Sie sind auf Seite 1von 13

Lakes & Reservoirs: Research and Management 1999 4: 121–133

A model to predict reservoir sedimentation


Pablo A. Tarela* and Angel N. Menéndez
Laboratorio de Hidráulica y del Ambiente, Instituto Nacional del Agua y del Ambiente, CC 21 (1802) Aeropuerto de Ezeiza,
Argentina

Abstract
An efficient mathematical model to predict the sedimentation process in a reservoir is presented. It is based on a parabolized
and laterally integrated form of the governing equations. For its numerical solution the finite element method is used. The
model formulation and numerical scheme are both explained. The model is validated through comparisons with empirical
curves that quantify sedimentation in a reservoir. The velocity and sediment concentration profiles in typical situations are
shown. Solid discharge longitudinal evolution, as well as stratification conditions, are studied. The formation and growth of
bottom structures are explained. It is shown that the reservoir bottom evolution depends strongly on the geometry of the
reservoir and the sediment size. It is also shown that the system acts as a filter for the coarse and fine fractions of the solid
discharge.

Key words
fate of sediments, reser voir bottom evolution, reser voir sedimentation.

INTRODUCTION with hydrodynamic parameters (Churchill 1948; Brune


The reservoir resulting from the construction of a dam in a 1953; Brown 1958). The distribution of sediment
river is a site for the sedimentation of solid particles trans- deposits was also addressed (Heinemann 1961; Graf 1983).
ported by the river, due to the decrease in the flow trans- Schoklitsch (1937) carried out a pioneering laboratory study.
port capacity. On the one hand, this sedimentation process In many experiments pronounced delta formations were
has engineering consequences because it leads to a reduc- observed (Graf 1983). Nowadays, a great amount of field
tion of the storage capacity of the reservoir (Graf 1984) and, data exists in the technical literature, but a large amount also
hence, of its efficiency. Flushing techniques (Chang et al. exists in unpublished reports (Graf 1983). A typical case is
1996; Lai & Shen 1996) are presently being studied and used the Lake Mead survey through the Colorado River (Lara &
as a way to control this effect. Sanders 1970).
In contrast, as a by-product of the human activities Several approachs were undertaken regarding compu-
upstream of the dam, the fine fraction of the incoming tational modelling. The simplest models use sediment
suspended sediments may carry sorbed pollutants. Its transport formulas and a one-dimensional (1-D) backwater
deposition may lead, then, to disturbing environmental con- profile calculation (Graf 1983). Two-dimensional vertical
sequences. Controlling this effect is a more complicated models solve the sediment concentration profiles, allowing
matter than the previous one, because we are dealing not for more precision in the near-bed particle exchange flux
only with the quantity of deposited sediments but also with calculation. However, existing 2-D models do not address
the quality of these deposits. In other words, one has to pre- specifically the present problem (van Rijn 1987; Lai & Shen
dict the fate of the sorbed pollutants. To this end, relatively 1996). Fully 3-D models were developed recently in relation
precise modelling techniques must be used. to sedimentation in water intakes (Olsen 1991), or estuarine
The first attempts to predict sedimentation in reservoirs and coastal sedimentation (Lin & Falconer 1996). The main
led to empirical curves relating the reservoir capacity loss disadvantage with 3-D models is the still high compu-
tational cost, because they involve very different spatial
and time scales.
In the present paper a 2-D vertical model for reservoir
*Email: ptarela@ina.gov.ar sedimentation is developed and tested. Through a lateral
Accepted for publication 14 April 1999. integration of the equations of motion, some 3-D effects are
122 P. A. Tarela and A. N. Menéndez

also accounted for. The model includes hydrodynamic and (7) Departure from local equilibrium conditions is weak,
sedimentation modules. With the aim of obtaining an effi- so the bottom shear stress is related to the mean flow
cient calculation tool, a parabolic formulation is posed, which velocity through Chezy’s formula.
allows a marching calculation procedure. Starting with the Navier–Stokes equations, the foregoing
The model is validated by comparing its predictions with assumptions lead to the following dimensionless equations
known empirical sedimentation curves. It is later used to of motion (Tarela 1995).
analyze the evolution of the sediment concentration profiles
∂bu ∂bw
and the resulting bottom structures. ∂x 1 ∂z 5 0 (1)

HYDRODYNAMIC MATHEMATICAL MODEL ∂u ∂u ∂u ∂p


Problem schematization
St–1
∂t ∂x ∂z (
1 u 1w 5 Fr–2 a–1sinu –
∂x
+ ) (2)

∂ ∂u
Figure 1 shows a schematized geometry of the problem that a–1Re–1 v
∂x v ∂z ( )
is used later for test calculations. Water flows from a rec-
tangular prismatic access channel into a uniformly diverging
∂w ∂w ∂w ∂p
width zone, ending at the dam location. The bottom slope is St–1
∂t
1u 1w 5 2a–2Fr–2 cosu 1
∂x ∂z (∂z
+ ) (3)
initially uniform and the bottom roughness is homogeneous. ∂ ∂u ∂ ∂w
A simulation of different cases is undertaken through
a–1Re–1 v (
∂x H ∂z )
+ 2a–1Re–1 v
∂z v ∂z ( )
variation of the geometric parameters within definite
where t is time, x and z are longitudinal and vertical cartesian
practical ranges.
coordinates, respectively (the z axis is measured from the
Although the results presented in this paper are related
bottom, positive in the upward direction), u and w are the
to the schematized geometry shown in Fig. 1, the mathem-
(turbulent mean) laterally averaged horizontal and vertical
atical model is general enough to be used in particular prac-
velocities, respectively, p is the laterally averaged pressure,
tical problems. The scope and limitations of the model are
u is the inclination of the bottom line, b is the local width
discussed in the next section.
and nH and nV are horizontal and vertical eddy viscosities,

Model formulation
The basic hypotheses of the hydrodynamic model are the
following.
(1) The fluid is incompressible and the sediment con-
centration is low. Hence the fluid density can be considered
as a constant.
(2) The reservoir divergence is relatively weak, therefore
no flow separation at the lateral boundaries occurs and a
main direction of motion can be distinguished throughout.
(3) Reynolds stresses are modelled using the eddy vis-
cosity concept.
(4) The reservoir length is much larger than the water
depth. Thus the vertical velocity gradients are much higher
than the horizontal ones, and the diffusion in the longitudinal
direction may be neglected (boundary layer-type approxi-
mation).
(5) The free surface is associated with a hydrostatic
pressure distribution (consistent with the previous
hypothesis). Hence, it is calculated as a backwater curve
(Henderson 1971) and imposed as a rigid lid for the
computation of the spatial distribution of the velocity and
pressure profiles.
(6) Lateral dimensions are small in comparison to the
reservoir length. Thus, a lateral integration of the equations
of motion is performed. In addition, the flow section is quasi- Fig. 1. Problem schematization. (a) Vertical view; (b) plane view.
rectangular. Bo, channel width; Bd, reservoir width at the dam location.
Reservoir sedimentation model 123

respectively. Note that the x axis is considered to be locally of the pressure field is taken into account through the free
parallel to the bottom line; hence, the z axis has an inclin- surface computation, calculated previously and imposed as
ation u with respect to the gravity direction. a rigid lid.
Non-dimensionalization has been performed introducing A fictitious bottom, located above the actual one, is taken
the reference magnitudes shown in Table 1, where r is the as a mathematical boundary where the following conditions
fluid density, g is the acceleration of gravity, k is the von are imposed:
Karman constant and f is the friction coefficient (f 5 g1/2/C;
u u
C 5 Chezy coefficient). The resulting dimensionless para- u(z1)5 k* coswb ln(z1) , w(z1) 5 k* sinwb ln(z1) (7)
meters are the following: Strouhal number, St 5 UT/L;
that is, a logarithmic velocity profile and the impenetrability
aspect ratio, a 5 H/L; Froude number, Fr 5 U/(gH)1/2; tur-
condition, with z1 5 zb/zo ~ 30 (White 1974) and wb 5 ∂zb/∂x,
bulent Reynolds number, Re 5 UH/n0 5 1/kf.
where zb is the coordinate of the fictitious border and zo is
The vertical eddy viscosity is modelled according to
related with the friction factor through the relation
Kerssens’s criterion (van Rijn 1987), that is, a parabolic-
constant distribution of the form
ln ( 2zh )
0
z0
h
1
k
– – –
f
1
3
50 (8)
1
vv 5 hu* c z(z9) (4) in order to be consistent with Eqn 6. The only limitation of
4
Eqn 8 is that it assumes nearly vertical side walls (see
hypothesis 6). Note that zb ~ 30zo defines the effective

{
z9 z9 < h
(
1– 1–2
h )
2
if
2 roughness height. Hence, the fictitious bottom is compatible
z(z9) 5 (5) with the boundary used for the suspended sediment model
1 if z9 ≥ h (see the following).
2
The introduction of a fictitious bottom avoids the resolu-
where z9 5 z – z0; z0 represents the virtual height where the
tion of the problem within the near-wall region, where the
horizontal velocity is null, h and u* are the dimensionless
velocity gradient is very high, and would thus require a too
local depth (referred to the channel depth) and shear
fine discretization.
velocity (referred to the channel velocity), respectively, and
On the free surface z9 5 h the boundary conditions are:
c 5 h/R is a correction factor due to the stream finite width,
where R is the hydraulic radius. Note that, according to ∂zfs ∂u
u –w=0 , =0 , p=0 (9)
hypothesis 7, one has ∂x ∂nfs

u* = f < u > (6) with zfs 5 h 1 z0 being the vertical coordinate of the free
surface and nfs being the outward normal. These equations
where <u> is the mean flow velocity (vertical average of u). mean, respectively, that tangential shear stresses are absent
The horizontal eddy viscosity is taken as the laboratory (i.e. no winds are present), and that the free surface is a
value nH 5 0.23 hu* (Fisher 1967). streamline and an iso-pressure line.
Note that the absence of horizontal diffusion means that Finally, as boundary (initial) conditions at the upstream-
the equation system (1)–(3) is parabolic in the velocities. most section, the vertical distributions of (the laterally inte-
Hence, the information regarding the velocity field propa- grated) velocities and pressure must be imposed. In the
gates only in the downstream direction. The elliptic nature unsteady case the initial depth and velocity components
should also be given.
Table 1. Reference magnitudes For the present paper, the time integration is solved
through a quasi-steady scheme; that is, a succession of
Magnitude Symbol Order of steady states of the system. Hence, we can take ∂/∂t 5 0 in
Time scale T Bottom changes
(1)–(3) for each steady state.
Longitudinal length L Reservoir length
Numerical resolution
Vertical length H 5 H0 Channel depth (Fig. 1)
System (1)–(3) for the steady case, with boundar y
Transversal length B 5 B0 Channel width (Fig. 1)
conditions (7) and (9), is solved using the finite element
Longitudinal velocity U Channel velocity
method, which is particularly suitable for non-cylindrical
Vertical velocity UH/L From continuity
evolution domains. Previously, these equations are trans-
Pressure rgH Hydrostatic value at
formed into its weak form. Technical details related with
the channel bottom
the numerical method have been reported elsewhere (Tarela
Eddy viscosity n0 5 kHf U
& Menéndez 1992; Tarela 1995; Tarela & Menéndez 1998).
124 P. A. Tarela and A. N. Menéndez

In the following, only the main features of the algorithm are irregular in the z direction, with smaller steps where the
discussed. velocity gradients are higher (typically, close to the bottom;
The parabolic character of system (1)–(3) allows its reso- Fig. 2a).
lution by a marching procedure, leading to a quite efficient In order to obtain a one-step marching procedure, the
computational procedure. Galerkin weighing functions are truncated by half in such a
A quadrilateral finite element with six nodes is used way that they are non-zero only within each vertical column
(Fig. 2). The velocity components are interpolated linearly (Fig. 2b). Note that this leads to total upwinding for the
in the longitudinal direction (where diffusion terms are longitudinal advective terms, so no additional treatment is
absent) and quadratically along the vertical (using the two necessary in order to avoid related numerical instability
extra nodes). The pressure is represented linearly in both problems. In this way, the numerical scheme works as a
directions. The presented finite element was specially devel- streamline upwind/Petrov–Galerkin (SUPG) method
oped for this problem (Tarela & Menéndez 1992). (Brooks & Hughes 1982).
The finite element grid consists of vertical columns, Due to the different physical phenomena that are domin-
namely lines perpendicular to the marching direction. ant in the horizontal and vertical directions, the elements
The column width is variable, allowing for densification in aspect ratio must be quite small. In fact, in a time interval t
zones with significant free surface curvature. The mesh is the longitudinal convection length is lx ~ ut while the verti-
cal diffusion length is lz ~ (nVt)1/2. Eliminating the arbitrary
time t, the length ratio becomes

lx ulx 1/2
lx
} ( )nV
; Rex
1/2
(10)

where the Reynolds number Rex is in general very large for


the scales of interest. Hence, associating lx and lz with the
grid column width (horizontal step) and the vertical step,
respectively, once lx is fixed (based on the longitudinal
length scale) then lz can be estimated through (10). This
provides automatically a ‘densification’ criterion close to the
wall, where Rex increases due to the fast decrease of nV.
The non-linearity of the problem is treated through a fixed-
point iterative method (with a tolerance of 10–6 in the rela-
tive errors as a convergence criterion).
In a reservoir sedimentation problem the time step must
be only small enough to obtain a good resolution of the agra-
dation process.

Model validation
The validation of the hydrodynamic model was carried out
through a comparison with a parametric model (van Rijn
1987). van Rijn’s model is heuristic, but their parameters
were empirically adjusted based on experimental data.
Primarily, it solves an equation for the horizontal free surface
velocity. Figure 3 presents results from both models for the
case of a typical reservoir in a steady situation; the com-
parison is related to the free surface velocity. The agreement
between them is considered satisfactory.

Numerical experiments
Numerical experiments with the hydrodynamic model were
performed. Steady conditions were considered; then, equi-
librium conditions were imposed as initial conditions at the
Fig. 2. (a) Schematization of the calculation grid; (b) weighting upstream access channel section. The reservoir diverging
function for the velocity at an internal node. angle b was varied from 0° (2-D case) to 10° (for larger
Reservoir sedimentation model 125

angles flow separation is expected and, consequently, the heuristically to account for backwater effects. Hd was taken
model is not applicable). as 8H. The inclination u was used as a free parameter in the
The channel width and depth were given the following range 2.5 3 10–5 ≤ u ≤ 1.3 3 10–3. In this way, the dimension-
values: B 5 100 m and 1000 m; and H 5 5 m. less numbers were varied within the following ranges:
The reservoir length was related to the maximum water 2.5 3 10–5 ≤ a ≤ 1.3 3 10–3, 0.0014 ≤ Fr ≤ 0.6 and 118 ≤ Re ≤
depth, Hd, at the dam location, through the expression 275. The last condition is equivalent to 0.036 ≤ f ≤ 0.085.
L 5 1.2Hd/sinu, where the factor 1.2 was introduced Figure 4 presents a typical solution for the evolution of the
horizontal velocity profile for the 2-D case (u 5 0°), scaled
with the local equilibrium surface velocity (i.e. the one cor-
responding to the velocity profile of a uniform flow with the
same water depth and inclination). The initial equilibrium
profile evolves in such a way that its upper half lies below
the local equilibrium profile and vice versa for its lower half,
except very close to the bottom. It is interesting to remark
that the crossing of the velocity profiles just above the mid-
dle depth coincides with experimental observations (van Rijn
1987).
Figures 5 and 6 show the longitudinal distribution of the
surface velocity components, relative to the longitudinal
equilibrium value (and further scaled with the Reynolds
number in the case of the vertical component), for the 2-D
case and different hydrodinamic conditions. Note that, with
the chosen normalization, water flows from right to left.
It is observed that, starting with equilibrium conditions at
the channel, a relatively short transition region exists,
after which a quasi-linear behaviour is attained for both
Fig. 3. Comparison between (– – –), van Rijn and (–––), the components.
present model. Water flows from right to left. Lower indexes indicate The pressure distribution remains essentially hydrostatic.
channel (0) and dam (d) location, respectively. The geometric and
hydrodynamic parameters are the following: H0 5 5 m; Hd 5 40 m;
B0 5 100 m; b 5 10°; u 5 1.25 3 10–3; Fr 5 0.6; Re 5 120.

Fig. 5. Horizontal surface velocity evolution for different


Fig. 4. Horizontal velocity evolution. ueq(h), local equilibrium conditions. ueq(h), local equilibrium surface velocity. (b 5 0°;
surface velocity. (f 5 0.084; Fr 5 0.6; b 5 0°; B/H 5 20.) B/H 5 20.)
126 P. A. Tarela and A. N. Menéndez

The longitudinal velocity profile evolution for the diverg- 7 in that region). However, the backflow region disappears
ing walls case is illustrated in Figs 7 and 8 for b 5 1° and 7°, for the larger channel width, as shown in Fig. 9.
respectively. It is observed that, contrary to the 2-D case,
the upper part of the velocity profile now lies above the equil- SEDIMENT TRANSPORT MATHEMATICAL
ibrium profile. The crossing point of the profiles remains MODEL
located at ~ 60% of the depth from the bottom. Note that back- Model formulation
flow close to the bottom may appear near the dam for high The model is based on the following assumptions.
diverging angles (this would actually invalidate hypothesis (1) Only the suspended transport mode is considered;
that is, the bed load for coarse material is not taken into

Fig. 8. Horizontal velocity profile evolution for b 5 7° and


Fig. 6. Vertical surface velocity evolution for different conditions.
B/H 5 20. (f 5 0.084; Fr 5 0.6.)
(b 5 0°; B/H 5 20.)

Fig. 7. Horizontal velocity profile evolution for b 5 1°. (f 5 0.084; Fig. 9. Horizontal velocity profile evolution for b 5 7° and
Fr 5 0.6; B/H 5 20.) B/H 5 200. (f 5 0.084; Fr 5 0.6.)
Reservoir sedimentation model 127

account. Nevertheless, if the bed load transport is signifi- E 5 ((a–1 L) cos xb ws – cos wb w)(Seq–s) d > 62 mm (15)
cant then a formula to predict it can be easily added to the
where ∂ is the (dimensional) mean diameter. In the present
model.
paper seq is calculated using van Rijn’s formula (van Rijn
(2) The particle size distribution can be characterized
1987), arising from an empirical–stochastic approach that
entirely by the mean diameter; that is, it is described by a
assumes a normal distribution for the effective bed shear
unique parameter. Alternately, the fall velocity associated
stress:
with the mean diameter can be used.
(3) The suspended sediment concentration is low (below d <T 3/2>
seq 5 l (16)
2000 mg L–1). Hence, the sediment is passively transported zb D*3/10
in the particulate phase by the fluid; that is, it has no influ- In Eqn 16 l ~ 3 3 10–2; D* is the dimensionless particle para-
ence on the hydrodynamics. meter defined as
(4) In step with hypothesis 4 of the hydrodynamic model,
1/3
the diffusion in the longitudinal direction may be neglected.
The dimensionless transport equation for the statistically
D*5 d [ (rrn–r)g ]
s
2
(17)

averaged and laterally integrated concentration distribution where n is the molecular fluid viscosity and rS the sediment
s is (Tarela 1995) particle density; T is the state of transport parameter, which
specifies when resuspension takes place. T is a function of
∂s ∂s ∂s the bottom shear stress, which is considered as normally dis-
St–1 1 (u 1 (Lsinu)ws) 1 (w – (a–1Lcosu)ws) 5
∂t ∂x ∂z
tributed (see van Rijn 1987 for more details), usually with a
∂ ∂s
a–1Re–1 (
n
∂z V ∂z ) (11) constant standard deviation. In the present problem the stan-
dard deviation of this distribution was related to local val-
where L 5 Ws/U, and ws and Ws are the dimensionless and ues through the expresion s9 5 0.4ru*2.
reference fall velocities, repectively. Since the concentration In the case of silt, the physicochemical cohesive forces are
is low the fall velocity is calculated using Stokes’ formula for significant when particles are deposited. Taking into account
an isolated particle falling in a fluid at rest (Batchelor 1980). this effect, and the fact that this is a decelerating flow, a non-
If high organic content, flocculation or high concentration resuspension boundary condition was imposed; namely
are present, the fall velocity must be estimated accordingly
E50 4 mm ≤ d ≤ 62 mm (18)
(Teisson 1992; Ziegler & Nisbet 1995).
Note that, owing to negligible particle interaction, in Eqn The probability of deposition itself is expresed as
11 it is assumed that there is no difference between water (Partheniades 1990)
and sediment particles diffusion.
On the free surface no particle flux is allowed:

– (Re–1 a–1)nv
∂s
∂nfs
1 (cos wfs w – (a–1L) cos xfs ws) s 5 0 (12)
Pd 5
{ 0
t
1– td*
t* > td

t* < td
4 mm ≤ d ≤ 62 mm (19)

where t* is the local bottom shear stress and td its critical


where wfs > ∂zfs/∂x and xfs is the angle between the outward value for deposition. Eqn 19 means that when the current
normal and gravity directions. strength is high (t* > td) no particle can stick to the bottom.
At the fictitious bottom (located at z1 5 30) the following There are no precise results about the values of td, which
general relation between the resuspension rate and the con- depends on hydrodynamic, chemical and biochemical con-
centration is imposed: ditions and on sediment properties. Typical values are in the
range 0.06 ≤ td ≤ 1.1 N m–2 (Hjulstrom 1935; Ziegler & Nisbet
∂s
– (Re–1 a–1)nv 5 ((a–1L) cos xb ws – cos wb w) 1995). In the present paper a constant value td 5 0.07 N m–2
∂nb
(1–Pd)s 1 E (13) is employed (Menéndez et al. 1997).
Note that when t* > td Eqns (13), (18) and (19) show that
where E is the erosion rate and Pd is the ‘probability of depo-
the rate of resuspension balances with the rate of deposition;
sition’, defined as the proportion of near-bed sediment that
that is, no effective settling occurs.
reaches the bed and sticks to it (Partheniades 1990). For
If conditions for flocculation are not attained (i.e. salinity
coarse material (sand and gravel) there is no sticking;
and organic contents are low), clay can be treated as silt;
that is,
namely, the sediment transport model for silt can be
Pd 5 0 d > 62 mm (14)
considered as valid for the whole range of fine sediments
and the erosion rate can be taken as being proportional to (d ≤ 62 µm).
the bottom concentration decrement below its equilibrium When the agradation process begins, the bottom is modi-
value seq: fied according to
128 P. A. Tarela and A. N. Menéndez

dzb 1 dQS defined as the ratio between the period of retention and the
5– (20)
dt (1– p)b dx mean water velocity through the reservoir. The period of
where p is the porosity of the deposited material and Qs is retention is equal to the reservoir capacity divided by the
the solid discharge. Eqn 20 is the expression for the mass average daily inflow to the reservoir.
balance. The hydrosedimentological conditions of a reach of the
The model formulation is closed when the initial distribu- Bermejo River were taken for the purpose of model vali-
tion of sediment concentration at the upstream section is dation. Bermejo River runs eastward through northern
specified. For the simulation of unsteady cases, the initial Argentina, being a tributary of the Paraná River. Its friction
concentration profile and the profile at the upstreammost coefficient can be taken as f 5 0.11 and its mean hydro-
section for every time must be specified. dynamic stage can be characterized by Fr 5 0.14. A represen-
tative measured granulometric curve of the suspended load
Numerical resolution is shown in Fig. 10 (Toniolo 1995). Note that only fine sedi-
Eqn 11 is decoupled from system (1)–(3). Hence, it is solved ments are relevant. The mean vertical concentration is
separately once the velocity field is known. The mixed-type ~ 8000 mg L–1. Although this value is four times higher than
boundary conditions (12) and (13) are naturally incorporated the upper limit imposed by hypothesis 3, no modifications
to the weak form of Eqn 11. Due to its parabolic nature, the of the sediment diffusion rate or the fall velocity were
same marching procedure used to compute the hydrody- introduced.
namics fields is applicable to solve the suspended sediment Different reservoir geometries were considered by vary-
transport equation. ing the diverging angle and its extension (up to 135 km long,
The sediment concentration field is interpolated using the corresponding to a 25-m-high dam). As initial conditions, the
full six-node finite element. The element size and grid den- local equilibrium velocity and hydrostatic pressure profiles
sification defined for the hydrodynamic model allows corresponding to the mean stage (H 5 6.03 m, B 5 208 m)
enough resolution for the sediment transport equation. were used. The associated sediment concentration was taken
For the time evolution of the system, the bottom topog- as uniform.
raphy is updated using Eqn 20 and the hydrodynamic con- Separate calculations were made for a series of different
ditions are recalculated. In each time step the agradation sediment diameters within the range of interest. Figure 11
process is calculated up to a time such that the variation in shows Brune’s curves and the model results for b 5 4° and
the bottom level can be considered negligible relative to the several reservoir lengths. Although many calculated points
local water depth. Usually the maximum change allowed in lie within Brune’s band, significant deviations are observed,
the bottom level is the order of 0.01 h. especially for the coarser grains in the upper range of the
Typical computer runs involve ~ 500 and 200 nodes in the capacity–inflow relation.
horizontal and vertical directions, respectively, and ~ 500
time steps. Then, the total number of degrees of freedom to
compute is of the order of 2 3 108 (> 200 million). This shows
the significance of the parabolic approximation and the con-
sequent marching procedure technique used, to the calcu-
lations.

Model validation
The validation of the sedimentation model was made by com-
paring its predictions with the Brune and Churchill empir-
ical methods for reservoir sedimentation (Shen 1971),
arising directly from field measurements. They provided
graphs where the sedimentation can be estimated based on
geometric and hydraulic conditions.
The trap efficiency curve by Brune represents sediment
trapped in the reservoir as a function of the ratio between
the capacity of the reservoir and the inflow rate. In the case
of Churchill’s method, the percentage of incoming silt
passing through the reservoir is presented as a function of
the sedimentation index of the reservoir. This index is Fig. 10. Granulometric curve for the Bermejo River.
Reservoir sedimentation model 129

Now, the empirical curves actually represent situations MODEL PREDICTIONS


where the whole sediment distribution is considered. Hence The suspended sediment transport model was used to inves-
these calculations were taken as representative of subranges, tigate some details of the evolution of meaningful quantities
and the associated results were combined according to their along the reservoir, as shown in the following sections.
respective weight (inferred from Fig. 10) in order to obtain
the net sedimentation. In this way, Fig. 12 was obtained for Sediment concentration profiles and
three different diverging angles. Note that a relatively good ’stratification’
agreement is observed (the improvement was expected due Figure 15 illustrates a typical evolution of the sediment
to the relative lower weight of the coarser fraction). concentration profile for coarse material, starting upstream
The same model results were used to compare with with an equilibrium distribution. Geometric parameters
Churchill’s curves, as shown in Figs 13 and 14. From the last were: b 5 0°, B 5 1000 m, H 5 5 m, Hd 5 40 m and u 5 2.6 3
one it is observed that the calculated results fall close to the 10–3 (simulated reservoir length was calculated as shown,
‘fine sediment’ curve, especially for the lower range of sed- resulting in L 5 18.5 km). Note that, due to the fact that the
imentation indexes. particle source is located at the bottom, the concentration
The obtained results are considered good enough to is maximum there. It is observed that the growth of the water
assess the validity of the proposed model to describe the sed- depth and the loss to sedimentation produce a continuous
imentation process in reservoirs. In addition, in order to sat- decrease of the concentration values.
isfy a theoretical requirement, the main practical advantage From Fig. 15 it is observed that the major part of the
of the model over the empirical approach is the model’s abil- particles tends to concentrate in the neighbourhood of the
ity to identify trends within an otherwise dispersed cloud of bottom, generating a sort of ‘stratification’. To establish a
points. parameter that measures the stratification layer thickness,
the volumetric solid flux below height r is calculated as
r
fs (x,r) 5 ∫ u(x,z) s(x,z) b(x,z) dz (21)
zb

Fig. 11. Comparison between Brune’s curve and model results for Fig. 12. Comparison between Brune’s curve and model results for
different particle diameters. different reservoir geometries.
130 P. A. Tarela and A. N. Menéndez

Thus, the proportion f of suspended load in a layer of height iour of Qs for the fully 2-D case (b 5 0°) and a particular
df at location x is hydrodynamic condition. At the reservoir head a relatively
fast deposition of suspended coarse material occurs. The
fs (x,df)
f(x) 5 (22) characteristic decay length is controlled by the hydro-
fs (x,h)
dynamic and geometric conditions and, to a lesser degree,
Figures 16 and 17 show the evolution of d90% (f 5 0.9) for
coarse material and different hydrodynamic and geometric
conditions. As expected, the layer thickness decreases when
moving along the reservoir, although some local overshoot
may occur for the b > 0° case at the reservoir head. The
trends towards a more stratified flow must be interpreted
with care, as simultaneously a fast decrease of the total
suspended sediment volume is taking place. Note that the
stratification becomes more significant for larger grain diam-
eters and for lower Fr values. In Figs 16 and 17 the curves
corresponding to b 5 4° (dashed lines) stop before the solid
discharge gets so close to zero that the round-off errors
become dominant.

Solid discharge
The total volumetric solid flux Qs is given by Eqn 21 taking
r 5 h; that is, Qs (x) 5 Fs(x; h). Figure 18 presents the behav-

Fig. 14. Comparison between Churchill’s curves and model results


for different reservoir geometries.

Fig. 13. Comparison between Churchill’s curves and model results Fig. 15. Evolution of sediment concentration profiles for b 5 0°;
for different particle diameters. d 5 170 mm. (f 5 0.084; Fr 5 0.6; B/H 5 200.)
Reservoir sedimentation model 131

by the size of the particles. The fine sediment solid dis- Growth of bottom structures
charge, on the other hand, remain uniform until the con- The decay of the volumetric solid flux along the reservoir
dition t* 5 td is fulfilled, decreasing from that point on. The has, as a counterpart, a sedimentation process at the reser-
decay length looks much more sensitive to the particle voir. After some time, the change in bottom level becomes
size. relevant in relation with the local depth. Hence, a recalcu-
For b > 0°, an overshoot of Qs for coarse sediment may lation of the hydrodynamics and the sediment transport for
appear at the reservoir head, depending on hydrodynamic the new domain has to be undertaken, which leads to new
conditions, as shown in Fig. 19. This is related to the local sedimentation rates. In this way, the bottom evolution can
decrease of the water depth, which produces a velocity be predicted.
increase and, consequently, an effective resuspension of To illustrate this procedure, a case is presented corre-
bottom particles which add up to the suspended load. sponding to a man-made reservoir with the following char-
Eventually, the overshooting effect tends to disappear when acteristics: length 5 18 km, width 5 1 km and water height
the reservoir head bottom structure grows. 5 5 m at the upstreammost section; width 5 6 km and water
height 5 40 m at the dam section; and initial bottom slope
5 2.5 3 10–3. All cross-sections are prismatic, with vertical
lateral walls. The mesh has 100 elements in the vertical and
360 elements in the horizontal directions. The sediment size
distribution is represented through two particle diameters:
40 µm (silt) and 100 µm (sand).
Figure 20 presents a typical growth pattern. Sand particles
are deposited at the head of the reservoir, like all non-
cohesive sediments. This is due to the sudden expansion and
the corresponding decrease of the flow velocity. The rate of
decrease of the sediment load depends on the reservoir
divergence. The deposited sand forms a submerged delta
that grows downstream, raising the water level.
The fine particles, on the other hand, remain in suspen-
sion until the shear velocity drops below the critical depo-
sitional value. From that point on the sediment load
decreases monotonically. Owing to the silt deposition, a
Fig. 16 Density layers evolution for f 5 0.084, Fr 50.60 and three second bottom structure appears and progresses towards
particle diameters. B/H 5 200; (–––), b 5 0°; (– – –), b 5 4°. the dam. Close to the dam the silt deposit builds up and the

Fig. 17. Density layers evolution for f 5 0.060, Fr 5 0.17 and three Fig. 18. Volumetric solid flux evolution for b 5 0°. Qs,o 5 volumetric
particle diameters. B/H 5 200; (–––), b 5 0°; (– – –), b 5 4°. solid flux at the channel. (f 5 0.06; Fr 5 0.05; B/H 5200.)
132 P. A. Tarela and A. N. Menéndez

reservoir decreases its capacity. This selection between sand presents a flat surface with a slope equals to twice the
and silt deposits is qualitatively in agreement with obser- initial river slope.
vations (Shen 1971).
In Fig. 21 the evolution of the sand delta at different time CONCLUSIONS
steps is showed. The form, height and extension of the delta The numerical simulation of the sedimentation process in a
are changing continuously and its apex travels downstream. reservoir can be undertaken through a computationally effi-
Preliminary results show that the main characteristics of cient mathematical model based on a 2-D vertical (laterally
these deltas (foreset and topset slopes, location of the apex integrated) analysis and a parabolic approximation. This
relative to free surface along the reservoir) are in good allows not only the calculation of the sedimentation rate, but
agreement with observations (Shen 1971). In addition, it was also leads to the prediction of the time evolution of bottom
observed that the apex height grows faster in the earlier deposit structures, which is fundamental when analyzing the
times. The advancement speed decreases with time. fate of pollutants.
The growth of the silt structure is presented in Fig. 22. Model results show that the longitudinal rate of change
A regressive deposit is observed. For this particular case of the reservoir surface width (3-D effect) and the suspended
the apex appears on the dam. The last bottom structure sediment particle diameter control the sedimentation rate.

Fig. 21. Three stages in the head of reservoir (sand) delta


evolution. (a) 2 years; (b) 4 years; (c) 6 years.
Fig. 19. Volumetric solid flux evolution for b 5 4°. (f 5 0.06; Fr 5
0.05; B/H 5200.)

Fig. 20. Vertical view of bottom deposits for a 6-year simulation Fig. 22. Three stages in the near dam (silt) bottom structure
period. evolution. (a) 2 years; (b) 6 years; (c) 11 years.
Reservoir sedimentation model 133

The distinction between coarse and fine suspended sedi- Lai J. & Shen H. W. (1996) Flushing sediments through
ment behaviour, through the consideration of a critical depo- reservoirs. J. Hydr. Res.4,3237–55.
sitional bottom shear stress for the latter, manifests in a Lara J. M. & Sanders J. I. (1970) The 1963–64 Lake Mead
filter-like response. In fact, the bulk of the coarse sediment Survey, REC-OCE-70–21. US Bureau of Reclamation,
is deposited near the reservoir head while the fine sediment, Denver.
which is the one potentially contaminated, is transported Lin B. & Falconer R. A. (1996) Numerical modeling of three-
downstream within the reservoir, and accumulates at the dimensional suspended sediment for estuarine and
dam site. coastal waters. J. Hydr. Res.4.3
Menéndez A. N., Bombardelli F. A. & Tarela P. A. (1997)
REFERENCES Estudio de la Afectación del Régimen de Sedimentación y
Batchelor G. K. (1980) An Introduction to Fluid Dynamics. del Transporte de Contaminantes y Temperatura en el
Cambridge University Press, UK. Proyecto de Profundización del Puerto de Buenos Aires,
Brooks A. N. & Hughes T. R. J. (1982) Streamline Report LHA 01–162–97, National Water and Environment
Upwind/Petrov–Galerkin formulations for convective Institute (INA), Argentina.
dominated flows with particular emphasis on the incom- Olsen N. R. B. (1991) A three-dimensional numerical model
pressible Navier–Stokes equations. Comput. Methods for simulation of sediment movements in water intakes.
Appl. Mech. Eng. PhD thesis, University of Trondheim, Trondheim,
3 2, 199–259. Norway.
Brown C. B. (1958) Sediment Transportation, Engineering Partheniades E. (1990) Estuarine sediment dynamics and
Hydraulics. Wiley, New York, USA. shoaling processes. In: Herbich J. B. (ed.) Handbook of
Brune G. M. (1953) Trap efficiency of reservoirs. Am. Coastal and Ocean Engineering. Gulf Publishing Co.,
Geophys. Union Trans. 4,3 Houston, Texas.
51–64. Schoklitsch A. (1937) Handbuch des Wasserbaues. Springer,
Chang H. H., Harrison L. L., Lee W. & Tu S. (1996) Vienna, Austria. (English translation by S. Shulits, 1930.)
Numerical modeling for sediment-pass-through reser- Shen H. W. (ed.) (1971) River mechanics, vol. II. H. W. Shen,
voirs. J. Hydraul. Eng. 1 2 2, Colorado, USA.
381–8. Tarela P. A. (1995) Modelación Matemática del Fenómeno
Churchill M. A. (1948) Discussion of ‘Analysis and use of de Sedimentación en Embalses, Fellowship Report.
reservoir sedimentation data’, by L. C. Gottschalk. National Council of Scientific Research (CONICET),
Proceedings of the Federal Inter-Agency Sedimen- Argentina.
tation Conference, Denver, Colorado 1947, pp. Tarela P. A. & Menéndez A. N. (1992) Un modelo hidrod-
139–140. inámico para flujo estratificado a superficie libre. In:
Fisher H. B. (1967) The mechanics of dispersion in natural Venere M. (ed.). Mecánica Computacional, vol. 13.
streams. J. Hydraul. Div. ASCE Bariloche, Argentina.
3.9 Tarela P. A. & Menéndez A. N. (1998) Hydrosedimentologic
Graf H. W. (1983) The hydraulics of reservoir sedimentation. model to predict reservoir sedimentation. In: Oñate E. &
Water Power Dam Construct. Idelsohn S. R. (eds). Computational Mechanics: New
5,345–52. Trends and Applications. CIMNE, Barcelona, Spain.
Graf H. W. (1984) Storage losses in reservoirs. Water Power Teisson C. (1992) Cohesive suspended sediment transport:
Dam Construct.6,337–41. Feasibility and limitations of numerical modelling. J. Hydr.
Heinemann H. G. (1961) Sediment Distribution in Small Res. 2 9, 755–69.
Flood-Retarding Reservoirs. US Department of Agri- Toniolo H. (1995) Analisis sedimentológico del río Bermejo.
culture, ARS pp. 42–4. Report, Facultad de Ingeniería y Ciencias Hídricas,
Henderson F. M. (1971) Open Channel Flow. Macmillan Co., Universidad Nacional del Litoral, Santa Fe, Argentina.
New York, USA. van Rijn L. C. (1987) Mathematical model of morphological
Hjusltrom F. (1935) Studies of morphological activity of processes in the case of suspended sediment transport.
rivers as illustrated by the River Fyris. Bull. Geol. Inst. PhD thesis, Delft University of Technology, Delft, The
Upsala 2 5, 221–527. Netherlands.

Das könnte Ihnen auch gefallen