You are on page 1of 18

Advances in Water Resources 113 (2018) 55–72

Contents lists available at ScienceDirect

Advances in Water Resources


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

River banks and channel axis curvature: Effects on the longitudinal T


dispersion in alluvial rivers

Stefano Lanzoni ,a, Amena Ferdousib, Nicoletta Tambronic
a
Department of Civil, Environmental and Architectural Engineering, University of Padua, Italy
b
Faculty of Engineering and Technology, Eastern University, Dhaka, Bangladesh
c
Department of Civil, Chemical and Environmental Engineering, University of Genova, Genova, Italy

A R T I C L E I N F O A B S T R A C T

Keywords: The fate and transport of soluble contaminants released in natural streams are strongly dependent on the spatial
Alluvial rivers variations of the flow field and of the bed topography. These variations are essentially related to the presence of
Dispersion the channel banks and to the planform configuration of the channel. Large velocity gradients arise near to the
Meandering rivers channel banks, where the flow depth decreases to zero. Moreover, single thread alluvial rivers are seldom
straight, and usually exhibit meandering planforms and a bed topography that deviates from the plane con-
figuration. Channel axis curvature and movable bed deformations drive secondary helical currents which en-
hance both cross sectional velocity gradients and transverse mixing, thus crucially influencing longitudinal
dispersion. The present contribution sets up a rational framework which, assuming mild sloping banks and
taking advantage of the weakly meandering character often exhibited by natural streams, leads to an analytical
estimate of the contribution to longitudinal dispersion associated with spatial non-uniformities of the flow field.
The resulting relationship stems from a physics-based modeling of the flow in natural rivers, and expresses the
bend averaged longitudinal dispersion coefficient as a function of the relevant hydraulic and morphologic
parameters. The treatment of the problem is river specific, since it relies on an explicit spatial description,
although linearized, of the flow field that establishes in the investigated river. Comparison with field data
available from tracer tests supports the robustness of the proposed framework, given also the complexity of the
processes that affect dispersion dynamics in real streams.

1. Introduction homogeneous and extends far inside the equilibrium region, the ad-
vection-diffusion equation prescribes that the variance of C in the along
Estimating the ability of a stream to dilute soluble pollutants is a stream direction s* increases linearly with time and any skewness, in-
fundamental issue for the efficient management of riverine environ- troduced by velocity shear close to the contaminant source (i.e., in the
ments. Rapidly varying inputs of contaminants, such as those associated advective zone) or by the initial distribution of contaminant, begins to
with accidental spills of toxic chemicals and intermittent discharge slowly decay, eventually leading at any instant to a Gaussian distribu-
from combined sewer overflows, as well as temperature variations tion of C(s*) (Chatwin and Allen, 1985). The coefficient of apparent
produced by thermal outflows, generate a cloud that spreads long- diffusivity K* governing this behavior, usually denoted as dispersion
itudinally affecting the fate of the pollutant. coefficient, is much greater than the coefficient for longitudinal diffu-
The classical treatment of longitudinal transport in turbulent flows sion by turbulence alone.
relies on the study put forward by Taylor (1954) for pipe flows, and Many engineering and environmental problems concerning the fate
extended to natural channels by Fischer (1967). Taylor’s analysis in- and transport of pollutants and nutrients are tackled resorting to the
dicates that, far enough from the source (in the so called equilibrium one-dimensional advection-diffusion approach (Botter and Rinaldo,
region), the cross-sectionally averaged tracer concentration, C, satisfies 2003; Revelli and Ridolfi, 2002; Rinaldo et al., 1991; Schnoor, 1996;
a one-dimensional advection-diffusion equation, embodying a balance Wallis, 1994) and, therefore, require a suitable specification of the
between lateral mixing and nonuniform shear flow advection (Fickian longitudinal dispersion coefficient K*. Also within the context of much
dispersion model). Under the hypothesis that the velocity field is sta- more refined models developed to account for the formation of steep
tistically steady and the investigated channel reach is geometrically concentration fronts and elongated tails caused by storage and delayed


Corresponding author.
E-mail address: stefano.lanzoni@unipd.it (S. Lanzoni).

https://doi.org/10.1016/j.advwatres.2017.10.033
Received 26 July 2017; Received in revised form 27 October 2017; Accepted 28 October 2017
Available online 31 October 2017
0309-1708/ © 2017 Elsevier Ltd. All rights reserved.
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

Notations T0* [s] longitudinal dispersion time-scale.


T1* [s] differential advection time-scale.
A*[m2] local cross sectional area. T2* [s] transverse mixing time-scale.
Au* [m2] mean value of the cross sectional area in the reach. t*[s] time.
B*[m] free surface half width of the channel. U*[m/s] depth averaged longitudinal velocity.
Bc* [m] half width of the central channel region. Uu* [m / s] mean value of the cross sectionally averaged longitudinal
C = c [/] cross sectionally averaged concentration velocity in the reach.
c[/] depth averaged concentration u*[m/s] longitudinal component of the velocity.
cfu[/] friction coefficient uf* [m / s] local friction velocity.
C [m−1] channel curvature ufc* [m / s] scale for the friction velocity in the central region of the
D*[m] local flow depth. channel.
Dc* [m] flow depth of the central region of the channel. * [m / s]
ufu scale for the friction velocity under uniform flow condi-
Du* [m] cross-sectionally averaged flow depth. tions.
Dz* [m] flow depth measured normally to the bed. V*[m/s] depth averaged transverse component of velocity.
* [m]
dgr sediment grain size. z*[m] upward directed axis.
dr[/] discrepancy ratio. β[/] half free surface width to uniform depth ratio.
et* [m2 / s] depth averaged transverse eddy diffusity βc[/] half central region width to uniform depth ratio.
g[m/s2] gravitational acceleration. βf[/] cross section shape parameter measuring the steepness of
H*[m] elevation of the water surface with respect to an hor- the bank.
izontal reference plane γ[/] relative importance of transverse mixing and nonuniform
Hu* [m1.5] cross sectional average of D*1.5 transport
(hs, hσ)[/] metric coefficients. Δ[/] immersed relative sediment density
K*[m2/s] mean value of the longitudinal dispersion coefficient in δ[/] relative variation rate of the cross section in the transverse
the reach. direction.
Ku* [m2 / s] dispersion coefficient scale. ϵn* [m2 / s] transverse mixing coefficient contribution due to disper-
* [m2 / s] transversal mixing coefficient for a straight channel.
knu sion.
(ks*, kn*)[m2 / s] longitudinal and transversal mixing coefficient. λ[/] dimensionless meander wave number
kt* [m2 / s] transverse dispersion coefficient. ν[/] curvature ratio
L*[m] average intrinsic meander length. νT* [m2 / s] turbulent viscosity.
Lc* [m] contaminant cloud length ξ*[m] pseudo-lagrangian coordinate
n*[m] horizontal coordinate normal to s*. φ angle that the vertical forms with the normal to the bed
P0* [m] wetted perimeter of each bank region of the channel. ρ[kg/m3] water density.
Q*[m3/s] flow discharge. ρs[kg/m3] sediment density.
R 0* [m] twice the minimum value of the radius of curvature. σ*[m] transverse curvilinear coordinate.
r*[m] local radius of curvature. ζ*[m] coordinate normal to the bed.
S[/] longitudinal channel slope. τ*u[/] Shields parameter
s*[m] longitudinal curvilinear coordinate coinciding with the
channel axis.

release of pollutant in dead zones (see, among many others, Bencala averaged flow velocity within the reach of interest and κi (i = 0, 1, 2)
and Walters, 1983; Czernuszenko et al., 1998 and Bear and are suitable constants, specified in Table 1. Note that in Table 1 we have
Young, 1983), a reliable estimate of longitudinal dispersion in the main just reported the main formulas; for sake of completeness a wider list of
stream (quantified by K*) is fundamental to properly account for che- relationships is given in the Supplementary Information.
mical and biological processes acting in different channel regions The empirical parameters that are introduced in these relations to
(Lees et al., 2000). address the complexity embedded in the mixing process still make the
Several procedures have so far been proposed to estimate the quantifying of K* a challenging task. In many cases, the proposed
longitudinal dispersion coefficient from either tracer data predictor provides only a rough estimate, and the discrepancy between
(Rutherford, 1994) or velocity measurements at a number of cross the predicted values of K* and those determined from tracer tests is
sections (Fischer, 1967). These approaches are usually expensive and quite high. Among the many reasons responsible for this high scatter,
time consuming. The lack of experimental data which characterizes one may be the prismatic character assumed as the basis of the Fickian
many applications, as well as the necessity of specifying K* when car- solution (Wang and Huai, 2016). Nevertheless, natural channels are
rying out preliminary calculations, has thus stimulated the derivation of
various semi-empirical and empirical relationships (Deng et al., 2002;
Table 1
2001; Disley et al., 2015; Etamad-Shahidi and Taghipour, 2012;
Values attained by the constants of the generalized formula (1) and by the associated
Fischer, 1967; Iwasa and Aya, 1991; Kashefipour and Falconer, 2002; Li mean value of the discrepancy ratio dr (defined in Section 4.2) for various predictors
et al., 2013; Liu, 1977; Noori et al., 2017; Sahay and Dutta, 2009; Sattar available in literature, namely: (1) Fischer et al. (1979); (2) Seo and Cheong (1998); (3)
and Gharabaghi, 2015; Seo and Cheong, 1998; Shucksmith et al., 2011; Liu (1977); (4) Kashefipour and Falconer (2002); (5) Iwasa and Aya (1991); (6)
Wang and Huai, 2016; Zeng and Huai, 2014), which can be cast in the Deng et al. (2001); (7) Wang and Huai (2016). The dimensionless transverse eddy dif-
fusivity, et, and mixing coefficient, kt, are defined in Section 2.
general form:
(1) (2) (3) (4) (5) (6) (7)
β κ1
K * = κ0 B * Uu*,
( cfu ) κ2 κ0 0.044 9.1 0.72 10.612 5.66 0.06/(et + kt ) 22.7
(1) −0.38 −1.0 −0.64
κ1 1.0 1.0 0.5 0.67
κ2 1.0 0.428 −0.5 1.0 −1.0 1.0 0.16
where β is the ratio of half channel with, B*, to mean flow depth, Du*, cfu < dr > 1.05 1.24 1.10 1.05 1.04 0.63 1.13
is the friction coefficient, Uu* is the mean value of the cross-sectionally

56
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

usually characterized by a complex bed topography, which strongly Godfrey and Frederick (1970), which includes detailed measurements of
affects the flow field and, hence, the longitudinal dispersion flow depth, longitudinal velocity, and the temporal evolution of the
(Guymer, 1998), but is only roughly accounted for in the various ap- tracer concentration at different cross sections, as well as estimates of K*
proaches. In addition, in some cases a suitable tuning of the empirical based on the method of moments. These concentration data are here
parameters is needed in order to achieve a good agreement with the reanalyzed by considering the Chatwin’s method (Chatwin, 1980), which
experimental data (Deng et al., 2001). indicates if and where a Fickian model likely applies, and the routing
Many of the existing expressions for predicting the longitudinal method, based on the Hayami solution (Rutherford, 1994). We anticipate
dispersion in rivers have been developed by minimizing the error be- that the proposed framework provides estimates of K* that are in rea-
tween predicted and measured (through tracer tests) dispersion coef- sonable good agreement with the values computed from tracer tests.
ficients. These relations generally differ in terms of the relevant di- The paper is organized as follows. The mathematical problem is
mensionless groups (determined through dimensional analysis) and of formulated in Section 2, with particular emphasis on the typical tem-
the optimization technique (e.g., nonlinear multi regression, genetic poral and spatial scales which allow to set up a rational perturbative
and population-based evolutionary algorithms) used to calibrate the framework and eventually determine the overall structure of the dis-
coefficients of the predictor (Disley et al., 2015; Kashefipour and persion coefficient in alluvial channels. The analytical solutions of the
Falconer, 2002; Noori et al., 2017; Sahay and Dutta, 2009; Seo and depth averaged flow field used to compute K* are described in
Cheong, 1998). More recently, the dataset provided by the dispersion Section 3. The comparison with available field data is reported in
coefficients measured in the field has been used for training and testing Section 4, while Section 5 is devoted to the discussion of the results.
artificial neural networks or bayesian networks (Alizadeh et al., 2017). Finally, Section 6 summarizes the concluding remarks.
A less few attempts were devoted to derive analytical relationships by
substituting in the triple integral ensuing from Fischer analysis of shear 2. Formulation of the problem
flow dispersion the flow field that, under the uniform-flow assumption,
establishes in stable straight channels (Deng et al., 2001) and in We consider the behavior of a passive, non-reactive contaminant
meandering rivers (Deng et al., 2002). In the present contribution we which (e.g., due to an accidental spill) is suddenly released in an allu-
follow this latter approach, which has the advantage of being river vial channel with a compact cross section and, in general, a meandering
specific, i.e., to relate the dispersion coefficient to the shear flow dis- planform. The river cross section bed is assumed to vary slowly in the
persion that actually takes place in the river under investigation. The transverse direction as the banks are approached. This assumption al-
improvement with respect to the contributions by Deng et al. (2001, lows for solving the flow field by adopting a closure model of turbu-
2002) is essentially related to the morphodynamics-based modelling of lence in which the turbulent viscosity νT* is a function of the local flow
the flow that establishes in alluvial rivers. In the case of straight rivers, condition (see Section 3.1). The channel has fixed banks, a constant free
rather than using the general hydraulic geometry relationship for stable surface width 2B*, a longitudinal mean slope S, and conveys a constant
cross sections, we propose a specific treatment of the shear flow effects discharge Q* (hereafter a star superscript will be used to denote di-
by dividing the cross section into a central flat-bed region and two mensional variables). The reach averaged value of the flow depth is Du*.
gently sloping banks computing the flow field therein. In the case of The corresponding cross-section area is Au* = 2B *Du*, while the cross-
meandering rivers, the flow field outside the boundary layers that form sectionally averaged mean velocity is Uu* = Q*/ Au*. The erodible
near to the banks is solved explicitly, although in a linearised way, channel bed is assumed to be made up of a uniform cohesionless se-
accounting parametrically for the secondary flow circulations induced diment with grain size dgr * , density ρs, and immersed relative density
by streamline curvatures, and computing the bed topography by solving Δ = (ρs − ρ)/ ρ , with ρ the water density. Moreover, we denote by
the two-dimensional sediment balance equation. β = B */ Du* the half width to depth ratio, ufu * = (gDu*S )1/2 the friction
The aim of the present contribution is thus to develop physics- velocity (with g the gravitational constant), and cfu = (ufu * / Uu*)2 the
based, analytic predictions of the longitudinal dispersion coefficient, friction coefficient. These two latter quantities are influenced by the
accounting for the cross-sectional morphology occurring in alluvial bed configuration, which can be either plane or covered by bedforms
rivers. More specifically, we intend to relate the estimates of K* to the such as ripples and dunes, depending on the dimensionless grain size
relevant hydraulic, geometric and sedimentologic parameters (flow * / Du*, and the Shields parameter, τ u = ufu
dgr = dgr *2/(Δgdgr* ).
*
discharge, bed slope, representative sediment size, bank geometry)
governing the steady flow in an alluvial river. First, we apply to the 2.1. The 2-D dimensionless advection-diffusion equation
flow field which establishes in sinuous, movable bed channels the
perturbative procedure developed by Smith (1983), that accounts for The problem can be conveniently studied introducing the curvi-
the fast variations of concentration induced across the section by irre- linear orthogonal coordinate system (s*, n*, z*) shown in Fig. 1a, where
gularities in channel geometry and the presence of bends. This meth- s* is the longitudinal curvilinear coordinate coinciding with the channel
odology, introducing a reference system moving downstream with the axis, n* is the horizontal coordinate normal to s*, and z* is the upward
contaminant cloud and using a multiple scale perturbation technique, directed axis. The two-dimensional advection-diffusion equation for the
allows the derivation of a dispersion equation relating entirely to shear depth-averaged concentration c(s*, n*, t*) reads (Yotsukura, 1977):
flow dispersion the along channel changes in the cross-sectionally
∂c ∂c ∂c ∂ ⎛ D* ∂c ⎞
averaged concentration. Next, we take advantage of the weakly hs D * + D*U * + hs D*V * = ⎜ks* ⎟

meandering character of many natural rivers to clearly separate the ∂t * ∂s * ∂n* ∂s * ⎝ hs ∂s * ⎠


contributions to longitudinal dispersion provided by the various sources ∂ ⎛ ∂c ⎞
+ kn* hs D*
of nonuniformities. At the leading order of approximation, corre- ∂n* ⎝ ∂n* ⎠ (2)
sponding to the case of a straight channel, we consider the differential
where t* denotes time, D* is the local flow depth, U* and V* are the
advection related to the presence of channel banks, solving the flow
depth-averaged longitudinal and transverse components of the velocity,
field by means of a rational perturbation scheme (Tubino and
and ks* and kn* are longitudinal and transverse mixing coefficients which
Colombini, 1992). At the first order of approximation, we introduce the
account for the combined effect of vertical variations of velocity and
correction to K* due to the presence of bends by using the hydro-
turbulent diffusion. Moreover, hs is a metric coefficient, arising from the
morphodynamic model of Frascati and Lanzoni (2013).
curvilinear character of the longitudinal coordinate, defined as:
The proposed methodology is finally validated through the compar-
ison with the tracer test data collected in almost straight and in mean- n*
hs = 1 + = 1 + ν n C,
dering rivers. Among others, we consider the dataset provided by r* (3)

57
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

where r*(s*) is the local radius of curvature of the channel axis, as- establishes in the meandering channel. The derivation chain rule im-
sumed to be positive when the center of curvature lies along the ne- plies that:
gative n*-axis, ν = B */ R 0* is the curvature ratio, n = n*/ B * is the di-
∂ ∂ ∂ ∂ ∂ ∂
mensionless transverse coordinate, C = R 0*/ r * is the dimensionless = +A , = − Uu*
∂s * ∂s * ∂ξ * ∂t * ∂t * ∂ξ * (6)
channel curvature, and R 0* is twice the minimum value of r* within the
meandering reach. where A = A */ Au*. Consequently, for an observer moving with velo-
In meandering channels the cross-sectionally averaged concentra- city Uu* (i.e., with the advected pollutant cloud) the dilution of the
tion undergoes relatively small and rapidly changing gradients, asso- concentration associated with longitudinal dispersion is accounted for
ciated with the spatial variations of the flow field along the bends, and a by the coordinate ξ* and occurs at a length scale comparable with the
slower evolution due to longitudinal dispersion. In order to deal with length of the contaminant cloud, Lc* . It then results that
the fast concentration changes acting at the meander scale, it proves c = c (s *, n*, ξ *, t *), and Eq. (2) can be rewritten as:

∂c ∂c ∂ ⎛ ∂c ⎞ ∂c ∂c ∂ ⎛ D* ∂c D* ∂c ⎞
D*U * + hs D*V * − kn* hs D* = D* (hs Uu* − AU *) − hs D * + ⎜k *
s + A ks* ⎟
∂s * ∂n* ∂n* ⎝ ∂n* ⎠ ∂ξ * ∂t * ∂s * ⎝ hs ∂s * hs ∂ξ * ⎠
∂ ⎛ D* ∂c D* ∂c ⎞
+A ⎜ k* + A ks* ⎟
s
∂ξ * ⎝ hs ∂s * hs ∂ξ * ⎠ (7)

convenient to introduce a pseudo-lagrangian, volume following co- In order to better appreciate how transverse mixing, differential
ordinate, ξ*, which travels downstream with the contaminant cloud and advection, longitudinal dispersion and spatial changes in bed topo-
accounts for the fact that the cross-sectionally averaged velocity Q*/A* graphy contribute to dilute the pollutant concentration, Eq. (2) is made
is not constant along the channel (Smith, 1983). This coordinate is dimensionless introducing the following scaling:
defined as:
s * = L*s, ξ * = Lc*ξ , n* = B *n, D* = Du*D , (8)
1 s*
ξ* =
Au*
∫0 A * ds * − Uu* t *
(4) L* ⎞
t * = T0*t , (U *, V *) = Uu* ⎛U , V , (ks*, kn*) = knu
* (ks, kn )
⎝ B* ⎠ (9)
where the integral on the right side is the water volume from the origin
of the coordinate system to the generic coordinate s*, while where L* is the average intrinsic meander length within the in-
B* B*
* is the transverse mixing coefficient for
vestigated reach (see Fig. 1a), knu
A* = ∫−B* D* dn*, A* = ∫−B* hs D* dn*. (5) a straight channel configuration, and T0* is the typical timescale at
which longitudinal dispersion operates within the contaminant cloud.
Clearly, A * and A* can vary along s* as a consequence of the var- Besides the timescale T0* = Lc* 2 / Ku* (where Ku* is a typical dispersion
iations of section geometry induced by the bed topography that coefficient), other two timescales, T1* = Lc* / Uu* and T2* = B * 2/ knu *,

Fig. 1. Sketch of a meandering channel and notations. a) Plan view. b) Typical cross-section in a neighborhood of the bend apex. Note the scour at the outer bank and the deposition
caused by the point bar at the inner bank. c) Average cross-section, typically occurring nearby the inflection point of the channel axis.

58
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

characterize the processes (longitudinal dispersion, differential advec- The parameter γ describes the relative importance of transverse mixing,
tion and transverse mixing) that govern the concentration dynamics of which tends to homogenize the contaminant concentration, and non-
the pollutant cloud. In order to ensure that they are well separated uniform transport at the bend scale, which, on the contrary, enhances
(Fischer, 1967; Smith, 1983), we introduce the small parameter concentration gradients. It is readily observed that γ = ϵ−1λ /2π , where
* the dimensionless meander wavenumber λ = 2πB */ L* typically ranges
knu
ϵ= , between 0.1 and 0.3 (Leopold et al., 1964). The product γϵ then turns
B *Uu* (10)
out of order O (10−2) and, hence, gives rise to higher order terms in the
and recall the relationships usually adopted to predict the transverse perturbation analysis described in the next Section.
and longitudinal mixing coefficients knu * and Ku*.
The rate of transverse mixing is determined by turbulent diffusion, 2.2. The longitudinal dispersion coefficient
quantified by the depth averaged transverse eddy diffusivity et*, and
vertical variations in the transverse velocity, quantified by the trans- The presence of different spatial and temporal scales can be handled
verse dispersion coefficient kt* (Rutherford, 1994). Both coefficients employing a multiple scale technique (Nayfeh, 1973). To this purpose
* Du* and, consequently, the transverse mixing coefficient can
scale as ufu we expand the concentration c = c (s, n, ξ , t ) as:
be expressed as:
c = c0 + ϵ c1 + ϵ 2 c2 + ⋯ (16)
* = (et + kt ) ufu
knu * Du* (11)
We substitute this expansion into (14), and consider the problems
Recalling that * Du* = B * Uu* cfu / β ,
ufu it results that arising at various orders of approximation:
ϵ = (et + kt ) cfu / β . Experimental observations in straight rectangular
O (ϵ 0) L c0 = 0 (17)
flumes indicate that et usually falls in the range (0.10 - 0.26), with a
mean value equal to 0.15 (Rutherford, 1994). On the other hand, for ∂c0
large rivers the transverse dispersion coefficient kt has been related to O (ϵ) L c1 = D (hs − UA )
∂ξ (18)
the mean flow velocity and the channel width through a relation of the
form (Smeithlov, 1990): ∂c1 ∂c
O (ϵ 2) L c2 = D (hs − UA ) − hs D 0 ,
1.38 ∂ξ ∂t (19)
⎡ 1 ⎛ Uu* ⎞ ⎛ 2B * ⎞ ⎤
kt = ⎢ ⎜ ⎟⎜ ⎟

* ⎝ Du* ⎠
3520 ⎝ ufu coupled with the requirements that ∂ci/ ∂n = 0 (i= 0, 1, 2) at the channel
⎣ ⎠ ⎦ (12)
banks, where the normal component of the contaminant flux vanishes.
Observing that the ratio cfu / β attains values of order O (10−2) and The partial differential Eqs. (17)–(19) provide a clear insight into
O (10−3) in gravel and sandy rivers (Hey and Thorne, 1986; Parker, the structure of the contaminant concentration. Recalling that, for a
2004), it results that the parameter ϵ is indeed small. steady open channel flow, the depth-averaged (i.e., two-dimensional)
According to the semi-empirical relationship developed by continuity equation, written in dimensionless form, reads:
Fischer et al. (1979), Ku* = 0.044 (B *Uu*)2 /(ufu
* Du*) . This functional de-
pendence is confirmed by the dispersion data reported by ∂ (UD) ∂ (hs VD)
+ = 0,
Rutherford (1994), indicating that the dimensionless ratio Ku*/ B *Uu* ∂s ∂n (20)
mostly falls in the range 0.14 − 36, with mean 4.4 and standard de- we integrate (17) across the section and find that c0 does not depend on
viation 5.0. We can then write: s, n and, hence, it is not affected by the fluctuations induced by flow
*
knu k + kt cfu meandering. Eq. (18) suggests a solution of the form c1 = g1 (s, n) ∂c0/ ∂ξ ,
= n ∼ ϵ2, with g1 a function describing the nonuniform distribution across the
Ku* 0.044 β 2 (13)
section of the contaminant concentration. Similarly, Eq. (19) indicates
Consequently, T1*/ T0* = ϵ and T2*/ T0* = ϵ2,
provided that B */ Lc* = ϵ 2, that c2 = g2 (s, n) ∂2c0/ ∂ξ 2 . The depth-averaged concentration then re-
that is the contaminant cloud has reached a length of order of hundred sults:
of meters or kilometers, depending on the width of the channel section.
∂c0 ∂ 2c
This result implies that the three time scales are well separated, i.e. c (s, n, ξ , t ) = c0 (ξ , t ) + ϵ g1 (s, n) + ϵ 2 g2 (s, n) 20 + O (ϵ3)
T0* < T1* < T2*. In other words, the longitudinal dispersion operates on a ∂ξ ∂ξ (21)
timescale much slower than the timescale characterizing transverse and clearly discriminates the slower evolution due to longitudinal dis-
mixing which, in turn, is much faster than nonuniform advection persion, embodied by the terms c0, ∂c0/∂ξ, ∂2c0/∂ξ2, from the small and
(Fischer, 1967; Smith, 1983). rapidly varying changes associated with the spatial variations of the
The derivation of the longitudinal dispersion coefficient takes ad- flow field, described by the functions g1 and g2.
vantage of the small character of the parameter ϵ, ensuring the se- Integrating (21) across the section and along a meander, the cross-
paration of the three timescales. Substituting the dimensionless vari- sectionally averaged concentration can be approximated as
ables (8) and (9) into Eq. (7), we obtain: c = c0 + O (ϵ3) provided that < gi > =0 (i = 1, 2 ), where:
∂c ∂c ∂ ∂ γ D ∂c D ∂c ⎞ 1 1 1 1 s + 1/2
− ϵ 2 hs D + ϵ 2 ⎜⎛γ + ϵA ⎞⎟ ⎛⎜ ks
Lc = ϵ D (hs − AU )
∂ξ ∂t ⎝ ∂s ∂ξ ⎠ ⎝ ϵ hs ∂s
+ A ks ⎟
hs ∂ξ ⎠ c=
2A
∫−1 hs D c dn, gi =
2A
∫−1 hs D gi dn, gi = ∫s−1/2 gi ds

(14) (22)
where the differential operator L reads: It is important to note that only averaging (21) along the entire
meander length ensures that the arbitrary constant embedded in gi does
∂ ∂ ∂ ⎛ ∂
L = γ D ⎛U + hs V ⎞ − hs D k n ⎞ not actually depend on s.
⎝ ∂s ∂n ⎠ ∂n ⎝ ∂n ⎠ (15)
We are now ready to derive the advection-diffusion equation, gov-
The additional parameter γ = ϵLc* / L* arises because of the presence erning the evolution of the cross-sectionally averaged concentration c,
of two spatial scales. The spatial variations of c associated with long- and the related longitudinal dispersion coefficient. We sum together
itudinal dispersion at the scale of the contaminant cloud are described Eqs. (18) and (19), integrated across the section and along a bend, and
1
by the slow variable ξ, whereas the comparatively small and rapidly require that the flux of contaminant vanishes, i.e., ∫−1 DUgi dn = 0
changing variations in concentration across the flow associated with (i=1,2), a condition needed in order to eliminate secular terms which
stream meandering are accounted for through the fast variables s, n. would lead c2 to grow systematically with s. We eventually obtain:

59
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

∂c0 ∂ 2c 0 [U (s, n), D (s, n), A (s )] = [U0 (n), D0 (n), 1]


=K + O (ϵ)
∂t ∂ξ 2 (23)
+ ν [U1 (s, n), D1 (s, n), A1 (s )]
where: + ν 2 [U2 (s, n), D2 (s, n), A2 (s )] + O (ν 3)
(28)
1 1 h
K= A 2K , K=
2A

−1
D ⎛ s − U ⎞ g1 dn,
⎝ A ⎠ (24)
0
where the unperturbed O(ν ) state corresponds to a straight channel.
Similarly, we expand in terms of ν the function g1 and the dimensionless
transverse mixing coefficient kn:
while the function g1(s, n), describing the cross-sectional distribution of
the concentration, results from the solution of the O(ϵ) equation: [kn (s, n), g1 (s, n)] = [kn0 (n), g10 (n)] + ν [kn1 (s, n), g11 (s, n)] + O (ν 2)
(29)
L g1 = D (hs − UA ), (25)
The longitudinal dispersion coefficient in meandering channels is
with the requirements that ∂g1/ ∂n = 0 at the channel banks, where the determined by substituting (28) and (29) into (24), and recalling (3).
normal component of the contaminant flux vanishes, and g1 = 0 . We obtain:
Before proceeding further, some observations on the expression (24)
K = K 0 + ν K1 + ν 2 K2 + O (ν 3) (30)
are worthwhile. In accordance with Fischer (1967), the contribution to
longitudinal dispersion provided by vertical variations of the velocity where
profile (embodied by the terms of (14) containing ks) is of minor im- 1 1
portance. Longitudinal dispersion is essentially governed by shear flow K0 =
2
∫−1 (1 − U0) D0 g10 dn (31)
dispersion induced by the nonuniform distribution across the section of
both the contaminant concentration, accounted for through the func- 1 1

tion g1(s, n), and the flow field, quantified by D (hs − UA ). This latter
K1 =
2
∫−1(1 − U0) < D0 g11 + D1 g10 > dn

term, however, differs from the much simpler term (1 − U ) that would 1 1

arise in the classical treatment pursued by Fischer (1967), as a con-


+
2
∫−1<nC − U1 − U0 A1 > D0 g10 dn
(32)
sequence of the fact that here the mean flow velocity can in general
1 1
vary along the channel, as accounted for through the volume-following K2 =
2
∫−1(1 − U0) < D0 g12 + D1 g11 + D2 g10 > dn
coordinate ξ. In addition, it is important to observe that the bend
1 1
averaged coefficient K is always positive while, in the presence of river +
2
∫−1<(nC − U1 − U0 A1) (D0 g11 + D1 g10) > dn
reaches characterized by rapid longitudinal variations of the flow field, 1 1
the coefficient K can also attain negative values, thus favoring spur- −
2
∫−1<U2 + U1 A1 + U0 A2 > D0 g10 dn
(33)
ious instabilities (Smith, 1983).
Finally, it is useful to relate the local and the bend averaged dis- It is immediately recognized that the leading order contribution K0
persion coefficients, K and K, to the local dispersion coefficient D that corresponds to the classical solution obtained by Fischer (1967). It
arises when considering only the fast coordinate s. Decomposing the accounts for dispersion effects which arise in a straight uniform flow as
concentration c and velocity U* as the sum of their cross-sectionally a consequence of the transverse gradients experienced by U0 and the
averaged values, c , U *, plus the corresponding fluctuations c′, U′*, the concentration distribution in the bank regions (Fig. 1). Note that ne-
classical one-dimensional advection-dispersion equation results: glecting these effects is equivalent to set U0 = D0 = 1, such that K 0 = 0 .
It is also easy to demonstrate that the O(ν) correction K1 is identically
∂c Q* ∂c 1 ∂ B* zero. Indeed, A1 = 0, and the various integrals involve products of even
∂t *
+
A * ∂s *
=−
A * ∂s *
∫−B* D*U ′*c′dn* (26) (1 − U0, D0, g10) and odd (D1, U1, g11, n) functions that, integrated
across a symmetrical section, yield a zero contribution. Finally, the O
Setting c′ = (B *2Uu*/ knu
* ) g1 ∂c / ∂s *, after some algebra it can be de- (ν2) term K2 includes the effects of the near bank velocity and con-
monstrated that: centration gradients, mainly represented by the first integral on the
right hand side of Eq. (33), and those due to the complex structure of
(Uu* B *)2 the flow field, the bed topography and the spatial distribution of the
D* = D,
*
knu concentration induced by the meandering stream. The former con-
1 1
tribution to K2 is likely of minor importance when dealing with wide
D =
2 A −1

D (U − U ) g1 dn,
(27) and shallow sections (i.e., with large β), as often occurs in alluvial rivers
and, in the following will be neglected in order to keep the model at the
and, consequently, K = D and K = A 2D . lower level of complexity. In fact, as it will be seen in the next Section,
the solution of the flow field in a meandering channel is available in
closed form only by neglecting the boundary layers that form near to
2.3. Structure of the longitudinal dispersion coefficient in meandering the banks (Frascati and Lanzoni, 2013).
channels The functions g1i(s, n) (i = 0, 1) that describe the cross sectional
distribution of c are obtained by solving the partial differential equa-
Natural channels are seldom straight. In the general case of a tions that arise by substituting from (28) and (29) into (25). They read:
meandering planform configuration, the problem can be faced by
taking advantage of the fact that the curvature ratio ν appearing in (3) ∂g1i ∂ ⎛ ∂g
γD0 U0 − D0 kn0 1i ⎞ = f1i (s, n) i = 0, 1
⎜ ⎟

is typically a small parameter, ranging in the interval 0.1 − 0.2 ∂s ∂n ⎝ ∂n ⎠ (34)


(Leopold et al., 1964). This evidence is widely used to describe the flow
field in meandering channels (Seminara, 2006) and to model their long and are subject the constraints that ∂g1i / ∂n = 0 at the walls and
term evolution (Frascati and Lanzoni, 2010; 2013). It implies that the 1

flow field and the bed topography of a meandering channel can be


∫−1 D0 < g1i > dn = b1i
determined by studying the relatively small perturbations associated with
1
with deviations from a straight channel configuration. We then in- b10 = 0, b11 = − ∫−1 <g10 (D1 + D0 nC ) > dn. (35)
troduce the expansions:

60
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

The forcing terms f1i are obtained recalling the expression of the σ* is the transverse curvilinear coordinate aligned along the cross-sec-
metric coefficient hs, and read: tion profile (with origin at the channel axis), and ζ* is the coordinate
normal to the bed (pointing upward) (see Fig. 1c). The curvilinear
f10 = D0 (1 − U0) (36) nature of σ* is accounted for through the metric coefficient:
dg10 ζ * ∂2D0* ∂D *
2
f11 = (nC − U1 − U0 A1) + D1 (1 − U0) − γ D0 V1 hσ = 1 + , cos φ = 1−⎛ 0⎞ ,
∂n cos φ ∂σ *2
⎜ ⎟

⎝ ∂σ * ⎠ (39)
∂ ⎡ dg
− (kn1 + nCkn0) D0 + kn0 D1) 10 ⎤ with D0* (σ *) the local value of the flow depth, φ the angle that the

∂n ⎣ dn ⎥⎦ (37)
vertical forms with ζ*, and Dz* = D0*/cos φ the flow depth measured
where kni = Di , having assumed that kn* = (et + kt ) ufu * D* (Deng et al., normally to the bed (Fig. 1).
2001). The uniform character of the flow implies that, on average, the flow
The solution of the boundary value problem given by (34) and its characteristics do not vary in time and along the direction s*. Hence,
constraints is in general given by the sum of a homogeneous solution, denoting by u*(σ*, ζ*) the corresponding component of the velocity, the
common to any order of approximation, and a particular solution re- longitudinal momentum equation, averaged over the turbulence, reads
lated to the forcing term f1i. The homogeneous solution can be written (Appendix A):
in term of Fourier series, and generally depends on the transverse dis-
tribution of concentration at the injection section. However, in the case ∂ ⎛ νT* ∂u* ⎞ ∂ ⎛ ∂u* ⎞
S hσ g + ⎜ ⎟ + ⎜ν *h σ ⎟ = 0,
T
of a sudden release of contaminant treated here, it tends to decrease ∂σ * ⎝ hσ ∂σ * ⎠ ∂ζ * ⎝ ∂ζ * ⎠ (40)
exponentially with the coordinate s and, hence, vanishes far down- where νT* is the eddy-viscosity used to express the turbulent Reynolds
stream of the input section (Smith, 1983). This condition is equivalent stresses through the Boussinesq eddy viscosity approximation.
to impose that the O(ϵ) and O(ϵ2) pollutant fluxes vanish In general, the channel cross section is assumed to consists of
1
( ∫−1 DUgi dn = 0 for i = 1, 2) , as required in the derivation of Eq. (23). (Fig. 1c): i) a central region of width 2Bc* and constant depth depth Dc*,
Finally, note that, for the uniform flow in a straight channel, the and ii) two bank regions, each one characterized by a width (B * − Bc* )
along channel gradient of g10 is identically zero and (34) yields the and wetted perimeter P0*. In natural channels the flow depth is usually
classical relation (Fischer, 1967; 1973; Rutherford, 1994): much smaller than the wetted perimeter and, consequently, the di-
mensionless parameter
∫−1 ⎡⎢ D01kn0 ∫−1
n n1
g10 (n) = D0 (U0 − 1) dn2 ⎤ dn1 + α 0
⎣ ⎥
⎦ (38) Du*
δ=
P0* + Bc* (41)
where the constant α0 allows g10 to satisfy the integral condition (35),
but does not give any contribution to K0. is small. We will take advantage of this for solving Eq. (40). To this aim,
we introduce the scaling:
3. Depth averaged flow field in alluvial channels
ζ* σ* D* u*
ζ= , σ= , D0 = 0 , u = ,
The characteristics of the steady flow that establishes in alluvial Dz* (σ *) P0* + Bc* Du* *
ufu (42)
channels are determined by the form of the cross section and the
planform configuration of the channel. The governing two-dimensional νT* uf*
νT = , uf = ,
equations of mass and momentum conservation are in general obtained Du*ufu
* *
ufu (43)
by depth-averaging the corresponding three-dimensional equations,
and by accounting for the dynamic effects of secondary flows induced where uf* = (gD0*S )1/2 is the local value of the friction velocity, related
by curvature and of the boundary layers that form near to the channel to the local bed shear stress τb* by the relation uf* = (τb*/ ρ)1/2 . Note that,
banks. having normalized ζ* with Dz*, it turns out that:
The complexity of the problem prevents the derivation of general
∂ 1 ∂ ∂ 1 ⎛⎜ ∂ − ζF1 ∂ ⎞⎟,
solutions in closed form. Also numerical solutions are non straightfor- = , =
∂ζ * Dz* ∂ζ ∂σ * P0* + Bc* ⎝ ∂σ ∂ζ ⎠ (44)
ward, owing to the difficulty of modeling secondary circulations
(Bolla Pittaluga and Seminara, 2011). However, the governing equa- with
tions can be linearized in the presence of gently sloping channel banks
and meandering channels with wide and long bends, such that the flow 1 ∂D0 ⎡ δ 2 D0 (∂2D0 / ∂σ 2) ⎤
F1 = 1+ .
field can be solved by perturbing the uniform flow solution in terms of ⎢
D0 ∂σ ⎣ 1 − δ 2 (∂D0 / ∂σ )2 ⎥
⎦ (45)
two small parameters δ and ν. We resort just to these solutions, which
Substituting (42) and (43) into (40), the dimensionless longitudinal
have the pratical advantage to explicitily account, although in a sim-
momentum equation results:
plified form, for the effects exerted on the base flow field by the bank
shape and the channel axis curvature. We then estimate analytically the 1 ∂ ⎡ ∂u ⎤ ∂ ∂ ν ∂ ∂
longitudinal dispersion coefficient through (31) and (33). hσ νT + δ 2 ⎛⎜ − ζF1 ⎞⎟ ⎡ T ⎜⎛ − ζF1 ⎟⎞ u ⎤ + hσ = 0
Dz2 ∂ζ ⎢ ∂ζ ⎥ ∂σ ∂ζ ⎢ hσ ∂σ ∂ζ⎠ ⎥
⎣ ⎦ ⎝ ⎠⎣ ⎝ ⎦
In the following, we first derive the cross-sectional distribution of
the longitudinal velocity U0 in a straight channel with gently sloping (46)
banks (small δ). Next, we briefly recall the structure of U1 in wide and Under the assumption that the transverse slope of the channel bank
long meander bends (small ν) with either an arbitrary or a regular varies slowly, such that the normals to the bed do not intersect each
distribution of the channel axis curvature. other, it is possible to express the dimensionless eddy viscosity νT as:
νT (ζ ) = uf Dz N (ζ ), (47)
3.1. Straight channels
The simplest model for the function N (ζ ) is that introduced by
The uniform turbulent flow field that establishes throughout a cross Engelund (1974), whereby N = 1/13. In the following, we will adopt
section of a straight channel can be conveniently studied introducing this scheme which allows for an analytical solution of the problem, and,
the local orthogonal coordinate system (s*, σ*, ζ*), where s* is the as shown by Tubino and Colombini (1992), leads to results that agree
longitudinal (in this case straight) coordinate (directed downstream), both qualitatively and quantitatively with those obtained with a more

61
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

accurate model for the function N (ζ ) . Under the assumption of a consider only values of βf ≥ erf −1 (0.999) = 2.32675, corresponding to
constant N , a slip condition has to be imposed at the bed, such that: Dc* < D0* (0) ≤ 0.999Dc* . Note also that βf is related to the parameter δβc
through the relation:
⎡ Dz ⎤
u ζ = 0 = uf ⎢2 + 2.5 ln ⎜⎛ ⎟⎞ ⎥. 2
d
⎝ gr ⎠ ⎦ (48) ⎡ erf −1 (0.999) ⎤
⎣ δ βc = ⎢1 −
βf ⎥
The other two boundary conditions to be associated to Eq. (46) ⎣ ⎦ (53)
require that the dimensionless shear stress vanishes at the water surface where βc = Bc*/ Du* and having assumed D0* (Bc*) ≃ 0.999Dc* . As βf in-
and equals uf at the bed: creases also δβc increases, resulting in progressively steeper cross sec-
2
tions (Fig. 2a). Note that increasing values of δ imply higher bank
⎡ν ⎛⎜ 1 ∂u − δ ∂u ⎞⎟ ⎤ ν ∂u ⎤
= 0, ⎡ T = uf2. slopes. In the limit of βf = 2.32675, it results δβc = 0, corresponding to
⎢ T D0 ∂ζ ⎥ ⎢
⎣ ⎝ h σ ∂σ ⎠ ⎦ζ = 1 ⎣ Dz ∂z ⎥
⎦ζ = 0 (49) Bc* = 0 (no central region), while as βf → ∞, the classical rectangular
cross-sectional configuration (P0* = 0 ) is recovered (Fig. 2a).
The presence of the small parameter δ allows the expansion of the
Interestingly, the distributions of uf1(σ) shown in Fig. 2b indicate an
flow variables as:
increase of the friction velocity uf, with respect to the uniform flow, in
(u, uf ) = (u 0 , uf 0) + δ 2 (u1, u f1) + O (δ 4 ). (50) the steeper portion of the bank, and a corresponding decrease in the
part of the bank adjacent to the central region. This trend, due to the
The cross sectional distribution u(σ, ζ) of the longitudinal velocity is longitudinal momentum transfer from the center of the cross section
obtained by substituting this expansion into Eqs. (46), (48) and (49), by (where flow velocities are higher) to the banks, implies that the channel
collecting the terms with the same power of δ2, and by solving the re- can transport sediments even though the bank toes are stable. Note also
sulting differential problems (see Appendix A). Integrating u along the that uf1(σ) vanishes towards the center of the cross section, where the
normal ζ to the bed, the local value U0(σ) of the depth-averaged long- bottom is flat, and at the outer bank boundary, where D0 tends to zero.
itudinal velocity results:
U0 = U00 + δ 2 U01 + O (δ 4 ) (51) 3.2. Meandering channels
where U00 is a function of the local flow depth D0(σ) and the relative
The flow field that takes place in a meandering channel with a
grain roughness dgr, while U01 depends also on ∂D0/∂σ (i.e., the local
compact cross section is strictly related to the secondary flow circula-
slope) and ∂2D0/∂σ2 (Appendix A).
tions driven by the curvature of streamlines and the deformation of the
The cross sectional distribution of U0 needed to compute the long-
channel bed, which generally exhibits larger scours in the correspon-
itudinal dispersion coefficient is then determined by specifying the re-
dence of the outer bank of a bend (Seminara, 2006). Although nu-
lative bed roughness dgr and, more importantly, the across section
merical models have the advantage to overcome the restrictions af-
distribution of the flow depth D0(σ). In the absence of experimental
fecting theoretical analyses (e.g., linearity or weak non-linearity,
data, we need to describe the bank geometry. Here, we propose to
simplified geometry) they still require a large computational effort to
handle empirically the problem assuming a transverse distribution of
correctly include the effects of secondary helical flow and to reproduce
the flow depth of the form:
the bed topography of movable bed channels (Bolla Pittaluga and
D0* (σ ) = Dc* erf [βf ( 1 − σ )], σ ∈ [−1, 1], Seminara, 2011; Eke et al., 2014). That is why linearized models have
(52)
been widely adopted to investigate the physics of river meandering
with βf a shape parameter measuring the steepness of the banks. Note (Seminara, 2006), the long-term evolution of alluvial rivers (Bogoni
that, according to (52), erf(βf) should be equal to 1 in order to ensure et al., 2017; Frascati and Lanzoni, 2009; Howard, 1992), and the pos-
that D0* (0) = Dc* . The latter requirement is fulfilled only asymptotically, sible existence of a scale invariant behavior (Frascati and
for βf tending to infinity. For this reason, in the following we will Lanzoni, 2010). These models, owing to their analytical character, not

Fig. 2. (a) Cross-sectional bed profile for various βf and for δ = 0.1 and δ = 0.2 . (b) Cross-section distributions along the σ coordinate of the O (δ 2) corrections provided to the di-
mensionless friction velocity uf1, for the values of βf considered in plot a) and for dgr = 0.02 .

62
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

only provide insight on the basic mechanisms operating in the process in excess of the transverse mixing distance, ∼ U0*B *2/ kn*. Consequently,
under investigation, but also allow to develop relatively simple en- the position s0 of the injection section can be set arbitrarily far up-
gineering tools which can be profitably used for practical purposes. stream, taking s0 = −∞. Physically, this is equivalent to assume that the
In the following we refer to the linearized hydro-morphodynamic solution depends only on values of p (s − χ ) upstream of s over a dif-
model developed by Frascati and Lanzoni (2013) that, in the most fusion length scale. Indeed, the integral with respect to the dummy
general case, can manage also mild along channel variations of the variable χ decays as exp(−μm2 χ / γ ), and hence depends on the values of p
cross section width. The model is based on the two-dimensional, depth- closest to s.
averaged shallow water equations, written in the curvilinear co- The solution (57) is in general valid for an arbitrary, although
ordinates s, n and, owing to the large aspect ratio β usually observed in slowly varying, spatial distribution of the channel axis curvature. It
natural rivers, neglects the presence of the near bank boundary layers. takes a particularly simple form in the schematic case of a regular se-
The flow equations, ensuring the conservation of mass and momentum quence of meanders with the axis curvature described by the sine
and embedding a suitable parametrization of the secondary flow cir- generated curve C (s ) = e 2π ı s + c. c. (Leopold et al., 1964), where i is the
culations, are coupled with the two-dimensional sediment balance imaginary unit, and c.c. denotes complex conjugate. In this case the
equation, complemented with the relation describing the rate of sedi- flow field reads (Blondeaux and Seminara, 1985):
ment transport. The solution of the resulting set of partial differential
equations takes advantage of the fact that, in natural channels, the U1 (n) =[du0 n + du1 sinh (Λ1n) + du2 sinh (Λ2n)] e 2πis + c. c.
curvature ratio ν is small (ranging in the interval 0.1–0.2), and assume D1 (n) =[dd0 n + dd1 sinh (Λ1n) + dd2 sinh (Λ2n)] e 2πis + c. c. (59)
that flow and topography perturbations originating from deviations of
The constant coefficients duj , ddj (j = 0, 1, 2), Λ1, Λ2 (reported in the
the channel planform from the straight one are small enough to allow
Supplementary Information) depend on β, dgr, τ*u, and λ. The above
for linearization. In the case of a constant width rectangular section (for
relationships indicate that both the flow depth and the velocity tend to
which Dc* = Du*), the dimensionless flow field yields:
increase towards the outside channel bank. The deepening of the outer
(U , D) = (1, 1) + ν (U1, D1) + O (ν 2) (54) flow that takes place in a movable bed, in fact, pushes the thread of
high velocity towards the outside bank, unlike in the fixed bed case,
where where the predicted thread of high velocities is located along the inside
∞ of the bend.
U1 (s, n) = ∑ u cm sin(Mc n) The forcing term f11 = nC − U1 can be written as f11 = p (s ) q (n) + c. c. ,
m=0

with p (s ) = e 2πis and q (n) =n − du0 n − du1 sinh (Λ1n) − du2 sinh (Λ2n) . It
D1 (s, n) = (h1 C + h2 C ′ + h3 C ″) n + ∑ d cm sin(M n) follows that:
m=0 (55) ∞
1 bm
Here, C (s ) is the local curvature of the channel, C ′ (s ) and C ″ (s ) its
g11 (s, n) =
γ
∑ 2π ı + M 2/ γ
cos[M (n + 1)] e 2π ı s + c. c.
m=0 (60)
first and second derivatives, hi and di (i = 1, 3) are constant coefficients,
M = (2m + 1) π /2, and u cm (s ) , d cm (s ) are functions of the longitudinal where bm are constant coefficients (see Supplementary Information).
coordinate s: Substituting (60) into (33) and recalling (24) we finally obtain the re-
4 4 lationship giving the bend averaged O(ν2) correction to the longitudinal
s

⎣ cj0 ∫0
u cm = ∑ ccmj e λ cmj s + Acm ∑⎡g C (ξ ) e λ cmj (s − ξ ) dξ + gcj1 C⎤ dispersion coefficient associated with a regular sequence of meanders:
j=1 j=1
⎦ ∞
4 5 bm b͠ m
d j − 1u
m c d j − 1C K = ν2 ∑ M2
M 4 + (2πγ )2
dcm = ∑ dmj + Am ∑ dmj m=0 (61)
ds j − 1 ds j − 1
j=1 j=1 (56)
where a tilde denotes complex conjugate.
We refer the interested reader to Frascati and Lanzoni (2013) for
further details about the model, its derivation and implementation,
4. Comparison with tracer field data
while all the coefficients needed to compute U1 and D1 are reported in
the Supplementary Information. It is worthwhile to note that the re-
4.1. The considered dataset
levant dimensionless parameters (geometric, hydraulic and sedi-
mentological) needed as input data to the model are the width to depth
In order to test the validity of the proposed theory, we need in-
ratio β, the dimensionless grain size dgr, and the Shields parameter for
formation not only on the dispersion coefficient and the average hy-
the uniform flow conditions, τ*u.
drodynamic properties of the considered river reach, but also on the
The expressions (54) are used to compute the forcing term f11,
planform shape of the channel, on the geometry of the cross sections
needed to solve the boundary value problem (34) for g11. Note that by
and, possibly, on the cross sectional velocity distribution. Despite the
substituting (54) into (38) yields g10 = 0 (owing to the neglecting of
numerous tracer experiments carried out on river dispersion (McQuivey
bank effects). The particular solution of (34) is obtained by writing the
and Keefer, 1974; Nordin and Sabol, 1974; Seo and Cheong, 1998;
forcing term as f11 = p (s ) q (n) (i.e., separating the variables through
Yotsukura et al., 1970), only a few report also this type of information.
Fourier series), and by introducing the appropriate Green function
In particular, the data collected by Godfrey and Frederick (1970) in-
(Morse and Feshbach, 1953). We obtain:
clude the time distribution of the local tracer concentration C at a

1 s − s0 2 number of monitoring sections and the cross-section distributions of the
g11 (s, n) =
γ
∑ (−1)m + 1 cos[μ 2m + 1 (n + 1)] ∫0 pm (s − χ ) e−μ2m + 1 χ / γdχ
flow depth, D*(n*), and of the vertical profiles of the longitudinal ve-
m=0
locity u*(n*, z*). This dataset therefore provides all the information
(57)
needed to assess the robustness of the present modeling framework. In
where μm = mπ/2, each test a radiotracer (gold-198) was injected in a line source across
2 the stream. About 15 ml of the tracer, a highly concentrated solution of
pm (s ) = (−1)mC (s ) − u cm (s ), gold chloride in nitric and hydrochloric acid, was diluted to a volume of
Mc2 (58)
2 l. The injection was made at a uniform rate over a 1-min period. The
and s0 denotes the position of the injection section. By assumption, the concentration of radionuclide used in each test was proportional to the
length scale over which the contaminant cloud has evolved, Lc* , is well discharge (about 2.6 GBq m −3 s −1). The concentrations near to the

63
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

stream centerline were observed by a scintillation detector. The resol- trapezoidal cross section was shaped by the flow.
ving time for the entire system was found to be 50 s. The error due to For a given test i, we used the data collected by Godfrey and
the resolving time was about 5–10%. Frederick (1970) to compute at each monitored cross section j the area
Among the five river reaches considered by Godfrey and Aij* and the total wetted perimeter Pij*. We then used the cross sectional
Frederick (1970), three exhibit almost straight planforms: the Copper velocity data uij* (n*, z *) to compute the depth averaged velocity Uij* (n*)
Creek below gage (near Gate City, Va), the Clinch River above gage and the flow discharge Qij*. All the relevant quantities deduced from the
(hereafter Clinch River a.g., near Clinchport, Va), and the Clinch River experimental dataset are collected in Table 2 reported in the Supple-
below gage (hereafter Clinch River b.g., near Speers Ferry, Va). The mentary Information. In particular, the values of the mean slope are
other two river reaches, the Powell River near Sedville (Tenn) and the those provided directly by Godfrey and Frederick (1970), while the
Copper Creek above gage (near Gate City, Va) have meandering plan- friction velocities have been estimated under the hypothesis of a locally
forms. This data set has been integrated with the estimates of K* ob- uniform flow field.
tained from tracer tests carried out in other five straight streams and The dispersion coefficients by Godfrey and Frederick (1970) have
eight meandering rivers, namely the Queich, Sulzbach and Kaltenbach been obtained by applying the method of moments. However, the
rivers (Noss and Lorke, 2016), the Ohio, Muskegon, St. Clair and Red presence of a relatively long tail in the temporal distribution of the
Cedar rivers (Shen et al., 2010), the Green-Duwamish River concentration (Fig. 4a) and the sensitivity of small concentrations to
(Fischer, 1968), the Missouri River (Yotsukura et al., 1970), the Lesser measurement errors limit the accuracy of this method
Slave River (Beltaos and Day, 1978), and the Miljacka River (Rutherford, 1994). For this reason, we have recalculated the dispersion
(Dobran, 1982). Fig. 3 shows the planform configurations of the in- coefficients by considering the Chatwin’s method (Chatwin, 1980),
vestigated reaches, extracted from topographic maps, while the geo- which has also the advantage to give an indication whether a mon-
metrical, hydraulic and sedimentologic parameters of each stream are itoring section is located or not whitin the equilibrium region, where a
reported in Table 2. In particular, the curvature ratio ν and the wave- Fickian dispersion model can be applied. Fig. 4b shows an example of
number λ have been determined from the spatial distribution of the application of the method to the tracer data collected in the Clinch
channel axis curvature through the automatic extraction procedure Creeek (test T10). The Chatwin method introduces the transformed
described by Marani et al. (2002). The mean grain size estimates have variable C (t *), defined as:
been obtained on the basis of information available from literature
(Beltaos and Day, 1978; Godfrey and Frederick, 1970; Yotsukura et al., *
Cmax tmax
 =± t *ln
C
1970), from the USGS National Water Information System (http:// C t* (62)
www.waterdata.usgs.gov/nwis ), or from direct inspection (Dobran
2007, personal communication). In addition to real meandering stream where C = C (s j*, t *) is the temporal distribution of the cross-sectionally
data, Table 2 reports the laboratory data characterizing the longitudinal averaged concentration measured the at jth cross section, Cmax is the
dispersion experiment carried out by Boxall and Guymer (2007) in a corresponding peak concentration, tmax * is the peaking time, and
flume with a sine generated meandering planform and a sand bed that the + and - signs apply for t * ≤ tmax * and t * > tmax
* , respectively. In the
transformed plane C , t *, a temporal distribution of tracer concentration
was artificially fixed by chemical hardening after the initially uniform
following a Gaussian behavior should plot as a straight line. The slope

Fig. 3. a) Plan view of the river reaches investigated by Godfrey and Frederick (1970) and location of the monitored cross sections. b) Planforms of further meandering streams (Beltaos
and Day, 1978; Dobran, 1982; Fischer, 1968; Noss and Lorke, 2016; Shen et al., 2010; Yotsukura et al., 1970) considered for testing the present theoretical approach.

64
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

Table 2
Tracer tests considered to assess the present theoretical framework. Definitions are as follows: B*, half cross-section width; Q*, flow discharge; S, longitudinal channel slope; δ, relative
variation rate of the cross section in the transverse direction= Du*/(P0* + Bc*) ; ν, curvature ratio = B */ R 0*, with R 0* twice the minimum radius of curvature of the channel axis within a
meandering reach; λ, dimensionless meander wavenumber, = 2πB */ L*, with L* the intrinsic meander length. All the quantities are averaged along the investigated river reach.

River B* Q* S u*fu Planform δ ν λ


(m) (m3/s) (%) (m/s)

Clinch River a.g.a 17.3 6.8 0.03 0.045 straight 0.032 0 –


Clinch River b.g.b 30 9.1,85,51 0.04 0.05,0.085,0.076 straight 0.04,0.07,0.07 0 –
Copper Creek a.g.c 8.5 1.5,8.5 0.13 0.08,0.104 straight 0.06,0.09 0 –
Copper Creek b.g.d 8.5 0.9 0.30 0.104 meandering 0.044 0.11 0.04
Powell Rivere 17.2 4 0.03 0.052 meandering 0.047 0.15 0.036
Green-Duwamishf 20.0 12 0.02 0.049 meandering 0.07 0.13 0.090
Lesser Slaveg 25.4 71 0.01 0.055 meandering 0.17 0.2 0.063
Missourih 90 950 0.01 0.055 meandering 0.06 0.05 0.04
Miljackai 5.7 1 0.11 0.055 meandering 0.08 0.09 0.05
Exp. Flumej 0.5 0.025 0.12 0.031 meandering 0.19 0.08 0.157
Queich 1k 1.52 0.21 0.12 0.048 meandering 0.25 0.13 0.32
Queich 2k 0.95 0.25 0.19 0.068 straight 0.42 0. –
Sulzbach 1k 1.315 0.16 0.32 0.079 meandering 0.25 0.11 0.21
Sulzbach 2k 0.72 0.16 0.26 0.025 straight 0.6 0. –
Kaltenbachk 1.01 0.15 0.52 0.103 straight 0.4 0. –
Muskegonl 35 48.41 0.6 0.24 straight 0.06 0. –
Ohiol 235 1405 0.007 0.061 meandering 0.04 0.05 0.1
St Clairl 276.6 5000 0.088 0.083 straight 0.06 0 –
Red Cedarl 6.33/12.34 2.7/19.8 0.2 0.11/0.14 meandering 0.19/0.15 0.01 0.02/0.04

a
Godfrey and Frederick (1970), test T5;
b
Godfrey and Frederick (1970), tests T2, T7, T10;
c
Godfrey and Frederick (1970), tests T1, T6;
d
Godfrey and Frederick (1970), test T3;
e
Godfrey and Frederick (1970), test T4;
f
Fischer (1968);
g
Beltaos and Day (1978);
h
Yotsukura et al. (1970);
i
Dobran (1982);
j
Boxall and Guymer (2007);
k
Noss and Lorke (2016);
l
Shen et al. (2010);

− 0.5 (Q*/ Aj* ) / K j* and the intercept 0.5s j*/ K j* of this line allow one to distribution. Conversely, a departure from the linear trend is evident in
estimate the dispersion coefficient K j* and the cross sectionally aver- the correspondence of the tails, indicating a deviation from the Fickian
aged velocity Q*/ Aj*. behavior.
The values of K j* estimated by considering the linear part of C (t *)
The data suggest that, for all the sections, only the rising limb and
for all the data collected by Godfrey and Frederick (1970) turn out
the near peak region of the concentration time distribution are ap-
invariably smaller than those calculated according to the method of
proximately linear, and hence can be described by a Gaussian
moments (see Table 3 of the Supplementary Information). In order to

12
a) measure
10 Hayami
8
C

6
4
2
0
0 2000 4000 6000 8000 10000 12000 14000 16000
*
t [sec]

200 18
c)
150 b)
K* [m s ]measured

16 Chatwin
100 14 calculated
C [sec ]

Hayami
0.5

50 12
0 10
-50 8
2 -1
^

-100 6
-150 4
-200 2
0 3000 6000 9000 12000 0
1 2 3 4 5 6
t *[sec] section
Fig. 4. a) Temporal distributions of the concentration measured by Godfrey and Frederick (1970) in the tracer test T10 carried out along the Clinch River b.g.: black circles indicate the
measured concentration; continuous lines denote the concentration profiles predicted by applying the Hayami’s calibration method. b) The Chatwin’s transformation is applied to the data
shown in a): black circles indicate the measured data; continuous straight lines are regression lines fitted to concentrations near the peak. c) Comparison between the dispersion
coefficients estimated by means of the present theoretical approach and Chatwin and Hayami methods.

65
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

assess the sensitivity of these estimates to the method used to derive other hand, the average velocity estimated with the two methods are
them, we applied also the routing method based on the Hayami solution very similar (Fig. 5b). In the following, we will consider the estimates of
(Rutherford, 1994). This method, provided that the dynamics of the the dispersion coefficients provided by the Chatwin method when re-
cross-sectionally averaged concentration is Gaussian, takes advantage ferring to the measured values of K*.
of the superimposition of effects to determine the temporal distribution
of C in a section, given the local values of U*, K* and the concentration- 4.2. Comparison with straight river dispersion data
time curve in an upstream section. The resulting solution has the ad-
vantage that it can be used to route downstream a given temporal Before pursuing a comparison between observed and predicted
distribution of concentration without invoking the frozen cloud ap- dispersion coefficients, it is worthwhile to test the reliability of the flow
proximation. In fact, for moderately large values of s* and t* it gives a field model described in Section 3.1. Fig. 6 shows the cross sectional
concentration profile similar to that provided by the classical Taylor distribution of the flow depth (left panels) and depth averaged velocity
solution. For all the monitored sections, except the first one, it is thus (right panels) measured by Godfrey and Frederick (1970) in six loca-
possible to estimate the values of U* and K* which ensure the best tions along the Clinch River b.g. (test T10). The theoretically predicted
agreement between the measured and predicted concentration profiles. velocities, shown in Fig. 6, have been obtained either by introducing
Fig. 5 shows the results of the application of the Chatwin and into equation (51) the observed flow depth, or by considering the
Hayami methods. In some cases the Hayami method tends to yield simplified cross sectional geometry described by Eq. (52) and selecting
larger values of K*. Possible reasons of this behaviour are the poor the value of βf which better interpolates the measured depth profile.
fitting of the routed solutions and the significance of the tail, due to the The agreement between measured and computed velocity distributions
entrapment and retarded release of the tracer into dead zones, ab- is in general reasonably good (correlation coefficient, RU2 = 0.81).
sorption on sediment surfaces, hyporheic fluxes. Nevertheless, the most The comparison between the estimates of K* obtained from the
significant differences between the two approaches generally occur in tracer data of Godfrey and Frederick (1970) and those predicted by
sections where the estimate deviates significantly (longer error bars in inserting in Eq. (31) the flow field described by Eq. (51) are shown in
Fig. 5a) from the average value in the considered river reach, i.e. where Fig. 7a. This figure also shows in white squares a comparison between
the dynamic of the tracer cloud is likely influenced by some localised the predicted dispersion coefficients and those estimated from mea-
effects, such as irregularities along the channel sides, or the channel surements by Noss and Lorke (2016) and Shen et al. (2010) for the
bed, determining the retention of a certain amount of tracer. On the Queich, Sulzbach, Kaltenbach and Muskegon Rivers.
The theoretical estimates are reasonably good, with about 70% of
predictions ensuring an error smaller than ± 30% (dotted lines in
Fig. 7a). Overall, the model tends to underestimate the dispersion
coefficient for the larger values of K*, which, usually occur in the most
distant sections from the injection, where the measured concentration
profiles are particularly flat.
In any case, the present estimates of K* are definitely more accurate
than those provided by other predictors available in literature
(Fig. 7c), as documented by the values of the discrepancy ratio,
* / Kmeas
dr = log (K pred * ), plotted in Fig. 7b) and d). To give a visual per-
ception of the goodness of the different models, Fig. 7 shows the data
relative to the models displaying the best (smaller mean discrepancy
ratio) and worst (larger mean discrepancy ratio) performance according
to Table 1 (a Figure reporting the predictions of all models is provided
in the Supplementary Information). Note that the coefficient κ0 of the
formula proposed by Deng et al. (2001) and reported in Table 1, has
been here reduced by 1/15. Indeed, this coefficient was originally de-
termined by introducing a multiplicative empirical constant ψ (= 15,
according to Deng et al. (2001)) in order to achieve a better agreement
with the observed dispersion coefficients. These coefficients, at least in
the specific case of the data provided by Godfrey and Frederick (1970),
were calculated through the method of moments that, as discussed
above, tends to overestimate K* with respect to the Chatwin or the
routing methods. Nevertheless, even by reducing the value of κ0, the
predictions of K* obtained from the formula by Deng et al. (2001) are
significantly less accurate (mean discrepancy ratio < dr > =0.63) than
those resulting from the present theoretical approach (< dr > =0.19).
Even worse results are attained when considering other predictors (see
Table 1).
It is important to stress that the present methodology, being physics
based, does not need the introduction of any fitting parameter. The
input data are simply the flow discharge, the free surface channel
width, the longitudinal slope, the friction velocity (strictly associated
with the sediment grain size and the type of bed configuration, i.e.,
Fig. 5. a) The dispersion coefficients predicted by the Hayami method are plotted versus
the values provided by the Chatwin method. The error bar measures the scatter of the
plane or dune covered), and the cross-sectional distribution of the flow
local value of the dispersion coefficient provided by the Haymi method with respect to the depth or, alternatively, its simplified analytical description (Eq. (52)).
average value of each river reach. b) The average velocity values of each cross section Finally, we observe that including the higher order effects that the
predicted by the Hayami and Chatwin methods are plotted versus the measured value. presence of the channel banks exert on the transverse gradient of U0
The continuous line denotes the perfect agreement; the dashed lines correspond to a (associated to the O(δ2) contribution in Eq. (51)) always leads to im-
± 50% error.
prove the estimate of K* (< dr > =0.190, instead of 0.188).

66
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

the topography variations induced by alternating bends (possible de-


partures being related to the presence of bedforms not accounted for in
the model). The overall comparison appears reasonably good also in
terms of depth integrated longitudinal velocities (RU2 = 0.93) and the
theoretically predicted profiles yield a better performance with respect
to those calculated according to the approximate method proposed by
Smith (1983) (RU2 =0.91).
Fig. 9a shows the comparison between the bend averaged values of
the longitudinal dispersion coefficient estimated from the measures
carried out by Godfrey and Frederick (1970) in the Copper Creek and in
the Powell River and those predicted by either the leading term K0
(Eq. (31)) entailing a straight channel, or by considering also the cor-
rection ν2 K2 (Eq. (33)), accounting for the presence of river bends. This
correction turns out to pick up the right order of magnitude and, on the
whole, ensures a degree of accuracy greater than that attained when
neglecting curvature effects (< dr > =0.32 instead of < dr > =0.76). As
expected, when treating the river as straight, the predicted values of K 0*
are systematically lower than those observed in the field. In is worth-
while to note that the points corresponding to cross sections T3-S1, T3-
S2 and T4-S1, for which the theory tends in any case to overestimate
K*, are quite close to the injection section and, therefore, likely fall
outside the zone where a Fickian dispersion model holds.
The ability of the present theoretical framework to give robust es-
timates of K* is confirmed by Fig. 9b, reporting the predicted values of
K* against those resulting from the tracer test data for all the considered
meandering streams. On the whole, the effect of the curvature is to
slightly improve the degree of accuracy (< dr > =0.22, instead of 0.29),
although sometimes the theoretical coefficients turn out to be lower
than those observed in the field. This can be partly explained with the
fact that the predicted O(ν2) correction does not account for the near
bank velocity gradients associated with the presence of a boundary
layer and has been obtained on the basis of a linearized treatment of the
flow field, which tends to underestimate the intensity of both secondary
circulations and transverse bed deformations forced by the meandering
stream. Clearly, a number of other processes act in the field to make
Fig. 6. Cross sectional distributions of the flow depth D* and of the depth averaged ve- dispersion not entirely Fickian, contributing to the data scatter. We
locity U0* across six sections of the Clinch River b.g.. Black circles correspond to the data
return later on this issue. Finally, note that for the considered set of
measured by Godfrey and Frederick (1970) in test T10. Continuous lines represent the
smoothed cross section described by (52) and the corresponding velocity profiles
rivers, the results remain basically unaltered ( < dr > =0.225, instead
(RD2 = 0.95 ; RU2 = 0.84 ). Dotted lines represent the velocity profile predicted by sub- of 0.22) when, instead of considering the observed spatially varying
stituting into Eq. (51) the actual flow depth distributions (RU2 = 0.81). curvature signal, we consider a sequence of regular meanders with
maximum curvature equal to the inverse of the mean minimum radius
of curvature within the river reach.
4.3. Comparison with meandering river dispersion data

* , additional
In the case of meandering streams, besides Q*, B*, S, ufu 5. Discussion
input information to the present model is the spatial distribution of
channel axis curvature. These data are used to determine the di- The rational perturbative framework developed in the previous
mensionless parameters β, τ*u, dgr, ν, as well as the along channel dis- sections, based on a suitable scaling of the two-dimensional advection-
tribution of the channel axis curvature C (s) needed to compute the diffusion equation and on the introduction of a reference system, tra-
flow field through Eqs. (55). The expressions of U1(s, n) and D1(s, n) are veling downstream with the contaminant cloud and accounting for the
then employed to compute f11 and to solve the problem (34) for g11, along channel variability of the cross-sectionally averaged velocity,
and, ultimately, to obtain the O(ν2) correction (33) to the longitudinal provides a clear picture of the processes affecting the spreading of a
dispersion coefficient. contaminant in alluvial rivers.
In order to test the reliability of the flow field model described in The velocity gradients that characterize the near bank regions of
Section 3.2, a comparison with the flow depths (left panels) and depth natural streams, where the flow depth progressively vanishes, influence
averaged velocities (right panels) measured by Boxall and the longitudinal dispersion at the leading order of approximation
Guymer (2007) at the apex and the cross-over sections of an experi- (equation (31)), corresponding to a straight channel planform. Sec-
mental meandering channel is reported in Fig. 8. The theoretically ondary circulations driven by centrifugal and topographical effects ty-
predicted velocities have also been compared with the velocity profiles pical of meandering channels provide a second order correction
calculated according to Smith (1983) for the theoretical flow depth (equation (33)). The presence of a secondary helical flow enhances
distributions (continuos lines on the left panels): transverse velocity gradients which, in turn, tend to increase the
longitudinal dispersion coefficient. On the contrary, the increased
D*0.5Uu*Du* transverse mixing promoted by secondary currents (e.g., Boxall et al.,
U* =
Hu* (63) 2003) would lead to a reduction of longitudinal dispersion. This be-
havior is summarized in the analytical relation (61), obtained by con-
where Hu* is the cross-sectional average of D*1.5. The cross sectional sidering a regular sequence of sine generated bends.
shapes predicted by the present model reproduce correctly (RD2 =0.90) Bend effects are explicitly accounted for through the dependence on

67
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

Fig. 7. On the left: Comparison between the dispersion coefficients predicted theoretically and those estimated from tracer test data (Godfrey and Frederick, 1970) in almost straight
channels by applying the Chatwin method. a) Present theoretical approach. c) Deng et al. (2001)’s (ψ=1) and Seo and Cheong (1998)’s predictors. White squares in a) are the dispersion
coefficients evaluated according to the present theoretical approach versus the values estimated from measurements by Noss and Lorke (2016) and Shen et al. (2010) for the Queich,
Sulzbach, Kaltenbach and Muskegon Rivers. The continuous line denotes the perfect agreement; the dotted lines correspond to a ± 30 % error. On the right, b) and d): values of the
discrepancy ratio associated with the data plotted in figures a) and c). The continuous line denotes the perfect agreement; the dashed lines correspond to a ± 50 % error.

ν2 while the characteristics of the flow field and the bottom topography function of the aspect ratio β for either plane (Fig. 10a) or dune-covered
affect the coefficients bm. Moreover, as observed by Fischer (1969) and bed (Fig. 10b). In both cases, for given values of the dimensionless
Smith (1983), the ratio γ of cross-sectional mixing timescale to long- parameters ν, dgr and γ, the bend averaged longitudinal dispersion
itudinal advection timescale, accounts for the frequency of alternating coefficient increases with the Shields parameter, τ*u. On the other hand,
bends along the meandering reach. In the case of long enough bends, i.e for a given τ*u, the values of K corresponding to quite different di-
such that γ is much smaller than 1, the term (2πγ)2 at the denominator mensionless grain sizes dgr exhibit a relatively narrow range of varia-
of (61) can be neglected with respect to M4. On the contrary, if γ in- tions, as shown in Fig. 10c. Finally, Fig. 10d demonstrates that K tends
creases, K tends to decrease. This is the case of short bends, for which to increase significantly when approaching the resonant conditions
the changes in the flow field associated with alternating curves are too (see, e.g., Lanzoni and Seminara, 2006). Nevertheless, it must be re-
fast to allow cross-sectional mixing to eliminate concentration gra- called that the meandering flow field in a neighborhood of the resonant
dients. state cannot be described by the linear model adopted here, but it
Fig. 10 shows two typical examples of the variations of K as a would require a weakly nonlinear approach.

Fig. 8. Cross sectional distributions of the flow depth D* and of the depth averaged velocity U* across two sections of the experimental meandering channel of Boxall and Guymer (2007).
Black circles correspond to the data measured by Boxall and Guymer (2007). Continuous lines represent the theoretical cross section (D0 + νD1), described by (52) and (59), and the
corresponding velocity profiles. Dotted lines represent the velocity profile predicted according to Smith (1983) for the theoretical flow depth distributions (continuous lines on the left
panels).

68
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

0.15
0.13
0.11
0.09

0.07

τ∗

0.6
0.4
0.3
0.2

Fig. 9. Comparison between the dispersion coefficients predicted by equations (31), (33)
and those estimated from the tracer test carried out in meandering rivers: a) Copper River
b.g. and Powell River; b) Copper b.g., Powell, Green-Duwamish, Lesser Slave, Missouri, β
Miljacka, Queich, Sulzbach, Ohio and Red Cedar rivers, and in the experimental flume of
Fig. 10. The theoretical values of bend averaged longitudinal dispersion coefficient K
Boxall and Guymer (2007). The sources of data are reported in Table 2. The continuous
predicted by (61) are plotted versus the aspect ratio β for ν = 0.1 and kn0 = 0.225 . a)
line denotes the perfect agreement; the dashed lines correspond to a ± 50 % error.
λ = 0.1, dgr = 0.01, plane bed;. b) λ = 0.1, dgr = 0.001, dune covered bed; c) τ u = 0.09 ;
*
λ = 0.1, plane bed; d) τ
*u = 0.09, ds = 0.01, plane bed.
In general, the linearized treatment of the flow field set as the basis
of the present theoretical framework holds for relatively wide bends
treatment of one-dimensional solutions from an instantaneous point
(small ν), long enough meanders to ensure slow longitudinal variations
source (Hunt, 2005).
of the flow field (small λ), small intensity of the centrifugally driven
secondary flow, a condition met for small values of ν /(β cfu ), and small
amplitude of bed perturbations with respect to the straight configura-
6. Conclusions
tion (small ν τ*u/ cfu and λβ τ*u ) (Bolla Pittaluga and Seminara, 2011;
Frascati and Lanzoni, 2013). These intrinsic limitation of the theory can
We set a physics-based theoretical framework to estimate the
partly explain the deviations of the predicted dispersion coefficients
longitudinal dispersion coefficient on the basis of the hydro-morpho-
from the values estimated from tracer test data. Other physical pro-
dynamic modeling of the flow field and the bed topography that es-
cesses however concur to the scatter of data. The bed configuration
tablish in alluvial rivers. The rational perturbative framework has been
predicted by the considered hydro-morphodynamic model stems from
developed on the basis of a suitable scaling of the two-dimensional
the imposed flow discharge, corresponding to that actually observed
advection-diffusion equation, and by the introduction of a reference
during the tracer tests. Nevertheless, this discharge can differ from the
system moving with the contaminant cloud with a velocity that varies
formative discharge producing the actual river bed configuration. The
according to the river cross sectional geometry. This framework pro-
presence of regulation works and human activities (e.g., sediment
vides a clear picture of the processes affecting the spreading of a con-
mining, dredging) can modify the bed topography and, consequently,
taminant in natural streams, that can be summarized as follows.
the structure of the flow field controlling shear flow dispersion. Finally,
The longitudinal dispersion dynamics in alluvial rivers is controlled
the presence of bedforms, width variations, islands, and dead zones
by velocity shear at the banks and secondary circulations driven by
concurs to a non-perfectly Fickian behavior, enhancing the rate of
centrifugal and topographical effects. In particular, the helical flow
dispersion and causing the long tails usually observed in concentration-
associated to these circulations enhances relatively small and rapidly
time curves. The Gaussian solution resulting from a Fickian approach to
changing velocity and concentration gradients, both in the transverse
dispersion can then be used to describe only the upper portion of the
and in the longitudinal directions, which in general lead to an increase
concentration-time curves, as indicated by the tracer data plotted using
of the longitudinal dispersion coefficient. Nevertheless, the planform
Chatwin’s transformation. Other approaches are needed to fit entirely
shapes of meandering channels are usually characterized by relatively
these curves, such as the transient storage model that accounts for the
small values of the curvature ratio ν, implying that the increased
effects of temporary entrapment and subsequent re-entrainment of
transverse mixing, also promoted by secondary flows, affects the con-
pollutants (Cheong and Seo, 2003), the adoption of a fractional ad-
centration distribution only at higher orders of approximation.
vection-dispersion equation (Deng et al., 2004), or the asymptotic
Another consequence of the small values typically attained by ν is

69
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

the possibility to separate the contribution to shear flow dispersion averaged longitudinal dispersion coefficient. The residual scatter can be
provided by near bank velocity gradients of the unperturbed straight partly explained by the linearized character of the hydro-morphody-
configuration (equation (31)) from that induced by streamline curva- namic model used to compute K*. Flow nonlinearities, enhancing both
tures and by the alternating sequence of bars and pools which estab- transverse mixing and shear flow dispersion, induce opposite effects on
lishes in the perturbed meandering configuration (equation (33)). The longitudinal dispersion. Other possible causes of the departures be-
former contribution can be accounted for analytically for gently sloping tween predicted and estimated coefficients are associated with the not
channel banks. entirely Fickian behavior of the dispersion process, whereby the con-
The longitudinal dispersion coefficient, averaged over the meander centration-time curves decay more slowly than if they were Gaussian.
length in order to deal with longitudinal variations of the flow field,
depends on the relevant bulk hydrodynamic and morphologic di-
mensionless parameters, β, dgr, τ*u, λ, ν and γ. The latter parameter, Acknowledgments
accounting for the ability of cross-sectional mixing to adapt to along-
channel flow changes, could lead to a reduction of the longitudinal This work has been partially supported by Cariparo within the
dispersion coefficient in the presence of a sequence of relatively short project “GIS-based integrated platform for Debris Flow Monitoring,
bends (Eq. (61)). Modeling and Hazard Mitigation” and by the University of Padua
The comparison with field data obtained from tracer tests indicates within the project (Progetto di Ateneo 2010) “Morphodynamics of
that the proposed approach provides robust estimates of the reach marsh systems subject to natural forcings and climatic changes”.

Appendix A. Cross-sectional distribution of a uniform turbulent flow in a straight channel

Let us consider the longitudinal momentum equation averaged over the turbulence, written terms of the local curvilinear orthogonal coordinate
system (s*, σ*, ζ*) (Lanzoni and D’Alpaos, 2015):

∂u* ∂u* v * ∂u* ∂u* ∂H * 1 ⎡ ∂ (hσ Tss* ) ∂Tσs* ∂ (hσ Tζs*) ⎤ v 2 + Tσσ / ρ ∂hσ
+ u* + + w* = −g + + + −
∂t * ∂s * hσ ∂σ * ∂ζ * ∂s * ⎢
ρ hσ ⎣ ∂s * ∂σ * ∂ζ * ⎦ ⎥ hσ ∂s * (A.1)
where hσ is the metric coefficient associated with the curvilinear transverse coordinate σ, u*, v*, w* are the components of the velocities along the
three coordinate axes, H* is the elevation of the water surface with respect to an horizontal reference plane, g is the gravitational constant, ρ is the
* are components of the turbulent Reynolds stress tensor. In the case of uniform flow conditions, as those occurring in
water density and Tss*, Tσs* , Tζs*, Tσσ
a straight channel with a compact cross section (Fig. 1c), the relevant variables do not vary in time and along the main flow direction s*.
Expressing the components Tσs and Tζs of the Reynolds stress tensor through the Boussinesq eddy-viscosity approximation:
νT* ∂u* ∂u*
Tσs = ρ Tζs = ρ νT*
hσ ∂σ * ∂ζ * (A.2)
equation (A.1) simplifies to:

∂ ⎛ νT* ∂u* ⎞ ∂ ⎛ ∂u* ⎞


S hσ g + ⎜ ⎟ + ⎜ν *h σ ⎟ = 0,
T
∂σ * ⎝ hσ ∂σ * ⎠ ∂ζ * ⎝ ∂ζ * ⎠ (A.3)
where S = −∂H */ ∂s * is the longitudinal water surface slope that, for uniform flow conditions, coincides with the bed slope and the energy slope, and
νT* is the turbulent eddy viscosity.
Under the assumption of a constant vertical distribution of νT*, the longitudinal slip-velocity at the bottom must satisfy the condition:

⎡ Dz* ⎤
u* = uf* ⎢2 + 2.5 ln ⎜⎛ ⎟⎞ ⎥
d *
ζ *= 0 ⎣ ⎝ gr ⎠ ⎦ (A.4)
where uf* = (gD0*S )1/2 is the local value of the friction velocity. In addition, the shear stress must vanish at the water surface and take the value ρuf*2 at
the bed, namely:

⎡ν * ⎛⎜ ∂u* − 1 ∂Dz* ∂u* ⎟⎞ ⎤ = 0, ⎡νT*


∂u* ⎤
= uf*2 .
⎢ T ∂ζ * hσ2 ∂p* ∂p* ⎠ ⎥ ⎢
⎣ ⎝ ⎦ζ *= Dz* ⎣ ∂ζ * ⎥ ⎦ζ *= 0 (A.5)
In terms of the dimensionless variables (42) and (43), the problem described by equations (A.3)–(A.5) leads to equations (46), (48) and (49).
The solution of this boundary value problem is obtained by expanding u(ζ, σ) and uf(σ) in terms of the mall parameter δ:
(u, uf ) = (u 0 , uf 0) + δ 2 (u1, u f1) + O (δ 4 ). (A.6)
Substituting this expansion into equations (46), (48), (49), and collecting the terms with the same power of δ , we obtain a sequence of ordinary
2

differential problems that can be readily solved in closed form. After some algebra we find:

• O(δ ) 0

13 ζ 2
5 D
u 0 (ζ , σ ) = ⎜⎛− + 13 ζ + 2 + ln 0 ⎟⎞ D0
⎝ 2 2 dgr ⎠ (A.7)

u f0 (σ ) = D0 (A.8)

70
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

• O(δ ) 2

2
⎧ ⎡ ⎛ 45 D 7 25 ⎛ D0 ⎞ ⎞ 1 ⎛ 7 5 D ⎞ ⎛7 5 D ⎞
u1 (ζ , σ ) = D 0 ⎢ ⎜ ln ⎜⎛ 0 ⎟⎞ + + ln ⎜ ⎟ + ⎜− − ln ⎛⎜ 0 ⎟⎞ ζ 2 + ⎜ + ln ⎜⎛ 0 ⎟⎞ ⎟ ζ
⎨ ⎢ 8 ⎝ dgr ⎠ 2 16 ⎝ dgr ⎠ ⎟ 13 8 16 ⎝ dgr ⎠ ⎟ 4 8 ⎝ dgr ⎠
⎩⎣⎝ ⎠ ⎝ ⎠ ⎝ ⎠
2
5 ⎛ D0 ⎞ 1 ζ ζ4 ζ3 3 ζ 2 ⎞ ⎤ ∂2D0 ⎡ ⎛ 205 ⎛ D0 ⎞ 33 25 ⎛ D0 ⎞ ⎞ 1
+ ln ⎜ ⎟ + + 13 ⎛ − ⎜ + − ⎥ D0 + ⎢⎜ ⎟ ln ⎜ ⎟ + + ln ⎜ ⎟
8 ⎝ dgr ⎠ 2 ⎝ 4 16 4 8 ⎠⎥ ∂σ 2
⎢ 16 d
⎝ ⎠gr 4 8 ⎝ dgr ⎠ ⎟ 13
⎦ ⎣⎝ ⎠
5ζ2 ⎛ 33 5 D ⎞ 7 5 D ζ2 ζ4 ζ ⎤ ∂D 2⎫
− +⎜ + ln ⎜⎛ 0 ⎟⎞ ⎟ ζ + + ln ⎜⎛ 0 ⎟⎞ + 13 ⎛− − + ⎞⎥ ⎛ 0 ⎞
⎜ ⎟

16 8 4 ⎝ dgr ⎠ 4 8 ⎝ dgr ⎠ ⎝ 8 16 4 ⎠ ⎝ ∂σ ⎠ ⎬
⎝ ⎠ ⎦ ⎭ (A.9)

2
D0 ⎡ ⎛ 5 D ∂2D0 ⎛ 59 + 5 ln D0 ⎞ ⎛ ∂D0 ⎞ ⎤
u f1 (σ ) = ⎢ ⎜5 + ln 0 ⎟⎞ D0 + ⎜ ⎟
13 ⎝ 8 dgr ⎠ ∂σ 2 ⎝8 4 dgr ⎠ ⎝ ∂σ ⎠ ⎥ (A.10)
⎣ ⎦

It is worthwhile to note that, at the leading order of approximation, the friction velocity is proportional to the square root of the local flow depth
(equation (A.8)), as it occurs under uniform flow conditions, while the first order correction (equation (A.11)) quantifies the effects due to the cross
slope and curvature of the cross-section bed profile.
The local value of the depth-averaged longitudinal velocity U0 (σ ) = U00 (σ ) + δ 2U01 (σ ) is determined by integrating u0 and u1 along the normal ζ.
It results:
* ⎛ 19
ufu D
U00 = ⎜ + 2.5ln 0 ⎟⎞ D 0
Uu* ⎝ 3 dgr ⎠ (A.11)

*
ufu ⎧ ⎡ 781 395 ⎛ D0 ⎞ 25 2 ⎛ D0 ⎞ ⎤ ⎡ 17437 465 ⎛ D0 ⎞ 25 2 ⎛ D0 ⎞ ⎤ 2 ⎫
U01 = D0 ⎢ + ln ⎜ ⎟ + ln ⎜ ⎟ D0 D0 ,σσ + ⎢ + ln ⎜ ⎟ + ln ⎜ ⎟ D0 ,σ
Uu* ⎨ 390 312 ⎝ dgr ⎠ 208 ⎝ dgr ⎠ ⎥ 3120 208 ⎝ dgr ⎠ 104 ⎝ dgr ⎠ ⎥ ⎬
⎩⎣ ⎦ ⎣ ⎦ ⎭
*
ufu D 0 ⎧⎡ 5 5 D ⎤ 1 σ ⎫
+ + ln ⎛⎜ 0 ⎟⎞ D0 ,σ ⎥ + ⎡ ∫0 D0 ,2σ dσ − D0 ,σ ⎤
Uu* 2 ⎨⎢ 2 4 ⎝ dgr ⎠ ⎢
⎣ D0 ⎥
⎦⎬
⎩⎣ ⎦ ⎭
Finally, we convert the coordinates σ to the corresponding Cartesian coordinates n by observing that:

1 σ ∂D 2 1 ⎡ δ2 σ 2
⎛ ∂D0 ⎞ dσ ′ + O (δ 4 ) ⎤
n (σ ) =
βδ
∫0 1 − ⎛δ 0 ⎞ dσ ′ =
⎝ ∂σ ′ ⎠ ⎢
βδ ⎣
σ−
2
∫0 ⎝ ∂σ ′ ⎠ ⎥
⎦ (A.12)

Supplementary materials

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.advwatres.2017.10.033.

References Chatwin, P.C., 1980. Presentation of longitudinal dispersion data. J. Hydraul. Proc. ASCE
106, 71–83.
Chatwin, P.C., Allen, M., 1985. Mathematical models of dispersion in rivers and estuaries.
Alizadeh, M.J., Shahheydari, H., Kavianpour, M.R., Shamloo, H., Barati, R., 2017. Ann. Rev. Fluid Mech. 17, 119–149.
Prediction of longitudinal dispersion coefficient in natural rivers using a cluster-based Cheong, T.S., Seo, I.W., 2003. Parameter estimation of the transient storage model by a
bayesian network. Environ. Earth. Sci. 76 (86). http://dx.doi.org/10.1007/s12665- routing method for river mixing processes. Water Resour. Res. 39 (4), 1074. http://
016-6379-6. dx.doi.org/10.1029/2001WR000676.
Bear, T., Young, P.C., 1983. Longitudinal dispersion in natural streams. J. Environ. Eng. Czernuszenko, W., Rowinski, P.M., Sukhodolov, A., 1998. Experimental and numerical
109 (5), 1049–1067. validation of the dead-zone model for longitudinal dispersion in rivers. J. Hydraul.
Beltaos, S., Day, T.J., 1978. A field study of longitudinal dispersion. Can. J. Civ. Eng 5, Res. 36 (2), 269–280.
572–585. Deng, Z.Q., Bengtsson, L., Singh, V.P., Adrian, D.D., 2002. Longitudinal dispersion coef-
Bencala, K.E., Walters, R.A., 1983. Simulation of solute transport in a mountain pool-and- ficient in single-channel streams. J. Hydr. Eng., ASCE 128 (10), 901–916.
riffle stream: a transient storage model. Water Resour. Res. 19 (3), 718–724. Deng, Z.Q., Singh, V.P., Bengtsson, L., 2004. Numerical solution of fractional advection-
Blondeaux, P., Seminara, G., 1985. A unified bar-bend theory of river meanders. J. Fluid dispersion equation. J. Hydr. Eng., ASCE 130 (5), 422–431.
Mech. 157, 449–470. Deng, Z.Q., Slingh, V.P., Bengtsson, L., 2001. Longitudinal dispersion coefficient in
Bogoni, M., Putti, M., Lanzoni, S., 2017. Modeling meander morphodynamics over self- straight rivers. J. Hydr. Eng. 919–927. http://dx.doi.org/10.1061/(ASCE)0733-
formed heterogeneous floodplains. Water Resour. Res 53, 5137–5157. http://dx.doi. 9429(2001)127:11(919).
org/10.1002/2017WR020726. Disley, T., Gharabaghi, B., Mahboubi, A., McBean, E., 2015. Predictive equation for
Bolla Pittaluga, M., Seminara, G., 2011. Nonlinearity and unsteadiness in riv- longitudinal dispersion coefficient. Hydrol. Process 29, 161–172.
ermeandering:areview of progress in theory and modelling. Earth Surf. Processes Dobran, B.H., 1982. Dispersion in montainous natural streams. J. Environ. Eng. Div.,
Landforms 36 (1), 20–38. http://dx.doi.org/10.1002/esp.2089. ASCE 108 (EE3), 502–514.
Botter, G., Rinaldo, A., 2003. A scale effect on geomorphologic and kinematic dispersion. Eke, E.C., Parker, G., Shimizu, Y., 2014. Numerical modeling of erosional and deposi-
Water Resour. Res. 39 (10). tional bank processes in migrating river bends with self-formed width: morphody-
Boxall, J.B., Guymer, I., 2007. Longitudinal mixing in meandering channels: new ex- namics of barp ush and bank pull. J. Geophys. Res. Earth Surf. 119, 1455–1483.
perimental data set and verification of a predictive technique. Water Res. 41, Engelund, F., 1974. Flow and bed topography in channels bends. J. Hydraul. Div. ASCE
341–354. 100 (HY11), 1631–1648.
Boxall, J.B., Guymer, I., Marion, A., 2003. Transverse mixing in sinuous natural open Etamad-Shahidi, A., Taghipour, M., 2012. Predicting longititudinal dispersion coefficient
channel flows. J. Hydraul. Res. 41, 153–165. in natural streams using m5’ model tree. J. Hydraul. Eng. 138 (6), 542–554.

71
S. Lanzoni et al. Advances in Water Resources 113 (2018) 55–72

Fischer, H.B., 1967. The mechanics of dispersion in natural streams. J. of Hydraul. Div., Nayfeh, A.H., 1973. Perturbation Methods. John Wiley, New York, pp. 1–425.
ASCE 93 (HY6), 187–216. Noori, R., Ghiasi, B., Sheikhian, H., Adamowski, J.F., 2017. Estimation of the dispersion
Fischer, H.B., 1968. Methods for predicting dispersion coefficients in natural streams, coefficient in natural rivers using a granular computing model. J. Hydraul. Eng. 143
with application to lower reaches of the green and duwamish rivers, washington, u.s. (5). 04017001-1-12
Geol. Survey Prof. Pap. 27. 582–A Nordin, C.F., Sabol, G.V., 1974. Empirical data on longitudinal dispersion in rivers. U.S.
Fischer, H.B., 1969. The effects of bends on dispersion in streams. Water Resour. Res 5, Geol. Surv. Water Res. Invest. 20–74. Washington, D.C
496–506. Noss, C., Lorke, A., 2016. Roughness, resistance and dispersion: relationships in small
Fischer, H.B., 1973. Longitudinal dispersion and turbulent mixing in open channel flows. streams. Water Resour. Res. 52, 2802–2821. http://dx.doi.org/10.1002/
Annu. Rev. Fluid Mech. 5, 59–78. 2015WR017449.
Fischer, H.B., List, E.J., Koh, R.C.Y., Imberger, J., Brooks, N.H., 1979. Mixing in inland Parker, G., 2004. 1d sediment transport morphodynamics with applications to rivers and
and coastal waters. Academic Press, New York. turbidity currents. http://cee.uiuc.edu/people/parkerg/morphodynamicse-book.
Frascati, A., Lanzoni, S., 2009. Morphodynamic regime and long-term evolution of htm.
meandering rivers. J. Geophys. Res., Earth Surf. 114, F02002. http://dx.doi.org/10. Revelli, R., Ridolfi, L., 2002. Influence of suspended sediment on the transport processes
1029/2008JF001101. of nonlinear reactive substances in turbulent streams. J. Fluid Mech. 472, 7–331.
Frascati, A., Lanzoni, S., 2010. Long-term river meandering as a part of chaotic dynamics? Rinaldo, A., Marani, A., Rigon, R., 1991. Geomorphological dispersion. Water Resour.
A contribution from mathematical modelling. Earth Surf. Process. Landforms 35 (7), Res. 27 (4), 513–525.
791–802. Rutherford, J.C., 1994. River Mixing. John Wiley, New York.
Frascati, A., Lanzoni, S., 2013. A mathematical model for meandering rivers with varying Sahay, R., Dutta, S., 2009. Prediction of longitudinal dispersion coefficient in natural
width. J. Geophys. Res., Earth Surf. 118, 1641–1657. http://dx.doi.org/10.1002/jgrf. rivers using genetic algorithm. Hydr.Res. 40 (6), 544–552.
20084. Sattar, A.M., Gharabaghi, B., 2015. Gene-expression models for prediction of longitudinal
Godfrey, R.G., Frederick, B.J., 1970. Stream dispersion at selected sites. U.S. Geol. Surv. dispersion coefficient in streams. J Hydrol. 524, 587–596.
Prof.Paper. 433-K, Washington D.C. Schnoor, J.L., 1996. Environmental modeling. John Wiley, New York.
Guymer, I., 1998. Longitudinal dispersion in sinuous channel with changes in shape. J. Seminara, G., 2006. Meanders. J. Fluid Mech. 554, 271–297. http://dx.doi.org/10.1017/
Hydraul. Eng., ASCE 124 (1), 33–40. S0022112006008925.
Hey, R.D., Thorne, C.R., 1986. Stable channels with mobile gravel beds. J. Hydraul. Eng., Seo, I.W., Cheong, T.S., 1998. Predicting longitudinal dispersion coefficient in natural
ASCE 112 (8), 671–689. streams. J. Hydr. Eng. ASCE 124 (1), 25–32.
Howard, A.D., 1992. Modelling channel migration and floodplain sedimentation in Shen, C., Niu, J., Anderson, E.J., Phanikumar, M.S., 2010. Estimating longitudinal dis-
meandering streams. In: Carling, P., Petts, G.E. (Eds.), Lowland Floodplain Rivers: persion in rivers using acoustic doppler current profilers. Adv. Water Res. 33,
Geomorphological Perspectives. John Wiley & Sons Ltd, pp. 1–41. 615–623. http://dx.doi.org/10.1029/2011WR010547.1.
Hunt, B., 2005. Asymptotic solutions for one-dimensional dispersion in rivers. J. Hydraul. Shucksmith, J.D., Boxall, J.B., Guymer, I., 2011. Determining longitudinal dispersion
Eng. 132 (1). http://dx.doi.org/10.1061/(ASCE)0733-9429(2006)132:1(87). coefficients for submerged vegetated flow. Water Resour. Res. 47, W10516. http://
Iwasa, Y., Aya, S., 1991. Predicting longitudinal dispersion coefficient in open-channel dx.doi.org/10.1029/2011WR010547.1.
flows. Proc. Int. Symp. Environ. Hydr., Hong Kong. pp. 505–510. Smeithlov, B.B., 1990. Effect of channel sinuosity on river turbulent diffusion. Yangtze
Kashefipour, S.M., Falconer, R.A., 2002. Longitudinal dispersion coefficient in natural River 21 (11), 62.
channels. Water Res. 36, 1596–1608. Smith, R., 1983. Longitudinal dispersion coefficients for varying channels. J. Fluid Mech.
Lanzoni, S., D’Alpaos, A., 2015. On funneling of tidal channels. J. Geophys. Res. Earth 130, 299–314.
Surf. 120. http://dx.doi.org/10.1002/2014JF003203. Taylor, G.I., 1954. The dispersion of matter in turbulent flow through a pipe. Proc. Royal
Lanzoni, S., Seminara, G., 2006. On the nature of meander instability. J. Geophys. Res. Soc. London, Ser. A 223, 446–468.
111, F04006. http://dx.doi.org/10.1029/2005JF000416. Tubino, M., Colombini, M., 1992. Correnti uniformi a superficie libera e sezione lenta-
Lees, M.J., Camacho, L.A., Chapra, S., 2000. On the relationship of transient storage and mente variabile. XXIII Convegno di Idraulica e Costruzioni Idrauliche. D.375-D.386
aggregated dead zone models of longitudinal transport in streams. Water Resour. Res. (In Italian)
36 (1), 213–224. Wallis, S.G., 1994. Simulation of solute transport in open channel flow. Mixing and
Leopold, L.B., Wolman, M.G., Miller, J.P., 1964. Fluvial Processes in Geomorphology. transport in the environment, edited by K. J. Beven, P. Chatwin, and J. Millbank.
Freeman, New York. John Wiley, New York, pp. 89–112.
Li, X., Liu, H., Yin, M., 2013. Differential evolution for prediction of longitudinal dis- Wang, Y., Huai, W., 2016. Estimationg the longitudinal dispersion coefficient in straight
persion coefficients in natural streams. Water Resour. Manag. 27, 5245–5260. natural rivers. J. Hydraul. Eng. 142 (11). 04016048-1-11
Liu, H., 1977. Predicting dispersion coefficient of streams. J. Envir. Eng. Div., ASCE 103 Yotsukura, N., 1977. Derivation of solute-transport eqiations for a turbulent natural-
(1), 59–69. channel flow. J. Res. U.S. Geol. Survey 5 (3), 277–284.
Marani, M., Lanzoni, S., Zandolin, D., Seminara, G., Rinaldo, A., 2002. Tidal meanders. Yotsukura, N., Fischer, H.B., Sayre, W.W., 1970. Measurement of mixing characteristics of
Water Resour. Res., ASCE 38 (11), 1225. http://dx.doi.org/10.1029/ the missouri river between sioux city, iowa, and plattsmouth, nebraska. U.S. Geol.
2001WR000404. Surv. Prof. Paper. 1899-G, Washington, D.C.
McQuivey, R.S., Keefer, T.N., 1974. Simple method for predicting dispersion in rivers. J. Zeng, Y., Huai, W., 2014. Estimation of longitudinal dispersion coefficient in rivers. J.
Envir. Engrg. Div. 100 (4), 977–1011. Hydro. Environ. Res. 8, 2–8.
Morse, P.M., Feshbach, H., 1953. Methods of theoretical physics. Pergamon, New York.

72

You might also like