You are on page 1of 18

Advances in Water Resources 104 (2017) 127144

Contents lists available at ScienceDirect

Advances in Water Resources


journal homepage: www.elsevier.com/locate/advwatres

Sediment heterogeneity and mobility in the morphodynamic


modelling of gravel-bed braided rivers
Umesh Singh a,b, Alessandra Crosato a,c,, Sanjay Giri d, Murray Hicks e
a
UNESCO-IHE, Westvest 7, 2601 DA Delft, The Netherlands
b
University of Trento, Department of Civil, Environmental and Mechanical Engineering, via Mesiano 77, 38123 Trento, Italy
c
Delft University of Technology, Faculty of Civil Engineering and Geosciences, Stevinweg 1, 2628 CN Delft, The Netherlands
d
Deltares, River Engineering & Morphology, Inland Water System, Rotterdamseweg 185, 2629 HD Delft, The Netherlands
e
National Institute of Water & Atmospheric Research (NIWA), PO Box 8602, 8440 Christchurch, New Zealand

a r t i c l e i n f o a b s t r a c t

Article history: The effects of sediment heterogeneity and sediment mobility on the morphology of braided rivers are
Received 24 March 2016 still poorly studied, especially when the partial sediment mobility occurs. Nevertheless, increasing the
Revised 1 February 2017
bed sediment heterogeneity by coarse sediment supply is becoming a common practice in river restora-
Accepted 5 February 2017
tion projects and habitat improvement all over the world. This research provides a step forward in the
Available online 14 February 2017
identication of the effects of sediment sorting on the evolution of sediment bars and braiding geome-
Keywords: try of gravel-bed rivers. A two-dimensional morphodynamic model was used to simulate the long-term
Gravel-bed braided rivers developments of a hypothetical braided system with discharge regime and morphodynamic parameters
Morphodynamic modelling derived from the Waimakariri River, New Zealand. Several scenarios, differing in bed sediment hetero-
Delft3D geneity and sediment mobility, were considered. The results agree with the tendencies already identied
Graded sediment in linear analyses and experimental studies, showing that a larger sediment heterogeneity increases the
Variable discharge
braiding indes and reduces the bars length and height. The analyses allowed identifying the applicability
Meyer Peter and Mller formula
limits of uniform sediment and variable discharge modelling approaches.
2017 Published by Elsevier Ltd.

1. Introduction tated bars form the habitat even for mammals. Shape statistics as-
sociated with bars such as shore-line length are used to dene in-
Braided rivers are characterized by a dynamic system of di- terface between aquatic and terrestrial habitats, in ecological mod-
verging and converging channels separated by bars (Lane, 1957; els (Tockner et al., 2006; van der Nat et al., 2003).
Leopold and Wolman, 1957; Miall, 1977; Bertoldi et al., 2009a). Several human interventions in the past century, such as ow
Gravel bed braided rivers are common in pro-glacial environments alterations, gravel mining, catchment scale sediment management
(e.g. Rust, 1972) and piedmont areas (e.g. Surian and Rinaldi, 2003) etc have converted braided reaches into transition or single thread
where the discharge is strongly variable and sediments are het- rivers (e.g. Piegay et al, 2006; Surian and Rinaldi, 2003). As such,
erogeneous. Although gravel-bed braided rivers are valued less for Braided river ecosystems are among the most endangered in the
the aesthetic point of view (Lay et al., 2013), they have high eco- world (Sadler et al., 2004). Flow augmentation, such as ow and
logical importance. Within a braided river corridor, a complex and ood pulses, (Bertoldi et al., 2010; Tockner et al., 20 0 0) and coarse
dynamic display of aquatic, amphibian and terrestrial habitat el- sediment supply (Binder, 2004; Liebault et al, 2008; Piegay et al,
ements are present which allows co-existence of many species 20 09; Rinaldi et al., 20 09; Surian et al., 20 09) are some widely
within the river corridor (e.g. Tockner et al., 2006). Bars are units practiced measures to restore and manage gravel bed braided
which provide terrestrial habbitat within the active river corridor. rivers. However, understanding of natural complexity and dyn-
They also serve as resting places for aviary species whereas vege- maics of braided rivers are required for developing their sustain-
able management schemes (Ward et al., 2001; Peigay et al, 2006).
Non-uniform bed material composed of mixtures of gravel and

Corresponding author. sand render the morphodynamic study of gravel-bed braided rivers
E-mail addresses: umesh.singh@unitn.it (U. Singh), a.crosato@unesco-ihe.org even more complex. Most laboratory studies on gravel-bed braided
(A. Crosato), sanjay.giri@deltares.nl (S. Giri), murray.hicks@niwa.co.nz (M. Hicks).

http://dx.doi.org/10.1016/j.advwatres.2017.02.005
0309-1708/ 2017 Published by Elsevier Ltd.
128 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

rivers have focused in braiding mechanism (Ashmore, 1982; 2. Model description


Ashmore, 1991a), channel morphology (Hundey and Ashmore,
2009; Kasparak et al, 2015) and vertical grain sorting dur- 2.1. Delft3D
ing braided channel evolution (Leduc et al., 2015). Lanzoni and
Tubino (1999) extended the classical linear bar theory by intro- The numerical model was developed using the physics-based,
ducing sediment non-uniformity. Their analysis shows that non- fully-non-linear, open-source software Delft3D [www.deltares.nl].
uniform sediment results in a reduction of alternate bars ampli- The model solves the unsteady shallow-water equations with hy-
tude and length. Lanzoni (20 0 0a,b) performed several ume ex- drostatic approximation (Lesser et al., 2004) in two or three di-
periments on alternate bar development with uniform and non- mensions (2DH). Flow, sediment transport and bottom updating
uniform sediment to verify his previous theoretical work. He ob- are computed at small time steps (Roelvink, 2006).
served that sediment sorting indeed causes a reduction of the bar Bars in river channels create ow bifurcations and conuences
amplitude, but has no consistent effects on bar length. Teramoto and induce curvature of the streamlines. For this, it is impor-
and Tsujimoto (2006) studied the effects of sediment size hetero- tant that the model takes into account the effects of the spi-
geneity of bed materials on short-term development of multiple ral motion that arises in curved ow (Mosselman and Le, 2015).
bars by means of a numerical model and a linear stability analysis. Schuurman and Kleinhans (2011) showed that with a relatively
They observed that the bar mode increases if sediment heterogen- coarse computational grid a 2DH model, with parameterized spi-
ity increases. ral motion, gives results on the large scale bar pattern statistics
Linear and weakly non-linear analyses describe the initial that are comparable with the results of a fully 3D model. So, to
stages of the river bed development starting from a at bed (e.g. limit computational time, we used a depth-averaged (2DH) model
Callander, 1969; Engelund, 1970; Engelund and Skovgaard, 1973; with a parameterization of two relevant 3D effects of the spiral
Parker, 1976; Seminara and Tubino, 1989; Colombini et al., 1987; motion (cf. Blanckaert et al., 2003): the redistribution of the main
Schielen et al., 1993). For large width-to depth ratios, the initial ow velocity in transverse direction due to the secondary-ow
stages are characterized by the formation of many small migrating convection and the correction of the sediment transport direction,
bars, resulting in high braiding intensity. Long-term development which would otherwise coincide with the direction of the depth-
are characterized by the decrease of the braiding intensity, which averaged ow velocity vector. This approach has already been suc-
is due to progressive bar merging, resulting in larger and less mo- cessfully used to model the morphological behaviour of braided
bile bars. These developments can be described by fully-non-linear rivers by Jagers (2003), Marra (2008), Crosato and Saleh (2011),
models (e.g. Enggrob and Tjerry, 1999; Nicholas, 2013; Schuurman Schuurman et al. (2013) and the long-term evolution of alternate
et al., 2013), since the non-linear terms in the equations describ- bars by Crosato et al. (2011,2012).
ing the river morphodynamic evolution play a crucial role for bar The model includes a wetting-drying procedure, for which all
merging. cells having water depth smaller than a certain depth (in our case
To summarize, the effects of sediment heterogeneity is studied 5 cm) are considered dry. To simulate the widening of channels be-
on long-term evolution of alternate bars and the effects on braid- tween bars that become exposed during low water ows, a simple
ing intensity is studied only at the initial stage of the bed evolu- erosion formulation is applied at the margin between wet and dry
tion. So there exists a gap in understanding on what type of effects cells, according to which the model assigns a part of the erosion
sediment heterogeneity has on the long-term morphodynamics of occurring inside wet cells to the adjacent dry cells (van der Wegen
bars in gravel-bed braided rivers. Based on the results of previ- and Roelvink, 2008; Schuurman et al., 2013).
ous work, we can expect the long-term morphological evolution of Non-uniform sediment processes are modelled by: (i) subdivid-
gravel-bed braided rivers to be affected by sediment sorting pro- ing the sediment mixture into a number of sediment fractions, (ii)
cesses (Powell, 1998). We can also expect to nd important dif- applying a transport formulae and a mass conservation equation
ferences depending on whether the bed sediment is always fully for each separate fraction, (iii) applying hiding-exposure correc-
mobile (all fractions are mobile) or only partially mobile (only the tions for the critical shear stress of each fraction, (iv) considering
smaller fractions are mobile), since partial mobility may result in an active transport layer participating in sedimentation and ero-
local bed armouring and bar stabilization (e.g. Hunziker and Jaeggi, sion, and (v) considering a book-keeping layer or substratum which
20 02; Parker, 20 07). Numerical modelling of braided rivers with has become inactive due to sedimentation (Mosselman, 2005; Sloff
erodible banks by Sun et al. (2015) shows that bed armouring in and Ottevanger, 2008) based on Hiranos (1971) model. The use of
the channels close to the banks reduces bank erosion and that book-keeping layers (Sloff et al., 2001) allows the model to register
higher sediment heterogeneity increases these effects, resulting in the composition of the deposited sediment and makes the model
narrower channels. more robust (Blom, 2003).
The main objectives of this research were to identify the ef- Each fraction is dened by the minimum and the maxi-
fects of sediment heteroginity on the evolution of bar charachter- mum sediment size of the fraction, the model then assumes a
istics and braiding degree of gravel-bed braided rivers, for partial log-uniform distribution between the extremes of each fraction,
and full sediment mobility conditions. We analysed the effects of whereas the median diameter represents the fraction in the com-
sediment sorting by varying the sediment grain size distribution putation.
in a hypothetical straight river channel with non-erodible banks To properly simulate sorting processes, we adapted the
by means of a two-dimensional (2D) fully-non-linear morphody- Meyer Peter and Mller (MPM) (1948) formula by including
namic model. The model settings are based on the Waimakariri Parker et al. (1982) hiding-exposure formulation. The transport of
River (Fig. 1) near Christchurch (New Zealand) to compare results each individual fraction was predicted as follows:
broadly against a prototype natural river, but without the inten-  
tion of reproducing the evolution of this river in particular. We qs,i = pi 8 gD350,i (i 0.047 )3 (1)
carried out the sensitivity analyses between a variable hydrograph
and a constant discharge regime and several other model param- where qs, i is the sediment fraction transport per unit width,
eters, which allowed us to make important choices on their val- pi is the percentage of occurrence of the ith sediment fraction,
ues and analyse their effects on numerical modelling of gravel-bed  = ( s )/ is the relative sediment density with s = density of
braided rivers. sediment and = density of water, D50,i is the median sediment
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 129

Fig. 1. Braided reach of the Waimakariri River near Cross-Bank. Typical compound bar is highlighted by red circle. Flow is from left to right.

diameter of the ith sediment fraction, = (C/C90 )3/2 is the ripple fraction in the subsurface layer psl,i , so,
factor in which C is the Chezys roughness, C90 = 18log 10 (12h/D90 ) 
pa,i i f z/ t>0
corresponds to grain Chezys coecient, D90 is the 90th percentile p f,i = psl,i i f z/ t<0
(4)
grain diameter and h is the water depth.
In the adapted MPM formula (Eq. (1)), i = u2 (C2 D50, i ) 1 is The formulation by Koch and Folkstra (1980) extended by
the Shields number, u is depth-averaged velocity and is Parker Talmon et al. (1995) was used to model the effect of bed slope
et al s hiding-exposure coecient: on sediment transport direction:
   1
= (Dm /D50,i ) (2) 1 zb 1 zb
tan i = sin t cos t (5)
where Dm is the mean sediment particle size and is the Parker
fi y fi x
et al.s exponent Where, i is the angle between the transport direction of ith sed-
Hunziker and Jaeggi (2002) and Wong and Parker (2006) re- iment fraction with the depth-averaged ow direction; t is the
analysed data sets used by Meyer Peter and Mller (1948), and angle between the direction of the near bed ow and the depth
proposed several modications. Hunziker and Jaeggi (2002) ex- averaged ow direction calculated as
tended the MPM formula to non-uniform sediment, decreased
the constant from 8 to 5 and slightly increased the value of tan t = A(u2 + v2 )1/2 (6)
the critical Shields number to 0.05 (instead of 0.047). Wong and In Eq. (4), A is the coecient weighing the inuence of helical
Parker (2006) and Huang (2010) found that in most cases the ow on the direction of bed shear stress; u and v are the depth-
form-drag correction is unnecessary. Wong and Parker re-evaluated averaged ow velocity, in x and y direction, respectively. In Eq. (3),
the formula by modifying the constant and the value of the critical fi is a dimensionless parameter weighing the inuence of the grav-
Shields number to 3.97 and 0.0495, respectively. ity pull along the inclined bed for the ith sediment fraction. It is
Hiranos (1971) single-layer model was used for the conserva- given by
tion of mass of the individual sediment fractions:
fi = Ash i sh (D50,i /h )Csh (D50 /D50,i )Dsh
B
(7)
( pa,i ) z
(1 ) + (1 ) p f,i + Mor f ac
 t
 t
where Ash, Bsh and Csh and Dsh are calibration parameters. D50/h
( pi qs ) ( pi qs ) in the case of uniform sediment is implemented as an analogy to
+ =0 (3) a similar parameter for the effect of bedforms, D50 /D50,i is imple-
x y
mented to incorporate the effects of hiding and exposure (Sloff and
Where is the porosity, taken as 40% (Jansen et al., 1979), pa,i is Mosselman, 2012).
the percentage of occurrence of the ith fraction in the active layer,
pf,i is the percentage of occurrence of the ith fraction in the sedi- 2.2. The Waimakariri river
ment ux between the active and sub-surface layers, pi is the per-
centage of occurrence of the ith fraction in the transported mate- The Waimakariri originates from the Southern Alps in New
rial, is the active layer thickness and z is elevation of the inter- Zealand and ows into the Pacic Ocean. In its lower course
face between active and sub-surface layer, Morfac is the morpho- through the Canterbury plains, the Waimakariri shows a typical
logical acceleration factor used to accelerate long-term morpholog- braided nature, except at Lower Gorge, where it is constricted by
ical development. It is important to note that pf, i is equal to pa,i in rocky banks. More downstream, the river width is constricted rst
conditions of sediment deposition and equal percentage of the ith by terraces in Pleistocene fan deposits and then, after the river
130 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

has emerged onto its own Holocene fan, by embankments and


groins designed to protect the city of Christchurch from ood-
ing (Hicks et al., 2002). Several kilometers from the coast, the
river narrows and undergoes a gravel-sand transition where it en-
counters the low gradient Holocene coastal plain. A 13 km long
reach, belonging to the part that is conned by embankments,
is selected as the real river example to our modelling exercise
(Fig. 1). The upstream end of the reach is located at 4327 46
latitude and 17218 51 longitude; the downstream end is located
at 4326 44 latitude and 17228 56 longitude, at a site known
as "Cross-Bank" (after Griths, 1979). In the selected reach, the
river forms two mildly-curved bends (averaged sinuosity 1.1); the
river width varies between 600 m and 1000 m, but it is most often
around 900 m. The river width is calculated as the sum of width
of individual channels and non-vegetated or sparsely vegetated
bars.
All topographic features required for this study were de-
rived from a 5 m resolution DEM, based upon LIDAR and asso-
ciated image-based bathymetry surveys conducted in July 2003, Fig. 2. Flow duration curve of the Waimakariri River near State Highway Bridge
with root-mean-square error on bed levels of 0.2 to 0.3 m showing the uniform ow (black diamond) used in the model setup and the non-
(Hicks et al., 2008). From the longitudinal bed-level proles of uniform ow runs (continuous black line) used for sensitivity analysis. The cross
centre line, right and left banks we obtained an average bed indicates the discharge below which sediment mobility becomes negligible at uni-
form ow conditions.
slope of 0.005, which is close to the one previously estimated by
Carson and Griths (1989) (0.0048) for the 30 km upstream of
Cross-Bank.
Arial photographs taken during the 2003 survey and Google
Earth images show that the bed topography is dominated by large-
scale compound bars intersected by a number of small channels
and delimited by large morphodynamically-active channels (Fig. 1).
The reach-averaged bar amplitude, computed as the elevation dif-
ference between the top 10% and the smallest 10% detrended val-
ues of bed levels is 1.8 m. The reach-averaged length of the com-
pound bars is about 10 0 0 m. To derive the reach-averaged braiding
index of the river, we analysed 43 cross-sections extracted from
DEM at an interval of 250 m. From the measured bed topography,
we counted the bars higher than 0.5 m from the adjacent channel
bottom and then derived the corresponding bar mode (Crosato and
Mosselman, 2009). The braiding index was found to vary along the
reach, ranging between 5 and 18, which is mainly due to chan-
nel width variations, the reach-averaged value being equal to ap-
proximately 10. In an almost straight, 900 m wide, 1.6 km long sub-
reach, the braiding index is 10.7.
For this study, we used the discharge time series measured
from 1967 to 2006 at the State Highway 1 Bridge, about 11 km Fig. 3. Grain size distributions of Waimakariri River and the different sediment het-
downstream from Cross-Bank, without any transformations. This erogeneity scenarios. Each curve of the scenarios is represented by ve fractions.
The black dots correspond to the median grain size of each fraction. The uniform
is justied by the absence of signicant tributaries in the inter-
sediment scenario is represented by a grain size of 19 mm (average of mean diam-
vening reach and by minimal loss of discharge into the gravel bed eters).
(Nicholas, 20 0 0). The typical annual hydrograph exhibits two dis-
tinct wet seasons in autumn and spring, the mean river ow being
120 m3 /s and the mean annual peak ow 1520 m3 /s (Carson and a coarse grid. In the attempt to accommodate accuracy of the re-
Griths, 1989). The lowest value of gravel-entraining discharge sults with the necessity to restrain computational time, the grid
in the Waimakariri River is 90 m3 /s (Carson and Griths, 1989), was made by rectangular grid cells 10 m long in transverse direc-
which is exceeded only 50 days per year (Fig. 2). tion and 20 m long in longitudinal direction, which resulted in 90
Bed material was sampled at three locations by North Canter- cells in the transverse direction. Model stability was assured by a
bury Catchment Board (NCCB) in 1979. Sediment sizes distribution computational time-step of 3 seconds.
at the middle part of the reach is shown in Fig. 3. In general, the A Mannings coecient equal to 0.031 m1/3 /s (assuming a di-
bed material can be classied as poorly sorted (geometric standard mensionless conversion factor k = 1) was derived based on direct
deviation 6.75), with median size of 25.4 mm. measurements of discharge, water levels, wetted area, perimeter
and longitudinal slope of a 4 km-long part of the selected reach by
2.3. Model setup Hicks et al. (2002). From this, our model computed the equivalent
Chzys coecient as C = (h1/6 )/n, where C is Chezys coecient, h
The model domain is a 15,0 0 0 m long and 900 m wide straight is the local ow depth and n is Mannings coecient, at every time
channel with xed banks having a longitudinal bed slope of 0.005. step and for every grid cell. In this way, the value of Chzys coef-
Schuurman and Kleinhans (2011) show that the large-scale statis- cient becomes larger (smaller ow resistance) as the water depth
tics of the bar pattern are independent of computational grid size, increases. The diffusion due to 2D turbulence was specied by a
but a ne grid describes the shape of compound bars better than constant horizontal eddy viscosity parameter (Lesser et al., 2004),
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 131

a value of 1 m2 /s was used based on previous experience from 2D sediment grain size distribution of the Waimakariri River repre-
modelling of the Dutch Rhine River branches (e.g. Kleinhans et al., sents the base-case scenario. Then two granulometric curves with
2008). the same median diameter, D50 (27 mm), but different size ranges
Carson and Griths (1989) predicted the long-term gravel were generated. The largest sediment sizes were increased and the
transport rate of the Waimakariri River near the reference area smallest sizes decreased, although maintaining the smallest limit
reasonably well using the standard MPM formula. For this, we of 0.0625 mm, in the curve representing more poorly sorted sedi-
used the original formulation, without taking into account sug- ment, with higher size heterogeneity than the base case (Table 1,
gested adjustments (Huang, 2010; Hunziker and Jaeggi, 2002; poorly sorted scenarios, and Fig. 3). The largest sediment sizes
Wong and Parker, 2006). Parker et al.s exponent with the value were decreased and the smallest sizes increased for the well sorted
of 0.75 was used to model hiding and exposure effect (Eq. (2)). sediment scenarios, with less size heterogeneity than the base-case
This approach was chosen after comparison with several other (Table 1, well sorted scenarios, and Fig. 3). The geometric standard
approaches, because it provided sediment transport rate results deviation of poorly sorted scenarios, base case and well sorted sce-
that best resembled the ones of Wilcock and Crowe (2003)s for- narios are 6.88, 5.38 and 2.35 respectively. Finally, the study also
mula (not implemented in the numerical model). This surface- includes the scenarios with uniform sediment (no size heterogene-
based formula was used as a reference because it had been de- ity), having particle diameter equal to the average geometric mean
rived experimentally for graded sediments in conditions close to diameter Dm (19 mm) of the mixtures
real rivers and validated on the eld with some modications by We obtained the computational scenarios P0, P1, P2 and P3 in
Gaeuman et al. (2009). which P0 represents the base-case scenario; P1 (highest size het-
Both active and book-keeping layers were given a thickness erogeneity) is the scenario with (more) poorly sorted sediment if
of 1.25 m, corresponding to approximately the reach-averaged half compared to P0; P2 (lowest heterogeneity) is the scenario with
maximum bar amplitude (calculated as the elevation difference be- well sorted sediment if compared to P0. Because not all sediment
tween the highest point of the bar and deepest point in the chan- fractions are always mobile, these scenarios represent partial mo-
nel). Schuurman et al. (2013) showed that the morphological evo- bility conditions. Finally, P3 is the uniform sediment scenario, for
lution of braided rivers was marginally affected by the morpho- which the sediment is necessarily (fully) mobile.
logical acceleration factor that is imposed to speed up morphody- For these scenarios, the sediment transport rate was computed
namic computations. So, a morphological acceleration factor (mor- using the MPM formula, which refers to the mean sediment di-
fac) equal to 25 was used to accelerate long-term morphological ameter of the mixture, with its standard critical threshold value of
evolution (Eq. (3)). 0.047 (Eq. (1)). The hiding-exposure coecient for each individual
To model the effects of bed slope on sediment transport fraction, , was computed using Eq. (2). We checked the sediment
direction, we considered to use the values recommended mobility of each fraction, considering that a sediment fraction is
by Sloff and Mosselman (2012), Mosselman (2005) and mobile if:
Talmon et al. (1995) based upon the modelling of real rivers:
( /i ) 1 with i = 0.047 (8)
Ash = 9, Bsh = 0.5, Csh = 0.3 and Dsh = 0 (Eq. (7)). Dsh parameter takes
into account the effects of hiding and exposure on the effects Where = ripple factor and = Shields number.
of bed slope on sediment transport direction which is poorly We computed also the non-linearity of sediment transport with
understood. We only implemented the hiding and exposure effects respect to ow velocity as (Crosato and Mosselman, 2009):
in the sediment transport predictor (Section 2.1) and neglected 1
b = 3 ( 0.047 ) (9)
its effects here by imposing Dsh = 0, to reduce the number of the
calibration parameters. This parameter rapidly increases if the sediment is close to the
The discharge of 700 m3 /s was provided at the inlet. The nor- conditions of initiation of motion. Table 1 shows that at the begin-
mal water level corresponding to the inlet discharge and initial bed ning of the simulations, with a at channel-bed and uniform ow
slope of the channel was provided as the outlet boundary. An equi- distribution, only the nest two fractions were mobile as well as
librium sediment transport corresponding to the local hydraulic uniform sediment of simulation P3. Based on the large values of b,
condition close to upstream boundary was prescribed at the the second sediment fraction was close to initiation of motion.
inlet. To study the effects of full sediment mobility versus partial sed-
The simulations were carried out for each scenario starting iment mobility on the morphological evolution, we dened an-
from an almost at bed. A small initial bed level undulation was other set of four computational scenarios with reduced threshold
imposed to trigger morphodynamic instability and quickly obtain of sediment motion (0.02 instead of 0.047) in Eqs. (1), (8) and
bar formation. At the end of the computations, the system is sup- (9) maintaining all other parameters unchanged. In this way, we
posed to have reached a "dynamic equilibrium". In this study, the obtained comparable runs.
dynamic equilibrium is characterized by a braiding index that is The full mobility conditions are represented by scenarios F0, F1,
constant. Considering this, after having performed some prelimi- F2 and F3 which are the counterparts of scenarios P0, P1, P2 and
nary computational tests, the duration of the simulations was set P3, respectively. The uniform sediment scenarios, F3 and P3, are in
to either 10 years (full mobility) or 14 years (partial sediment mo- fact both fully mobile. Their difference lies basically in the thresh-
bility scenarios). Since the Waimakariri River is morphodynami- old used for initiation of motion and in their sediment transport
cally active only 50 days per year, this was also the duration of rates. Table 1 provides the sediment characteristics of all sediment
the computations representing one single year. Considering that scenarios and the mobility check of each fraction for these scenar-
we used a value of the morphological factor, morfac, equal to 25, ios.
10 years of morphological evolution were covered by 20 computa-
tional days and 14 years by 28 days. 2.4. Sensitivity analyses
Several scenarios were dened based on sediment mixture
characteristics and sediment mobility conditions to assess the ef- The results of morphodynamic simulations of gravel-bed
fects of sediment size heterogeneity on the morphology of braided braided rivers are strongly affected by a number of model param-
rivers. Altering the size heterogeneity of the bed sediment, we de- eters and discharge regime. Considering previous experiences, we
ned four scenarios differing in bed sediment composition. The analysed the effects of varying the values of these important pa-
rameters by means of sensitivity analyses.
132 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

Table 1
Subdivision in 5 sediment fractions. Mobility of sediment fractions of P1, P0 and P2 (partial mobility) and F1, F0 and F2 (full mobility) scenarios at the start of the
computations with a (virtually) at channel bed and uniform ow of 700 m3 /s.

Sed. Dmin Dmax Percentage of Mean size median size Dmax /Dmin Partial mobility Full mobility

fraction (mm) (mm) volume (%) Dm (mm) D50 (mm) i ( )/ i b i ( )/ i b

Poorly sorted scenarios P1/F1 (highest heterogeneity)


1 0.0625 3 15.62 0.76 0.4 48.0 3.67 37.0 1.74 2.1 6 0.74 4.95 4
2 3 13 23.18 6.82 6 4.3 0.25 5.0 0.23 1.1 40 0.10 2.54 5
3 13 56 24.91 29.44 27 4.3 0.06 1.7 0.08 0.8 0.03 1.76 7
4 56 106 18.5 78.36 77 1.9 0.02 0.8 0.04 0.6 0.02 1.36 11
5 106 250 17.79 167.83 163 2.4 0.01 0.4 0.02 0.5 0.01 1.13 27
Base-case scenarios P0/F0 (Waimakariri River)
1 0.0625 4.75 15.62 1 0.5 76.0 2.5 24.7 1.16 2.2 6 0.49 5.07 4
2 4.75 19.05 23.18 10 10 4.0 0.14 2.9 0.14 1.1 59 0.06 2.48 5
3 19.05 38.10 24.91 27 27 2.0 0.05 1.3 0.06 0.8 0.03 1.91 6
4 38.10 76.20 18.50 55 54 2.0 0.03 0.8 0.04 0.7 0.02 1.61 8
5 76.20 152.40 17.79 110 108 2.0 0.01 0.5 0.02 0.6 0.01 1.35 12
Well-sorted scenarios P2/ F2 (smallest heterogeneity)
1 2.0 0 0 0 8.00 15.62 4.3 4 4.0 0.28 4.3 0.20 1.4 11 0.09 3.23 4
2 8.00 24.00 23.18 14.6 14 3.0 0.08 1.7 0.18 1.0 477 0.03 2.36 5
3 24.00 29.00 24.91 26.4 26 1.2 0.04 1.1 0.05 0.9 0.02 2.01 6
4 29.00 45.00 18.5 36.4 36 1.6 0.03 0.8 0.04 0.8 0.02 1.86 6
5 45.00 80.00 17.79 60.8 60 1.8 0.02 0.6 0.03 0.7 0.01 1.64 8
Uniform sediment scenario P3/F3 (no heterogeneity)
1 19 19 100 19 19 1 0.08 1 0.047 1.78 7 0.020 4 4

The Waimakiriri River discharge is strongly variable, as in every We included the effects of bed slope parameterization on
natural gravel-bed braided system. Generally in this type of rivers, the bed load direction using the formulation by Koch and Folk-
high ows occur infrequently and for short periods of time. The stra (1980) extended by Talmon et al. (1995) (Eq. (3)). This equa-
Waimakariri is morphodynamically active only for about 50 days tion requires calibration based on the optimization of the follow-
per year, when the discharge is larger than 90 m3 /s (Carson and ing parameters: Ash , Bsh , Csh and Dsh (Eq. (5)). Increasing the value
Griths, 1989). To avoid strong discharge variations causing model of Ash directly lowers the bed slope effects and this is also ob-
instability, we designed the computational yearly discharge hydro- tained by decreasing the value of Csh . However, unlike Ash (param-
graph based upon the ow duration curve. The schematized hy- eter studied already by Schuurman et al., 2013), Csh imposes dif-
drograph is then repeated for the number of years that character- ferent bed slope effects for each sediment fraction (Sloff and Mos-
ize the duration of the simulation. This type of approach has been selman, 2012). Dealing with graded sediment and focusing on the
already successfully used for long-term morphodynamic mod- effects of sediment size heterogeneity, we decided to investigate
elling (e.g. Yossef and Sloff, 2012). In the computational variable- the effects of varying Csh . This coecient weighs the effects of
discharge hydrograph (Fig. 2), the minimum value of the discharge bed slope for different sediment fractions, through the sediment
(300 m3 /s) corresponds to an averaged low ow and the maximum size to local water depth ratio. D50 /h ratio is larger on bar tops
value (2750 m3 /s) corresponds to an averaged ow event occurring and smaller in channels. Since Csh < 1, increasing D50 /h results in
for one day per year. smaller values of fi (Eq. (5)) and increasing effects of bed slope on
For variable discharge, the value of morfac used in our model sediment transport direction (Eq. (3)). We performed three runs
decreases as the value of the discharge increases so that the high- simulating the long-term morphological evolution starting from an
est value of morfac (200) is applied for the smallest value of almost at bed with different values of this parameter: 0.2, 0.3 and
the discharge (300 m3 /s), for which the sediment transport rate is 0.4. We considered the base-case scenario under full mobility con-
very small, and no acceleration is applied to the highest discharge ditions with Ash = 9 and Bsh = 0.5 (Talmon et al., 1995), whereas we
(morfac = 1), since during peak ow conditions the sediment trans- omitted the contribution of sediment sorting on the effects of bed
port rate is high and the morphological changes are relatively fast. slope by imposing Dsh = 0. The value of Ash and Bsh is commonly
A number of scientists (e.g. Leopold and Wolman, 1957; used as 1.5 and 0.5 respectively if Csh and Dsh are ignored (e.g.
Fredse, 1978; Hey and Thorne, 1986; van den Berg, 1995; Parker Sloff and Mosselman, 2012)
et al., 2007) suggested using the bankfull discharge or the mean Sloff and Mosselman (2012) showed that the thickness of the
annual ood (eg. Antropovskiy, 1972; Bray, 1982) to represent the active bed layer is important for the simulation of both sed-
formative conditions of a river. We selected the value of the con- iment sorting and bed topography. Hiranos (1971) method re-
stant discharge based on the best t between the Waimakariri quires the denition of the thickness of the layer of sediment
River morphology and the results of our model. The discharge of that is actively participating in the sediment transport process.
700 m3 /s results in the best t and can be considered to represent The sediment underneath comes into action in cases of scour-
the "formative discharge". ing. The thickness of the active layer, which is kept constant
Gravity diverts bed load along the downslope direction during the computations, plays a sensitive role in the perfor-
(e.g. Koch and Flokstra, 1980), which decreases bar amplitudes mance of the model (Blom, 2008; Sloff and Ottevanger, 2008).
and produces sediment sorting (summarized by Powell, 1998). Hirano assumed the active layer thickness to be equal to the
Schuurman et al. (2013) showed that the morphological evolution maximum grain size. Later, Armanini and Di Silvio (1988), Parker
of braided rivers is affected by the effects of transverse bed slope (1991) and Ribberink (1987) suggested using half the amplitude
parameterization on bed load direction. These effects are especially of dunes on the river bed, whereas Blom (2008) suggested us-
important for bar formation (see also Mosselman and Le, 2015). ing the entire dune amplitude. This approach is conrmed by
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 133

Sloff and Ottevanger (2008), Mosselman (2012) and Sloff and Mos- The braiding index is here represented by the bar mode, m. It
selman (2012), who show that in modelling real rivers (Dutch is derived using the method developed by Schuurman et al. (2013),
Rhine branches), the active layer thickness requires much higher which consists of counting the number of parallel channels cross-
values than half the dune amplitude. ing the river cross-sections and then computing the reach-averaged
Although a few studies indicate the presence of dune-like struc- value. Channels are identied by a bed level below the cross-
tures in the channels of gravel bed rivers (e.g. Dinehart, 1992), sectional averaged one. This method includes both active and in-
there is no evidence of dune formation in the reference river. active channels. The bar mode is then computed from the num-
Moreover, our study focuses on the reworking of bars. For this, an ber of channels, since the number of channels in the cross-section
active layer equal to half of or the entire bar amplitude may re- increases by 0.5 if m increases by 1 (e.g. Crosato and Mossel-
sult in more realistic results. To study the effects of active layer man, 2009).
thicknesses on river braiding, we carried out a sensitivity analysis. The bed topography of braided rivers consists of a complex net-
We performed several morphological simulations starting from an work of mid-channel bars and ephemeral channels. Bars are sub-
almost at bed for the base-case scenario under full mobility and merged during high ow and exposed during low ows. In this
partial mobility conditions. The smallest thickness of 0.2 m corre- context, there is no clear denition or threshold level to distin-
sponds to three times D90 . The thickness of 1.25 m is close to half guish bars from channels.
the maximum bar amplitude, dened as the difference in level be- Hypsometric curves have been mostly used to represent the u-
tween the highest bar top and deepest channel bed level. We also vial forms and processes within a river basin (e.g. Willgoose and
investigated the thicknesses of 2, 5 and 10 m to study the effects Hancock, 1998) or to characterize and evaluate the development
of large active layer thickness. We included relatively large thick- of the morphological features in estuaries and tidal inlets (e.g.
ness of 10 m to investigate the case in which the active layer cor- Marciano et al., 2005; Wang et al., 2002). The use of hypsomet-
responds to the entire alluvial bed. The sub-surface was divided ric curves is not common in river morphodynamics, but they ap-
into a number of book-keeping layers with the same thickness as pear as a suitable tool to assess the complex topography of braided
the active layer to make the model more robust. rivers in statistical terms. The advantage of using this method is
that it does not require a threshold limit to dene bars and chan-
2.5. Analysis methods nels, whereas it shows the extension of areas above or below cer-
tain levels. Different bar/channel distributions, as well as differ-
To compare different scenarios and detect temporal trends, we ent bar amplitudes, result in different hypsometric curves. So, the
analysed the bed topography distribution at specic times and hypsometric curves can be effectively used to compare the topo-
the temporal evolution of reach-averaged cross-sectional sediment graphical characteristics of river reaches and modelling results. We
transport rate, braiding index, reach-averaged bar amplitude and derived the hypsometric curve from the bed topography after de-
bar length. We also analysed the median grain size sediment dis- trending by initial bed slope. In our hypsometric curves, the de-
tribution to understand the effects of sediment sorting processes trended bed elevation is given in the y-axis, where it is repre-
on the channel evolution. A 10 km reach was chosen for the analy- sented by the distance between the bed and a reference plane. The
sis after removing 2.5 km reach each at the upstream and down- percentage of area having a smaller value is given in the x-axis.
stream boundary to eliminate boundary effects. Egozi and Ash- Negative values are characteristic of the low areas (channels) and
more (2008) state that the sample-reach length should be at least positive value belongs to the high areas (bars).
10 times the channel width at the channel forming discharge. Their To represent the reach-averaged bar amplitude, we adopted the
criteria are based upon a laboratory experiment but serves as the method by Schuurman et al. (2013), in which the bar top is dened
basic criteria for rivers as well. Since, width of the river in the as the highest 10% and the channel deepest point as the lowest
model is 900 m we consider that 10 km reach is sucient for the 10% of detrended bed elevation (in the hypsometric curves the lev-
analysis. els corresponding to 10% and 90% of area, respectively). The reach-
Considering that the bar merging process results in progressive averaged difference between these two levels is then assumed to
reduction of braiding index, we assumed that the river bed topog- be the representative bar amplitude.
raphy can be considered as fully-developed when the braiding in- To compute the reach-averaged bar length, bars were assumed
dex reaches a stable value. For this, the temporal evolution of the to be represented by all areas having water depth smaller than half
braiding index and reach averaged cross-sectional sediment trans- the reach-averaged water depth at a discharge of 700 m3 /s. After
port rate were used as a measure to establish whether morphody- having mapped all bars, we measured their length. Since at ev-
namic equilibrium is achieved. ery moment in time several bars of different size are present, we
Egozi and Ashmore (2008) provide a list of indices that computed the percentage of surface occupied by all bars having
are used to represent the index of river braiding. The "chan- the same length which was then assumed to be their weight. The
nel count index" appears as a promising one, since it is eas- weighted average was then used as representative reach-averaged
ily quantiable and comparatively less sensitive to river-stage ef- bar length.
fects. Bertoldi et al. (2009b), however, suggested counting only
the channels which are morphologically active, which results in 3. Results
a stage-dependent "active braiding index". The "bar mode"-m-
(Engelund and Skovgaard, 1973) is another effective parameter to 3.1. Main results
dene the river braiding intensity. This is the number of parallel
alternate-bar rows that are needed to reproduce the topographic Fig. 4 shows evolution of water depth distribution in P0 sce-
pattern of the river channel: m = 1 corresponds to a channel with nario. Deeper parts (blue) represent channels and shallow parts
alternate bars, typical of meandering rivers; m = 2 corresponds to (white) represent bars. Starting from the near atbed topography,
a channel with central bars, typical of rivers in transition between 0 years, several small amplitude bars appeared in the river bed at
meandering and braiding; and m > 2 corresponds to a channel with the initial stage of the morphological evolution, 0.25 years. Smaller
multiple bars, typical of braided rivers. Bar mode is similar to the bars then merged together and formed larger bars as the develop-
channel count index, but is derived by counting the number of ment progressed (2.5 years). Bars continued to merge and become
faces of bars wetted by a channel rather than the channels alone larger and ow concentrated in few deep channels (10 years). Sim-
(Kleinhans and van den Berg, 2011). ilar morphological evolution trends were observed in other partial
134 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

Fig. 4. Temporal evolution of water depth distribution in base case scenario (P0). Flow is from left to right.

Fig. 5. Water depth distribution after 10 years of morphological evolution for the full-mobility scenarios (right column) and for the partial-mobility scenarios (left column).
Constant discharge equal to 700 m3 /s. Colour bar: water depths in meters. Flow is from left to right. P1/F1: poorly sorted; P0/F0: base case; P2/F2: well sorted and P3/F3:
uniform sediment scenarios.

mobility scenarios (P1, P2 and P3) and full mobility scenarios (F1, evolution progressed which denoted the coarsening of the channel
F0 and F3). Fig. 5 shows water depth distribution after 10 years of bed and possibility of channel bed armouring. Similar trends were
morphological evolution. Full-mobility eventually resulted in much observed in P1 and P2 scenarios as well (Fig. 7). P1 and P2 scenar-
deeper channels and in bed topographies that are highly irregular ios had the largest and smallest sediment size representing coarse
(some irregularities might be produced by local model instability), part of their sediment size distribution curve respectively (Fig. 3
whereas partial mobility resulted in more realistic channel-bar pat- and Table 1). So, the largest and smallest D50 was observed in the
terns (compared to the Waimakariri River, Fig. 1). channels of scenario P1 and P2 respectively.
Fig. 6 shows median sediment grain size (D50 ) distribution dur- Unlike in P0 scenario, D50 distribution did not show any con-
ing the morphological evolution in the partial mobility scenario-P0 sistent pattern among channels and bars in F0 scenario. However,
(left) and the full mobility scenario-F0 (right). In the P0 scenario, larger D50 were observed in some parts of the deeper channels.
larger D50 were observed in the channels whereas ner D50 on Similar trends were also observed in F1 and F2 scenarios (Fig. 7).
top of the bars. D50 increased in channel bed as the morphological Since F1 scenario had highest sediment heterogeneity and the F2
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 135

Fig. 6. Median grain size distribution during morphological evolution for the full-mobility scenario: P0 (right column) and for the partial-mobility scenario: F0 (left column).
Flow is from left to right.

Fig. 7. Medium sediment grain size (D50 ) distribution after 10 years of morphological evolution for the full-mobility scenarios (right column) and for the partial-mobility
scenarios (left column). Flow is from left to right. P1/F1: poorly sorted; P0/F0: base case; P2/F2: well sorted.

scenario had the lowest sediment heterogeneity, the variation in development (Fig. 11a). High growth rates of bar length occurred
D50 size distribution was also observed highest in the F1 and low- in the latest phases of the development, to the point that uniform
est in the F2 scenario. sediment ended up with the longest bars. Similar results were ob-
The temporal evolution of the reach-averaged sediment trans- tained for the averaged bar amplitude (Fig. 10, column a): in the
port rate is shown in Fig. 8 (partial mobility scenarios) and Fig. 9 early stages of the development, the bar amplitude increased with
(full mobility scenarios) shows the transport rates of each sedi- the sediment heterogeneity, but later this trend reversed, with uni-
ment fraction, as well as the total sediment transport rates. As ex- form sediment (P3) resulting in the highest bars. Summarizing,
pected, reduction of the threshold for sediment motion in the full uniform sediment eventually produced the largest bars (20 0 0 m
mobility scenarios results in higher sediment transport rates and long with an amplitude of 3.2 m), resulting in the most unreal-
relatively higher contributions of the coarser fractions to the to- istic morphology if compared to the reference sub-reach of the
tal transport. Note that P0 and P1 have similar sediment transport Waimakariri River, whose bars are, on average, 10 0 0 m long and
rates (Fig. 9) and that the smallest size fractions are responsible 1.8 m high. The excessive bar development can be explained by the
for the highest bed material loads. The uniform sediment scenario fact that uniform sediment is always characterized by full mobility
results in the lowest rates. (one single mobile fraction), whereas the other scenarios always
Oscillations of the sediment transport rates are typical features had one or more sediment fractions that are not mobile. Scenar-
of braided rivers (Ashmore, 1991b) in which channel excavation ios P1 (poorly sorted) and P0 (base-case) always resulted in sim-
and consequent ow concentration result in increased sediment ilar developments, with nal braiding indices gradually stabilizing
transport rates. Comparing the evolution of the sediment transport near the value of 12. These two scenarios produced the most re-
rates with the evolution of the braiding indices (Fig. 10a, dotted alistic results: a reach-averaged bar amplitude of 1.9 m and a bar
line), we can observe that the stabilization of the sediment trans- length close to 10 0 0 m (Fig. 11a). Note that these two scenarios dif-
port rate does not represent the equilibrium, since the develop- fer in sediment size heterogeneity, being P1 more heterogeneous
ment and redistribution of the bars and channels also occur when than P0, but behave similarly, because the sediment fraction 5 was
the total sediment transport rates is relatively constant. not mobile (Fig. 8e). Based on braiding index and bar amplitude, as
Comparing the partial-mobility scenarios, we observed that uni- well as total sediment transport, the two scenarios seemed to have
form sediment (P3) showed an initial temporal lag in the bed to- reached a condition of morphodynamic equilibrium at the end of
pography development with respect to the heterogeneous sedi- the computations, i.e. after 10 years. The bar length (Fig. 11a), how-
ment cases, which is caused by the much lower sediment transport ever, was still slightly growing, which is a sign of progressive on-
rates (Fig. 8). A time lag could be observed also for the bar length going bar merging.
136 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

Fig. 8. Temporal evolution of the reach-averaged cross-sectional sediment transport rate of a) sediment fraction 1 (nest), b) fraction 2, c) fraction 3, d) fraction 4, e) fraction
5 (coarsest fraction) and f) sum of all fractions for the sediment heterogeneity scenarios P1, P0, P2 and P3 (partial mobility). P1: poorly sorted; P0: base case; P2: well sorted
and P3: uniform sediment scenarios.

In the full-mobility scenarios we can observe that the poorly- 3.2. Model sensitivity
sorted sediment (highest heterogeneity), F1, resulted in the lowest
bars and the highest braiding index, whereas well-sorted sediment Bars and braiding properties were found to be highly sensitive
(smallest heterogeneity), F2, produced the highest bars and the to the transverse bed slope effects parameterization. The results of
smallest braiding index. The results of the base-case scenario, F0, the sensitivity analysis showed that higher values of Csh , produc-
lay in between (Fig. 10, column b). Surprisingly, also uniform sedi- ing relatively larger bed slope effects, developed larger and deeper
ment, F3, produced values close to the base-case scenario. Hiding- channels, lower braiding index, and higher bars (Fig. 12). This could
exposure effects, not present with uniform sediment, might be re- be observed by analysing the hypsometric curves of the bed to-
sponsible for this, since they produced increased mobility of the pography after 10 years of morphological evolution (Fig. 12a). In-
coarsest fractions and decreased mobility of the smallest fractions creasing the Csh value from 0.3 to 0.4 decreased the braiding index
in the other non-uniform sediment scenarios. The temporal varia- up to 23% and increased the bar amplitude up to 23% during the
tions in bar length are more confused (Fig. 11b), but in general the morphological evolution. Similarly, lower values of Csh , producing
most poorly sorted sediment (highest heterogeneity), F1, resulted relatively smaller bed slope effect, lead to smaller channels, lower
in the longest bars and the scenarios with the low sediment het- bars and more braiding. Reducing the value of Csh from 0.3 to 0.2
erogeneity, F2 and F3, in the shortest. However, at the end of the increased the braiding index up to 35% and decreased the bar am-
computations all sediment scenarios resulted in similar bar lengths plitude up to 19%.
(1.52.0 km). Fig. 13 shows nal hypsometric curves, the temporal evolution
In general, as expected, the full mobility conditions resulted in of braiding index and bar amplitude resulting from different values
more intense morphodynamic variations than the partial mobility of the active layer thickness. The results showed that full and par-
conditions. This is reected in the smaller nal braiding indices, tial mobility conditions led to different morphological behaviours.
an indication of more advanced bar merging. Full mobility condi- In general, for the same active layer thickness full mobility con-
tions produced results that differ considerably from the measured ditions resulted in deeper channels and higher bars, but smaller
ones: larger bar amplitudes: 4.86.0 m against the measured 1.8 m, braiding indices.
as well as bar lengths: 20 0 0 m against 10 0 0 m. Uniform sediment, The volume of sediment fraction available for erosion from the
scenario F3, produced results which fall between the results of the channel bed is directly proportional the active layer thickness. In
heterogeneous scenarios. Similar results were obtained for uniform the case of thin active layers (0.2 m), the volume of ne sediment
sediment also by Nicholas (2013). fraction is less, so they were eroded immediately from the channel
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 137

Fig. 9. Temporal evolution of the reach-averaged cross-sectional transport rate of a) sediment fraction 1 (nest), b) fraction 2, c) fraction 3, d) fraction 4, e) fraction 5
(coarsest fraction) and f) sum of all fractions for the sediment heterogeneity scenarios F1, F0, F2 and F3 (full mobility). F1: poorly sorted; F0: base case; F2: well sorted and
F3: uniform sediment scenarios.

bed. Since, the coarser sediment fractions were not mobile in par- plitude is still growing, although at a very small rate (Fig. 14c).
tial mobility scenario they were left in the channel bed leading to Approaching equilibrium, effects of the single discharge peaks be-
formation of armour layer which resulted in strongly retarded bed come negligible for both braiding index and bar amplitude.
development. This is observable from the hypsometric curves. Comparison between the computed hypsometric curves and the
The volume of ne sediment fractions increases if the ac- measured one (Waimakariri River) shows that the constant dis-
tive layer thickness is increased. The larger amount of ne sed- charge scenario produces the most realistic distribution of bed
iment fractions increased the bed mobility thus increasing the topography. The braiding indices obtained with the variable dis-
morphological evolution at the initial phase. As the morphologi- charge hydrograph (BI = 8) and the constant discharge regime
cal evolution progressed, bars merged and ow was concentrated (BI = 12) show that the constant discharge results in a braid-
in the channels, which increased the mobility of the larger sed- ing index that is closer to the measured one, the braiding in-
iment fractions in the channel bed thus leading to a more com- dex in the 900 m wide straight part of the Waimakariri River
plex morphology, with signicantly higher braiding indices and bar being BI = 10.7. The constant discharge scenario also results in
amplitudes. the most realistic (averaged) bar amplitude: 1.9 m, against the
Since all of the sediment fractions were mobile with full sedi- 4 m obtained with variable discharge, the reference measured one
ment mobility scenarios, the results of the morphological simula- being 1.54 m.
tions were less dependent on the active layer thickness. Neverthe- The water depth distributions obtained at the end of the com-
less, the 0.2 m thickness generally resulted in the smallest braid- putations are shown in Fig. 15. It is possible to observe that the
ing indices and lowest bars and the 10 m thickness in the highest constant discharge produced a denser and more realistic network
braiding indices and bars. of channels than the variable discharge regime (compare to Fig. 1,
Fig. 14 shows the hypsometric curves of bed topography, braid- showing the Waimakariri River).
ing index and bar amplitude compared between those obtained
from the variable-discharge hydrograph and a constant discharge 4. Discussion
of 700 m3 /s. With variable discharge, an almost equilibrium braid-
ing index, reaching the value of 8 is achieved after 810 years We carried out a numerical analysis in this study using a fully
(Fig. 14b). However, at the end of the computations the bar am- non-linear morphodynamic model Delft3D. The numerical model
138 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

a) partial mobility b) full mobility

16 16
P1 F1
P0 F0
14 14
P2 F2
P3 F3

braiding Index
braiding index

12 12

10 10

8 8

6 6

4 4
0 5 10 15 0 5 10
time (years) time (years)
6 6
P1
P0
5 5
P2
bar amplitudet (m)

bar amplitude (m)

P3
4 4

3 3

2 2
F1
1 1 F0
F2
F3
0 0
0 5 10 15 0 5 10
time (years) time (years)

Fig. 10. Top to bottom: hypsometric curves at the end of the computations; temporal evolution of braiding index; and temporal evolution of bar amplitude, computed as
the elevation difference between the top 10% and the smallest 10% detrended values of bed levels. a): partial mobility scenarios (P1, P0, P2 and P3) and b): full mobility
scenarios (F1, F0, F2 and F3). P1/F1: poorly sorted; P0/F0: base case; P2/F2: well sorted and P3/F3: uniform sediment scenarios.

was set up with the morphodynamic variables obtained from the river bed at the upstream boundary, to mimic propagation of bar
Waimakariri River, New Zealand, without the intention to repro- at the inlet, such that more dynamic bar conguration is achieved
duce any specic river. The numerical runs were carried out in during the morphological evolution of braided channel. An equi-
a straight channel with xed banks starting from almost at bed librium sediment transport corresponding to the local hydraulic
(containing small perturbations to trigger bar development). The condition close to upstream boundary was prescribed at the inlet
morphological evolution of bar was described by the appearance in this study. It allowed to impose uctuations compared to con-
of several small amplitude bars in the river bed at the initial stage stant sediment feed at the inlet. However imposing re-circulating
followed by bar merging to form larger bars at the later stage of sediment transport boundary condition, which takes into account
the development, conrming the general evolution trend reported sediment transport uctuation due to propagation of bar, at the
in the literature (Fujita, 1989; Enggrob and Tjerry, 1999; Nicholas inlet may result in more dynamic bars than those observed in this
et al., 2013; Schuurman et al., 2013). The trend of evolution of study.
braiding index and bar height approaching to a stable value indi- This study shows that braiding index increases and bar am-
cate that the morphological evolution is approaching to an equilib- plitude decreases if the sediment heterogeneity is increased. In-
rium condition. terventions affecting the river bed sediment, like gravel augmen-
The morphodynamic evolution of river bars depends on tation, alter both the sediment heterogeneity and the median
the sediment transport boundary condition at the inlet. grain size. Assuming full sediment mobility, Crosato and Mos-
Mendoza et al. (2016) show that in a nite domain, the sedi- selman (2009) show that larger median diameters, reecting in
ment boundary conditions which allow periodicity or uctuations larger sediment transport non-linearity (Eq. (9)), result in increased
at the upstream inlet result in more dynamic and realistic bar braiding indices (bar modes) and reduced bars (wave) lengths.
morphology. Nicholas et al. (2013) used periodic uctuation of
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 139

4000
a) a)
3500

3000
bar length (m)

2500

2000

1500

1000

500

0
0 5 10
time (years) 16

14 b)
4000
12
b)

braiding index
3500
10

3000 8
bar length (m)

2500 6

4
2000
2
1500
F0,0.2 F0,0.3 F0,0.4
0
0 2 4 6 8 10
1000
time (years)
500
7
0
0 5 10 6
c)
time (years)
bar amplitude (m)

5
Fig. 11. Temporal evolution of bar length. a) partial mobility scenarios b) full mo-
bility scenarios. Open square and black dashed line represent P1/F1 scenario, black
lled circle and black continuous line represent P0/F0 scenario, grey lled triangle 4
and grey dashed line represent P2/F2 scenario and grey lled circle and grey con-
tinuous line represent P3/F3 scenario. 3

1
These additional effects should be taken into account when plan-
ning any interventions that alter the bed sediment. F0,0.2 F0,0.3 F0,0.4
0
We analysed the effects of transverse bed slope effect 0 2 4 6 8 10
parametrization on the morphological evolution of braided river time (years)
by altering value of Csh . Braiding intensity and bar amplitude were
observed to be affected by the value of Csh . Braiding intensity de- Fig. 12. a) hypsometric curves of bed topography after 10 years of morphologi-
cal evolution b) temporal evolution of braiding index and c) temporal evolution of
creased with increase in transverse bed slope effects (higher Csh ),
bar amplitude obtained with the values of coecient Csh = 0.02; 0.03 and 0.04. The
which are in agreement with Fig. 18 of Schuurman et al. (2013), computations refer to the base-case scenario with full mobility conditions.
who performed a similar analysis, by altering the coecient Ash
with a constant discharge and full sediment mobility. They made
analysis with active braiding index which only takes into account Transverse slope effect parameters is chosen as calibrating pa-
the channels transporting sediment. The transverse bed slope ef- rameter to adjust patterns of bars and pools in numerical mod-
fect relation implemented by Schuurman et al. (2013) does not in- elling application on real rivers (Sloff and Mosselman, 2012). We
clude the Csh parameter. do not aim to model individual bars and channels of any specic
The bar amplitude increased if the bed slope effect increased. river reach in this study. So the value of Csh = 0.3, chosen based
Schuurman et al. (2013) observed opposite trend in the bar height. on modelling experience of real rivers (e.g. Talmon et al., 1995), is
A lower braiding index corresponds to a smaller number of deeper reasonable for this study. Nicholas et al. (2013) argue that generic
channels and higher bars than a higher braiding index. The small- understanding of morphological evolution of bars and channels can
est braiding index is found for the highest bed slope effects, which be obtained from the numerical model without precisely calibrat-
explains why the bar amplitude is the highest for this value of Csh . ing the parameters of the transverse bed slope effects.
140 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

a) partial mobility b) full mobility

16 16
0.2 1.25 2
15 15 5 10 m

14 14

braiding index
braiding index

13 13

12 12

11 11

10 10

0.2 1.25 2 9
9
5 10 m
8 8
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3

time (years) time (years)

4 4
0.2 1.25 2
3.5 3.5
5 10 m
bar amplitude (m)
bar amplitude (m)

3 3

2.5 2.5

2 2

1.5 1.5

1 1

0.5
0.2 1.25 2
0.5
5 10 m
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
time (years) time (years)
Fig. 13. Temporal evolution of channel bed topography (from top to bottom): hypsometric curves, braiding index and bar amplitude obtained for different values of the
active layer thickness in case of a) partial sediment mobility and b) full sediment mobility. The values 0.2 to 10 correspond to the different thicknesses of the active layer in
meters.

Thickness of active layer strongly affected the morphological The comparison between model simulations and measured data
development in the partial mobility scenario. Based on the results, shows that, contrary to expectations, the constant-discharge ap-
we can observe that the active layer thickness must be much larger proach leads to results that are much more realistic than the
than the largest grain size to model the morphological evolution of variable-discharge approach. The reasons for this might lie in the
gravel-bed braided river at partial mobility conditions . Since bars use of the MPM sediment transport formula. This formula is not
are reworked during morphological development by braiding pro- suitable for cases with high ow velocity (as during discharge
cesses, Leduc et al. (2015) recommend to use active layer thick- peaks), since high ow velocities result in Shields parameters that
ness with respect to bar amplitude, in the morphological study of are much larger than the critical value for sediment motion ( >
braided channels. So, active layer thickness of about half of the > crit with crit = 0.047 if the ripple factor is assumed equal to 1)
maximum bar amplitude was used for the numerical modelling in (Mosselman, 2005, p.78; Siviglia and Crosato, 2016). In this case,
this study. according to the MPM formula, the sediment transport capacity
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 141

becomes independent of grain size, which is not realistic. Secondly,


MPM predicts at high mobility that the relative amount or concen- a)
tration of sediment in motion does not increase if bed shear stress
increases. Third, MPM predicts at high mobility that narrowing of
the river bed leads to steeper slopes, contrary to common obser-
vation.
The formula should therefore always be applied within its ap-
plicability range, in compliance with the conditions of the lab-
oratory experiments under which it was derived: s /u > 1,
Dm > 0.4 mm and < 0.2, with Dm = mean sediment grain size,
S = fall velocity of sediment, = ripple factor (in this study
= 1), and u = u g/C 2 = bed shear velocity (Meyer Peter and
Mller, 1948).
During computations, the Shields parameter exceeds the value
of 0.2 in both runs with variable and constant discharge. In the
simulation with constant discharge, the value of 0.2 is exceeded
only locally and sporadically, inside the deepest channels (link to
Supplementary material 1). Instead, in the simulations with vari- 16
able discharge, peak conditions systematically result in excessive
Shields parameters, reaching the value of 0.5, in a large area of the
14 b)
computational domain (link to Supplementary material 2). 12

10

braiding index
5. Conclusions
8
We carried out a numerical analysis to study the effects of sed-
iment heterogeneity on the long-term morphological evolution of 6
gravel-bed braided systems. To obtain realistic simulations, the nu-
4
merical model was set up with the morphodynamic variables ob- constant Q variable Q
tained from the Waikariri River, New Zealand. 2
Larger sediment heterogeneity resulted in increased braiding in- river
dices, conrming the ndings of Teramoto and Tsujimoto (2006), 0
0 2 4 6 8 10
as well as reduced bar size (amplitude and length), conrming the
ndings of Lanzoni and Tubino (1999). With full sediment mo-
time (years)
bility, however, the highest sediment heterogeneity produces the
6
longest bars. Lanzoni (20 0 0b) had contradictory experimental re- constant Q
sults on the effects of sediment heterogeneity on bar length. A
5 variable Q
c)
possible explanation could be that these effects depend on sed-
bar amplitude (m)

iment mobility, since bars become shorter with partial sediment river
mobility and longer with full sediment mobility, if the sediment 4
heterogeneity increases. It is important to note that Lanzoni and
Tubino (1999) and Lanzoni (20 0 0b) focused on alternate bars. This 3
work shows that their conclusions are valid also for multiple bars,
even after the bar merging process. 2
The results of this study indicate that uniform sediment can-
not be an acceptable assumption to study the long-term response 1
of natural gravel-bed braided rivers, because these rivers are gen-
erally characterized by partial sediment mobility. In this case, uni- 0
form sediment is found to produce an unrealistic bed topography, 0 2 4 6 8 10
characterized by bars that are too long and too high with respect time (years)
to those observed in nature. Moreover, due to the relatively small
sediment transport rates resulting in delayed bed adaptations, the Fig. 14. a) hypsometric curves at the end of the computations for constant and
variable discharge and hypsometric curve of the Waimakariri River, b) temporal
uniform sediment scenario requires longer computational times to
evolution of braiding index obtained with constant and variable discharge (base-
reach the conditions of morphodynamic equilibrium. case) and c) temporal evolution of bar amplitude, computed as the difference be-
Instead, the uniform sediment approach appears acceptable in tween the top 10% and the smallest 10% detrended values of bed levels.
case of full-mobility conditions. However, even in this case, the
short-term and mid-term evolution, governed by the bar merging
process, might present important differences. Full mobility condi-
tions are mostly found in sand-bed rivers or in very large gravel- Waimakariri River. The reason could be found in the use of the
bed braided rivers, like the Congo at Kinshasa (Peters, 1978). The Meyer Peter and Mller formula (1948) at all ow conditions. This
long-term morphological response of braided rivers like the Ya- transport formula is not suitable if the value of the Shields param-
muna, in Bangladesh, characterized by sandy (full-mobile) bed ma- eter falls outside its applicability range (Mosselman, 2005; Siviglia
terial, can therefore be studied with the assumption of uniform and Crosato, 2016). Given the lack of sediment transport formu-
sediment (Ashworth et al., 20 0 0; Jagers, 2003). las covering all coarse sediment transport conditions, from incip-
In this study, contrary to expectations, the "formative" con- ient motion (low discharges) to extreme mobility (peak ows), a
stant discharge has produced results that are much more realistic solution to this problem could be the application of different for-
than the variable discharge, compared to the morphology of the mulas during the morphological computations, selecting the most
142 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

Fig. 15. Water depth distribution at the end of the computations showing the network of channels: a) with constant discharge (700 m3 /s) and b) with variable discharge at
the same ow conditions (700 m3 /s). Base-case sediment scenario. Flow direction from left to right.

appropriated one based on the instant value of Shields parameter Bertoldi, W., Zanoni, L., Tubino, M., 2010. Assessment of morphological changes
or bed shear stress. induced by ow and ood pulses in a gravel bed braided river: the Taglia-
mento River (Italy). Geomorphology 114 (3), 348360. http://dx.doi.org/10.1016/
j.geomorph.2009.07.017.
Binder, W., 2004. Restoration of rivers and oodplains in Bavaria. In: 3rd European
Acknowledgements
Conference on RIver Restoration. Zagreb, pp. 1721.
Blanckaert, K., Glasson, L., Jagers, H.R.A., Sloff, C.J., 2003. Quasi-3D simulation
We would like to express our deep gratitude to Filip Schuur- of ow in sharp open-channel bends with equilibrium bed topography. In:
man and Maarten Kleinhans for allowing us to use their codes to Snchez-Arcilla, A., Bateman, A. (Eds.), River, Coastal and Estuarine Morpho-
dynamics: RCEM 2003, 1-5 Sept. 2003, Vol. I. Barcelona, Spain. IAHR, pp.
compute braiding indices and bar characteristics. We thank Arthur 652663.
Mynett and Erik Mosselman for their valuable guidance and avail- Blom, A., 2003. A Vertical Sorting Model of Rivers with Non-Uniform Sediment and
ability, as well as Astrid Blom, Ton Hoitink, Ralph Schielen and Dunes Phd Thesis. University of Twente, Netherlands.
Blom, A., 2008. Different approaches to handling vertical and streamwise sorting in
Howard Southgate for their enthusiastic participation at discus- modeling river morphodynamics. Water Resour. Res. 44 (3), W03415. http://dx.
sions. We would also like to thank the reviewers for their review doi.org/10.1029/20 06WR0 05474.
of the manuscript which helped to improve the paper. Our special Bray, D.I., 1982. Regime equations for gravel-bed rivers. In: Hey, R.D., Bathurst, J.C.,
Thorne, C.R. (Eds.), Gravel-Bed Rivers: Fluvial Processes, Engineering and Man-
thanks to the Environment Canterbury who kindly supplied the LI- agement. John Wiley and Sons, Chichester, England, pp. 517552.
DAR and ow data for the Waimakariri River. Hickss input was Callander, R.A., 1969. Instability and river channels. J. Fluid Mech. 36 (03), 465480.
supported by the New Zealand Ministry for Science and Innova- http://dx.doi.org/10.1017/S0 0221120690 01765.
Carson, M.A., Griths, G.A., 1989. Gravel transport in the braided Waimakariri
tion through Contract C01X0308. Singhs input was funded by the
river: mechanisms, measurements and predictions. J. Hydrol. 109 (34), 201
Netherlands Fellowship Program (NFP) and NUFFIC. 220. http://dx.doi.org/10.1016/0022-1694(89)90016-4.
Colombini, M., Seminara, G., Tubino, M., 1987. Finite-amplitude alternate bars. J.
Fluid Mech. 181, 213232. http://dx.doi.org/10.1017/S0 0221120870 02064.
Supplementary materials Crosato, A., Mosselman, E., 2009. Simple physics-based predictor for the number of
river bars and the transition between meandering and braiding. Water Resour.
Res. 45 (3), 114. http://dx.doi.org/10.1029/20 08WR0 07242.
Supplementary material associated with this article can be Crosato, A., Saleh, M.S., 2011. Numerical study on the effects of oodplain vegetation
found, in the online version, at doi:10.1016/j.advwatres.2017.02. on river planform style. Earth Surf. Process. Landforms 36 (6), 711720. http://
005. dx.doi.org/10.1002/esp.2088.
Crosato, A., Mosselman, E., Beidmariam Desta, F., Uijttewaal, W.S.J., 2011. Experimen-
tal and numerical evidence for intrinsic nonmigrating bars in alluvial channels.
References Water Resour. Res. 47 (3), W03511. http://dx.doi.org/10.1029/2010WR009714.
Crosato, A., Desta, F.B., Cornelisse, J., Schuurman, F., Uijttewaal, W.S.J., 2012. Exper-
Armanini, A., Di Silvio, G., 1988. A one-dimensional model for the transport of a imental and numerical ndings on the long-term evolution of migrating alter-
sediment mixture in non-equilibrium conditions. J. Hydraul. Res. 26 (3), 275 nate bars in alluvial channels. Water Resour. Res. 48 (6), W06524. http://dx.doi.
292. http://dx.doi.org/10.1080/00221688809499212. org/10.1029/2011WR011320.
Ashmore, P.E., 1982. Laboratory modelling of gravel braided stream morphology. Dinehart, R.L., 1992. Evolution of coarse gravel bed forms: Field measurements at
Earth Surf. Process. Landforms 7 (3), 201225. http://dx.doi.org/10.1002/esp. ood stage. Water Resour. Res. 28 (10), 26672689. http://dx.doi.org/10.1029/
3290070301. 92WR01357.
Ashmore, P.E., 1991a. How do gravel-bed rivers braid? Can. J. Earth Sci. 28, 326341. Egozi, R., Ashmore, P., 2008. Dening and measuring braiding intensity. Earth Surf.
http://dx.doi.org/10.1139/e91-030. Process. Landforms 33 (14), 21212138. http://dx.doi.org/10.1002/esp.1658.
Ashmore, P., 1991b. Channel morphology and bed load pulses in braided, gravel- Engelund, F., 1970. Instability of erodible beds. J. Fluid Mech. 42 (02), 225244.
bed streams. Geogr. Ann. Ser. A, Phys. Geogr. 73, 37. http://dx.doi.org/10.2307/ http://dx.doi.org/10.1017/S0 022112070 0 01210.
521212. Engelund, F., Skovgaard, O., 1973. On the origin of meandering and braiding in
Antropovskiy, V.I., 1972. Quantitative criteria of channel macroforms. Soviet Hydrol. alluvial streams. J. Fluid Mech. 57 (02), 289302. http://dx.doi.org/10.1017/
477484. S0 0221120730 01163.
Ashworth, P.J., Best, J.L., Roden, J.E., Bristow, C.S., Klaassen, G.J., 20 0 0. Morphological Enggrob, H.G., Tjerry, S., 1999. Simulation of morphological characteristics of a
evolution and dynamics of a large, sand braid-bar, Jamuna River, Bangladesh. braided river. In: Proceedings of the IAHR Symposium on River, Coastal and Es-
Sedimentology 47, 533555. http://dx.doi.org/10.1046/j.1365-3091.20 0 0. tuarine Morphodynamics, Vol. I. Genova, pp. 585594. 6-10 Sept.
00305.x. Fredse, J., 1978. Meandering and braiding of rivers. J. Fluid Mech. 84 (04), 609624.
Bertoldi, W., Gurnell, A., Surian, N., Tockner, K., Zanoni, L., Ziliani, L., Zolezzi, G., http://dx.doi.org/10.1017/S0 0221120780 0 0373.
2009a. Understanding reference processes: linkages between river ows, sed- Fujita, Y., 1989. Bar and channel formation in braided streams. In: Ikeda, S.,
iment dynamics and vegetated landforms along the Tagliamento River, Italy. Parker, G. (Eds.), River Meandering. American Geophysical Union, Washington
River Res. Appl. 25 (5), 501516. http://dx.doi.org/10.1002/rra.1233. D.C., pp. 417462. http://dx.doi.org/10.1029/WM012p0417.
Bertoldi, W., Zanoni, L., Tubino, M., 2009b. Planform dynamics of braided streams.
Earth Surf. Process. Landforms 34 (4), 547557. http://dx.doi.org/10.1002/esp.
1755.
U. Singh et al. / Advances in Water Resources 104 (2017) 127144 143

Gaeuman, D., Andrews, E.D., Krause, A., Smith, W., 2009. Predicting fractional bed Miall, A.D., 1977. A review of the braided-river depositional environment. Earth-Sci.
load transport rates: Application of the Wilcock-Crowe equations to a regulated Rev. 13 (1), 162. http://dx.doi.org/10.1016/0012- 8252(77)90055- 1.
gravel bed river. Water Resour. Res. 45 (6), W06409. http://dx.doi.org/10.1029/ Mosselman, E., 2005. Basic equations for sediment transport in CFD for uvial
20 08WR0 07320. morphodynamics. In: Computational Fluid Dynamics. John Wiley & Sons, Ltd.,
Griths, G.A., 1979. Recent Sedimentation History of the Waimakariri River, New pp. 7189. http://dx.doi.org/10.10 02/0470 015195.ch4.
Zealand. J. Hydrol. (New Zealand) 18 (I), 628. Mosselman, E., 2012. Modelling Sediment Transport and Morphodynamics of
Hey, D., Thorne, C.R., 1986. Stable channels with mobile gravel beds. J. Hydraul. Gravel-Bed Rivers. In: Gravel-Bed Rivers. John Wiley & Sons, Ltd, Chichester, UK,
Eng., ASCE 112 (8), 671689. http://dx.doi.org/10.1061/(ASCE)0733-9429(1986) pp. 101115. http://dx.doi.org/10.1002/9781119952497.ch9.
112:8(671). Mosselman, E., Le, T.B., 2015. Five common mistakes in uvial morphodynamic
Hicks, D.M., Duncan, M.J., Walsh, J.M., Westaway, R.M., Lane, S.N., 2002. New views modelling. Adv. Water Resour. http://dx.doi.org/10.1016/j.advwatres.2015.07.025.
of the morphodynamics of large braided rivers from high-resolution topo- Nicholas, A.P., 20 0 0. Modelling bedload yield in braided gravel bed rivers. Geomor-
graphic surveys and time-lapse video. In: The structure, Function and manage- phology 36 (12), 89106. http://dx.doi.org/10.1016/S0169-555X(0 0)0 0 050-7.
ment Implications of Fluvial Sedimentary Systems, Proceedings of International Nicholas, A.P., 2013. Modelling the continuum of river channel patterns. Earth Surf.
Symposium. Alice Springs, Australia. IAHS Publ., pp. 373380. Process. Landforms 38, 11871196. http://dx.doi.org/10.1002/esp.3431.
Hicks, D.M., Duncan, M.J., Lane, S.N., Tal, M., Westaway, R.M., 2008. Contemporary Nicholas, a.P., Ashworth, P.J., Sambrook Smith, G.H., Sandbach, S.D., 2013. Numeri-
morphological change in braided gravel-bed rivers: new developments from cal simulation of bar and island morphodynamics in anabranching megarivers.
eld and laboratory studies, with particular reference to the inuence of ripar- J. Geophys. Res. Earth Surf. 118 (4), 20192044. http://dx.doi.org/10.1002/jgrf.
ian vegetation. In: Habersack, H., Piegay, H., Rinaldi, M. (Eds.), Gravel-bed Rivers 20132.
VI: From Process Understanding to River Restoration. Elsevier, pp. 557584. Parker, G., 1976. On the cause and characteristic scales of meandering and
Chapter 21. braiding in rivers. J. Fluid Mech. 76 (03), 457480. http://dx.doi.org/10.1017/
Hirano, M, 1971. River bed degradation with armouring. Trans. Japanese Soc. Civil S0022112076000748.
Eng. 3, 194195. Parker, G., 1991. Selective sorting and abrasion of river gravel. I: theory. J. Hy-
Huang, H.Q., 2010. Reformulation of the bed load equation of Meyer and Peter and draul. Eng. 117 (2), 131147. http://dx.doi.org/10.1061/(ASCE)0733-9429(1991)
Mller in light of the linearity theory for alluvial channel ow. Water Resour. 117:2(131).
Res. W09533. http://dx.doi.org/10.1029/20 09WR0 08974. Parker, G., 2007. Transport of gravel and sediment mixtures. In: Garcia, M.H. (Ed.).
Hundey, E.J., Ashmore, P.E., 2009. Length scale of braided river morphology. Water In: Sedimentation Engineering: Theories, Measurements, Modeling and Practice,
Resour. Res. 45 (8). http://dx.doi.org/10.1029/20 08WR0 07521, n/a-n/a. ASCE Manual Rep. Eng. Practice, 110, pp. 165264. http://dx.doi.org/10.1061/
Hunziker, R.P., Jaeggi, M.N.R., 2002. Grain sorting processes. J. Hydraul. Eng. 9780784408148.ch03, Chapter 3.
128 (12), 10601068. http://dx.doi.org/10.1061/(ASCE)0733-9429(2002)128: Parker, G., Klingeman, P.C., Mc Lean, D.G., 1982. Bedload and size distribution in
12(1060). paved gravel-bed streams. J. Hydraul. Div., ASCE 108 (HY4), 544571.
Jagers, H.R.A., 2003. Modelling Planform Changes of Braided Rivers PhD thesis. Parker, G., Wilcock, P.R., Paola, C., Dietrich, W.E., Pitlick, J., 2007. Physical basis
Tewnte University, Enschede, the Netherlands. for quasi-universal relations describing bankfull hydraulic geometry of single-
Jansen, PPh., van Bendegom, L., van den Berg, J., de Vries, M., Vanen, A., 1979. Prin- thread gravel bed rivers. J. Geophys. Res. 112 (F4), F04005. http://dx.doi.org/10.
ciples of River Engineering; the Non-Tidal Alluvial River. Pitman, London. 1029/20 06JF0 0 0549.
Kasprak, A., Wheaton, J.M., Ashmore, P.E., Hensleigh, J.W., Peirce, S., 2015. The rela- Peters, J.J., 1978. Discharge and sand transport in the braided zone of the Zaire es-
tionship between particle travel distance and channel morphology: Results from tuary. Netherlands J. Sea Res. 12 (3-4), 273292.
physical models of braided rivers. J. Geophys. Res. Earth Surf. 120 (1), 5574. Pigay, H., Grant, G., Nakamura, F., Trustrum, N., 2006. Braided river management:
http://dx.doi.org/10.10 02/2014JF0 03310. from assessment of river behaviour to improved sustainable development. In:
Kleinhans, M.G., van den Berg, J.H., 2011. River channel and bar patterns explained Braided Rivers. Blackwell Publishing Ltd., Oxford, UK, pp. 257275.
and predicted by an empirical and a physics-based method. Earth Surf. Process. Pigay, H., Alber, A., Slater, L., Bourdin, L., 2009. Census and typology of braided
Landforms 36, 721738. http://dx.doi.org/10.1002/esp.2090. rivers in the French Alps. Aquat. Sci. 71 (3), 371388. http://dx.doi.org/10.1007/
Kleinhans, M.G., Jagers, H.R.A., Mosselman, E., Sloff, C.J., 2008. Bifurcation dynam- s0 0 027-0 09-9220-4.
ics and avulsion duration in meandering rivers by one-dimensional and three Powell, D.M., 1998. Patterns and processes of sediment sorting in gravel-
dimensional models. Water Resour. Res. 44, W08454. http://dx.doi.org/10.1029/ bed rivers. Progress Phys. Geograph. 22 (1), 132. http://dx.doi.org/10.1177/
20 07WR0 05912. 030913339802200101.
Koch, F.G., Flokstra, C., 1980. Bed level computations for curved alluvial channels. Ribberink, J.S., 1987. Mathematical Modelling of One-Dimensional Morphological
Paper presented at XIXth Congress of the International Association for Hydraulic Changes in Rivers with Non-Uniform Sediment Phd Thesis. Delft University of
Research 2-7 Feb. 1981. Technology, Delft, The Netherlands.
Lane, E.W. (1957), A study of the shape of channels formed by natural streams ow- Rinaldi, M., Simoncini, C., Pigay, H., 2009. Scientic design strategy for promot-
ing in erodible material., edited, Missouri River Division Sediment Series No. 9. ing sustainable sediment management: the case of the Magra River (Central-
U.S. Army Engineer Division, Missouri River Corps of Engineers, Omaha, NE. Northern Italy). River Res. Appl. 25 (5), 607625. http://dx.doi.org/10.1002/rra.
Lanzoni, S., 20 0 0a. Experiments on bar formation in a straight ume: 1. Uniform 1243.
sediment. Water Resour. Res. 36 (11), 33373349. http://dx.doi.org/10.1029/ Roelvink, J.A., 2006. Coastal morphodynamic evolution techniques. Coastal Eng. 53
20 0 0WR90 0160. (23), 277287. http://dx.doi.org/10.1016/j.coastaleng.2005.10.015.
Lanzoni, S., 20 0 0b. Experiments on bar formation in a straight ume: 2. Graded Rust, B.R., 1972. Structure and process in a braided river. Sedimentology 18 (34),
sediment. Water Resour. Res. 36 (11), 33513363. http://dx.doi.org/10.1029/ 221245. http://dx.doi.org/10.1111/j.1365-3091.1972.tb0 0 013.x.
20 0 0WR90 0161. Sadler, J.P., Bell, D., Fowles, A., 2004. The hydroecological controls and conservation
Lanzoni, S., Tubino, M., 1999. Grain sorting and bar instability. J. Fluid Mech. 393, value of beetles on exposed riverine sediments in England and Wales. Biol. Con-
149174. http://dx.doi.org/10.1017/S0 0221120990 05583. serv. 118 (1), 4156. http://dx.doi.org/10.1016/j.biocon.20 03.07.0 07.
Le Lay, Y.F., Pigay, H., Rivire-Honegger, A., 2013. Perception of braided river land- Schielen, R., Doelman, A., de Swart, H.E., 1993. On the non-linear dynamics of free
scapes: Implications for public participation and sustainable management. J. En- bars in straight channels. J. Fluid Mech. 252, 325356. http://dx.doi.org/10.1017/
viron. Manage. 119, 112. http://dx.doi.org/10.1016/j.jenvman.2013.01.006. S0 0221120930 03787.
Leduc, P., Ashmore, P., Gardner, J.T., 2015. Grain sorting in the morphological active Schuurman, F., Kleinhans, M.G., 2011. Self-formed braided bar pattern in a numerical
layer of a braided river physical model. Earth Surf. Dyn. 3 (4), 577585. http:// model. In: Proceedings River, Coastal and Estuarine Morphodynamics. Tsinghua
dx.doi.org/10.5194/esurf- 3- 577- 2015. University Press, pp. 16471657.
Leopold, L.B., Wolman, M.G., 1957. River channel patterns: braided, meandering, and Schuurman, F., Marra, W.A., Kleinhans, M.G., 2013. Physics-based modeling of large
straight. In: Professional Paper. U.S. Geological Survey, pp. 3985. edited, pp. iv. braided sand-bed rivers: bar pattern formation, dynamics and sensitivity. J. Geo-
Lesser, G.R., Roelvink, J.A., van Kester, J.A.T.M., Stelling, G.S., 2004. Development and phys. Res.: Earth Surf. 118, 25092527. http://dx.doi.org/10.10 02/2013JF0 02896.
validation of a three-dimensional morphological model. Coastal Eng. 51 (89), Seminara, G., Tubino, M., 1989. Alternate bars and meandering: free, forced and
883915. http://dx.doi.org/10.1016/j.coastaleng.2004.07.014. mixed interactions. In: Ikeda, S., Parker, G. (Eds.). In: River Meandering, Wa-
Libault, F., Pigay, H., Frey, P., Landon, N., 2008. Tributaries and the management ter Resour. Monogr. Ser., vol. 12. AGU, Washington, D. C., pp. 267320. http://
of main-stem geomorphology. In: River Conuences, Tributaries and the Fluvial dx.doi.org/10.1029/WM012p0267.
Network. John Wiley & Sons, Ltd, Chichester, UK, pp. 243270. Siviglia, A., Crosato, A., 2016. Numerical modelling of river morphodynamics: latest
Marciano, R., Wang, Z.B., Hibma, A., de Vriend, H.J., Dena, A., 2005. Modeling of developments and remaining challenges. Adv. Water Resour. http://dx.doi.org/
channel patterns in short tidal basins. J. Geophys. Res. 110 (F1), F01001. http:// 10.1016/j.advwatres.2016.01.005.
dx.doi.org/10.1029/20 03JF0 0 0 092. Sloff, C.J., Ottevanger, W., 2008. Multiple-layer graded-sediment approach: improve-
Marra, W.A., 2008. Dynamics and Interactions of Bars in Rivers and the Relation ment and implications. In: Altinakar, K., Gogus, Tayfur, Kumcu, Yildirim (Eds.),
between Bars and a Braided River Pattern B.S. Thesis. Faculty of Geosciences, River Flow 2008, Taylor and Francis/Balkema. Proceedings, pp. 14471456.
Dep. of Physical Geography, Utrecht, Netherlands. Sloff, C.J., Jagers, H.R.A., Kitamura, Y., Kitamura, P., 2001. 2D Morphodynamic mod-
Mendoza, A., Abad, J.D., Langendoen, E.J., Asce, M., Wang, D., Tassi, P., El, K., elling with graded sediment. In: Proceedings of 2nd IAHR Symposium on River,
Abderrezzak, K. 2016. Effect of sediment transport boundary conditions on Coastal and Eastuarine Morphodynamics 10-14 September, 2001. Obiro, Japan.
the numerical modeling of bed morphodynamics, 112, doi:10.1061/(ASCE)HY. Sloff, K., Mosselman, E., 2012. Bifurcation modelling in a meandering gravel-sand
1943-790 0.0 0 01208. bed river. Earth Surf. Process. Landforms 37, 15561566. http://dx.doi.org/10.
Meyer Peter, E., Mller, R., 1948. Formulas for bed load transport. In: Proceedings of 1002/esp.3305.
2nd IAHR Congress. Stockholm, Sweden, pp. 3964.
144 U. Singh et al. / Advances in Water Resources 104 (2017) 127144

Sun, J., Lin, B., Yang, H., 2015. Development and application of a braided river model van der Wegen, M., Roelvink, J.A., 2008. Long-term morphodynamic evolution of
with non-uniform sediment transport. Adv. Water Resour. http://dx.doi.org/10. a tidal embayment using a two-dimensional, process-based model. J. Geophys.
1016/j.advwatres.2014.12.012. Res. 113, C03016. http://dx.doi.org/10.1029/20 06JC0 03983.
Surian, N., Rinaldi, M., 2003. Morphological response to river engineering and man- Wang, Z.B., Jeuken, M.C.J.L., Gerritsen, H., de Vriend, H.J., Kornman, B.A., 2002. Mor-
agement in alluvial channels in Italy. Geomorphology 50 (4), 307326. http:// phology and asymmetry of the vertical tide in the Westerschelde estuary. Conti-
dx.doi.org/10.1016/S0169- 555X(02)00219- 2. nental Shelf Res. 22 (17), 25992609. http://dx.doi.org/10.1016/S0278-4343(02)
Surian, N., Mao, L., Giacomin, M., Ziliani, L., 2009. Morphological effects of differ- 00134-6.
ent channel- forming discharges in a gravel-bed river. Earth Surf. Process. Land- Wilcock, P.R., Crowe, J.C., 2003. Surface-based transport model for mixed-size
forms 1107 (March), 10931107. http://dx.doi.org/10.1002/esp. sediment. J. Hydraul. Eng. 129 (2), 120128. http://dx.doi.org/10.1061/(ASCE)
Talmon, A.M., Struiksma, N., Van Mierlo, M.C.L.M., 1995. Laboratory measurements 0733-9429(2003)129:2(120).
of the direction of sediment transport on transverse alluvial-bed slopes. J. Hy- Willgoose, G., Hancock, G. , 1998. Revisiting the hypsometric curve as an indicator
draul. Res. 33 (4), 495517. http://dx.doi.org/10.1080/00221689509498657. of form and process in transport-limited catchment. Earth Surface Processes and
Teramoto, A., Tsujimoto, T., 2006. Effects of size heterogeneity of bed materi- Landforms 23 (7), 611623. http://dx.doi.org/10.1002/(SICI)1096-9837(199807)
als on mechanism to determine bar mode. In: Parker, G., Garcia, M. (Eds.), 23:7611::AID- ESP8723.0.CO;2- Y.
River, Coastal and Estuarine Morphodynamics RCEM 2005. Taylor & Francis, Ward, J.V., Tockner, K., Uehlinger, U., Malard, F., 2001. Understanding natural pat-
pp. 433444. terns and processes in river corridors as the basis for effective river restoration.
Tockner, K., Malard, F., Ward, J.V., 20 0 0. An extension of the ood pulse Regul. Rivers Res. Manag. 17, 311323. http://dx.doi.org/10.1002/rrr.646.
concept. Hydrol. Process. 14 (1617), 28612883. http://dx.doi.org/10.1002/ Wong, M., Parker, G., 2006. Reanalysis and correction of bed-load relation of meyer
1099-1085(20 0 011/12)14:16/172861::AID-HYP1243.0.CO;2-F. peter and mller using their own database. J. Hydraul. Eng. 132 (11), 11591168.
Tockner, K., Paetzold, A., Karaus, U., Claret, C., Zettel, J., 2006. Ecology of braided http://dx.doi.org/10.1061/(ASCE)0733-9429(2006)132:11(1159).
rivers. In: Braided Rivers. Blackwell Publishing Ltd, Oxford, UK, pp. 339359. Yossef, M.F.M., Sloff, K., 2012. Detailed modelling of river morphological response to
van den Berg, J.H., 1995. Prediction of alluvial channel pattern of perennial rivers. climate change scenarios. In: Murillo Munoz, R.E. (Ed.), River Flow 2012, Volume
Geomorphology 12 (4), 259279. http://dx.doi.org/10.1016/0169-555X(95) 1. CRC Press/Balkema, Leiden, the Netherlands, pp. 845852.
0 0 014-V.
van der Nat, D, Tockner, K., Edwards, P.J., Ward, J.V., Gurnell, A.M., 2003. Habitat
change in braided ood plains (Tagliamento, NE-Italy). Freshw. Biol. 48 (10),
17991812. http://dx.doi.org/10.1046/j.1365-2427.2003.01126.x.

You might also like