosion along a small mountain stream. Nikko Volcanic
Area, Central Japan, Permafrost and Periglacial Processes, 17(4), 325–339. https://doi.org/10.1002/ppp.569
Zech, Y., Soares-Frazão, S., Spinewine, B., & Le Grelle, N. (2008). Dam-break induced sediment movement: Experimental approaches and
numerical modelling. Journal of Hydraulic Research, 46(2), 176–190. https://doi.org/10.1080/00221686.2008.9521854
Zeng, H., Tang, C., Cheng, Q., Zhu, C., Lin, Z., Yin, L., & Shi, B. (2022). Desiccation-induced curling of mud layers: Field observations and
experimental insights. Engineering Geology, 296, 106458. https://doi.org/10.1016/j.enggeo.2021.106458
Zhang, K., Gong, Z., Zhao, K., Wang, K., Pan, S., & Coco, G. (2021). Experimental and numerical modeling of overhanging riverbank stability.
Journal of Geophysical Research: Earth Surface, 126(10), e2021J–e6109J. https://doi.org/10.1029/2021JF006109
Zhang, L. L., Zhang, J., Zhang, L. M., & Tang, W. H. (2011). Stability analysis of rainfall-induced slope failure: A review. Proceedings of the
Institution of Civil Engineers-Geotechnical Engineering, 164(5), 299–316. https://doi.org/10.1680/geng.2011.164.5.299
Zhang, Z., Shu, A., Zhang, K., Liu, H., Wang, J., & Dai, J. (2019). Quantification of river bank erosion by RTK GPS monitoring: Case studies
along the Ningxia-Inner Mongolia reaches of the Yellow River, China. Environmental Monitoring and Assessment, 191(3), 140. https://doi.
org/10.1007/s10661-019-7269-7
Zhang, W., Xu, Y., Hoitink, A. J. F., Sassi, M. G., Zheng, J., Chen, X., & Zhang, C. (2015). Morphological change in the Pearl River Delta, China.
Marine Geology, 363, 202–219. https://doi.org/10.1016/j.margeo.2015.02.012
Zhao, K., Gong, Z., Xu, F., Zhou, Z., Zhang, C. K., Perillo, G., & Coco, G. (2019). The role of collapsed bank soil on tidal channel evolu-
tion: A process-based model involving bank collapse and sediment dynamics. Water Resources Research, 55(11), 9051–9071. https://doi.
org/10.1029/2019WR025514
Zhao, K., Gong, Z., Zhang, K., Wang, K., Jin, C., Zhou, Z., et al. (2020). Laboratory experiments of bank collapse: The role of bank height
and near-bank water depth. Journal of Geophysical Research: Earth Surface, 125(5), e2019JF005281. https://doi.org/10.1029/2019JF005281
Zhao, K., Lanzoni, S., Gong, Z., & Coco, G. (2021). A numerical model of bank collapse and river meandering. Geophysical Research Letters,
48(12), e2021G–e93516G. https://doi.org/10.1029/2021GL093516
Zhou, Z., Coco, G., Townend, I., Olabarrieta, M., Van Der Wegen, M., Gong, Z., et al. (2017). Is “morphodynamic equilibrium” an oxymoron?
Earth-Science Reviews, 165, 257–267. https://doi.org/10.1016/j.earscirev.2016.12.002
Zhou, Z., Coco, G., van der Wegen, M., Gong, Z., Zhang, C., & Townend, I. (2015). Modeling sorting dynamics of cohesive and non-cohesive
sediments on intertidal flats under the effect of tides and wind waves. Continental Shelf Research, 104, 76–91. https://doi.org/10.1016/j.
csr.2015.05.010
Zhou, Z., Ye, Q., & Coco, G. (2016). A one-dimensional biomorphodynamic model of tidal flats: Sediment sorting, marsh distribution, and
carbon accumulation under sea level rise. Advances in Water Resources, 93, 288–302. https://doi.org/10.1016/j.advwatres.2015.10.011
Zolezzi, G., Luchi, R., & Tubino, M. (2012). Modeling morphodynamic processes in meandering rivers with spatial width variations. Reviews of
Geophysics, 50(4), RG4005. https://doi.org/10.1029/2012rg000392

ZHAO ET AL. 51 of 51

You might also like