You are on page 1of 29

9.

16 River Meandering
JM Hooke, University of Liverpool, Liverpool, UK
r 2013 Elsevier Inc. All rights reserved.

9.16.1
9.16.2
9.16.3
9.16.3.1
9.16.3.1.1
9.16.3.1.2
9.16.3.1.3
9.16.3.2
9.16.3.2.1
9.16.4
9.16.4.1
9.16.4.2
9.16.4.3
9.16.4.3.1
9.16.4.3.2
9.16.4.3.3
9.16.4.4
9.16.4.5
9.16.5
9.16.5.1
9.16.5.2
9.16.5.3
9.16.6
9.16.6.1
9.16.7
References

Introduction
Research Phases and Topics
Approaches and Methods
Empirical Approaches
Field measurements and observations
Map and remote sensing evidence
Techniques of meander morphology and change analysis
Theoretical and Numerical Modeling Approaches
Experimental modeling
Empirical Evidence and Analysis
Morphology
Morphological Change
Meander Processes
Flow patterns and sediment movement
Bank erosion
Deposition and bar formation
Bedrock and Incised Meanders
Spatial Distribution and Controls on Characteristics
Theoretical and Conceptual Explanations
Fundamental Physical and Numerical Analyses
Conceptual Analyses
Experimental, Modeling and Numerical Analysis Results
Perspective and Synthesis
Future Research
Conclusions

260
262
264
264
264
265
265
265
267
268
268
269
272
272
275
275
276
276
277
277
279
280
281
282
282
283

Abstract
Various phases of progress and differing approaches to research on river meandering are identified. Since early quantification of equilibrium relations of meander morphology to discharge and sediment, research has been pursued empirically,
theoretically, and experimentally. The theoretical approaches have sought to provide fundamental explanations of meander
development and produced numerical simulations. Empirical work, using field, map, and air photo evidence, has demonstrated variations in meander morphology, and stability and the evolution of meanders over time to compound forms
and cut-offs; it has elucidated process mechanisms and interactions. Flume work has investigated the effects of particular
conditions. Research using these differing approaches is now converging.

9.16.1

Introduction

Meandering rivers are single channels with a sinuous planform comprising a series of loops, frequently depicted as
regular and simple in form and size, but in reality often having
some irregularity, asymmetry, and complexity (Figure 1). They
are differentiated from braided channel patterns, which have
multiple channels or multiple free bars within the course, and
straight channels, which have very low sinuosity, though this
lower limit of meandering is somewhat arbitrary. Thresholds

Hooke, J.M., 2013. River meandering. In: Shroder, J. (Editor in Chief),


Wohl, E. (Ed.), Treatise on Geomorphology. Academic Press, San Diego,
CA, vol. 9, Fluvial Geomorphology, pp. 260288.

260

and conditions for development of different types of pattern


have been much investigated. Meandering courses are found
not only in fluvial rivers and in bedrock channels but also in
tidal flows, on glaciers, in oceanic currents, and in submarine
and Martian environments. This implies intrinsic characteristics of fluid flows, but debate still surrounds the basic
question of why rivers meander. Discussion in this chapter is
confined to fluvial flows in rivers.
Meandering rivers are ubiquitous on the Earth and are the
most common type of channel pattern. They have long been a
source of fascination and esthetic delight (Figure 1) so curiosity
as to their formation and processes of development has driven
much research. Many meandering channels are mobile, with
the meander forms migrating downstream and the channels

Treatise on Geomorphology, Volume 9

http://dx.doi.org/10.1016/B978-0-12-374739-6.00241-4

River Meandering

261

(a)

(e)

(b)

(f)
(c)

(d)

(g)

Figure 1 Photographs illustrating the characteristics and variability of meander morphology, Images (a)(c) US National Fish and Wildlife
Service; Images (d)(k) Google Earth; Images (l)(n) J M Hooke.

shifting in position within floodplains, producing significant


contribution to landscape change. This occurs on varying
timescales but, where rates of movement are high, it can cause
practical problems for riparian activities and structures. Thus,

much research has also been driven by the practical need to


understand the processes, rates, and patterns of movement and,
if possible, to develop methods of predicting meander movement. Their sensitivity means they are also important indicators

262

River Meandering

(h)

(i)

(j)

(m)

(k)

(l)

(n)

Figure 1 Continued.

of environmental change and response and are therefore of


global significance. Meandering rivers are very important for
their ecology and biodiversity, offering a range of types of
habitat, and many of these channels and associated floodplains
are under threat from human activities. Conservation goals have
driven much recent research, and a trend to put back meanders
into artificially straightened river courses has required a depth of
understanding of the forms and dynamics of meandering river
channels. This trend to river restoration, or reinstatement
and recreation of meanders, is recognition of the inherent

occurrence and behavior of river meandering in nature. Common features of river meanders are depicted in Figure 2.

9.16.2

Research Phases and Topics

Research on river meanders has been undertaken in a range


of disciplines, including geomorphology (in geography and
earth sciences), engineering, mathematics and fluid dynamics,
physics, biology, and ecology. The aims and approaches tend

Down valley
bank
ave
c
n
Co
Pool

263

Cross valley

River Meandering

Ch
P
ffle
Ri

Vegetation

Riffle

Point

Rc

F lo

eg

a lw

ute
ar
tb
oin

Rc

Radius of curvature

Tn

Vegetation

Vegetation

bar

Channel
width

Pool
Conc

ave bank

Figure 2 Common features of meanders. Redrawn with permission from Lagasse, P.F., Zevenbergen, L.W., Spitz, W.J., Thorne, C.R., 2004.
Methodology for predicting channel migration. NCHRP Web-Only Document 67 (Project 24-16). Report prepared for TRB (Transportation
Research Board of the National Academies of the US) http://onlinepubs.trb.org/onlinepubs/nchrp/nchrp_w67.pdf

to differ but all have made contributions to geomorphological


understanding of these features.
Meandering channels were identified early as key components of long-term landscape development, in W.M. Davis
Cycle of Erosion (King and Schumm, 1980), where he provided
detailed description of how meanders develop from original
incised bedrock forms in youthful stages to alluvial forms in
open valleys in old age. Occurrence and sequences of cut-offs
were recognized. Practical problems associated with the actively
migrating meanders of the Mississippi River were an impetus for
some of the early scientific work. Friedkin (1945) undertook
experiments into meander dynamics, which are still much cited,
and Fisk (1944) studied the meander morphology in relation to
floodplain materials and geomorphological controls.
The major era of quantitative research on river meander
morphology and processes began in the 1950s, continuing
through the 1960s, with the seminal work of Leopold and
Wolman (1957, 1960), Leopold and Langbein (1966), Leopold
et al. (1964), and Schumm (1960, 1963). This work set the
foundations for modern fluvial geomorphology and the basic
understanding of river meanders. Morphometric and processform relations, for example, between meander wavelength
and discharge, were quantified. Meander morphology was
thought to fit a sine-generated curve, and theoretical propositions were made of why rivers attained this form, associated
with arguments of energy distribution and minimum variance.
Other work in the same period, now largely ignored, was that
of Dury (1954, 1955, 1958, 1960) who pioneered an approach,
now recognized as paleohydrology, in which he tried to
explain the formation of valley and misfit meanders and the
longer-term landscape development in relation to former
discharges.
Until the 1970s the assumption was mostly that meander
development took place over long timescales and that meanders tended to equilibrium forms. The occurrence of cutoffs was recognized from the work on the Mississippi and
from the widespread occurrence of relict channels or oxbow

lakes in floodplains (Figure 1). With a shift to research on


shorter-term processes in geomorphology from the late 1960s,
attention began to turn to dynamics of river meander changes,
based on empirical field and historical evidence. Research by
Brice (1973, 1974, 1977), Hickin (1974, 1978), Hickin and
Nanson (1975), Nanson (1980a), Lewin (1972, 1976, 1978),
Knighton (1972, 1973), Hooke (1977, 1979), and Thorne and
Lewin (1979) in the 1970s showed that meander changes were
detectable on timescales of a few years. This research demonstrated the greater complexity of morphology and the
continuous evolution of meanders, in contrast to previous
equilibrium analyses. Asymmetric and compound bends were
identified as part of evolutionary sequences of meander development. Processes of erosion and deposition were measured and analyzed, including seminal research in individual
meander bends (Jackson, 1975; Bridge and Jarvis, 1976;
Dietrich et al., 1979) and research on the sedimentology of
bars (Bluck, 1971; Allen, 1970). By the 1980s much geomorphological evidence of meander behavior had been produced, as demonstrated in the important collection of papers
published in 1984 (Elliott, 1984), though there was still some
tendency to assume that high channel instability was associated with human interference in catchments.
The amount of field-based and geomorphological empirical
work declined through the 1990s but has recently seen resurgence, aided especially by technological developments that
enable detailed and rapid morphological and flow measurements. However, during the 1980s, theoretical and modeling
work on river meanders increased enormously, much of it
emanating from engineering and fluid dynamics spheres. Much
debate was engendered, culminating in a major workshop that
produced the seminal volume on river meandering in 1989
(Ikeda and Parker, 1989). Further developments in numerical
simulation and modeling took place, but by 1995 Mosselman
(Mosselman, 1995) stated that there was still no easily available
software for modeling meander changes and, in a large review
in 2004, Lagasse et al. (2004) found that there was no standard

264

River Meandering

method of meander movement prediction that they could


apply. The last decade or more has seen a proliferation of
analytical and numerical modeling of increasing sophistication.
Theoretical analyses, reviewed by Seminara (2006), clarify that
meandering is related to fundamental instability mechanisms
associated with the interaction of the flow and erodible
boundary. Hardware experimental modeling has also regained
favor, with much engineering experimentation concentrated on
flow patterns and bed topography in models with rigid walls.
Early experiments undertaken by Friedkin (1945), Hooke
(1975), Parker (1976), Schumm and Kahn (1972), and others
demonstrated meander development and influence of factors,
but most attempts at producing mobile meanders encountered
practical difficulties in preventing braiding. Significant recent
developments have used vegetation to produce self-formed
meanders (Tal and Paola, 2007; Braudrick et al., 2009). An
important impetus to research in the last two decades has been
to develop techniques and tools with which to design new
meandering channels in river restoration schemes (Rinaldi and
Johnson, 1997; Kondolf, 2006).
A key question is why rivers meander. This has proved
difficult to answer and some researchers would dispute whether it has been achieved. A major review and summary of
arguments to date were provided by Rhoads and Welford
(1991). The major arguments that have been used over time
can be summarized as:

Coriolis force
Energy arguments (excess, minimization)
Bank erosion and sediment effect
Helical and secondary flows
Inherent property of turbulent flow
Interaction between flow and mobile channel
Bar theory
Bend theory.

Seminara (2006) has summarized the answer as any small


random perturbation of channel alignment eventually grows,
leading to a meandering pattern, as shown by bend instability
theory. From the bend theory it can be summarized that a
straight channel is intrinsically unstable, provided the banks
are erodible. Federici and Seminara (2003) state that bar
instability is recognized as the fundamental mechanism
underlying the formation of large-scale forms of rivers, but
this may not be considered adequate to explain nonalluvial
meanders. Other researchers consider the explanation lies in
fundamental characteristics of fluid flow.
The current situation in river meandering research is one in
which a large number of theoretical and numerical simulation
models have been proposed, with varying theoretical bases
and assumptions, but many are lacking field validation or
testing with real river meander planforms. Several key differences in approach and assumptions are evident between fluid
dynamicists who make various assumptions such as equal
width, and geomorphologists who recognize the spatial and
temporal variability of meanders and the influence of factors
such as gradient, bank resistance, and discharge. Although
theoretical and empirical analyses were not entirely divorced
in the past, increasing convergence of approach and comparison of results are occurring (Hooke et al., 2011).

9.16.3

Approaches and Methods

A major division can be made into theoretical/modeling approaches and empirical approaches to river meandering. The
modeling approaches comprise numerical/analytical models
and the experimental (hardware) models. Empirical approaches entail field measurements and use of historical and
remote sensing data. Some empirical research has produced
inductive or kinematic models and some of these observed
relations have been used in other numerical simulations.

9.16.3.1
9.16.3.1.1

Empirical Approaches
Field measurements and observations

Much of the field research on river meanders has been focused


on component processes within river bends and along channels, concentrating on flow processes, bank erosion, and, to a
lesser extent, deposition and bar formation. In the mid-1970s,
several important field studies were undertaken in which
processes within a bend were measured, including flow patterns, sediment movement, erosion, and sedimentation (e.g.,
Jackson, 1975; Bridge and Jarvis, 1976; Dietrich et al., 1979;
Thorne and Lewin, 1979). These still remain as some of the
prime forms of evidence for whole meander bends. From this
period until recently, field measurements of combined processes have been largely neglected, though notable exceptions
include the work of Frothingham and Rhoads (2003) in a
compound bend. Technological developments and the need
to validate the large modeling effort are now stimulating increased field efforts.
Flow patterns and characteristics were measured with current meters and flow tracers, but these have now been replaced
by a range of acoustic Doppler instruments, enabling 3D
flow direction and intensity to be measured rapidly (e.g., Frothingham and Rhoads, 2003; Dinehart and Burau, 2005). Bank
erosion has been measured directly by various techniques including erosion pins, resurvey, exposure measurement instruments, photogrammetry, and terrestrial scanning (Lawler,
1993). Analysis of associated conditions has entailed monitoring of pore water pressures, stability, bank composition, and
strength (e.g., Simon et al., 2000; Parker et al., 2008). Deposition has involved particle size measurements and its spatial and temporal variation, and detailed mapping of
sedimentary structures in bars. Changes in bed morphology
have been measured by standard topographic surveying techniques, or in deeper rivers, by bathymetry from boats. Technological developments are transforming the ability to collect
detailed topography and monitor morphological changes and
processes, particularly the advent of terrestrial scanning
and differential GPS (e.g., Brasington et al., 2000; Heritage and
Hetherington, 2007) and of side scan sonar for large rivers.
Measurement and monitoring of spatial sequences of meanders over longer periods of time (more than 23 years) have
used more extensive field methods or mainly remote sensing
sources. Meander movement and changes in morphology
are monitored by resurvey of bank lines, field mapping, ground
photography, and cross-sectional surveys. Some field measurements of meander movement entail use of sedimentary and
morphological evidence, for example meander scrolls,

River Meandering

vegetation succession (Hickin and Nanson, 1975; Lawler, 1993;


Malik, 2006) on timescales varying from a few years to the
Quaternary (Alonso and Garzon, 1994). Very little dating of full
meander sequences has yet taken place, but deposits in oxbow
lakes, palaeochannels, and floodplains have been dated, mostly
by 14C. Developments in OSL offer much potential for dating
channel movement and formation of features (Rittenour et al.,
2005; Rodnight et al., 2005; Rowland et al., 2005).

9.16.3.1.2

Map and remote sensing evidence

River meandering is primarily a planform characteristic, and


relatively long timescales are often needed to detect change in
meander position and morphology, so major sources of evidence and bases of analysis are maps, aerial photographs, and
other remote sensing imagery. Historical maps have been used
to extend back in time (Figure 3). Amongst the documentations of river meander changes, one of the most remarkable
was that of Dort (1978) in a compilation of historical courses
of the Kansas River and its major tributaries from 1857, recently republished with additional explanation by the
American Geographical Society (Dort, 2009). Similar historical compilations include those of the Po River, Italy, from the
twelfth century onwards by Braga and Gervasoni (1989).
Quantitative use of historical maps was pioneered by Hooke
(1977) (Hooke and Kain, 1982; Trimble, 2008), and errors
associated with use of historical map have been analyzed by
Hooke and Redmond (1989), Downward (1995) and Gurnell
et al. (2003). Methods of use of maps and imagery within a
GIS environment have been developed (Gurnell et al., 1994;
Leys and Werritty, 1999; Hughes et al., 2006). Aerial photographs have the advantage of also showing a range of associated features such as bars, meander scrolls, and vegetation
cover as well as bank lines (e.g., Brice, 1974; Hickin and
Nanson, 1975; Gurnell, 1997) (Figure 1) and digital photogrammetry is now widely applied (Lane, 2000; Chandler et al.,
2002). Increased resolution and availability of satellite imagery means their increasing application to river meandering
(e.g., Seker et al., 2005). LIDAR imagery of high resolution is
now becoming widely available and is proving invaluable,
though problems may be encountered in removing vegetation
cover and detecting details of bank topography. Analysis of
past maps and imagery are now regarded as essential in any
engineering or management study as a basis for assessing
channel stability/mobility.

9.16.3.1.3

Techniques of meander morphology and


change analysis

Techniques of characterizing and analyzing river meander


courses identified by Hooke (1984) were: bend parameters,
curve fitting, spectral analysis, graphical analysis, and modeling. Major technological developments have taken place since
then, but many of the principles still pertain. Several parameters are needed to characterize the form adequately, to
encompass scale or size, shape or sinuosity, and irregularity
(Ferguson, 1975).
The standard method of characterizing meander morphology is the assumption of a regular wave form and the
measurement of standard parameters as shown in Figure 4.
Many variants on the detail of measurements exist, for example
whether for bank lines or centerlines, from single meanders, or

265

meander trains. Wavelength is the standard measure of scaling


the size of meanders. (Dimensionless wavenumber (pw/L) is
used in fluid dynamics literature.) Meander amplitude is a
measure of cross-valley breadth. Sinuosity is a measure of the
degree of wiggliness, measured for individual bends or, more
often, for a series of meanders. Radius of curvature (r) is a
measure of size by fitting circles to individual meander bends
and r/width is used as a measure of shape or curvature. These
parameters are usually measured directly on plots of meander
courses but are subjective and can be difficult to apply on the
more irregular forms that often occur in nature.
Efforts at automating meander planform analysis have
been made, most using digitized centerlines of courses
(Figure 5). Hooke (1977), Hooke and Kain (1982), Trimble
(2008) and ONeill and Abrahams (1986) used curvature to
identify points of inflection and subdivide bends objectively.
Howard and Hemberger (1991) developed measures of
sinuosity, wavelength, curvature moments, asymmetry, and
pattern irregularity and showed that natural meander
morphology is different from model simulations. Andrle
(1996) assessed problems of measurement of sinuosity, due to
scale and wavelength, because of subjectivity of bend definition, and tested an angle measure technique. Similarly,
Lancaster and Bras (2002) suggested the use of different scales
of sinuosity measurement. Coulthard and Van De Wiel (2006)
have developed a method of radius of curvature definition
within raster modeling, as opposed to the usual use of vector
data. An alternative approach, early on, for characterizing
meandering was that of spectral analysis (Speight, 1965), but
this was found to be rather insensitive. Digitization of courses
allows curvature sequences and characteristics to be assessed,
though values are influenced by digitizing interval. Developments in wavelet analysis may offer much potential. A major
development recently by Guneralp and Rhoads (2008, 2009)
is in characterizing curvature more continuously by fitting
cubic splines to digitized courses.
Various classifications of typologies of change in meander
morphology have been produced (Table 1; Hooke, 1997).
Components of change include shortening, extension, translation, rotation, enlargement, cut-off, and complex change,
and have been identified on various rivers, for example, the
Ganga (Swamee et al., 2003) and the Luanga (Gilvear et al.,
2000). Bank lines are usually superimposed to detect channel
changes and digitization in GIS enables quantitative analysis.
Approaches to prediction of meander changes have mostly
involved meander modeling, but an empirical approach was
used in a large project undertaken by the US Transportation
Research Board (Lagasse et al., 2004) in which they compiled
historical aerial photograph evidence for three dates from 141
meander sites containing 1503 meander bends on 89 rivers in
the US. Various methods were investigated; the method of
linear extrapolation of trends from simple circle fitting was
adopted, but this is not ideal for nonlinear and compound
development.

9.16.3.2

Theoretical and Numerical Modeling Approaches

An enormous explosion of effort devoted to modeling river


meanders has occurred in the last few years, some of it aimed at

55
54

50

48

NW
SE
100 m

57
0 0.25 0.5

1 mile

53

(b)

Scale 1:55.000

Saline river - parts of sheets 1 and 2, between Salina and Culver

0 0.25 0.5

45

49
2001
1984
1970

1907
1871
1840

Republican river - parts of sheets 3 and 4, vicinity of Morganville

1 mile

Scale 1:55.000

(c)

NW

Solomon river - parts of sheets 1 and 2, vicinity of Bennington

0 0.25 0.5

Radnor
bridge

SE

1 mile

Scale 1:55.000

NW

(a)

Smoky hill river - parts of sheets 3 and 4, between Solomon and Salina

1840
1870
1910
1947
1968
1984

1 mile

Scale 1:55.000

Holly
bank

Swettenham
bridge

Twemlow
bend

0 0.25 0.5

SE

NW

SE
Historical courses
500 m

(d)

Figure 3 Examples of meander changes from compilations of historical evidence. (a) Historical compilations of the courses of Republican, Saline, Solomon and Smoky Hill Rivers in Kansas, USA.
Reproduced with permission from Dort, W., 2009. Historical Channel Changes of the Kansas River and Its Major Tributaries. American Geographical Society, New York, 80 pp. (b)(d) Historical
compilations of the River Dane and Bollin, NW England (Hooke and Harvey, 1983; Hooke, 1987).

River Meandering

Color key
1
2
3
4
5
6
7
8
9
10
11
12

266

Tithe map 1840


O.S. map 1870
O.S. map 1907
Air photos 1947
O.S. map 1968
Air photos 1984

River Meandering

267

nk
ba

Thalweg

nd
Be ius
rad

Cross over

Surface
width W

Rm

Meander width MB

ex

Location of
point bar
Amplitude A

e
av

Co
nv

Co
nc

Meander width ML

nk
ba

Axis of bend

(a)

Meander
wavelength

Mean radius
of curvature

Concave
bank

Convex
bank

Wave
amplitude

Point of
inflection

Axis of
bend

Meander
belt axis

(b)

Figure 4 Standardized parameters of meander morphology. (a) Single meander. Redrawn with permission from Knighton, D., 1998. Fluvial
Forms and Processes a New Perspective. Oxford University Press, Oxford. (b) Meander train. Redrawn from Huggett, R., 2003. Fundamentals of
Geomorphology. Routledge, London, 386 pp.

solving practical problems and providing a means of prediction


and design, others aimed at elucidating the bases of river meandering by testing fundamental theory. Models of increasing
sophistication and complexity have been developed, progressing
from simple 1D equations through to 2D representations and
then 3D modeling using computational fluid dynamics (CFD).
Mostly, the simpler models simulate meander behavior in a
meander train (sequence of meander loops) but, because of
computational demands, the more complicated models tend
only to produce simulations for single bends or limited reaches.
The theoretical modeling approaches are based on fundamental
physics (dynamical or analytical modeling) and formulations of
flow interactions with the boundary. Much of the modeling
builds hierarchically on previous work (Camporeale et al.,
2007). Much of it is linear and reductionist in approach, but
developments in nonlinear modeling are now taking place. A
major underlying theory is that bank erosion and bend movement are related to excess near-bank velocity. Many of the
models have been based on the original theory by Ikeda et al.
(1981), later modified by Johannesson and Parker (1989)
and others. Some other models are built on empirically
derived relations (kinematic modeling), much of it based on
movementcurvature relations. Some models simulate development of meanders from straight; others start with a meandering
form, usually a sine-generated curve, and simulate changes.

Models of individual processes, such as bank erosion, are now


being incorporated into meander movement simulations. Some
integrated models of flow pattern, sediment transport, bed
topography, and grain size sorting in bends have been developed
and compared with field evidence (e.g., Bridge, 1992).

9.16.3.2.1

Experimental modeling

Early efforts at hardware, experimental simulation of natural


behavior of meander movement and change in flumes encountered problems of creating enough resistance in the banks
to produce a persistent single, meandering channel. Many
produced low-sinuosity, self-formed channels with submerged
bars (e.g., Friedkin, 1945; Wolman and Brush, 1961; Ackers and
Charlton, 1970) (Figure 6(b)) but tended toward a braided
planform as the channel migrates (Parker, 1998). Attempts
were made by these and others (e.g., Smith, 1998) to solve the
problem by using clay mixtures with sand. A more recent
phase of experimentation is producing mobile meanders more
successfully (Peakall et al., 2007; Tal and Paola, 2007;
Braudrick et al., 2009). Braudick et al.s experiments created
self-maintaining meandering channels with cut-offs by using
vegetation in the form of alfalfa sprouts to provide the appropriate resistance (based on an idea from Tal and Paola)
(Figure 6(a)). In the flume experiments of Peakall et al. sedimentary structures of bars were reproduced. Most other

268

River Meandering

j + 1
Digital
points

j + 1

j + 1

(a)

1.74 radians


bf

Mean downvalley
direction

 = 
c
e

g
 = 

Direction (), radians

b
2

g
 = 1.74 sin ( 2 )

d

0.25
2

(b)

f
0.25

0.25

Relative channel distance (  )

Direction (), radians

Channel course
2

Curvature ()

1
(c)

9.16.4
9.16.4.1

f
c

2004, 2005; Abad and Garcia, 2008; Blanckaert, 2009; Termini,


2009; Zeng et al., 2008) rather than simple low curvature, sinegenerated forms. Most experiments are run with constant discharge and uniform sediment size. Many of the experimental
results are used for testing theoretical or numerical simulation
model results (e.g., Zolezzi et al., 2005; Shams et al., 2008). A
few of the numerical and flume models have been tested
against field data; for example, Ferguson et al. (2003) tested a
CFD model of flow in a high angle bend against field data.

Distance (km)

Figure 5 Digitization and curvature analysis. (a) Digitization points,


direction and angle difference of links. (b) Curvature and wavelength
calculation for a sine-generated course. (c) Direction and curvature for
an irregular course, several bends in length. Redrawn with permission
from Knighton, D., 1998. Fluvial Forms and Processes a New
Perspective. Oxford University Press, Oxford.

laboratory experiments have focused on flow patterns and/or


bed topography in meandering channels and most have used
fixed bank channels, and some fixed beds. Some experiments
use complex, high sinuosity channels or sharp-angle bends
(e.g., Whiting and Dietrich, 1993a, b; Blanckaert and de Vriend,

Empirical Evidence and Analysis


Morphology

The outer banks of meander bends are generally steep and


eroding, with a pool present in the bed at the apex, and sloping
bed topography from the point bar on the opposite inner bank
(Figures 1 and 2). Riffles or shallowings are present in the inflection regions of the bend, in the straight limbs between apices, and the cross sections and banks are more symmetrical
there. Meanders are usually depicted in reference books as
regular, symmetric sine-generated curves, as in Figure 4, a form
introduced by Langbein and Leopold (1966), who argued that it
represented the minimum variance form and most uniform
distribution of change along the curve. However, many actual
meanders were recognized by them (and earlier) as being
commonly asymmetric and skewed, then more planforms were
examined by Brice (1973) and Lewin (1972, 1978) amongst
others, so that by 1983 Carson and Lapointe (1983) said that
use of sine-generated curves should be abandoned as the
standard description of the equilibrium type of freely meandering rivers because most meanders were asymmetric. Yet
this form is still widely used as the basis for much modeling. A
particular shape of skewed and fattened, tortuous meander,
commonly occurring, has been named a Kinoshita curve after
its originator; convoluted meanders are also called goosenecks.
In a neglected study of the Jordan River, Schattner (1962)
named a range of bend forms and analyzed their spatial distribution. Carey (1969) identified smoothly curving caving bends
and abrupt angle bends but considered them to differ from true
meanders. Occurrence of sharp angled bends has been recognized by many researchers (e.g., Alvarado-Ancieta and Ettmer,
2008), and highly convoluted loops and cut-offs have been reported from a range of environments, including humid tropical
(e.g., Ebisemiju, 1993; Gilvear et al., 2000; Gautier et al., 2007).
Low sinuosity channels that are more angular and react to
local controls without developing true meanders are termed
wandering rivers or pseudo-meandering (Bartholdy and Billi,
2002). Confined and incomplete meandering are terms used to
refer to meandering that is constrained by valley width,
as is particularly common in many formerly glaciated environments (Lewin and Brindle, 1977; Nicoll and Hickin, 2010).
Subdivisions of meander types have been suggested, mostly
based on sinuosity classes (Schumm, 1963; Brice, 1982;
Church, 1992) (Figure 7). Meanders have been subject to
fractal analysis and found to have asymmetrical (self-affine)
fractal scaling (Nikora, 1991; Montgomery, 1996; Stolum,
1996). Montgomery (1996) showed that fractal dimension
and sinuosity are highly correlated.

River Meandering

269

Table 1 Classifications or typologies of change


Author

Model/classification

Kondratyev (1968)
Daniel (1971)

Graphical mode of change from simple bend through to cut-off.


Fitted a sine-generated curve to meanders. Identified five types of movement involving expansion, rotation
and translation.
Processes of change and relationship to bedforms in a five-stage model.
Classification of morphology defined by circle combinations and analysis of sequences of change.
Pattern of meander change developed from analysis of erosion pathlines.
Predictive model for determining rates of channel bend migration.
Identified six types of lateral activity:
downstream progression,
progression and cut-offs,
mainly cut-offs,
entrenched loop development,
irregular lateral activity, and
avulsion.

Keller (1972)
Brice (1974)
Hickin (1974)
Hickin and Nanson (1975)
Kellerhals et al. (1976)

Hooke (1977)

Hooke and Harvey (1983)

Used Daniels (1971) primary elements of movement in double and triple combinations to compose a suite
of 70 models of movement. Primary elements:
extension,
translation,
rotation,
enlargement,
lateral movement, and
complex change.
Seven categories of change defined:
simple migration,
confined migration,
growth (extension),
lobing and compound growth,
retraction and cut-off,
complex changes (islands, abandonment, etc.), and
stable bends, no change.

Brice (1984)

Classified changes in meander form according to mode of bank erosion.

Meanders are also frequently depicted as of uniform width,


and this is an assumption of many models; but Brice (1982)
showed that the most active meanders are wider at apices than
crossings, and Lagasse et al. (2004) in their large data compilation confirmed that variation in width around a meander
is an indication of activity. Carson (1986) also recognized that
overwidened sections are common and Luchi et al. (2010)
recently quantified the systematic variation in width around
active meander bends.
In the era of quantification of morphology and processform relations, many equations formulating the interrelations
of meander parameters (defined in Figure 4) or the relations
of meander forms to controls, mainly discharge, were produced (Leopold et al., 1964). Wavelength (ML) is a measure of
scale of meanders and is found generally to be 1014 times
channel width. It correlates with discharge, though the appropriate measure of discharge was much debated (Carlston,
1965), and with drainage area as a surrogate of discharge.
Pools and riffles are conventionally regarded as spaced at 57
channel widths, in phase with the wave form. Radius of
curvature is also closely related to channel width, averaging
23 widths and 0.2 of the wavelength. Amplitude is much less
closely related to width or discharge. Leopold and Wolman
(1960) produced power form equations of the morphological
relations and they have since been widely tested (Bridge,

2003). Type of channel pattern and degree of meander intensity is closely related to sediment load and to channel resistance (Schumm, 1960, 1963), with alluvial channels being
more sinuous in material with higher percentage silt-clay.

9.16.4.2

Morphological Change

It has long been recognized that meanders migrate downstream, as was well known from rivers such as the Mississippi.
It was also known that many meanders exhibit skewing and
asymmetry, but it was assumed that this was due to inhomogeneities in the floodplain material, producing differential
erosion rates in different parts of the meander train.
A major development in analysis of meanders was the
recognition of a common sequence of evolution of meander
form over time, stemming largely from the work of Brice (1973,
1974, 1984) in which he proposed the sequence of types in
Figure 8(a) and from the work of Hickin and Nanson (1975)
(Figure 8(b)) on the Beatton river in Canada, using the evidence of meander scrolls mapped from aerial photographs.
Some other sequences were produced, for example, Kellers
(1972) (Figure 8(c)) model, but elongation of meander limbs is
less widely applicable. Hickin (1978) generalized his observations into the very important conceptual model (Figure 9(a)),
which became the basis for much other empirical analysis and

270

River Meandering

(a)

(b)

(c)

Figure 6 (a) Photograph of flume experiment using alfalfa sprouts to simulate vegetation cover. Reproduced with permission from 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 of the United States of America 106(40), 1693616941. (b) Flume experiment showing development
of bars in a straight channel. Reproduced from Seminara, G., 2006. Meanders. Journal of Fluid Mechanics 554, 271297, with permission from
Cambridge University Press. (c) Bar development in a straightened channel for comparison. From Jaeggi, M. 1984. Formation and effects of alternate
bars. Journal of Engineering, American Society of Civil Engineers 110, 142156.

kinematic modeling. This shows a nonlinear increase in migration rate with curvature (decreasing r/w) to a maximum at a
critical curvature, identified as r/2w 23, and decrease in migration rates again beyond this critical curvature value. Hooke
(1997) compiled data from other rivers, showing similar
envelope curves to meander behaviors, though with some
tighter curves and higher rates at low r/w than the original
(Figure 9(b)), later also identified by Hudson and Kesel (2000)
on the Mississippi. Brice (1974), Hooke and Harvey (1983) and
others recognized the importance of compound forms, and
Hooke and Redmond (1992) and Hooke (1995b) produced the
qualitative model in Figure 8(d) of sequence through to neck
cut-off, which involves change in the position and number of
pools and riffles (Hooke and Harvey, 1983).
Debate continues on the mechanism and conditions, and
reasons for the compound development. Hooke and Harvey
(1983) showed that the compound development was associated with development of an additional riffle in the apex region as the pathlength increased beyond some critical value. It
was assumed that this leads to breakdown of secondary flow
patterns (Thompson, 1986). Lofthouse and Robert (2008)
identified a lengthening of pools with increased angular
deflection up to a critical length at which formation of
a new riffle took place. One of the few studies to examine

flow relations in a compound bend is that of Frothingham


and Rhoads (2003) (see below: flow patterns). From empirical
analysis, Guneralp and Rhoads (2009) demonstrate that
the spatial structure of the planform curvature effect on
migration rates depends on the complexity of the planform
geometry. Evidence on variation in rate of erosion and
migration through the meander evolution sequence is
somewhat contradictory, with Hickin and Nanson suggesting
the nonlinear increase to a curvature limit, supported by evidence from Hooke and Harvey (1983) but Furbish (1988,
1991) suggesting monotonic increase. Hooke and Yorke
(2010) now show acceleration of rate through the growth
phase and a slowing during the compound phase. Some of
the confusion is due to the lack of distinction between
bank erosion rates and extent of downvalley or cross-valley
movement.
Meander movement and change in morphology can produce cut-offs, with oxbow lakes in the abandoned channel.
Former channels at various stages of infill are very common
features of meandering river floodplains. The timing, process,
and distribution and effects of cut-offs have received much
attention, with a resurgence of interest in recent years, much of
this driven by ecological and conservation interests. Lewis and
Lewin (1983) identified five types of cut-off chute, neck,

River Meandering

Single phase, equiwidth channel, deep

(a)

Single phase, equiwidth channel

(b)

Single phase, wider at bends, chutes rare

(c)

Single phase, wider at bends, chutes common

(d)

Single phase, irregular width variation

(e)

Two-phase underfit, low water sinuosity

(f)

Two-phase bimodal bankfull sinuosity

(g)

Figure 7 Modified Brice classification of meander forms. Redrawn


from Lagasse, P.F., Zevenbergen, L.W., Spitz, W.J., Thorne, C.R.,
2004. Methodology for predicting channel migration. NCHRP Web-Only
Document 67 (Project 2416). Report prepared for TRB
(Transportation Research Board of the National Academies of the US)
http://onlinepubs.trb.org/onlinepubs/nchrp/nchrp_w67.pdf

mobile bar, mulitloop chute, and mulitloop neck; of these,


two main types are usually distinguished, chute and neck
cut-off.
Neck cut-offs (Figure 1) can be the endpoint of the meander evolution illustrated above (Figures 8 and 9). It was
assumed that these mostly take place in, and can be caused by,
high flow events. Some case studies have now shown that
actual cut-off may occur in moderate flow events if the bend
neck has narrowed sufficiently (Hooke, 1995a; Gautier et al.,
2007). Very few studies have followed the adjustments immediately after cut-off, but both Hooke (1995a) and Fuller
et al. (2003) have shown variable and high amounts of sedimentation, channel erosion, and morphological change

271

within the main channel in the period immediately succeeding cut-off. Plugs are formed in the upstream entrance to
the old channel very quickly and more slowly downstream,
but the actual dynamics and subsequent infill of the abandoned channel are related to angle and lake connectivity to
main channel and its dynamics (Citterio and Piegay, 2009).
Piegay et al. (2000) measured sedimentation rates in 39
oxbow lakes in SE France, producing a mean rate
02.57 cm yr1. This was linked to water depth and decrease
from entrance as a function of overbank frequency, which also
varies with channel geometry, Gautier et al. (2007) found on
the Beni that the functioning of abandoned branches is
strongly associated with the mobility of the main channel
rather than with flood intensity. The length of oxbow lakes is
closely related to the sinuosity of the channel (Constantine
and Dunne, 2008). Cut-offs are important in preservation of
flood history and environmental changes and some have been
dated as part of understanding floodplain chronologies (see
Chapter 9.32).
Controls and mechanisms of chute cut-offs are still
somewhat debated and not fully understood, though height of
floodplain, sediment supply, and discharge regime may be
factors. One mechanism is by formation of headcuts on the
downstream side of the bend and progression upstream (Gay
et al., 1998). Much longer-length channel avulsions can occur
on meandering rivers where the river breaks out of its course
to form a new channel down a length of valley. Explanation
and prediction of avulsion is still proving challenging (Slingerland and Smith, 1998; Aslan et al., 2005; Phillips, 2009) (see
Chapter 9.32). Both Peakall et al. (2007) and Braudrick et al.
(2009) have reproduced chute cut-offs in their hardware
models.
Overall, meandering systems are found to exhibit a wide
range of behaviors (e.g., Hooke, 2007; Seker et al., 2005;
Goswami et al., 1999), varying from meanders which are
evolving very slowly such as channels in bedrock or resistant
material, channels with highly vegetated banks or very low
slopes (e.g., Biedenharn et al., 1984; Rhoads and Miller, 1991;
Swanson, 1984) to those which are highly complex and dynamic. Timescales of the full sequence of meander evolution
have not been established for many rivers and more research is
needed. Dort (2009) suggested from the meander scroll and
cartographic evidence that meander development from initiation to cut-off on the Kansas River takes at most a few decades. Harmar and Clifford (2006) have suggested that
meander trains on the Lower Mississippi River are continually
formed and modified over a period of approximately 120
years. Hooke (2004) found that on the River Bollin bends had
evolved from low curvature to cut-off in 120150 years. The
periods for complete floodplain reworking ranged from 600 to
7000 years on streams in Devon, if current rates are simply
extrapolated (Hooke, 1980). Rates and timescales of floodplain reworking have been found to vary even in the same
region. For example, Mertes et al. (1996) calculated an alluvial
plain recycling time ranging from 1000 to 4000 years on the
SolimoesAmazon River, but Gautier et al. (2007) found that
it is about 10 times higher on the tributary Beni River. Beechie
et al. (2006) calculated a turnover recurrence interval of 60
years on meandering forested mountain river systems of the
Pacific Northwest, USA.

272

River Meandering

2
A

Compound symmetrical

Compound asymmetrical

1
B

Stages
4

Simple symmetrical

2
E

1
Simple asymmetrical
(a)

Erosin path lines

Meander scrolls

(b)

Initiation

Doubleheading

5
Riffle

Migration

Pool
Erosion
Primary flow
Depositional bar

Cutoff
Growth
Pool
Riffle
Erosion

(c)

Asymmetric shoal for


stages 12, point bar
for stages 3, 45

(d)

Figure 8 Models of sequences of meander development. (a) Sequence through from simple symmetric bends to asymmetric and compound.
Reproduced from Brice, J.C., 1974. Evolution of meander loops. Bulletin of Geological Society of America 85, 581586, with permission from GSA. (b)
Elongation, skewing and compound development, based on meander scroll bar evidence on Canadian rivers. Reproduced from Figure 7 in Hickin, E.J.,
1974. Development of meanders in natural river-channels. American Journal of Science 274(4), 414442. (c) Sequence of limb elongation with
additional pool and riffles in the limbs. Reproduced from Keller, E.A., 1972. Development of alluvial stream channels: a five-stage model. Geological
Society of America Bulletin 83, 15311540, with permission from GSA. (d) Sequence of migration, growth and compound development through to cutoff, involving development of additional riffle in apex. Reproduced from Hooke, J.M., 1995b. Processes of channel planform change on meandering
channels in the UK. In: Gurnell, A., Petts, G.E. (Eds.), Changing River Channels. Wiley, Chichester, pp. 87116.

9.16.4.3
9.16.4.3.1

Meander Processes
Flow patterns and sediment movement

Flow patterns in meander bends were characterized from early


on as a periodically reversing helical motion helicoidal flow,
with Leliavsky (1955) explaining that any small flow deviations
will cause a centrifugal force from one side of a bend to
the other, which will promote development of transverse flow.
Recognition of more complex patterns of secondary flows,

particularly the existence of an outer bank cell came from


the field measurements of Bathurst, Hey, and Thorne in the
mid-1970s (e.g., Bathurst et al., 1977; Thorne and Hey, 1979).
The essential characteristics of flow patterns in meander bends
are summarized by Markham and Thorne (1992) in
Figure 10(a).
Major features of flow from detailed field measurements are
exemplified from the work of Dietrich in a bend on Muddy
Creek, Wyoming (Figures 10(b) and 10(c)) (Dietrich, 1987).

0.5

Grwoth period

Termination stage

Migration rate (m year1)

1.0

273

Initiation stage

River Meandering

0
0

1.0

2.0

(a)

3.0
4.0
rm (w)

5.0

6.0

7.0

Erosion rate widths (m w1)

0.2
Beatton river (Hickin and Nanson)
R. Dane (Hooke)
Canadian rivers (Hickin and Nanson)
Red Rivers (Biedenham et al.)
R. Bollin

0.1

0.0
0

(b)

10

Rate of movement (width year1)

0.3

12

14

A
B

Chute cutoff

Neck cutoff
Migrating
Stabilizing

D
0.2

Bollin envelope curve


B

0.1

D
0.0
(c)

5
6
7
8
Bend curvature (r/w)

10

11

12

13

Figure 9 Relationship of rate of movement or erosion to curvature. (a) Conceptual generalization of meander development and accelerating rate
of movement with increase in bend curvature (decrease of r/w). Reproduced from Hickin, E.J., 1978. Hydraulic factors controlling channel
migration. In: Davidson-Arnott, R., Nickling, W. (Eds.), Research in Fluvial Geomorphology, Proceedings Fifth Guelph Symposium on
Geomorphology, pp. 5966. (b) Envelope curves of data of dimensionless movement rates in relation to bend curvature for several rivers.
Reproduced from Hooke, J.M., 1997. Styles of channel change. In: Thorne, C., Hey, R., Newson, M. (Eds.), Applied Fluvial Geomorphology for
River Engineering and Management. Wiley, Chichester, pp. 237268. (c) Hypothetical trajectories for different meander behavior over time.
Reproduced from Hooke, J.M., 2003. River meander behaviour and instability; a framework for analysis. Transactions of Institute of British
Geographers 28, 238253.

274

River Meandering

Outer bank
cell

Superelevated water surface

Outrward shoaling flow


across point bar

Characteristic
secondary
velocity profiles

Pathe lines of secondary flow


(a)

2
En Us w
g R

High

ax

High

Water surface elevation


(E)

Low

b

b
ax

+n

b

S = I Es
IN s

b pghS

b

ax

Es

Low

En

ma

Low

s
+n n

Es

High
En

+n

Bar exposed
at low flow

(b)

Break in bed slope

Thalweg
Near surface velocity
Near bed velocity

(c)

Figure 10 (a) Generalized pattern of flow in meander bends. Reproduced from Markham, A.J., Thorne, C.R., 1992. Geomorphology of gravelbed rivers. In: Billi, P., Hey, R.D., Thorne, C.R., Tacconi, P. (Eds.), Dynamics of Gravel Bed Rivers. Wiley and Sons, Chichester, pp. 433456.
(b) Channel curvature and bed topography effects on the boundary shear stress field. Reproduced from Figure 8.6 in Dietrich, W., 1987.
Mechanics of flow and sediment transport in river bends. In: Richards, K.S. (Ed.), River Channels: Environment and Processes. Blackwell, Oxford,
pp. 179227. (c) Flow field in equilibrium with bed topography in bends with well developed bars. Reproduced from Figure 8.9 in Dietrich, W.,
1987. Mechanics of flow and sediment transport in river bends. In: Richards K.S. (Ed.), River Channels: Environment and Processes. Blackwell,
Oxford, pp. 179227.

River Meandering

Water surface is superelevated round bends, the slope increasing


with the tightness of bends. This has an effect on the distribution
of boundary shear stress and the velocity fields. The centrifugal
force produces secondary circulation. Bed topography takes on
the form of bar-pool units (Figure 10(b)) with a break in bed
slope, commonly termed a bar or lobe front, extending across
the channel and around the front of the exposed point bar. The
maximum-velocity thread crosses from near the inner bank at
the bend entrance to the outer bank at the bend exit, crossing at
the zone of maximum curvature. In a simple bend, of the
morphology shown, the maximum shear stress is downstream of
the apex, resulting in bend migration. The zone of spiral motion
with outward flow near the surface and inward flow near the bed
is confined to the deepest 2030% of the cross-section. This
detailed flow pattern differs from that of Hey and Thorne (1975)
in exhibiting less cross-stream orientation, a shoaling-induced
outward flow over the bar and lack of dual cells in the crossings,
and with a much smaller outer bank cell. Much subsequent
research on flow patterns, particularly in flumes, has aimed to
understand the controls on these variations.
Bagnold (1960) identified that in tight bends flow
separation can take place, developing at the inner bank
downstream of the bend apex at first, then a second zone
developing at the outer bank in very tight bends. Hodskinson
and Ferguson (1998) made field measurements to test a numerical model of flow separation and showed that existence
and extent of concave bank flow separation can be significantly influenced by changes in bend planform, point bar
topography and upstream planform. In very tight (abrupt
angle) bends or where the outer bank is constrained outer,
concave-bank deposition and inner, convex-bank erosion can
occur (Andrle, 1994).
Understanding of the mechanisms of development of
compound loops and bend asymmetry have been helped by
the field measurements of flow pattern in a compound meander containing multiple pool-riffle structures made by Frothingham and Rhoads (2003). The downstream velocity field
is characterized by a high-velocity core that shifts slightly
outward as flow moves through individual lobes of the loop.
For some of the measured flows this core becomes submerged
below the water surface downstream of the lobe apexes. Skewinduced helical motion develops within the pools near lobe
apexes and decays over riffles where channel curvature is less
pronounced. Maximum rates of bank retreat generally occur
near lobe apexes where impingement of the flow on the outer
channel bank is greatest. However, maximum rates and loci of
bank retreat differ for upstream and downstream lobes of the
loop, leading to increasing asymmetry of loop geometry
over time.
The pattern and mechanisms of grain sorting in bends has
been examined in some field research (e.g., Jackson, 1975;
Bridge and Jarvis, 1976; Lapointe and Carson, 1986). Anthony
and Harvey (1991) showed that in a meandering river with
high sediment supply the flow and sediment transport patterns differed with discharge level. Clayton and Pitlick (2007),
building on previous work, found that fine and coarse fractions are differently routed in bends, with fine grains swept
inward over the point bar, and coarse grains routed outward
toward the pool. At bankfull, all sizes were transported but
size increased toward the outer bank.

9.16.4.3.2

275

Bank erosion

Some debate still surrounds the details of the influence of


the flow pattern on bank erosion, but the proposition by
Leliavsky (1966) and later developments by Parker et al.
(1983) and Hasegawa (1989) indicating that erosion in a
meander bend is primarily controlled by the excess near-bank
velocity are widely accepted and have become the basis for
much modeling (e.g., Howard, 1992).
Mechanisms and processes of bank erosion were reviewed
by Lawler et al. (1997). Major lateral movement of channels in
meanders occurs by bank erosion that is mainly associated
with high flows but influenced by the soil moisture status and
the pore water pressures. Composition and height of the bank
affect stability. Many floodplain banks of alluvial meanders
have a composite structure of a coarse, gravel base and finer
material above (Thorne and Tovey, 1981). Julian and Torres
(2006) found, by testing various flow properties, that the
amount of hydraulic erosion of cohesive riverbanks is dictated
by flow peak intensities. Erosion rates are highly influenced by
bank resistance, particularly bank material and vegetation (e.g.,
Schumm, 1963; Hooke, 1979), but vegetation effects have
proved difficult to quantify, though research is continuing.
Within-event evidence indicates that much of the bank
failure takes place by erosion of the lower gravel layer then
toppling (cantilever failure) of the fine layer (e.g. Pizzuto,
1984). In finer or more homogenous material and with high
pore water pressures the failure may be by sliding. Luppi et al.
(2009) found that the occurrence of mechanisms on the
Cecina R, Italy varied between the seven flow events in one
year. Slide failures were closely related to peak flow, and
cantilever failures occurred in late phases of the hydrograph. If
blocks remain at the base of the bank then they may reduce
subsequent erosion (Thorne and Tovey, 1981; Lapointe and
Carson, 1986). In active bends, bank erosion may take place
in several events a year (Hooke 1979, 1980). Larsen et al.
(2006) applied an empirical relation between erosion rate and
stream power to assess the effects of variable discharge on
lateral channel movement. The balance between bank erosion
and deposition varies with discharge, Pizzuto (1994) showing
from a large number of bank erosion measurements on the
Powder River, USA, that high flows were producing erosion
dominance. Pizzuto (2009) derived an empirical model from
field measurements of bank erosion in which he finds that
cohesive bank erosion has a strong stochastic component, but
that hydraulic erosion is responsible for 87% of all erosion.

9.16.4.3.3

Deposition and bar formation

Several different types of bars are recognized in river channels


(e.g., Church and Jones, 1982) but those most characteristic of
meandering channels are point bars, found on the inner side
of meander bends. These are termed fixed or steady bars. Bars
separated from the floodplain that migrate are termed free.
Formation of free alternate bars in straight channels is often a
precursor to formation of true meanders. The coexistence of
free and fixed bars was recognized from field observations by
Kinoshita and Miwa (1974). Single mid-channel bars are quite
common in active meandering channels and may indicate
transition to braiding, but these bars have been shown to
be significant in the sequence of development of some meanders (Hooke, 1986; Hooke and Yorke, 2010). Few studies

276

River Meandering

have quantified the timescale of development of bars,


but Church and Rice (2009) examined long-term evidence
of alternate type bars in the Fraser River and found that individual bars have a life history of about 100 years, except in
certain protected positions. A newly formed gravel bar quickly
assumes its ultimate thickness and relatively quickly approaches its equilibrium length. Growth continues mainly by
lateral accretion of unit bars. Hooke (1986, 2007b; Hooke and
Yorke, 2011) identified a life-cycle of 79 years average for
mid-channel bars on a small river in NW England.
The field measurements of Jackson (Jackson, 1975) and of
Bridge and Jarvis (1976) of flow and sediment movement in
bends were aimed at fuller understanding of the development
of sedimentary structures on point bars. The classical model is
of a simple fining upward sequence and lateral bar face structures (Allen, 1970), but they found sequences and structures
were more complex and spatially varied than this. Formation
of point bars and other bars contributes to floodplain construction (see Chapter 9.32). Point bar morphology is greatly
influenced by the w/d ratio (cross-sectional shape) of the
channel with high w/d channels having wide, flat-topped bars
(Dietrich, 1987). In laboratory experiments in sharper bends,
Kawai and Julien (1996) found that point bar morphology
depends on particle size and that point bars were considerably
smaller in coarse sand than fine sand. Differing views over the
major mechanisms of point bar deposition, whether due to
secondary flow circulation producing inward flow at the bed
on bars or whether due to deceleration of longitudinal flow
and distribution of shear stress are still apparent, and both
flume and field experiments are being used to investigate this.
In terms of dynamics, Bennett et al. (1998) measured temporal
variations in point bar morphology on a low w/d ratio channel
over 2 years in two bends and found erosion and coarser bars
in winter, but deposition and finer bar texture in summer. Net
changes at each site were synchronous and equal in magnitude.
In contrast, Hooke (2008) found spatial and temporal variability in depositional characteristics within a reach.
Counterpoint bars or concave bank benches/bars (Lewin,
1983; Page and Nanson, 1982) can form in meander bends
where separation takes place at the outer bend and/or rapid
migration takes place leaving a large zone of low velocity and
shear stress in which relatively fine material is deposited. Such
bar deposits can be much thicker than point bars (Burge and
Smith, 1999). More continuous bank benches have been
recognized (e.g., Shi et al., 1999) and can occur from dominance of deposition over erosion (Pizzuto, 1994).

9.16.4.4

Bedrock and Incised Meanders

Large bedrock meanders have long been recognized in the


landscape, including by W.M. Davis (King and Schumm, 1980).
Major valley features are termed either entrenched, in which the
valley sides are steep and symmetrical, or ingrown, in which
some lateral cutting has occurred, producing a more open
valley with slip-off spurs. Dury (Dury, 1954, 1955, 1958, 1960)
studied valley meanders, identifying them as fossil features
formed by much greater discharges in a previous hydrological
period, and now occupied by misfit channels with a smaller
scale of present meandering. The long-held assumption on

Figure 11 Photograph of the incised gooseneck meanders of the


San Juan River, Utah, USA.

valley meanders was that the meandering form was inherited by


superimposition or antecedence. Schumm (1977) found from
experiments on formation of incised meanders that they could
not be created from superimposition but they did occur by base
level lowering (analogous to vertical uplift) and he produced
the convoluted meanders typical of the famous San Juan goosenecks (Figure 11). Shepherd (1972) showed in flume experiments that the forms were influenced by the sediment load
transported. Local base level and gradient influence the incision
and morphology, as do tectonics (Harvey, 2007). A study of the
distribution and geometry of 600 km of incised meanders in
the central Colorado Plateau (Harden, 1990) found that most
meander cross sections in the area are relatively symmetrical,
but highly ingrown forms are also present and that, in general,
symmetric bends are associated with resistant bedrock units,
whereas ingrown forms develop in massive sandstone and in
highly erodible bedrock. Gradient significantly influences the
distribution of ingrown bends, with asymmetric meanders
concentrated in reaches of low average gradient; and this is a
greater influence than the lithology itself. In contrast, Tooth
et al. (2002) and McCarthy and Tooth (2004) consider that the
nature of the bedrock influences the nature and processes of
incision and bedrock meander formation. Complex sequences
of meanders can occur in mixed bedrock and alluvial reaches,
with subtle variations in sediment load (Tooth et al., 2004;
Marren et al., 2006), and sedimentary features are able to form
and persist over short to medium timescales even while bedrock erosion is ongoing. Mechanisms of formation of meanders
in bedrock still remain to be elucidated, but it has been found
that many bedrock meanders have the same features as alluvial
meanders (e.g., Kale, 2005), and can be highly complex (Zhang
et al., 2008) even exhibiting double-heading and compound
forms (Hooke, 2003). Cut-offs are common features of incised
bedrock meanders. (Bedrock channels are discussed more fully
in Chapter 9.28).

9.16.4.5

Spatial Distribution and Controls on


Characteristics

Research on the threshold of meandering with braiding


helps to explain the distribution of meandering, relating it to
stream gradient, discharge, sediment supply, and erodibility

River Meandering

of boundaries (see Chapter 9.17). Spatial variability


of morphology and stability can be high within river
reaches (Figures 1 and 3). Characteristics of meanders in
different zones within and between rivers are related to the
gradient and/or resistance of banks influenced by material
and/or vegetation and land use, and rock type and structure
(e.g., Timar, 2003; Zhang and Fu, 2003; Alvarado-Ancieta and
Ettmer, 2008; Zhang et al., 2008). Generally, it is found that
the most active meanders are in the steepest alluvial reaches
and those with highest stream power but with erodible material, often in the middle reaches of a river. Rates of lateral
movement have been analyzed particularly in relation to discharge and stream power and in relation to curvature. They
show a strong, square-root scale relationship to drainage area,
as a surrogate of discharge (Brice, 1984; Hooke, 1980; Lawler,
1993). Stream power is a strong influence (e.g., Lewin, 1983;
Hickin and Nanson, 1984; Richard et al., 2005; Larsen et al.,
2006) and Nicoll and Hickin (2010) recently found that it
explained 450% of variation in migration rate on a selection
of confined meanders in the Canadian prairies. Differing
channel materials influence the channel form and mobility
(e.g., Fisk, 1944; Biedenharn et al., 1984; Rhoads and Miller,
1991; Hudson and Kesel, 2000; Wolfert and Maas, 2007) as
does vegetation. Bank erosion and rates of lateral movement
have been shown to be generally higher on unvegetated or
grassed banks than forested (Beeson and Doyle, 1995; Burckhardt and Todd, 1998; Harmel et al., 1999) and vary with the
type of land use (e.g., Micheli et al., 2004). Individual
trees and vegetation can increase roughness and resistance
(Ebisemiju, 1994; Thorne and Furbish, 1995). Much field and
experimental work is now ongoing into vegetation effects.
Tectonics affect channel pattern and form of meanders
(Schumm, 2005) mainly by the influence on slope. For example, Gomez and Marron (1991) detected differential
morphology and changes upstream and downstream of uplift
axes. Tectonics can also cause lateral tilting of meander courses
(Nanson, 1980; Aswathy et al., 2008).
Although the autogenic nature of meander evolution and
the lack of inherent stabilization of many meanders is now
widely recognized, it is still acknowledged that the primary
controls on behavior and morphology are discharge, sediment
size and supply, bank resistance, and gradient, and thus if any
of these change then a response in the river channel is likely.
Many studies have aimed to identify and measure the responses of meandering channels to such allogenic alterations
(e.g., Bradley and Smith, 1984; Erskine et al., 1992; Klein,
1985; Uribelarrea et al., 2003; Tiegs and Pohl, 2005).
The problems caused by channel mobility are such that
very many meandering channels are now controlled by structures to prevent or reduce bank erosion and channel movement. Much engineering modeling has aimed to predict the
effects of proposed structures and some empirical studies have
examined the effects of modifications. Not only does the
prevention of bank erosion have effects at the site, but also the
reduction in sediment supply downstream has ramifications. A
common solution to the problem of channel mobility and
bank erosion, as well as to flooding, was that of channelization. Very many channels worldwide have been straightened.
The detrimental effects of straightening naturally meandering
channels are now widely acknowledged and have led to the

277

massive movement toward river restoration. Most of these


schemes involve remeandering the channels and that has
meant producing designs for the new channel. Some have
been based on empirical evidence of past morphology, others
on theoretical, experimental, or empirical modeling (Kondolf,
2006). Rinaldi and Johnson (1997) and Kondolf (2006)
provide analyses of the accuracy of empirical relations of meander parameters for use in meander restoration procedures.

9.16.5
9.16.5.1

Theoretical and Conceptual Explanations


Fundamental Physical and Numerical Analyses

The basic theoretical components of meandering and their


interactions are summarized by Camporeale et al. (2007) in
Figure 12, with the two key elements being curvature and an
erodible boundary. Derived linear models are based on the laws
of mass and momentum conservation governing fluid mechanics
within open channel flows with movable boundaries (bed and
banks) and therefore are held to fully account for meander
mechanics (flow properties, bed deformation, channel migration), although in a rather simplified way. In terms of the fundamental theory underlying much of the numerical modeling,
the basic premises for many of the models were developed by
Ikeda et al. (1981), Parker et al. (1983), Struiksma et al. (1985),
Parker and Andrews (1986), Johannesson and Parker (1989),
Odgaard (1989), and Ikeda and Parker (1989), based on earlier
fundamental work by Rozovski (1961), Leliavsky (1955, 1966),
Engelund (1974), and Kalwijk and de Vriend (1980), and others
on flow and bed topography in curved channels. Johanessen and
Parkers paper corrected problems of Ikeda et al.s model of not
satisfying sediment continuity; corrections also incorporated by
Blondeaux and Seminaras (1985) earlier paper. Leliavsky (1966),
Ikeda et al. (1981), and Hasegawa (1989) proposed that there is a
linear relation of channel migration rate to the excess near-bank
longitudinal velocity, a fundamental explanation that is widely
accepted. Nevertheless, the extent of longitudinal flow, secondary
flow, and topographic steering influences on meandering has
been the subject of much modeling and experimentation.
In Seminaras (2006) view, the major components of fundamental theory and modeling of meanders comprise: a
nonlinear planform evolution equation obtained by stipulating that the centerline of erodible channels moves in the
lateral direction with some lateral migration speed; an erosion
rule relating this erosion speed to the near-bank hydrodynamics; and a model of flow and bed topography in sinuous channels required to predict near-bank flow. Seminara
(2006) explains the mechanisms as follows:

The lateral pressure gradient associated with a lateral free surface


slope in a bend y[produces].. a secondary flow .. directed outward
close to the free surface and inward close to the bed. y.[which]
downstream, y.transfers momentum toward the outer bend...
Secondary flow transports sediments in the lateral inward direction
building up a rhythmic sequence of forced (point) bars and pools
respectively at the inner and outer bends of a train of meanders.
The establishment of a bar-pool pattern then gives rise to a further
topographically induced component of the secondary flow, which
drives an additional contribution to sediment transport and further
modifies the bed topography.

278

River Meandering

Erodible boundary
Curvature
(banks)

(bed)

Vortex-induced
stresses

Friction

Secondary currents
(driven by curvature)

Free bars
Gravity

Bedload

Transversal
flow field

Resonance
Shoaling
(point bars)
Secondary currents
(driven by topography)

Momentum
redistribution

Topographic
steering
+

Inward
bed stresses
+

Outward shifting of
bulk of the stream

Bank erosion
Phase lag in the longitudinal flow
phase lag in the secondary currents

Excess bank
stress

Downstream migration
and skewness

Figure 12 Components and interrelationship of factors influencing meander development. Redrawn from 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,
with permission from AGU.

Thus once a perturbation is induced this is amplified.


Since the early experiments (Friedkin, 1945), it has been
recognized that free bars form in alternate patterns in straight
channels that transport sediment (Figure 6(c)), and that
channels with erodible boundaries eventually evolve into a
curved channel pattern with fixed (steady) bars in each bend
(point bars). The formation of alternate free bars in straight
channels has been analyzed theoretically in bar theory;
Tubino et al. (1999) state that
The formation of bars has been conclusively explained in terms of
an inherent instability of erodible bed subject to a turbulent flow in
almost straight channels, which leads to the spontaneous development of bottom perturbations (free bars) migrating downstream.

Tubino and Seminara (1990) originally identified that free


bars in straight channels become suppressed and become fixed
bars as channel curvature increases, the threshold value at
which this occurs being a function of meander wavenumber
and sediment characteristics for given flow. Blondeaux and
Seminara (1985) identified a condition of resonance in which
the meander wavenumber and the width ratio of the channel
take values such as to force formation of free bars which
neither grow nor decay either in time or in space. Both theory
and experiments suggest that the formation of free bars is
controlled by a threshold of width/depth ratio of the channel
that depends on the Shields stress and roughness.
Fluvial bars may also be considered to arise as the result of
forcing effects of curvature or possibly width variations or
confluences. This complementary approach (Tubino et al.,

1999) is termed bend theory whereby meander development


is controlled by the nonlinear interaction between self-excited
(free) and curvature-driven (forced) bed responses. The major
bend theory was further developed by Seminara et al. (2001)
and Zolezzi and Seminara (2001) who produced theoretical
formulations to show that crossing the resonance barrier leads
to a reverse in directions of meander and bar perturbation
migration, the subresonant condition having downstream
influence on bar migration and bend form and superresonance having upstream influence. The w/d (or aspect ratio
w/2d) is a fundamental control on behavior. Simulations
using this theory have now produced compound bend evolution (e.g., Frascati and Lanzoni, 2009).
Much debate has long surrounded the role of alternate bars
in meander development. Mosselman et al. (2006) explain
overdeepening as essentially being due to the superimposition
of a steady alternate bar pattern on a streamwise uniform bed
topography. Both Sun et al. (1996) and Lancaster and Bras
(2002) argue that the influence of alternate bars on the initial
development of meander loops appears to be negligible. The
argument between free bars becoming fixed (bar theory) and
fixed (steady) bars being induced by the planform (bend
theory) can be termed as a difference between a temporal and
a spatial approach to the formation of steady bars (Crosato
and Mosselman, 2009). These authors have used a modeling
approach suited to fixed (steady) bars, which they argue is
more appropriate for real rivers than is a free migrating bar
method because most rivers have curvature.
A component of meandering and bar formation is grain
sorting in bends. Seminal theoretical work on this was by

River Meandering

Parker and Andrews (1985) in which two-dimensional bed


load transport of graded material was produced by incorporating effects of gravity on lateral slopes and of secondary flows.
The model indicated that the locus of the coarse sediment
transport shifts from the inner to the outer bank near the bend
apex. An integrated model of flow, sediment transport, bed
topography, and grain sorting was produced by Bridge (1992),
and further developments incorporating bank erosion were
produced by Darby et al. (2002) and Duan and Julien (2005).
Meander growth and meander migration arise from the
occurrence of a phase lag between bank erosion, produced by
the flow pattern, and channel curvature. It is argued that a lag
between curvature of the channel and flow produces a spatial
memory effect and the characteristic skewness found in many
bends. Seminara (2006) considers that nonlinear analytical
solutions of theoretical equations show that patterns emerging from observations of geomorphologists [such as sinegenerated and Kinoshita curves] are not purely empirical
correlations, but rather different approximations of an exact
periodic solution of the planform evolution equation related
to the harmonics of the curvature relations. Thus multilobe
meander loops have been explained by growth of higher order
harmonics (Seminara et al., 2001; Sun et al., 2001). Seminara
et al. (2001) also noted that no equilibrium solution has been
found, though they consider the evidence indicates that the
meander growth rate is found to decrease sharply in the late
stage of meander development, whereas the migration speed
of the meander train is found to decrease monotonically from
incipient formation to cut-off.
There are a number of limiting assumptions of the prevailing linear theory and models, and a current debate is the
extent to which processes and relations are linear or nonlinear.
Theories and models are now being developed which incorporate nonlinearities and reduce some of the present limiting assumptions (Pittaluga et al., 2009), though these are
still within a reductionist, mechanistic framework. Mosselman
et al. (2006) reviewed the history of the shifting interpretations of the overdeepening theory and demonstrate that
mathematical analyses of physical systems do not produce
straightforward answers or predictions, but inherently require
interpretations based on experimental evidence and a thorough understanding of the underlying physical processes. The
nature of the curvature influence in bends is still not fully
understood and Guneralp and Rhoads (2009) suggest that the
convolution functions used in theoretical models of meander
migration are hypothetical in nature and none has been
evaluated empirically using data on the spatial structure of the
curvaturemigration relation obtained from natural meandering rivers.

9.16.5.2

Conceptual Analyses

At a more conceptual level, underlying much of the interpretation of morphology and morphological changes in fluvial geomorphology is the theory of equilibrium and
adjustment. It is assumed that channel form, as measured by
variables such as meander wavelength and channel width, is
adjusted to the discharge and sediment load delivered to it.
This is evidenced by statistically stable relations from

279

empirical measurements. It is assumed that, as changes in


these basic external controls take place, for example, by
climate change or by land-use change, water control or abstraction, the channel form will adjust to a new equilibrium.
Numerous case studies have demonstrated such adjustment
and net change in meander characteristics. Changes in
channel form are therefore explained as adjustment to changed extrinsic factors. In the equilibrium state it is assumed
that meanders migrate but do not change in morphology,
once the period of development is complete, and that cut-offs
are largely due to variation in rates of movement associated
with differing resistance and constraints within the channel
system, mainly valley width. Firmly embedded in the
traditional theory of extrinsic controls and explanation of
meander morphology and allogenic causes of change are ideas
of major factors influencing river planform. Though discharge
largely controls scale of meanders, caliber and amount of
sediment load, gradient, channel resistance (including vegetation), structure and tectonics, valley size, and long-term
landscape development can all affect the details of
morphology.
Leopold and associates, in research in the 1960s, argued
that the quasi-equilibrium meander form that develops is related to the most probable state, that of minimization of
variance, which is a compromise between minimum total
work and uniform distribution of power expenditure, and that
the sine-generated form minimizes variance (Leopold, 1994).
Such arguments have been dismissed as teleological. Equilibrium behavior has been reexamined by Eaton et al. (2006)
in a conceptual model of meander initiation, within an
equilibrium and slope-minimizing context. They attempt to
explain why uniform shear stress distributions are nearly always unstable, once perturbed. They base this on the feedback
process between asymmetry of cross section shear stress distribution and local transport capacity. The initial perturbation
is related to turbulence on side walls.
The research on meander change in the 1970s and early
1980s showed that many meanders exhibit no stabilization to
an equilibrium form, rather that continuous evolution occurs,
in some cases through to cut-off. At first there was an assumption amongst geomorphologists that this was still an
adjustment to changed controls. Eventually, partly from the
empirical evidence and partly from theoretical modeling, it
has become more accepted that such a sequence is autogenic
and intrinsic. Hickin and Nanson (1975) suggested that the
increase in migration rate is nonlinear to a maximum at a
critical curvature, usually r 23, and decreases at higher
curvature and this has since been substantiated by other evidence (Hooke, 1997; Hudson and Kesel, 2000). The Hickin
and Nanson relation is empirical rather than theoretically
derived, but it has now been argued that it is predicted by
physics-based models (Crosato, 2009). The rising limb can be
explained by the decrease in relative lag distance between
near-bank flow velocity and forcing curvature as r/w increases.
The falling limb results from the decrease in local channel
curvature and near-bank flow velocity excess. Bagnold (1960)
argued that energy losses were at a minimum at r/w 23 and
Hooke (1975) had predicted from experiments that the most
stable meander geometry would occur at r/w 23 because of
the shear stress distribution.

280

River Meandering

Development of nonlinear dynamical or chaos theory and


ideas of self-organization have been applied to river meandering at a conceptual level (Stolum, 1996, 1998; Phillips,
2007). The idea of self-organized criticality (SOC) originates
from Bak (Bak, 1996; Bak et al., 1987; Bak and Chen, 1991).
Many composite systems naturally evolve to a critical state in
which a minor event starts a chain reaction that can affect any
number of elements in the system. It is a holistic theory, which
does not depend on microscopic mechanisms. Based on this,
Stolum (1996, 1998) developed a simulation model of meanders based on fluid mechanics theory and showed that meanders will tend toward a critical state of sinuosity and then
oscillate by occurrence of cut-offs. A straight course represents
order and a highly sinuous course chaos. In the model, bends
developed and cut-offs occurred to bring about self-organization. Meanders in an unconstrained state developed toward a
critical sinuosity limit of 3.14 (which is p). Beyond this limit,
cut-offs occur to maintain the steady state of the critical sinuosity. The system may at times undershoot or overshoot. In the
phase approaching criticality then occasional cut-offs occur, but
in the supercritical state clusters of meanders or avalanches
occur. Stolum (1998) also recognized a power-law distribution
of size of cut-offs, tested by an example of satellite imagery of
some rivers in Brazil. Hooke (2004) produced an example
which she suggests fits this framework, with clustering of cutoffs explained by the sinuosity state. Montgomery (1996) used
modeling and comparison with natural river reaches to plot
attractors and concluded that the rivers were not clearly chaotic.
It has been argued by others that cut-offs induce long-term
equilibrium and reduce complexity of sinuosity (e.g., Howard
and Knutson, 1984; Camporeale et al., 2005, 2008). Perucca
et al. (2005) and Frascati and Lanzoni (2010) suggest there is
no evidence of chaotic behavior in the long-term in meandering rivers. Not all rivers are highly mobile and Hooke
(2003, 2007) has suggested a framework in which high activity
may be reconciled with equilibrium behavior.
Theoretical and experimental models have mostly assumed
a temporally constant and spatially equal channel width.
Those that have allowed for variation have assumed that
movement of the outer bank line (erosion) and inner bank
(deposition) are balanced within short time periods. Alterations in width over time were assumed to be adjustments to
altered discharge or occurrence of flood events. Recent research suggests that the two banks may be operating more
independently (Lauer and Parker, 2008) though Nanson and
Hickin (1983) suggested that the annual lateral migration of
river bends is not a continuous process in space and time;
rather it is a process in which the concave bank rushes ahead
only to stop and wait until the convex bank catches up before
rushing ahead again.

9.16.5.3

Experimental, Modeling and Numerical Analysis


Results

Much of the flume experimentation is aimed at understanding


the influences of flow structure on magnitude and position of
scour and deposition in meander bends and extent of overdeepening (outer-bend pools being deeper than the equilibrium outer bend water depth). Differences persist in the

detailed flow structure produced by various studies, particularly


the intensity, size, and position of secondary flow cells. Many
researchers are examining the influence of curvature on flow
structure and processes (e.g., Blanckaert, 2009; Termini, 2009);
for example, the difference in flow structure in upstream and
downstream skewed Kinoshita bends (Abad and Garcia,
2009), or the conditions giving rise to development of outer
bank secondary cells (Blanckaert and de Vriend, 2004). Many
of the experiments use idealized configurations, including the
sine-generated curve, so questions still remain on the applicability to real meandering channels, and field testing is still
limited.
In more realistic simulations in which self-formed meanders
are allowed to develop, Braudrick et al. (2009) found that
elevated bank strength (provided by alfalfa sprouts) relative to
the cohesionless bed material and the blocking of troughs
(chutes) in the lee of point bars via suspended sediment deposition were the necessary ingredients to successful meandering. Varying flood discharge was not necessary but sand
supply may be essential. Pyrce and Ashmore (2005) confirmed
from flume experiments an inherent connection between the
loci of particle deposition and point bar formation, largely
controlled by the morphology of the channel with deposition
typically focused around the point bar apex, downstream of the
apex (contributing to downstream bar migration), and at the
bar head/riffle surface, but loci of deposition vary with stage of
flow and of development of the bend.
Many numerical models are now simulating the details of
specific component processes. Olsen (2003) and Ruther and
Olsen (2007) demonstrate that 3D CFD can now be used to
model features of meandering including migration and chute
cut-offs. Numerical modeling such as that of Chen and Duan
(2006) is now simulating migration of meandering channel
including downstream translation, lateral extension, expansion, and downstream and upstream rotation. They found that
low-sinuosity, free meanders migrate rapidly downstream. As
the sinuosity increases, downstream translation diminishes,
and meandering loops expand laterally with the head rotating
toward downstream and then upstream (as found in empirical
studies). The simulated results indicated that the gradient of
the longitudinal sediment transport rate is essential in modeling meandering evolution. Bank erosion rates and mechanisms are obviously very influential, and basal erosion and
bed load transport are particularly important (Duan and
Julien, 2005). Darby et al. (2002) show that cohesion of bank
material significantly influences equilibrium bed topography,
as does the bed material size in energetic rivers.
Simulation models from quite an early stage (e.g., Howard
and Knutson, 1984; Lancaster and Bras, 2002) implied that
multiloop bends may be sensitive to bank roughness, supporting field experimental evidence that roughness may influence meander morphology (Thorne and Furbish, 1995).
Though few models are able to produce compound forms and
neck cut-offs, the Zolezzi-Seminara (ZS) model has done so
(Seminara et al., 2001; Camporeale et al., 2007) (Figure 13).
Simulation models produce sequences of meanders but their
morphology may not fit with real meanders (Howard and
Hemberger, 1991), and Frascati and Lanzoni (2009) found that
even in the presence of the strong filtering action exerted by
cut-off processes, a close, although not yet complete, similarity

River Meandering

100

281

Flow direction

50
0
50
(a) 100

100

Flow direction

50
0
50
(b)

100

100

200

300

400
x

500

600

700

800

Figure 13 Example of output from numerical simulation model. Numerical simulations beyond cut-off: (a) subresonant case, b 9, t0.3,
ds0.005, E 10  8, dune covered bed. (b) superresonant case, b 13, t0.3, ds0.005, E10  8, dune-covered bed. Note the formation
of compound loops. Also note that sinuosity decreases as development proceeds. In the subresonant case: red 2.67; green 3.08; black thin1.80;
black thick 1.69. In the superresonant case: red 2.98; green 2.22; black thin 2.32; black thick 2.45. (Calculations kindly provided by S.
Lanzoni, 2005.) Reproduced from Seminara, G., 2006. Meanders. Journal of Fluid Mechanics 554, 271297, with permission from Cambridge
University Press.

with natural meandering planforms can be achieved only by


adopting a flow field model which yincludes superresonant
conditions. Both studies showed that a large number of morphological parameters are needed to characterize meander
morphology adequately, as Ferguson (1975) recognized.
Surprisingly, few comparisons of models have been made
to date, except the review by Camporeale et al. (2007), which
mainly considers the Ikeda et al. (1981), the Johanesson and
Parker (1989) and the Zolezzi and Seminara (2001) (ZS)
models, which build on each other. Flume experiments of
Zolezzi et al. (2005) confirm the theoretical positions that w/d
ratio of channels influence meander behavior and propagation of overdeepening. Recent analysis of data from 100
gravel-bed rivers also appear to support the ideas of controls
on subresonant and superresonant behavior and propagation
of change in downstream and upstream directions by the
bankfull aspect ratio (0.5 w/d) and the Shields stress (Zolezzi
et al., 2009). Dense vegetation and reduced gravel supply were
found to promote the subresonant regime. Frascati and Lanzoni (2009) compare synthetic long-term morphology produced from ZS derived modeling with observed patterns of
rivers and also find two distinct behaviors.
The other main group of numerical models is based on
kinematic or empirical relations, that is, not derived from
fundamental physical principles. These are mainly based on
migration or erosion rate relations to curvature, many of them
derived from Hickin and Nanson (1975). The original empirical relationship showed a nonlinear increase in migration
rate with increased curvature to some critical value (r/w 23)
beyond which the rate declined again. Guneralp and Rhoads
(2008) consider that the weakness of current theoretical
models aimed at predicting planform migration is that they
relate the rate of meander migration at a particular location to
the channel curvature at and upstream of that location. Using
a newly developed method of curvature analysis, Guneralp
and Rhoads (2009) demonstrated that the migrationcurvature

relation is much more complex and needs more sophisticated


characterization of the curvature. The results also indicate
that the spatial memory of the curvaturemigration relation
(i.e., the decay length of the spatial response) of simple bends
is longer than that of compound loops or multilobes. This
finding is in accord with the idea that simple bends migrate
downstream more than they migrate laterally, whereas compound loops migrate laterally more than they migrate downstream (Hooke and Harvey, 1983; Hooke and Yorke, 2010;
Seminara et al., 2001). The changing structure of the upstream
curvature effect on migration rates with increasing planform
complexity that Guneralp finds suggests that the relationship
between planform curvature and migration rates is highly
nonlinear and spatially variable.

9.16.6

Perspective and Synthesis

The following is a summary of some of the most important


conclusions on key aspects of river meanders. Some of these
have been known for a long time, but modern research has
substantiated the conclusions and provided greater detail on
mechanisms and controls.
Meanders vary in morphology, with a general scaling factor
but shape varying with bank and bed materials and resistance,
gradient, stage of bend evolution, tectonics, and other controls. Active meanders (rapidly changing) vary in width
through the bends; equal width meanders are a sign of stability. Discharge and sediment load are major controls and if
either is altered or varies then adjustment of meanders
morphology occurs, for example, in response to climate
change or human factors.
Commonly occurring sequences of meander development
and evolution have been identified, with bends tending to
develop from low curvature form through to high curvature
then compound forms and eventually being cut-off. Such

282

River Meandering

sequences are autogenic and inherent and can be observed in


timescales of decades on the most active meanders. Low
sinuosity meanders tend to migrate in a downstream direction,
whereas tighter bends grow cross valley. Rates accelerate, that is,
are nonlinear in relation to curvature, through to compound
stage but then may slow again.
Cut-offs may occur in various positions on meander loops
and may represent completion of the evolution sequence
(neck cut-offs) or may interrupt loop development (chute cutoffs). The overall role of cut-offs and the extent to which
meander behavior is chaotic is debated, but evidence of chaotic behavior is limited and cut-offs appear to maintain an
equilibrium of sinuosity in the long-term. However, perspective on this may depend on the rates of activity and the
timescale of analysis.
Flow configuration in meanders is generally helicoidal, but
the extent and complexity of development of secondary flow
cells and vortices is still much debated and does vary with
meander morphology, particularly curvature, transverse slope,
and width/depth ratio. Flow pattern influences meander
change because bank erosion is related mainly to the nearbank velocity. The secondary flow patterns bring sediment to
the inner bank where it forms point bars, but sediment of
different sizes show different trajectories, with coarser material
being carried outward and finer material inward.
Bank erosion occurs primarily by hydraulic erosion related
to peak flows. The mechanisms vary with bank composition,
height, state of moisture, or pore water pressures. Many banks
are composite and fail by cantilever failure. Rate of erosion is
related to near-bank velocity, which is influenced by bend w/d,
asymmetry, gradient, curvature, transverse slope, and to bank
resistance.
Alternate, freely migrating bars develop in straight channels
as a result of perturbation of flow and interaction with the
erodible boundary. Theory predicts that bars become fixed as
sinuosity develops and free (migrating) bars appear to be rare
in meandering channels. Theory suggests limits of sinuosity at
which they become fixed. Other bars are forced by curvature or
width, the most important being point bars, induced by
curvature. Fixed mid-channel bars can be common in meandering channels and influence bend development. They
may be widthinduced. Erosion and deposition may not be
equal at all times and thus width varies over time.
Debates on explanation by bar or bend theory can be resolved into the former being a temporal explanation and the
latter a spatial explanation. Thus in bar theory meanders
gradually develop, but in bend theory curves are already present. Theory of resonance indicates that changes may be
propagated in a downstream or upstream direction, depending primarily on w/d ratio, and sediment size in bends.
The debate on why rivers meander is regarded as largely
resolved from a theoretical point of view as a perturbation and
propagation of instability in flow against an erodible boundary in which there is a resonance at certain harmonics.

9.16.6.1

Future Research

Current debates still surround some issues and are the subject
of ongoing research or provide the agenda for future research.

These can be summarized as debates between: equilibrium or


evolution of morphology; extent and distribution of simple or
compound forms; the extent of linear or nonlinear relations;
the degree of allogenic compared with autogenic influences;
the influence of primary versus secondary flows in bend processes; and the designs and principles that should be used for
river management and restoration.
The behavior of bends in different environments, materials
and conditions and whether their morphology does stabilize
in some circumstances, with simple downstream migration, or
whether bends have an evolutionary trend and sequence of
development requires further investigation. It may be the case
that both stabilization and evolution are applicable but evident on different timescales so this needs to be researched
more fully. Not enough evidence is available of the dynamics
of change at various timescales and of the variation between
different environments and types of river. More dating of
sediments and of channel movement is needed to produce
trajectories and timescales of meander development and life
cycles, reworking of floodplains, and life cycles of bars.
Mechanisms of development of meander features need to be
studied in the field over decadal timescales that allow influence of variations in flow and conditions and of feedback
effects to be analyzed.
More investigation is needed of the mechanisms of development of compound forms and the interrelationships of
change in bed topography, locus of erosion and flow patterns,
particularly from field measurements. More discussion and
model development needs to incorporate nonlinear relations
and test their effects. Evidence for and understanding of
autogenic behavior in meandering needs to be further researched, and the extent to which morphology and changes in
form are inherent and predicted by theory and also the extent
to which they are influenced by local factors needs to be more
fully quantified. More field, empirical quantification of resistance, including that of vegetation, and its effects on migration rates and meander morphology is needed and further
research into the spatial scale of influence on forces, particularly of curvature and width variations.
Numerical and experimental models need to incorporate
feedback effects and adjustment of altered morphology. Major
developments in theoretical modeling have taken place, but
assumptions still need to be reduced, for example, fixed banks,
equal width, and steady flows. Some models are still far from
reality though they help to solve theoretical problems. Most
modeling is linear and reductionist, but nonlinear and holistic
approaches need to be more fully explored. The success of
some kinematic models may indicate that there is an emergent
behavior that is not apparent in reductionist approaches.
Development of realistic predictive models, incorporating the
spatial variability of controls such as gradient and resistance,
interaction of adjacent bends, autogenic sequences and feedbacks, and complexity is a major goal. All modeling requires
much more empirical and field validation.

9.16.7

Conclusions

Major developments have taken place in understanding river


meandering and the variation in characteristics and activities

River Meandering

that occur. Equilibrium and statistical relations of meander


morphology to major controls such as discharge are still
found to be applied at a general level. However, the complexity of some meander morphology and the widespread
occurrence of sequences of continuous evolution of some
meanders have now been recognized. The variation in mobility and morphology is seen to be related to major factors
such as stream power, gradient, and resistance of boundary
materials, and understanding of localized effects and feedback
mechanisms is increasing.
It is now largely accepted that fundamental bend theory
based on propagation of instabilities in channels explains the
development of meanders, though variations are still apparent
in the details of numerical modeling. The extent of nonlinearity
of behavior is still under discussion. Mathematical theory, processes and control of meandering have all been investigated
through experiments in flumes and model channels and by
numerical simulation but much more testing of models with
detailed geomorphological field evidence is still needed. Predictive models of specific meander behavior and movement are
still proving elusive, but understanding of dynamics of meander
mobility has progressed. Spatial variability of controls such as
gradient and resistance need to be incorporated. Future research
should bring together these approaches and provide further
insight; meander research is entering an exciting phase in which
this integration is being realized.

References
Abad, J.D., Garcia, M.H., 2008. Bed morphology in Kinoshita meandering channels:
Experiments and numerical simulations. River, Coastal and Estuarine
Morphodynamics: RCEM 2007 1 and 2, 869875.
Abad, J.D., Garcia, M.H., 2009. Experiments in a high-amplitude Kinoshita
meandering channel: 2. Implications of bend orientation on bed
morphodynamics. Water Resources Research, 45.
Ackers, P., Charlton, F.G., 1970. Slope and resistance of small meandering channels.
Proceedings of the Institution of Civil Engineers, Paper 73623, 349370.
Allen, J.R.L., 1970. Studies in fluviatile sedimentation a comparison of finingupwards cyclothems, with special reference to coarse-member composition and
interpretation. Journal of Sedimentary Petrology 40(1), 298323.
Alonso, A., Garzon, G., 1994. Quaternary evolution of a meandering gravel-bed river
in central Spain. Terra Nova 6(5), 465475.
Alvarado-Ancieta, C.C., Ettmer, B., 2008. Fluvial morphology and erosion in abrupt
curves of the Ucayali River, Peru. Ingenieria Hidraulica En Mexico 23(4), 6990.
Andrle, R., 1994. Flow structure and development of circular meander pools.
Geomorphology 9(4), 261270.
Andrle, R., 1996. Measuring channel planform of meandering rivers. Physical
Geography 17(3), 270281.
Anthony, D.J., Harvey, M.D., 1991. Stage-dependent cross-section adjustments in a
meandering reach of Fall River, Colorado. Geomorphology 4(34), 187203.
Aslan, A., Autin, W.J., Blum, M.D., 2005. Causes of river avulsion: insights from
the late Holocene avulsion history of the Mississippi River, USA. Journal of
Sedimentary Research 75(4), 650664.
Aswathy, M.V., Vijith, H., Satheesh, R., 2008. Factors influencing the sinuosity of
Pannagon River, Kottayam, Kerala, India: an assessment using remote sensing
and GIS. Environmental Monitoring and Assessment 138(13), 173180.
Bagnold, R.A., 1960. Some aspects of the shape of river meanders. US Geological
Survey Professional Paper 282-E.
Bak, P., 1996. How Nature Works. Springer (Copernicus), New York, 205 pp.
Bak, P., Chen, K., 1991. Self-organised criticality. Scientific American 264, 2633.
Bak, P., Tang, C., Wiesenfeld, K., 1987. Self-organised criticality: an explanation of
I/f noise. Physical Review Letters 59, 381384.
Bartholdy, J., Billi, P., 2002. Morphodynamics of a pseudomeandering gravel bar
reach. Geomorphology 42(34), 293310.

283

Bathurst, J.C., Thorne, C.R., Hey, R.D., 1977. Direct measurements of secondary
currents in river bends. Nature 269(5628), 504506.
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(12), 124141.
Beeson, C.E., Doyle, P.F., 1995. Comparison of bank erosion at vegetated and nonvegetated channel bends. Water Resources Bulletin 31(6), 983990.
Bennett, S.J., Simon, A., Kuhnle, R.A., 1998. Temporal variations in point bar
morphology within two incised river meanders, Goodwin Creek, Mississippi.
Water Resources Engineering 1 and 2, 14221427.
Biedenharn, D., Raphett, N., Montague, C., 1984. Long-term Stability of the
OuchitaRiver. In: Elliott, C.M. (Ed.), River Meandering, Proceedings of the
Conference Rivers 0 83, New Orleans, LA, pp. 126137.
Blanckaert, K., 2009. Saturation of curvature-induced secondary flow, energy losses,
and turbulence in sharp open-channel bends: laboratory experiments, analysis,
and modeling. Journal of Geophysical Research-Earth Surface, 114.
Blanckaert, K., de Vriend, H.J., 2004. Secondary flow in sharp open-channel bends.
Journal of Fluid Mechanics 498, 353380.
Blanckaert, K., De Vriend, H.J., 2005. Turbulence structure in sharp open-channel
bends. Journal of Fluid Mechanics 536, 2748.
Blondeaux, P., Seminara, G., 1985. A unified bar bend theory of river meanders.
Journal of Fluid Mechanics 157(AUG), 449470.
Bluck, B.J., 1971. Sedimentation in the meandering River Endrick. Scottish Journal
of Geology 7, 93138.
Bradley, C., Smith, D.G., 1984. Meandering channel response to altered flow regime
Milk River, Alberta and Montana. Water Resources Research 20(12), 19131920.
Braga, G., Gervasoni, S., 1989. Evolution of the Po river: an example of the
application of historic maps. In: Petts, G.E., Moller, H., Roux, A.L. (Eds.),
Historical Change of Large Alluvial Rivers: Western Europe. John Wiley and
Sons, Chichester, pp. 113126.
Brasington, J., Rumsby, B.T., McVey, R.A., 2000. Monitoring and modelling
morphological change in a braided gravel-bed river using high resolution GPSbased survey. Earth Surface Processes and Landforms 25(9), 973990.
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 of the United States of
America 106(40), 1693616941.
Brice, J.C., 1973. Meandering pattern of the White River in Indiana an analysis.
In: Morisawa, M. (Ed.), Fluvial Geomorphology. Binghamton State University,
New York, pp. 178200.
Brice, J.C., 1974. Evolution of meander loops. Bulletin of Geological Society of
America 85, 581586.
Brice, J.C., 1977. Lateral migration of the middle Sacramento River, California.
USGS Water Resources Investigation 77-43, 51 pp.
Brice, J.C., 1982. Stream Channel Stability Assessment. Federal Highway
Administration Report RHWA/RD-82/021, 41 pp.
Brice, J.C., 1984. Planform properties of meandering rivers. In: Elliott, C.M. (Ed.),
River Meandering, Proceedings of the Conference Rivers 0 83, New Orleans, LA,
pp. 115.
Bridge, J.S., 1992. A revised model for water flow, sediment transport, bed
topography, and grain-size sorting in natural river bends. Water Resources
Research 28, 9991013.
Bridge, J.S., 2003. Rivers and Floodplains. Blackwell, Oxford, 491 pp.
Bridge, J.S., Jarvis, J., 1976. Flow and sedimentary processes in meandering River
South Esk, Glen-Clova, Scotland. Earth Surface Processes and Landforms 1(4),
303336.
Burckhardt, J.C., Todd, B.L., 1998. Riparian forest effect on lateral stream channel
migration in the glacial till plains. Journal of the American Water Resources
Association 34(1), 179184.
Burge, L.M., Smith, D.G., 1999. Confined meandering river eddy accretions:
sedimentology, channel geometry and depositional processes. Fluvial
Sedimentology 28, 113130.
Camporeale, C., Perona, P., Porporato, A., Ridolfi, L., 2005. On the long-term
behavior of meandering rivers. Water Resources Research 41(12).
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.
Camporeale, C., Perucca, E., Ridolfi, L., 2008. Significance of cutoff in meandering
river dynamics. Journal of Geophysical Research-Earth Surface 113(F1).
Carey, W.C., 1969. Formation of floodplain lands. Journal of Hydraulic Engineering,
ASCE 95.
Carlston, C.W., 1965. The relation of free meander geometry to stream discharge
and its geomorphic implication. American Journal of Science 263, 864885.

284

River Meandering

Carson, M.A., 1986. Characteristics of high-energy meandering: rivers: the


Canterbury Plains, New Zealand. Geological Society of America, Bulletin 97,
886895.
Carson, M.A., Lapointe, M.F., 1983. The inherent asymmetry of river meander
planform. Journal of Geology 91(1), 4155.
Chandler, J., Ashmore, P., Paola, C., Gooch, M., Varkaris, F., 2002. Monitoring
river-channel change using terrestrial oblique digital imagery and automated
digital photogrammetry. Annals of the Association of American Geographers
92(4), 631644.
Chen, D., Duan, J.D., 2006. Simulating sine-generated meandering channel
evolution with an analytical model. Journal of Hydraulic Research 44(3),
363373.
Church, M., 1992. Channel morphology and typology. In: Calow, P., Petts, G.
(Eds.), The Rivers Handbook. Blackwell, Oxford, pp. 126143.
Church, M., Jones, D., 1982. Channel bars in gravel-bed rivers. In: Hey, R.D.,
Bathurst, J.C., Thorne, C.R. (Eds.), Gravel-Bed Rivers. Wiley, Chichester, pp.
291338.
Church, M., Rice, S.P., 2009. Form and growth of bars in a wandering gravel-bed
river. Earth Surface Processes and Landforms 34(10), 14221432.
Citterio, A., Piegay, H., 2009. Overbank sedimentation rates in former channel lakes:
characterization and control factors. Sedimentology 56(2), 461482.
Clayton, J.A., Pitlick, J., 2007. Spatial and temporal variations in bed load transport
intensity in a gravel bed river bend. Water Resources Research 43(2).
Constantine, J.A., Dunne, T., 2008. Meander cutoffs and the controls on the
production of oxbow lakes. Geology 36(1), 2326.
Coulthard, T.J., Van De Wiel, M.J., 2006. A cellular model of river meandering.
Earth Surface Processes and Landforms 31(1), 123132.
Crosato, A., 2009. Physical explanations of variations in river meander migration
rates from model comparison. Earth Surface Processes and Landforms 34(15),
20782086.
Crosato, A., Mosselman, E., 2009. Simple physics-based predictor for the number
of river bars and the transition between meandering and braiding. Water
Resources Research, 45.
Daniel, J.F., 1971. Channel movement of menadering Indiaina streams. USGS Prof.
Paper 732-A.
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), 1163.
Darby, S.E., Delbono, I., 2002. A model of equilibrium bed topography for meander
bends with erodible banks. Earth Surface Processes and Landforms 27(10),
10571085.
Dietrich, W., 1987. Mechanics of flow and sediment transport in river bends.
In: Richards, K.S. (Ed.), River Channels: Environment and Processes. Blackwell,
Oxford, pp. 179227.
Dietrich, W.E., Smith, J.D., Dunne, T., 1979. Flow and sediment transport in a sand
bedded meander. Journal of Geology 87(3), 305315.
Dinehart, R.L., Burau, J.R., 2005. Averaged indicators of secondary flow in repeated
acoustic Doppler current profiler crossings of bends. Water Resources Research
41(9).
Dort, W., Jr., 1978. Channel Migration Investigation, Historic Channel Change
Maps, Kansas River and Tributaries Bank Stabilization Component, Kansas and
Osage Rivers. Kansas Study, U.S. Army Corps of Engineers, Kansas City
District.
Dort, W., 2009. Historical Channel Changes of the Kansas River and its Major
Tributaries. American Geographical Society, New York, 80 pp.
Downward, S., 1995. Information from topographic survey. In: Gurnell, A., Petts,
G.E. (Eds.), Changing River Channels. Wiley, Chichester, pp. 303323.
Duan, J.G., Julien, P.Y., 2005. Numerical simulation of the inception of channel
meandering. Earth Surface Processes and Landforms 30(9), 10931110.
Dury, G.H., 1954. Contribution to a general theory of meandering valleys. American
Journal of Science 252(4), 193224.
Dury, G.H., 1955. Bed-width and wave-length in meandering valleys. Nature
176(4470), 3132.
Dury, G.H., 1958. Tests of a general theory of misfit streams. Transactions of the
Institute of British Geographers 25, 105118.
Dury, G.H., 1960. Misfit streams problems in interpretation, discharge, and
distribution. Geographical Review 50(2), 219246.
Eaton, B.C., Church, M., DavleS, T.R.H., 2006. A conceptual model for meander
initiation in bedload-dominated streams. Earth Surface Processes and Landforms
31(7), 875891.
Ebisemiju, F.S., 1993. The planimetric and geometric-properties of the channel
bends of low-energy streams in a forested humid tropical environment,
southwestern Nigeria. Journal of Hydrology 142(14), 319335.

Ebisemiju, F.S., 1994. The sinuosity of alluvial river channels in the seasonally wet
tropical environment-case-study of river elemi, southwestern Nigeria. Catena
21(1), 1325.
Elliott, C.M. (Ed.), 1984, River Meandering, Proceedings of the Conference Rivers
0
83, New Orleans, LA, 1036pp.
Engelund, F., 1974. Flow and bed topography in channel bends. Journal of the
Hydraulics Division, ASCE 100, 16311648.
Erskine, W., McFadden, C., Bishop, P., 1992. Alluvial cutoffs as indicators of former
channel conditions. Earth Surface Processes and Landforms 17, 2337.
Federici, B., Seminara, R., 2003. On the convective nature of bar instability. Journal
of Fluid Mechanics 487, 125145.
Ferguson, R.I., 1975. Meander irregularity and wavelength estimation. Journal of
Hydrology 26, 315333.
Ferguson, R.I., Parsons, D.R., Lane, S.N., Hardy, R.J., 2003. Flow in meander
bends with recirculation at the inner bank. Water Resources Research 39(11).
Fisk, H.N., 1944. Geological Investigation of the Alluvial Valley of the Lower
Mississippi River. Mississippi River Commission, Vicksburg, MI.
Frascati, A., Lanzoni, S., 2009. Morphodynamic regime and long-term evolution
of meandering rivers. Journal of Geophysical Research-Earth Surface,
114.
Frascati, A., Lanzoni, S., 2010. Long-term river meandering as a part of chaotic
dynamics? A contribution from mathematical modelling. Earth Surface Processes
and Landforms 35(7), 791802.
Friedkin, J.F., 1945. A Laboratory Study of the Meandering of Alluvial Rivers. US
Waterways Experiment Station, Vicksburg, Mississippi.
Frothingham, K.M., Rhoads, B.L., 2003. Three-dimensional flow structure and
channel change in an asymmetrical compound meander loop, Embarras River,
Illinois. Earth Surface Processes and Landforms 28(6), 625644.
Furbish, D.J., 1988. River-bend curvature and migration how are they related.
Geology 16(8), 752755.
Furbish, D.J., 1991. Spatial autoregressive structure in meander evolution.
Geological Society of America Bulletin 103(12), 15761589.
Fuller, I.C., Large, A.R.G., Milan, D.J., 2003. Quantifying channel development and
sediment transfer following chute cutoff in a wandering gravel-bed river.
Geomorphology 54(34), 307323.
Gautier, E., Brunstein, D., Vauchel, P., et al., 2007. Temporal relations between
meander deformation, water discharge and sediment fluxes in the floodplain of
the Rio Beni (Bolivian Amazonia). Earth Surface Processes and Landforms
32(2), 230248.
Gay, G.R., Gay, H.H., Gay, W.H., Martinson, H.A., Meade, R.H., Moody, J.A., 1998.
Evolution of cutoffs across meander necks in Powder River, Montana, USA.
Earth Surface Processes and Landforms 23, 651662.
Gilvear, D., Winterbottom, S., Sichingabula, H., 2000. Character of channel planform
change and meander development: Luangwa River, Zambia. Earth Surface
Processes and Landforms 25(4), 421436.
Gomez, B., Marron, D.C., 1991. Neotectonic effects on sinuosity and channel
migration, Belle Fourche River, western South-Dakota. Earth Surface Processes
and Landforms 16(3), 227235.
Goswami, U., Sarma, J.N., Patgiri, A.D., 1999. River channel changes of the
Subansiri in Assam, India. Geomorphology 30, 227244.
Guneralp, I., Rhoads, B.L., 2008. Continuous characterization of the planform
geometry and curvature of meandering rivers. Geographical Analysis 40(1),
125.
Guneralp, I., Rhoads, B.L., 2009. Empirical analysis of the planform
curvaturemigration relation of meandering rivers. Water Resources Research,
45.
Gurnell, A.M., 1997. Channel change on the River Dee meanders, 19461992, from
the analysis of air photographs. Regulated Rivers-Research & Management
13(1), 1326.
Gurnell, A.M., Downward, S.R., Jones, R., 1994. Channel planform change on the
River Dee meanders, 18761992. Regulated Rivers-Research & Management
9(4), 187204.
Gurnell, A., Peiry, J.-L., Petts, G., 2003. Using historical data in fluvial
geomorphology. In: Kondolf, G.M., Piegay, H. (Eds.), Tools in Fluvial
Geomorphology. Wiley, Chichester.
Harden, D.R., 1990. Controlling factors in the distribution and development of
incised meanders in the central Colorado plateau. Geological Society of America
Bulletin 102(2), 233242.
Harmar, O.P., Clifford, N.J., 2006. Planform dynamics of the Lower Mississippi
River. Earth Surface Processes and Landforms 31(7), 825843.
Harmel, R.D., Haan, C.T., Dutnell, R., 1999. Bank erosion and riparian vegetation
influences: Upper Illinois River, Oklahoma. Transactions of the Asae 42(5),
13211329.

River Meandering

Harvey, A.M., 2007. High sinuosity bedrock channels: response to rapid


incision examples in SE Spain. Revista Cuaternario & Geomorfologia 21(34),
2147.
Hasegawa, K. 1989. Studies on qualitative and quantitative prediction of meander
channel shift. In: Ikeda, S., Parker, G. (Eds.), River Meandering. AGU Water
Resources Monograph 12, pp. 215236.
Heritage, G., Hetherington, D., 2007. Towards a protocol for laser scanning in
fluvial geomorphology. Earth Surface Processes and Landforms 32(1), 6674.
Hey, R., Thorne, C., 1975. Secondary flows in river channels. Area 7, 191195.
Hickin, E.J., 1974. Development of meanders in natural river-channels. American
Journal of Science 274(4), 414442.
Hickin, E.J., 1978. Hydraulic factors controlling channel migration. In: DavidsonArnott, R., Nickling, W. (Eds.), Research in Fluvial Geomorphology, Proceedings
Fifth Guelph Symposium on Geomorphology, pp. 5966.
Hickin, E.J., 1978. Mean flow structure in meanders of Squamish River, BritishColumbia. Canadian Journal of Earth Sciences 15(11), 18331849.
Hickin, E.J., Nanson, G.C., 1975. The character of channel migration on the Beaton
River, Northeast British Columbia, Canada. Geological Society of America
Bulletin 86, 487494.
Hickin, E.J., Nanson, G.C., 1984. Lateral migration rates of river bends. Journal of
Hydraulic Engineering, ASCE 110(11), 15571567.
Hodskinson, A., Ferguson, R.I., 1998. Numerical modelling of separated flow in
river bends: model testing and experimental investigation of geometric controls
on the extent of flow separation at the concave bank. Hydrological Processes
12(8), 13231338.
Hooke, J.M., 1977. The distribution and nature of changes in river channel pattern.
In: Gregory, K.J. (Ed.), River Channel Changes. John Wiley, Chichester, pp.
265280.
Hooke, J.M., 1979. Analysis of the processes of river bank erosion. Journal of
Hydrology 42(12), 3962.
Hooke, J.M., 1980. Magnitude and distribution of rates of river bank erosion. Earth
Surface Processes 5(2), 143157.
Hooke, J.M., 1984. Changes in river meanders a review of techniques and results
of analyses. Progress in Physical Geography 8(4), 473508.
Hooke, J.M., 1995a. River channel adjustment to meander cutoffs on the River
Bollin and River Dane, northwest England. Geomorphology 14(3), 235253.
Hooke, J.M., 1995b. Processes of channel planform change on meandering
channels in the UK. In: Gurnell, A., Petts, G.E. (Eds.), Changing River Channels.
Wiley, Chichester, pp. 87116.
Hooke, J.M., 1986. The significance of mid-channel bars in an active meandering
river. Sedimentology 33(6), 839850.
Hooke, J.M., 1987. Changes in meander morphology. In: Gardiner, V. (Ed.),
International Geomorphology 1986 Part I. Wiley, Chichester, pp. 591609.
Hooke, J.M., 1997. Styles of channel change. In: Thorne, C., Hey, R., Newson, M.
(Eds.), Applied Fluvial Geomorphology for River Engineering and Management.
Wiley, Chichester, pp. 237268.
Hooke, J.M., 2003. River meander behaviour and instability; a framework for
analysis. Transactions of Institute of British Geographers 28, 238253.
Hooke, J.M., 2004. Cutoffs galore!: occurrence and causes of multiple cutoffs on a
meandering river. Geomorphology 61(34), 225238.
Hooke, J.M.M., 2007a. Complexity, self-organisation and variation in behaviour in
meandering Rivers. Geomorphology 91, 236258.
Hooke, J.M., 2007b. Spatial variability, mechanisms and propagation of change in
an active meandering river. Geomorphology 84, 277296.
Hooke, J.M., 2008. Temporal variations in fluvial processes on an active meandering
river over a 20-year period. Geomorphology 100(12), 313.
Hooke, J.M., Gautier, E., Zolezzi, G., 2011. River meander dynamics. Earth Surface
Processes and Landforms 36, 15501553.
Hooke, J.M., Harvey, A.M., 1983. Meander changes in relation to bend morphology
and secondary flows. In: Collinson, J., Lewin, J. (Eds.), Modern and Ancient
Fluvial Systems. International Association of Sedimentologists Special
Publications, Vol. 6, pp. 121132.
Hooke, J.M., Kain, R.J.P., 1982. Historical Change in the Physical Environment: A
Guide to Sources and Techniques. Butterworths, London.
Hooke, J.M., Redmond, C.E., 1989. Use of cartographic sources for analysis of
river channel change in Britain. In: Petts, G.E. (Ed.), Historical Changes on
Large Alluvial European Rivers. Wiley, Chichester, pp. 7993.
Hooke, J.M., Redmond, C.E., 1992. Causes and nature of river planform change.
In: Billi, P., Hey, R.D., Thorne, C.R., Tacconi, P. (Eds.), Dynamics of Gravel-Bed
Rivers. Wiley, Chichester, pp. 549563.
Hooke, J.M., Yorke, L., 2010. Rates, distributions and mechanisms of change in
meander morphology over decadal timescales, River Dane, UK. Earth Surface
Processes and Landforms 35(13), 16011614.

285

Hooke, J.M., Yorke, L., 2011. Channel bar dynamics on multi-decadal timescales in
an active meandering river. Earth Surface Processes and Landforms 36,
19101928.
Hooke, R.L.B., 1975. Distribution of sediment transport and shear-stress in a
meander bend. Journal of Geology 83(5), 543565.
Howard, A.D., 1992. Modelling channel migration and floodplain sedimentation in
meandering streams. In: Carling, P.A., Petts, G.E. (Eds.), Lowland Floodplain
Rivers. Wiley, Chichester, pp. 141.
Howard, A.D., Hemberger, A.T., 1991. Multivariate characterisation of meandering.
Geomorphology 4, 161186.
Howard, A.D., Knutson, T.R., 1984. Sufficient conditions for river meandering: a
simulation approach. Water Resources Research 20, 16591667.
Hudson, P.F., Kesel, R.H., 2000. Channel migration and meander-bend curvature in
the lower Mississippi River prior to major human modification. Geology 28(6),
531534.
Huggett, R., 2003. Fundamentals of geomorphology. Routledge, London, 386 pp.
Hughes, M.L., McDowell, P.F., Marcus, W.A., 2006. Accuracy assessment of
georectified aerial photographs: implications for measuring lateral channel
movement in a GIS. Geomorphology 74(14), 116.
Ikeda, S., Parker, G. (Eds.), 1989. River meandering. AGU Water Resources
Monograph 12, 485.
Ikeda, S., Parker, G., Sawai, K., 1981. Bend theory of river meanders 1. Linear
development. Journal of Fluid Mechanics 112(Nov), 363377.
Jackson, R.G., 1975. Velocity-bed-form-texture patterns of meander bends in lower
Wabash River of Illinois and Indiana. Geological Society of America Bulletin
86(11), 15111522.
Johannesson, H., Parker, G., 1989. Velocity redistribution in meandering rivers.
Journal of Hydraulic Engineering-ASCE 115(8), 10191039.
Julian, J.P., Torres, R., 2006. Hydraulic erosion of cohesive riverbanks.
Geomorphology 76, 193206.
Kale, V.S., 2005. The sinuous bedrock channel of the Tapi River, Central India: its
form and processes. Geomorphology 70(34), 296310.
Kalkwijk, J.P.T., deVriend, H.J., 1980. Computation of the flow in shallow river
bends. Journal of Hydraulic Research 18, 327342.
Kawai, S., Julien, P.Y., 1996. Point bar deposits in narrow sharp bends. Journal of
Hydraulic Research 34(2), 205218.
Keller, E.A., 1972. Development of alluvial stream channels: a five-stage model.
Geological Society of America Bulletin 83, 15311540.
Kellerhals, R.M., Church, M., Bray, I., 1976. Classification and analyisis of river
processes. Journal of the Hydraulics Division, ASCE 102, 813829.
King, P.B., Schumm, S.A. (Eds.), 1980. The Physical Geography of William Morris
Davis. Geo Books, Norwich.
Kinoshita, R., Miwa, H., 1974. River channel formation which prevents downstream
translation of transverse bars. Shinsabo 94, 1217.
Klein, M., 1985. The adjustment of the meandering pattern of the lower Jordan
River to change in water discharge. Earth Surface Processes and Landforms
10(5), 525531.
Knighton, A.D., 1972. Changes in a braided reach. Bulletin of the Geological
Society of America 83, 38133822.
Knighton, A.D., 1973. Riverbank erosion in relation to streamflow conditions, River
Bollin-Dean, Cheshire. East Midland Geographer 5, 416426.
Knighton, D., 1998. Fluvial Forms and Processes a New Perspective. Oxford
University Press, Oxford.
Kondolf, G.M., 2006. River restoration and meanders. Ecology and Society 11(2),
313330.
Kondratyev, N., 1968. Hydromorphic principles of computations of free
meandering and signs and indexes of free meandering. Soviet Hydrology 4,
309335.
Lagasse, P.F., Zevenbergen, L.W., Spitz, W.J., Thorne, C.R., 2004. Methodology
for predicting channel migration. NCHRP Web-Only Document 67 (Project 2416). Report prepared for TRB (Transportation Research Board of the National
Academies of the US) http://onlinepubs.trb.org/onlinepubs/nchrp/nchrp_w67.pdf
Lancaster, S.T., Bras, R.L., 2002. A simple model of river meandering and its
comparison to natural channels. Hydrological Processes 16(1), 126.
Lane, S.N., 2000. The measurement of river channel morphology using digital
photogrammetry. Photogrammetric Record 16(96), 937957.
Langbein, W., Leopold, L.B., 1966. River meanders theory of minimum variance.
US Geological Survey Professional Paper 422-H.
Lapointe, M.F., Carson, M.A., 1986. Migration patterns of an asymmetric
meandering river: the Rouge River, Quebec. Water Resources Research 22,
731743.
Larsen, E.W., Fremier, A.K., Girvetz, E.H., 2006. Modeling the effects of variable
annual flow on river channel meander migration patterns, Sacramento River,

286

River Meandering

California, USA. Journal of the American Water Resources Association 42(4),


10631075.
Lauer, J.W., Parker, G., 2008. Modeling framework for sediment deposition, storage,
and evacuation in the floodplain of a meandering river: theory. Water Resources
Research 44, 4.
Lawler, D.M., 1993. The measurement of river bank erosion and lateral channel
change a review. Earth Surface Processes and Landforms 18(9), 777821.
Lawler, D.M., Thorne, C.R., Hooke, J.M., 1997. Bank erosion, stability and retreat.
Chapter 7. In: Thorne, C., Hey, R., Newson, M. (Eds.), Applied Fluvial
Geomorphology for River Engineering and Management. Wiley, Chichester, pp.
137172.
Leliavsky, S., 1955, 1966. An Introduction to Fluvial Hydraulics. Dover Publications,
New York, 257 pp.
Leopold, L.B., 1994. A View of the River. Harvard University Press, Cambridge, MA,
298 pp.
Leopold, L.B., Langbein, W.B., 1966. River meanders. Scientific American 214(6),
60-&.
Leopold, L.B., Wolman, M.G., 1957. River Channel Patterns: Braided, Meandering
and Straight. US Geological Survey Professional Paper 282B, 85p.
Leopold, L.B., Wolman, M.G., 1960. River meanders. Geological Society of America
Bulletin 71, 769794.
Leopold, L.B., Wolman, M.G., Miller, J.P., 1964. Fluvial Processes in
Geomorphology. Freeman Press, San Francisco, 522 p.
Lewin, J., 1972. Late-stage meander growth. Nature Physical Science 240, 116.
Lewin, J., 1976. Initiation of bed forms and meanders in coarse-grained sediment.
Geological Society of America, Bulletin 87, 281285.
Lewin, J., 1978. Meander development and floodplain sedimentation: a case study
from mid-Wales. Geological Journal 13, 2536.
Lewin, J., 1983. Changes of channel patterns and floodplains. In: Gregory, K.J.
(Ed.), Background to Palaeohydrology. John Wiley & Sons, Chichester, UK, pp.
303319.
Lewin, J., Brindle, B.J., 1977. Confined meanders. In: Gregory, K.J. (Ed.), River
Channel Changes: Chichester, UK. John Wiley & Sons, pp. 221233.
Lewis, G.W., Lewin, J., 1983. Alluvial cutoffs in Wales and the Borderlands.
In: Collinson, J.D., Lewin, J. (Eds.), Modern and Ancient Fluvial Systems:
International Association of Sedimentologists. Special Publication, 6, pp. 145154.
Leys, K.F., Werritty, A., 1999. River channel planform change: software for historical
analysis. Geomorphology 29, 107120.
Lofthouse, C., Robert, A., 2008. Riffle-pool sequences and meander morphology.
Geomorphology 99(14), 214223.
Luchi, R., Hooke, J.M., Zolezzi, G., Bertoldi, W., 2010. Width variations and midchannel bar inception in meanders: River Bollin (UK). Geomorphology 119, 18.
Luppi, L., Rinaldi, M., Teruggi, L.B., Darby, S.E., Nardi, L., 2009. Monitoring and
numerical modelling of riverbank erosion processes: a case study along the Cecina
River (central Italy). Earth Surface Processes and Landforms 34(4), 530546.
Malik, I., 2006. Contribution to understanding the historical evolution of
meandering rivers using dendrochronological methods: example of the Mala
Panew River in southern Poland. Earth Surface Processes and Landforms
31(10), 12271245.
Markham, A.J., Thorne, C.R., 1992. Geomorphology of gravel-bed rivers. In: Billi,
P., Hey, R.D., Thorne, C.R., Tacconi, P. (Eds.), Dynamics of Gravel Bed Rivers.
Wiley and Sons, Chichester, pp. 433456.
Marren, P.M., McCarthy, T.S., Tooth, S., Brandt, D., Stacey, G.G., Leong, A.,
Spottiswoode, B., 2006. A comparison of mud- and sand-dominated meanders
in a downstream coarsening reach of the mixed bedrock-alluvial Klip River,
eastern Free State, South Africa. Sedimentary Geology 190(14), 213226.
McCarthy, T.S., Tooth, S., 2004. Incised meanders along the mixed bedrock-alluvial
Orange River, Northern Cape Province, South Africa. Zeitschrift Fur
Geomorphologie 48(3), 273292.
Mertes, L.A.K., Dunne, T., Martinelli, L.A., 1996. Channel-floodplain geomorphology
along the Solimoes-Amazon River, Brazil. Geological Society of America Bulletin
108, 10891107.
Micheli, E.R., Kirchner, J.W., Larsen, E.W., 2004. Quantifying the effect of riparian
forest versus agricultural vegetation on river meander migration rates, Central
Sacramento River, California, USA. River Research and Applications 20(5),
537548.
Montgomery, K., 1996. Sinuosity and fractal dimension of meandering rivers. Area
28(4), 491500.
Mosselman, E., 1995. A review of mathematical-models of river planform changes.
Earth Surface Processes and Landforms 20(7), 661670.
Mosselman, E., Tubino, M., Zolezzi, G., 2006. The overdeepening theory in river
morphodynamics: two decades of shifting interpretations. River Flow 2006 1 and
2, 11751181.

Nanson, G.C., 1980a. Point-bar and floodplain formation of the meandering


Beatton River, northeastern British-Columbia, Canada. Sedimentology 27(1),
329.
Nanson, G.C., 1980b. Regional trend to meander migration. Journal of Geology
88(1), 100108.
Nanson, G.C., Hickin, E.J., 1983. Channel migration and incision on the Beatton
river. Journal of Hydraulic Engineering ASCE 109(3), 327337.
Nicoll, T., Hickin, E., 2010. Planform geometry and channel migration of confined
meandering rivers on the Canadian prairies. Geomorphology 11, 3747.
Nikora, V.I., 1991. Fractal structures of river plan forms. Water Resources Research
27, 13271333.
Odgaard, A.J., 1989. River-meander model. 1. Development. Journal of Hydraulic
Engineering ASCE 115(11), 14331450.
Olsen, N.R.B., 2003. Three-dimensional CFD modeling of self-forming meandering
channel. Journal of Hydraulic Engineering ASCE 129(5), 366372.
O Neill, M.P., Abrahams, A.D., 1986. Objective identification of meanders and
bends. Journal of Hydrology 83, 337353.
Page, R.W., Nanson, G.C., 1982. Concave bank benches and associated floodplain
formation. Earth Surface Processes and Landforms 7, 529542.
Parker, C., Simon, A., Thorne, C.R., 2008. The effects of variability in bank material
properties on riverbank stability: Goodwin Creek, Mississippi. Geomorphology
101(4), 533543.
Parker, G., 1976. Cause and characteristic scales of meandering and braiding in
rivers. Journal of Fluid Mechanics 76(Aug 11), 457480.
Parker, G., 1998. River meanders in a tray. Nature 395(6698), 111112.
Parker, G., Andrews, E.D., 1985. Sorting of bed-load sediment by flow in meander
bends. Water Resources Research 21(9), 13611373.
Parker, G., Andrews, E.D., 1986. On the time development of meander bends.
Journal of Fluid Mechanics 162, 139156.
Parker, G., Diplas, P., Akiyama, J., 1983. Meander bends of high amplitude. Journal
of Hydraulic Engineering ASCE 109(10), 13231337.
Peakall, J., Ashworth, P.J., Best, J.L., 2007. Meander-bend evolution, alluvial
architecture, and the role of cohesion in sinuous river channels: a flume study.
Journal of Sedimentary Research 77(34), 197212.
Perucca, E., Camporeale, C., Ridolfi, L., 2005. Nonlinear analysis of the geometry
of meandering rivers. Geophysical Research Letters 32(3).
Phillips, J.D., 2007. Perfection and complexity in the lower Brazos River.
Geomorphology 91(34), 364377.
Phillips, J.D., 2009. Avulsion regimes in southeast Texas rivers. Earth Surface
Processes and Landforms 34(1), 7587.
Piegay, H., Bornette, G., Citteroi, A., Herouin, E., Moulin, B., Statiotis, C., 2000.
Channel instability as a control on silting dynamics and vegetation patterns
within perifluvial aquatic zones. Hydrological Processes 14, 30113029.
Pittaluga, M.B., Nobile, G., Seminara, G., 2009. A nonlinear model for river
meandering. Water Resources Research 45.
Pizzuto, J.E., 1984. Bank erodibility of shallow sandbed streams. Earth Surface
Processes and Landforms 9, 113124.
Pizzuto, J.E., 1994. Channel adjustments to changing discharges, Powder
River, Montana. Geological Society of America Bulletin 106(11),
14941501.
Pizzuto, J., 2009. An empirical model of event scale cohesive bank profile
evolution. Earth Surface Processes and Landforms 34(9), 12341244.
Pyrce, R.S., Ashmore, P.E., 2005. Bedload path length and point bar development
in gravel-bed river models. Sedimentology 52(4), 839857.
Rhoads, B.L., Miller, M.V., 1991. Impact of flow variability on the morphology of a
low-energy meandering river. Earth Surface Processes and Landforms 16(4),
357367.
Rhoads, B.L., Welford, M.R., 1991. Initiation of river meandering. Progress in
Physical Geography 15(2), 127156.
Richard, G.A., Julien, P.Y., Baird, D.C., 2005. Case study: modeling the lateral
mobility of the Rio Grande below Cochiti Dam, New Mexico. Journal of
Hydraulic Engineering ASCE 131(11), 931941.
Rinaldi, M., Johnson, P.A., 1997. Characterization of stream meanders for
stream restoration. Journal of Hydraulic Engineering ASCE 123(6),
567570.
Rittenour, T.M., Goble, R.J., Blum, M.D., 2005. Development of an OSL chronology
for Late Pleistocene channel belts in the lower Mississippi valley, USA.
Quaternary Science Reviews 24(2324), 25392554.
Rodnight, H., Duller, G.A.T., Tooth, S., Wintle, A.G., 2005. Optical dating
of a scroll-bar sequence on the Klip River, South Africa, to derive the
lateral migration rate of a meander bend. Holocene 15(6), 802811.
Rozovski, I.L., 1961. Flow of Water in Bends of Open Channels. National Science
Foundation, Washington.

River Meandering

Rowland, J.C., Lepper, K., Dietrich, W.E., Wilson, C.J., Sheldon, R., 2005.
Tie channel sedimentation rates, oxbow formation age and channel
migration rate from optically stimulated luminescence (OSL) analysis of
floodplain deposits. Earth Surface Processes and Landforms 30(9),
1161q1179.
Ruther, N., Olsen, N.R.B., 2007. Modelling free-forming meander evolution in a
laboratory channel using three-dimensional computational fluid dynamics.
Geomorphology 89(34), 308319.
Schattner, I., 1962. The Lower Jordan Valley. Hebrew University, Jerusalem.
Schumm, S.A., 1960. The shape of alluvial channels in relation to sediment type.
US Geological Survey Professional Paper 352-B.
Schumm, S.A., 1963. Sinuosity of alluvial rivers on the Great Plains. Geological
Society of America Bulletin 74, 10891100.
Schumm, S.A., 1977. The Fluvial System. John Wiley and Sons, New York, 338 pp.
Schumm, S.A., 2005. River Variability and Complexity. Cambridge University Press,
220 pp.
Schumm, S.A., Kahn, H.R., 1972. Experimental study of channel patterns.
Geological Society of America Bulletin. 83, 17001755.
Seker, D.Z., Kaya, S., Musaoglu, N., Kabdasli, S., Yuasa, A., Duran, Z., 2005.
Investigation of meandering in Filyos River by means of satellite sensor data.
Hydrological Processes 19(7), 14971508.
Seminara, G., 2006. Meanders. Journal of Fluid Mechanics 554, 271297.
Seminara, G., Zolezzi, G., Tubino, M., Zardi, D., 2001. Downstream and upstream
influence in river meandering. Part 2. Planimetric development. Journal of Fluid
Mechanics 438, 213230.
Shams, M., Ahmadi, G., Smith, D.H., 2008. Sensitivity of flow and sediment
transport in meandering rivers to scale effects and flow rate. Environmental
Engineering Science 25(5), 747756.
Shepherd, R.G., 1972. Incised river meanders evolution in simulated bedrock.
Science 178(4059), 409411.
Shi, C.X., Petts, G., Gurnell, A., 1999. Bench development along the regulated,
lower River Dee, UK. Earth Surface Processes and Landforms 24(2),
135149.
Simon, A., Curini, A., Darby, S.E., Langendoen, E.J., 2000. Bank and near-bank
processes in an incised channel. Geomorphology 35(34), 193217.
Slingerland, R., Smith, N.D., 1998. Necessary conditions for a meandering-river
avulsion. Geology 26(5), 435438.
Smith, C.E., 1998. Modeling high sinuosity meanders in a small flume.
Geomorphology 25(12), 1930.
Speight, J.G., 1965. Meander spectra of the Angabunga River. Journal of Hydrology
3, 115.
Stolum, H.H., 1996. River meandering as a self-organization process. Science
271(5256), 17101713.
Stolum, H.H., 1998. Planform geometry and dynamics of meandering rivers.
Geological Society of America Bulletin 110(11), 14851498.
Struiksma, N., Olesen, K.W., Flokstra, C., Devriend, H.J., 1985. Bed deformation in
curved alluvial channels. Journal of Hydraulic Research 23, 5779.
Sun, T., Meakin, P., Jossang, T., 2001. A computer model for meandering rivers
with multiple bed load sediment sizes 2. Computer simulations. Water
Resources Research 37(8), 22432258.
Sun, T., Meakin, P., Jossang, T., Schwarz, K., 1996. A simulation model for
meandering rivers. Water Resources Research 32(9), 29372954.
Swamee, P.K., Parkash, B., Thomas, J.V., Singh, S., 2003. Changes in the channel
pattern of River Ganga between Mustafabad and Rajmahal, Gangetic Plains since
18th century. International Journal of Sediment Research 18, 219231.
Swanson, J.E., 1984. Tazlina River meanders loop: a case history, In: Elliott, C.M.
(Ed.), River Meandering, Proceeding of the Conference Rivers 0 83, New Orleans,
LA, pp. 231239.
Tal, M., Paola, C., 2007. Dynamic single-thread channels maintained by the
interaction of flow and vegetation. Geology 35(4), 347350.
Termini, D., 2009. Experimental observations of flow and bed processes in largeamplitude meandering flume. Journal of Hydraulic Engineering ASCE 135(7),
575587.

287

Thompson, A., 1986. Secondary flows and the pool-riffle unit: a case study of the
processes of meander development. Earth Surface Processes and Landforms 11,
631641.
Thorne, C.R., Hey, R.D., 1979. Direct measurements of secondary currents at a river
inflection point. Nature 280(5719), 226228.
Thorne, C.R., Lewin, J., 1979. Bank processes, bed material movement and
planform development in a meandering river. In: Rhodes, D.D., Williams, G.P.
(Eds.), Adjustments of the Fluvial System. George Allen and Unwin, London, pp.
117137.
Thorne, C.R., Tovey, N.K., 1981. Stability of composite river banks. Earth Surface
Processes and Landforms 6(5), 469484.
Thorne, S.D., Furbish, D.J., 1995. Influences of coarse bank roughness on flow
within a sharply curved river bend. Geomorphology 12(3), 241257.
Tiegs, S.D., Pohl, M., 2005. Planform channel dynamics of the lower Colorado
River: 19762000. Geomorphology 69(14), 1427.
Timar, G., 2003. Controls on channel sinuosity changes: a case study of the Tisza
River, the Great Hungarian Plain. Quaternary Science Reviews 22(20),
21992207.
Tooth, S., Brandt, D., Hancox, P.J., McCarthy, T.S., 2004. Geological controls on
alluvial river behaviour: a comparative study of three rivers on the South African
Highveld. Journal of African Earth Sciences 38(1), 7997.
Tooth, S., McCarthy, T.S., Brandt, D., Hancox, P.J., Morris, R., 2002. Geological
controls on the formation of alluvial meanders and floodplain wetlands: the
example of the Klip River, eastern Free State, South Africa. Earth Surface
Processes and Landforms 27(8), 797815.
Trimble, S.W., 2008. The use of historical data and artifacts in geomorphology.
Progress in Physical Geography 32(1), 329.
Tubino, M., Repetto, R., Zolezzi, G., 1999. Free bars in rivers. Journal of Hydraulic
Research 37(6), 759775.
Tubino, M., Seminara, G., 1990. Free forced interactions in developing meanders
and suppression of free bars. Journal of Fluid Mechanics 214, 131159.
Uribelarrea, D., Perez-Gonzalez, A., Benito, G., 2003. Channel changes in the
Jararna and Tagus rivers (central Spain) over the past 500 years. Quaternary
Science Reviews 22(20), 22092221.
Whiting, P.J., Dietrich, W.E., 1993a. Experimental studies of bed topography and
flow patterns in large-amplitude meanders. 1. Observations. Water Resources
Research 29(11), 36053614.
Whiting, P.J., Dietrich, W.E., 1993b. Experimental studies of bed topography and
flow patterns in large-amplitude meanders. 2. Mechanisms. Water Resources
Research 29(11), 36153622.
Wolfert, H.P., Maas, G.J., 2007. Downstream changes of meandering styles in the
lower reaches of the River Vecht, the Netherlands. Netherlands Journal of
Geosciences Geologie En Mijnbouw 86, 257271.
Wolman, M.G., Brush, L.M., 1961. Factors controlling the size and shape of stream
channels in coarse, noncohesive sands. US Geological Survey, Professional
Paper 282-G, 183210.
Zeng, J., Constantinescu, G., Blanckaert, K., Weber, L., 2008. Flow and bathymetry
in sharp open-channel bends: experiments and predictions. Water Resources
Research 44(9).
Zhang, B., Ai, N.S., Huang, Z.W., Yi, C.B., Qin, F.C., 2008. Meanders of the Jialing
River in China: morphology and formation. Chinese Science Bulletin 53(2),
267281.
Zhang, B.S., Fu, J.M., 2003. Training direction for meandering lower Yellow River.
Proceedings of the First International Yellow River Forum on River Basin
Management, Vol. Ii, pp. 299305.
Zolezzi, G., Guala, M., Termini, D., Seminara, G., 2005. Experimental observations
of upstream overdeepening. Journal of Fluid Mechanics 531, 191219.
Zolezzi, G., Luchi, R., Luchi, R., Tubino, M., 2009. Morphodynamic regime of
gravel bed, single-thread meandering rivers. Journal of Geophysical ResearchEarth Surface, 114.
Zolezzi, G., Seminara, G., 2001. Downstream and upstream influence in river
meandering. Part 1. General theory and application to overdeepening. Journal of
Fluid Mechanics 438, 183211.

288

River Meandering

Biographical Sketch
Janet Hooke is Professor of Physical Geography in the School of Environmental Sciences at the University of
Liverpool, UK. She obtained her BSc from Bristol University and her PhD from Exeter University. She is a fluvial
geomorphologist specializing in research on river meanders and river channel changes, particularly the spatial and
temporal dynamics of morphological changes on event to historical timescales, the impact of hydrological
variations and the analysis of sediment processes and fluxes. In addition to publishing more than 80 research
papers, books, and book chapters, she has advised on environmental management of rivers and catchments,
especially in relation to erosion and flooding. She has managed more than 30 research projects, including
Coordinator of an EU project on combating land degradation. She works in both semiarid and humid fluvial
systems. In 2009, she was awarded the Busk Medal by the Royal Geographical Society of UK for her field research
into river systems and their conservation and for demonstrating the contribution of physical geography to
environmental management, conservation, and policy. She is a past Chair of the British Society for
Geomorphology.

You might also like