Water: Hydrodynamic Structure With Scour Hole Downstream of Bed Sills

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

water

Article
Hydrodynamic Structure with Scour Hole
Downstream of Bed Sills
Mouldi Ben Meftah * , Francesca De Serio , Diana De Padova and Michele Mossa
Department of Civil, Environmental, Land, Building Engineering and Chemistry, Polytechnic University of Bari,
Via E. Orabona 4, 70125 Bari, Italy; [email protected] (F.D.S.); [email protected] (D.D.P.);
[email protected] (M.M.)
* Correspondence: [email protected]; Tel.: +39-080-5963-508

Received: 2 December 2019; Accepted: 6 January 2020; Published: 9 January 2020 

Abstract: Experimental turbulence measurements of scour hole downstream of bed sills in alluvial
channels with non-cohesive sediments are investigated. Using an Acoustic Doppler Velocimeter
(ADV), the flow velocity-field within the equilibrium scour hole was comprehensively measured.
In this study, we especially focus on the flow hydrodynamic structure in the scour hole at equilibrium.
In addition to the flow velocity distribution in the equilibrium scour hole, the turbulence intensities,
the Reynolds shear stresses, the turbulent kinetic energy, and the turbulent length scales are analyzed.
Since the prediction of equilibrium scour features is always very uncertain, in this study and based
on laboratory turbulence measurements, we apply the phenomenological theory of turbulence to
predict the maximum equilibrium scour depth. With this approach, we obtain a new scaling of the
maximum scour depth at equilibrium, which is validated using experimental data, satisfying the
validity of a spectral exponent equal to −5/3. The proposed scaling shows a quite reasonable accuracy
in predicting the equilibrium scour depth in different hydraulic structures.

Keywords: scour; velocity field; turbulence; equilibrium scour depth; new scaling of scour depth

1. Introduction
Prediction of maximum scour depth downstream of hydraulic structure i.e., bridge piers and
abutments, sills, sluice gates, spillways, weirs, offshore platforms, wind turbines, etc., is of primary
concern for a wide range of engineering and environmental applications. This topic has drawn attention
and interest from many researchers for decades [1–14]. Despite these numerous studies, prediction
of equilibrium-scour hole characteristics always remains challenging because of the complexity of
the phenomenon and its dynamic sensitivity to structure and sediment properties. Most of these
studies [3–9,12–14] proposed different empirical formulae based on experimental/field measurements,
to predict the maximum eroded depth, its maximum length, and other properties. Ben Meftah and
Mossa [3], Tregnaghi et al. [14], and Lu et al. [15] observed that, based on laboratory measurements of
steady/unsteady flows, the scour downstream of a grade control structure evolves into three distinct
phases, including an initial phase, a developing phase, and an equilibrium phase. Tregnaghi et al. [14]
argued that the scour process usually reaches its equilibrium condition rapidly in live-bed conditions
and rather slowly in clear-water conditions. Lu et al. [15] indicated that the scour hole in non-cohesive
sediments is influenced by both the channel characteristics and the sediment properties, especially the
channel bed slope, the densimetric Froude number, the tailwater depth, and the sediment median size.
The enormous amount of studies conducted on this issue asserts that the scour hole profiles
are similar in shape, giving rise to a typical profile with appropriate scaling of the horizontal and
vertical coordinates. However, in spite of the great effort made by researchers, many different formulae
were derived to predict the scour profile at equilibrium. This large number of different empirical

Water 2020, 12, 186; doi:10.3390/w12010186 www.mdpi.com/journal/water


Water 2020, 12, 186 2 of 17

formulae, sometimes composed of complicated parameters, makes them less operational in practice
than expected. Moreover, most of these formulae are affected by large uncertainties and suffer some
limitations. According to Manes and Brocchini [16], the approaches based on dimensional analysis
suffer from two main shortcomings: One due to the experimental laboratory scale issues, hiding the
real shape of functional relations between non-dimensional groups at field scales, and the other one is
related to the fact that the empirical approach does not provide a theoretical framework to interpret
the experimental data and to understand the physics underlying such functional relations.
Recent studies [16–19] proposed very important and innovative approaches to predict localized
turbulent flow scours, applying the phenomenological theory of turbulence (PTT). This approach
hypothesizes that the scour process is controlled by the momentum transport generated by eddies
belonging to the dissipation and production spectrum ranges. By scaling the eddy-characteristic-lengths
of these spectrum-ranges with the equilibrium-scour dimensions and the characteristic-sediment-length,
researchers tried to derive general predictive formulae, by merging the PTT-theoretical aspects with
empirical observations, for the equilibrium maximum-scour depth at some hydraulic structure.
The main aim of this study is to contribute to this novel kind of approach, being the prediction
of the scour features based on experiments and theory still challenging due to the complexity of the
phenomenon. Therefore, in the present study, we first experimentally focus on the flow turbulence
measurements in a scour hole developed downstream of a grade control structure in sand-bed channels,
providing an integrated hydrodynamic picture of the scouring process. Successively, we propose a
new scaling of the maximum scour depth at equilibrium and validate it using the experimental data of
this study and some data collected from previous studies. Specifically, the proposed scaling approach
is easy to use, depending in particular on a densimetric Froude number and on a relative roughness.
Nevertheless, its application shows a quite reasonable accuracy in predicting the equilibrium scour
depth in different hydraulic structures. Therefore, our findings would contribute to improve the
understanding of the scouring mechanisms by applying the phenomenological theory of turbulence.

2. Experimental Set-Up
The experiments on the scour processes were carried out in a rectangular flume of closed-circuit
flow at the Hydraulic Laboratory of the Mediterranean Agronomic Institute of Bari (Italy). The flume
has glass sidewalls and a Plexiglass floor, allowing a good side view of the flow. It is 7.72 m long,
0.30 m wide, and 0.40 m deep. A pump of maximum discharge of 24 l/s was used to deliver water
from the laboratory sump to an upstream tank equipped with a baffle and lateral weir, maintaining a
constant head upstream of a movable slide-gate constructed at the inlet of the flume. The slide-gate
regulates channel flow-discharge. To create a smooth flow transition from the upstream reservoir to
the flume, a wooden ramp was installed at the inlet of the flume; it is 1.55 m long, 0.15 m thick and
of same channel width (Figure 1). At the outlet of the flume, water is intercepted by a stilling tank,
equipped with three vertical grids to stabilize water, and a triangular weir (V-notch sharp crested
weir) to measure discharge with relative uncertainty of ±8%. At the downstream end of the flume,
a movable gate made of Plexiglass and hinged at the channel bottom is used to regulate the flow depth.
In order to simulate grade control structures protecting riverbeds against erosion, in this study we
have used a series of sills consisting of PVC plates 0.30 m wide and 0.01 m thick. The sills were installed
on an experimental area extended 6 m along the channel, downstream of the wooden ramp. The sill
height decreases progressively going downstream from the wooden ramp, respecting a determined
initial slope S0 of 0.0086. Different configurations were investigated, the difference between them being
the distance, L, between sills. More details on the sills distribution are reported in Ben Meftah and
Mossa [3].
The flume bottom downstream of the wooden ramp and between the sills is covered with an
erodible bed material layer, consisting of almost uniform sand particles with a mean average size, d50 ,
of 1.8 mm and density of 2650 kg/m3 (see Ben Meftah and Mossa [3] for more information). Along the
experimental area, the sand layer was leveled respecting the maximum sill heights and that of the
upstream wooden ramp, forming the original bed of the channel with a slope S0 (Figure 1).
The data collected during each test included discharge, water surface elevation, flow depth,
temporally eroded bed profile, equilibrium bed profile, and scour dimensions (depth, length, position
of maximum depth). The profiles of the equilibrium eroded bed along the channel centerline, near
Water 2020, 12, 186 3 of 17
the channel sidewalls, and at an intermediate distance between the channel centerline and both
sidewalls (Figure 1) were measured, as the vertical distance between the initial bed elevation and the
bed at equilibrium,
experimental area,by
themeans
sand of a point
layer wasgauge
leveledofrespecting
±0.1 mm accuracy. The water
the maximum level profile
sill heights along
and that ofthe
the
channel
upstream centerline
woodenwas measured
ramp, formingusing an electrical
the original bed ofhydrometer
the channelwith
withan accuracy
a slope of ±0.11).
S0 (Figure mm.

Front view
Slide-gate Erodible material layer
Original bed Movable gate
z Expected scour
q x
hx
Wooden ramp

0.50 m 2m 3m 1m

First sill Top view


L.W.
y C.L.

0.30 m
x
.. q C
Channel axis C.R.
R.W.
Figure 1. General sketch of the laboratory flume with the initial condition and expected scour hole
Figure 1. General sketch of the laboratory flume with the initial condition and expected scour hole
(dashed profile) downstream of bed sills. hx = flow depth in the expected equilibrium scour at a
(dashed profile) downstream of bed sills. hx = flow depth in the expected equilibrium scour at a
downstream position x, L.W. = left wall side of channel, R.W. = right wall side, C = centerline (channel
downstream position x, L.W. = left wall side of channel, R.W. = right wall side, C = centerline (channel
axis), C.L. = centerline of left half of channel, and C.R. = centerline of right half, (x, y, z) = longitudinal,
axis), C.L. = centerline of left half of channel, and C.R. = centerline of right half, (x, y, z) = longitudinal,
transversal, and vertical directions, respectively.
transversal, and vertical directions, respectively.
The data collected during each test included discharge, water surface elevation, flow depth,
In
temporallyaddition to the
eroded bedmeasurements of thebed
profile, equilibrium scour geometric
profile, characteristics,
and scour dimensionsthe flow length,
(depth, velocity-fields
position
were also carried out in the scour hole at equilibrium condition. The
of maximum depth). The profiles of the equilibrium eroded bed along the channel centerline, velocity data were collected
near the
using a 3D Acoustic Doppler Velocimeter (ADV) system, developed by Nortek,
channel sidewalls, and at an intermediate distance between the channel centerline and both sidewalls with a sampling rate
of(Figure
25 Hz 1) at were
a time window of 70 s. The sampling volume of the ADV was located
measured, as the vertical distance between the initial bed elevation and the bed 5 cm below theat
transmitter probe. The ADV was used with a velocity range equal to ±0.30 m/s,
equilibrium, by means of a point gauge of ±0.1 mm accuracy. The water level profile along the channel a measured velocity
accuracy
centerline ofwas
±1%,measured
and a sampling
using anvolume
electricalof hydrometer
less than 0.25 cman
with
3. For high-resolution measurements,
accuracy of ±0.1 mm.
the manufacturer
In addition to recommends a 15 db signal-to-noise
the measurements ratio (SNR)
of the scour geometric and a correlation
characteristics, coefficient
the flow larger
velocity-fields
than
were70%.alsoThe acquired
carried out indata
thewere
scourfiltered
hole atbased on the Tukey’s
equilibrium condition.method and bad data
The velocity sampleswere(SNR < 15
collected
db and correlation coefficient < 70%) were also removed. Additional details
using a 3D Acoustic Doppler Velocimeter (ADV) system, developed by Nortek, with a sampling rate concerning the ADV-
system
of 25 Hz operations
at a timecan be found
window of 70in [20–27]. Flow velocity
s. The sampling volumemeasurements
of the ADVthrough the scour
was located 5 cmhole
below werethe
carried out for different configurations in both the longitudinal plane of
transmitter probe. The ADV was used with a velocity range equal to ±0.30 m/s, a measured velocity symmetry and in some
transversal
accuracy ofplanes.±1%, and a sampling volume of less than 0.25 cm3 . For high-resolution measurements,
The initial
the manufacturer experimental
recommends conditions and the geometric
a 15 db signal-to-noise ratio characteristics of scours,coefficient
(SNR) and a correlation related tolarger
this
study,
than 70%.are illustrated in Table
The acquired data1,were
where hc is the
filtered flowon
based depth over themethod
the Tukey’s crest of and
the sill
baddownstream
samples (SNR of
which the scour hole is measured, U c is the flow velocity over the sill (mean velocity in
< 15 db and correlation coefficient < 70%) were also removed. Additional details concerning the
correspondence
ADV-system operations of hc), zs is thebemaximum
can equilibrium
found in [20–27]. Flowscour depth
velocity from the original
measurements throughbedtheprofile,
scourhhole
s is

the flow depth at the position of maximum equilibrium scour depth, λ c = d50/hc is a relative roughness,
were carried out for different configurations in both the longitudinal plane of symmetry and in some
Ftransversal
dc = Uc/(Δgd50)0.5 is the densimetric Froude number for the approach flow over the sill, Δ = [(ρs − ρw)/
planes.
ρw] is the submerged relative density
The initial experimental of sediment
conditions and theparticles,
geometric ρwcharacteristics
is the water density, ρs is related
of scours, the sediment
to this
density, g is the gravitational acceleration, Re c = Uc hc/υ is the Reynolds number for the approach flow
study, are illustrated in Table 1, where h is the flow depth over the crest of the sill downstream of which
c
the scour hole is measured, Uc is the flow velocity over the sill (mean velocity in correspondence of hc ),
zs is the maximum equilibrium scour depth from the original bed profile, hs is the flow depth at the
position of maximum equilibrium scour depth, λc = d50 /hc is a relative roughness, Fdc = Uc /(∆gd50 )0.5 is
the densimetric Froude number for the approach flow over the sill, ∆ = [(ρs − ρw )/ ρw ] is the submerged
relative density of sediment particles, ρw is the water density, ρs is the sediment density, g is the
Water 2020, 12, 186 4 of 17

gravitational acceleration, Rec = Uc hc/υ is the Reynolds number for the approach flow over the sill,
Reg = Uc d50 /υ is the grain Reynolds numbers, and υ is the water kinematic viscosity. For the sake of
brevity, in this study we focus in detail on data of run T21 (Table 1) for the analysis of the turbulent
parameters, while we adopt all runs (T04–T22) in the scaling procedure.

Table 1. Initial experimental conditions and parameters of the investigated runs.

L hc Uc zs hs Fdc λ Rec Reg


Runs
(m) (m) (m/s) (m) (m) (-) (-) (-) (-)
T04 1 0.035 0.593 0.039 0.10 3.474 0.051 18,159 934
T05 1 0.048 0.689 0.088 0.15 4.035 0.038 28,517 1069
T06 1 0.029 0.522 0.028 0.08 3.060 0.062 12,497 776
T07 1 0.054 0.726 0.105 0.18 4.255 0.033 32,354 1078
T08 1 0.042 0.647 0.064 0.13 3.790 0.043 23,774 1019
T09 2 0.036 0.585 0.053 0.09 3.425 0.050 18,942 947
T10 2 0.042 0.634 0.057 0.12 3.717 0.043 25,981 1113
T11 2 0.048 0.688 0.070 0.13 4.033 0.038 32,217 1208
T12 2 0.054 0.736 0.081 0.15 4.312 0.033 39,743 1325
T13 2 0.060 0.762 0.090 0.17 4.466 0.030 46,880 1406
T14 4 0.034 0.576 0.076 0.09 3.374 0.053 20,563 1089
T15 4 0.042 0.647 0.090 0.12 3.790 0.043 29,208 1252
T16 4 0.048 0.691 0.112 0.14 4.048 0.038 37,309 1399
T17 4 0.030 0.533 0.065 0.08 3.125 0.060 18,000 1080
T18 3 0.029 0.531 0.050 0.07 3.113 0.062 17,721 1100
T19 3 0.036 0.586 0.061 0.10 3.433 0.050 24,784 1239
T20 3 0.042 0.653 0.071 0.12 3.827 0.043 32,236 1382
T21 3 0.048 0.686 0.084 0.14 4.019 0.038 38,688 1451
T22 3 0.052 0.762 0.094 0.17 4.463 0.035 46,543 1611

3. Results and Discussion

3.1. Velocity Fields


Since turbulence is the most important mechanism of sediment entrainment, causing a significant
increase in the shear stress around the base of a hydraulic structure, a large set of measurements of the
flow velocity field in the scour holes was carried out. Figure 2, as an example, shows a vector map of
the flow velocity, Vxz , in the scour hole downstream of the first bed sill, located 2 m downstream of the
wooden ramp (Figure 1). In Figure 2 the (x, z)-coordinates take origins at the first sill position and
the channel bottom, respectively. The Vxz -velocity is the resultant of the streamwise U and vertical W
time-averaged velocity components. The three profiles of the initial bed (solid line), the free-surface
flow (triangle down), and the bed at equilibrium condition (bullet) are also reported in Figure 2.
The random point cloud in Figure 2 represents the remaining amount of sediment between sills at
scour-equilibrium. All the data illustrated in Figure 2 were obtained in the plane of flow-symmetry
(y = 0).
Figure 2 clearly shows the flow velocity behavior through the scour hole. Three flow velocity
distribution regions can be recognized in Figure 2: (i) A first region, 1, where a sort of a free entering
jet flows, originated by the flow condition over the sill-crest; (ii) a second region, 2, characterized
by vortex formations (eddies) due to the jet diffusion, located near the bottom of the scour hole and
extended along the upstream scour-side; and (iii) a third region, 3, seeming less turbulent and taking
place downstream, outside the vortex region. Between the regions 1 and 2, a sort of a hydraulic jump
may occur, depending on the hydraulic conditions. The absence of velocity measurements in the upper
flow region is due to the limitation of the ADV-downlooking probe, being the uppermost 7 cm of the
flow could not be sampled. However, in the jet-like region 1, the acquired ADV-signal was very noisy,
which could be due to the strong jet-flow agitation interacting with the hydraulic jump.
velocity. They also observed that the lowest region of the scour hole is mainly covered by the
sediment of size d90, grain size for which 90% of sampled particles are finer.
In the region 3 (Figure 2), the flow velocity considerably increases, compared to region 2. This
increase is gradual in the downstream direction. Moreover, the velocity vectors tend to be more
horizontal and of almost comparable values. The flow redistribution in region 3 indicates a sort of
Water 2020, 12, 186 5 of 17
smooth transition flow from scour hole to the downstream tailwater flow. This smooth transitional
flow result in less flow turbulence, which is the subject of the next section.

Figure 2. Vector map of the flow velocity, V , in the scour hole at the plane of flow-symmetry (y = 0).
Figure 2. Vector map of the flow velocity, Vxz
xz, in the scour hole at the plane of flow-symmetry (y = 0).
The dashed line qualitatively indicates the separation between regions 1, 2, and 3.
The dashed line qualitatively indicates the separation between regions 1, 2, and 3.

The jet-like flow (region 1) plays a crucial role in the different phases of the scour development.
3.2. Turbulence Intensity Associated with Scour Hole
This is due to its high velocity, which leads to an increase of the jet potential erosive action on the bed
To get
channel. As thefurther
jet sizeinformation
increases overontime,
the the
scouring process,
jet begins in Figure
to gradually 3 erosive
lose its we plot the streamwise
potential. The state
turbulence
of equilibriumintensity,
occursU′, as athe
when function
path of of
thethe normalized
impinging vertical coordinate
jet becomes sufficiently Z/zlongs. Herein, U′ was
and its diffused
defined as the ratio of the standard deviation of the streamwise flow-velocity
velocity is reduced to values lower than the minimum value required for sediment movement [28]. component fluctuations
to theInaverage
the regionvelocity,
2 (FigureUc,2),
measured overreduction
a significant the sill-crest,
of theZflow
is the vertical
velocity position
occurs. from the original
Furthermore, the flow
bed profile shows
distribution (solid line in FigureA2)portion
two portions: at a given downstream
of negative velocity position
starting at x. the
Theposition
verticalof U′-profiles
maximum
correspond
scour depthtoand different
extendingdownstream
towards positions
the upward x/xsbed
= 0.33,
sill,0.67, 1.00, a1.33,
forming sort2.00, 2.33, andlocal
of clockwise 2.67, vortex,
where
xand
s is another
the x-position
portionfrom the grade-control
of positive velocity thatstructure (sill)
shifts the at which
flow the scour
downstream, attains
from its maximum
the position of the
maximum scour depth. Similar behaviors have been observed by Ghodsian et al. [29]. The authors
named as “weak flow” the portion of negative velocity and “strong flow” the portion of positive
velocity. They also observed that the lowest region of the scour hole is mainly covered by the sediment
of size d90 , grain size for which 90% of sampled particles are finer.
In the region 3 (Figure 2), the flow velocity considerably increases, compared to region 2.
This increase is gradual in the downstream direction. Moreover, the velocity vectors tend to be more
horizontal and of almost comparable values. The flow redistribution in region 3 indicates a sort of
smooth transition flow from scour hole to the downstream tailwater flow. This smooth transitional
flow result in less flow turbulence, which is the subject of the next section.

3.2. Turbulence Intensity Associated with Scour Hole


To get further information on the scouring process, in Figure 3 we plot the streamwise turbulence
intensity, U0 , as a function of the normalized vertical coordinate Z/zs . Herein, U0 was defined as the
ratio of the standard deviation of the streamwise flow-velocity component fluctuations to the average
velocity, Uc , measured over the sill-crest, Z is the vertical position from the original bed profile (solid
line in Figure 2) at a given downstream position x. The vertical U0 -profiles correspond to different
downstream positions x/xs = 0.33, 0.67, 1.00, 1.33, 2.00, 2.33, and 2.67, where xs is the x-position from
1, 2 and the upstream side of region 3. At these regions, which practically occupy the whole part of
the main scour hole, the values of turbulence intensities range between a minimum of 0.13 and a
maximum of 0.3. At x/xs = 0.67, U′ experiences the maximum measured values at Z/zs ≥ −0.7. This
position is located within the region 1, where the jet flow penetrates into the scour pool, generating
high2020,
Water levels
12,of
186turbulence. Figure 3 mainly indicates a tendency of U′ to reduce as going down towards 6 of 17
the scour bed. In region 2, at Z/zs < −0.7, U′ reduces by almost 50% as compared to region 1. The
significant reduction in turbulence intensity in region 2 is related to the sharp decrease in flow
the grade-control
velocity structure
in this region, (sill) at
as shown inwhich
Figurethe scour
2. At the attains its the
exit from maximum depth.
scour hole, Note
at x/x that, due to the
s > 1.33 in region 3,
flow symmetry, the spanwise velocity, V, is theoretically expected to be null and therefore
U′ shows the smallest values, as expected based on the flow velocity distribution (Section 3.1). it has not
At this
any physical
region, significance
U′ decreases in this
almost twiceplane.
compared to region 1 and by 60% compared to region 3.

U'
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
0.0
-0.1
x/xs
-0.2
0.33
-0.3 0.67
Z/zs

-0.4 1.00
-0.5 1.33
2.00
-0.6
2.33
-0.7
2.67
-0.8
-0.9
-1.0

Figure 3. Vertical profiles of streamwise turbulence intensity, U0 , at different downstream positions


Figure
x/x 3. Vertical profiles of streamwise turbulence intensity, U′, at different downstream positions
s along the plane of flow symmetry (y = 0).
x/xs along the plane of flow symmetry (y = 0).
Figure 3 shows that the maximum turbulence intensities take place at x/xs ≤ 1.33, at the regions 1,
2 andIn theFigure
upstream4 wesideplotof the
regionvertical turbulence
3. At these intensity,
regions, W′, profiles
which practically at the
occupy thesame
whole downstream
part of the
positions x/x = 0.33, 0.67, 1.00, 1.33, 2.00, 2.33, and 2.67. Herein, W′ is
main scour hole, the values of turbulence intensities range between a minimum of 0.13 and a maximum
s defined as the ratio of the
standard
of x/xs = 0.67,
0.3. At deviation 0
ofUtheexperiences
vertical flow-velocity
the maximum component
measured fluctuations
values at to Z/zUs c≥. Figure 4 clearly
−0.7. This shows
position is
a substantial
located withinreduction
the regionin1,W′ as compared
where the jet flow to penetrates
U′. For all into
the profiles,
the scourW′ ranges
pool, betweenhigh
generating a minimum
levels of
of 0.03 and Figure
turbulence. a maximum
3 mainly of 0.17 against
indicates a tendency U0 to
0.13 andof0.30, respectively,
reduce as going observed
downfor U′. Furthermore,
towards the scour bed.at
theregion
In different Z/zs < −0.7, Upositions
2, at downstream 0 reduces byx/xalmost
s, W′ decreases with decreasing
50% as compared to regionZ/z (i.e.,significant
1. sThe going down to the
reduction
equilibrium-scour
in turbulence intensity bed). in This
regiondecrease
2 is relatedis continuous
to the sharpanddecrease with ina flow
significant
velocityreduction rate as
in this region, as
shown in Figure 2. At the exit from the scour hole, at x/xs > 1.33 in region 3, U shows the smallest
compared to U′. This may be explained by the considerable reduction of the 0
vertical velocity flow
going towards
values, as expected the based
equilibrium bed profile,
on the flow velocitywhich effectively
distribution reduces
(Section 3.1). Atthethis
vertical
region, U0 decreases
flow-force that
could lift the sediments from the bottom. Similar
almost twice compared to region 1 and by 60% compared to region 3. to U′, W′ experiences maximum values in region 1
and at
In the upstream
Figure sidethe
4 we plot ofvertical
region 3. In region intensity,
turbulence 2 however, 0
W W′ exhibits
, profiles at the
the smallest values, comparable
same downstream positions
x/x s =
to those that
0.33, occurred
0.67, 1.00, in region
1.33, 3.
2.00, 2.33, and 2.67. Herein, W 0 is defined as the ratio of the standard

deviation of the vertical flow-velocity component fluctuations to Uc . Figure 4 clearly shows a substantial
reduction in W 0 as compared to U0 . For all the profiles, W 0 ranges between a minimum of 0.03 and a
maximum of 0.17 against 0.13 and 0.30, respectively, observed for U0 . Furthermore, at the different
downstream positions x/xs , W 0 decreases with decreasing Z/zs (i.e., going down to the equilibrium-scour
bed). This decrease is continuous and with a significant reduction rate as compared to U0 . This may be
explained by the considerable reduction of the vertical velocity flow going towards the equilibrium
bed profile, which effectively reduces the vertical flow-force that could lift the sediments from the
bottom. Similar to U0 , W 0 experiences maximum values in region 1 and at the upstream side of region
3. In region 2 however, W 0 exhibits the smallest values, comparable to those that occurred in region 3.
Water2020,
Water 2020,x,12, 186 PEER REVIEW
x FOR 7 of1617
7 of

W'
0.00 0.05 0.10 0.15 0.20
0.0
-0.1
x/xs
-0.2
0.33
-0.3 0.67
Z/zs

-0.4 1.00
-0.5 1.33
2.00
-0.6
2.33
-0.7
2.67
-0.8
-0.9
-1.0

Figure 4. Profiles of vertical turbulence intensity, W 0 , at different downstream positions x/xs along the
Figure 4. Profiles
plane of flow symmetry (y =
of vertical turbulence
0). intensity, W′, at different downstream positions x/xs along the
plane of flow symmetry (y = 0).
Figure 5 illustrates the vertical profiles of the normalized Reynolds shear stresses in the scour
hole.Figure
The profiles were the
5 illustrates taken at theprofiles
vertical downstreamof thepositions
normalized = 0.33, 0.67,
x/xsReynolds 1.00,
shear 1.33, 2.00,
stresses 2.33,
in the and
scour
2.67.The
hole. In Figure
profiles U0 W 0taken
5, were = −<uat0 wthe0 >/U 2 , where <u0 w0 > is the time-averaged stress over the length of
downstream
c positions x/xs = 0.33, 0.67, 1.00, 1.33, 2.00, 2.33, and
the time
2.67. series5,and
In Figure U′W′ (u=0 , −<u′w′>/U
w0 ) are thec2,velocity fluctuations
where <u′w′> of the streamwise
is the time-averaged and
stress oververtical component,
the length of the
0 0
respectively.
time series and The(u′,
valuesw′) of areU the
W clearly
velocityhave a heterogeneous
fluctuations of the distribution
streamwise atand the vertical
differentcomponent,
downstream
positions x/xs .The
respectively. Thisvalues
heterogeneityof U′W′ vertically
clearlydecreases
have a downward (towards
heterogeneous the scour bed).
distribution at the Furthermore,
different
Figure 5 shows that the heterogeneity of the Reynolds stresses is spatially
downstream positions x/xs. This heterogeneity vertically decreases downward (towards the scour variable in the scour hole.
s = 0.67, U WFigure
At x/x 0 0 varies5from a maximum value O(4 × of −2
10the) to a value stresses −3
O(10 ).isBoth the magnitude
bed). Furthermore, shows that the heterogeneity Reynolds spatially variable
0 gradually decrease with increasing x/x . For−2x/x > 1.33, there −3
U0 WU′W′
inand
thethe
scourvariation
hole. At rangex/xs of= 0.67, varies from a maximum value O(4s × 10 ) tos a value O(10 is).a
sharp
Both thedecrease
magnitude of the andReynolds
the variationstress magnitude
range of U′W′and the values
gradually of U0with
decrease W0 tend to an homogeneous
increasing x/xs. For x/xs
−3
>distribution,
1.33, there isfroma sharpa maximum
decrease of predictable
the Reynoldsaveraged value O(4 ×and
stress magnitude 10 the) tovalues
a value U′W′−3tend
of O(10 ). Figure
to an5
also shows that U 0 W0 exhibits the largest values in region 1 (see Figure 2), it decreases slightly in−3the
homogeneous distribution, from a maximum predictable averaged value O(4 × 10 ) to a value O(10 −3 ).
upstream
Figure 5 alsoside of region
shows that U′W′ 3, and it is significantly
exhibits reduced
the largest values in in region
region 2. At
1 (see the edge
Figure 2), itof the downstream
decreases slightly
inscour-side,
the upstream x/xs >side
1.33,ofUregion0 W0 shows the lowest values.
3, and it is significantly reduced in region 2. At the edge of the
downstream scour-side, x/xs > 1.33, U′W′ shows the lowest values.
Figure 5 also indicates that U′W′ always has positive values in the scour hole, along the plane of
flow symmetry (y = 0). This reflects a clear idea upon the frictional drag distribution and the vertical
effective momentum transfer associated with the scour hole structure in the plane of flow symmetry.
In the scour hole, the Reynolds shear stress is developed due to the formation of eddies of many
different length scales. In the immediate vicinity of the sediment scour bed, localized turbulent eddies
play an important role in removing a sediment particle from its stabilized position. According to Ali
and Dey [30], at the flow-bed interface, the eddies of greater size than the sediment diameter, d50 as
an example, weakly contribute to the vertical velocity component. By contrast, the eddies of smaller
size than the sediment diameter perfectly fit in the space between the particles, providing an effective
contribution to the vertical velocity component and therefore a substantial transfer of the vertical
momentum may occur. The increase of the vertical lift force generated by small eddies (with smaller
size than the sediment diameter) and the important horizontal component of momentum transmitted
by large eddies (with greater size than the sediment diameter) at the sediment bed, play a crucial role
in putting the sediment particles in suspension. Applying the phenomenological theory of turbulence
and a dimensional analysis, Ali and Dey [30] found a scale of the Reynold shear stress at the bed for
the incipient motion of sediment particles.
1 U
Figure 5. Vertical profiles of normalized Reynolds shear stress,

 0 W0 , in scour hole and at different
downstream positions x/xs along the plane
w u 'w  wU b  (y = 0).
2
of 'flow symmetry
b
2 (1)
Water 2020, 12, 186 8 of 17

Figure 5 also indicates that U0 W0 always has positive values in the scour hole, along the plane of
flow symmetry (y = 0). This reflects a clear idea upon the frictional drag distribution and the vertical
effective momentum transfer associated with the scour hole structure in the plane of flow symmetry.
In the scour hole, the Reynolds shear stress is developed due to the formation of eddies of many
different length scales. In the immediate vicinity of the sediment scour bed, localized turbulent eddies
play an important role in removing a sediment particle from its stabilized position. According to Ali
and Dey [30], at the flow-bed interface, the eddies of greater size than the sediment diameter, d50 as
an example, weakly contribute to the vertical velocity component. By contrast, the eddies of smaller
size than the sediment diameter perfectly fit in the space between the particles, providing an effective
contribution to the vertical velocity component and therefore a substantial transfer of the vertical
momentum may occur. The increase of the vertical lift force generated by small eddies (with smaller
size than the sediment diameter) and the important horizontal component of momentum transmitted
by large eddies (with greater size than the sediment diameter) at the sediment bed, play a crucial role
in putting the sediment particles in suspension. Applying the phenomenological theory of turbulence
and a dimensional analysis, Ali and Dey [30] found a scale of the Reynold shear stress at the bed for
the incipient motion of sediment particles.

(1+σ)
−ρw u0 w0 b ∼ ρw Ub 2 λ− 2

(1)

where the subscript b indicates the flow-bed interface position, Ub is the threshold velocity, defined
as the near-bed velocity that is marginally sufficient to initiate the particle motion at the bed surface,
λ(= d50 /hx ) is a relative roughness, hx is defined in Figure 1, σ is the spectral exponent of the turbulent
energy spectrum. By equating the Reynolds shear stress, obtained near the bed surface, and the bed
shear stress τb , we can obtain the threshold velocity Ub . τb can be related to the gravitational shear
stress τg as τb ~ τg = (ρs − ρw )gd50 θb , where θb is the threshold Shields parameter. θb is a function of
a particle parameter D = [(g∆d50 )/υ2 ]1/3 . Combining the different parameters together, one obtains a
scaling expression of the threshold densimetric Froude number, Fdb :
(1+σ)
ρ w Ub 2 λ − 2 ∼ (ρs − ρw ) gd50 f (D)
Ub (1+σ) 1 (2)
Fdb = √ ∼λ 4 f 2 (D)
∆gd50

It is worth mentioning that for a hydraulically rough flow regime the function f(D) tends to a
constant value.
Figure 6 depicts the vertical profiles of the time-averaged turbulent kinetic energy, K, normalized
by Uc 2 , at different downstream positions x/D from the bed sill. In the plane of flow symmetry
(y = 0), k = (<u02 > + <w02 >)/2, where the angle brackets indicate the average over the length of
the time series. In the energy inertial subrange, the energy cascade yields the ‘σ = −5/30 spectral
law. Since the energy equilibrium in this energy subrange is maintained by the balance between the
production and dissipation rate of the turbulent kinetic energy, the energy spectrum function is scaled as
E(κ) ~ ε2/3 κ−5/3 , where ε is the turbulent kinetic energyRdissipation rate and κ is the wavenumber. In this
case, the turbulent kinetic energy can be scaled as K ~ ε2/3 κ−5/3 dκ. Figure 6 clearly highlights a spatial
variation of K within the scour flow-field. In Figure 6, all the vertical profiles, except that at x/xs = 1.33,
show a decrease of K as going down towards the scour bed, but with different reduction rates that
depend on the downstream positions x/xs . At x/xs = 0.67, K experiences both the maximum values and
the maximum reduction rate. This implies that the jet-like region, indicated by region 1 in Figure 2, is a
location of maximum turbulent energy production. At x/xs = 1.33, K shows an increase with increasing
depth, it attains a maximum value O(3 × 10−2 ) at Z/zs = −0.6 and then begins to decrease going down to
the bed-flow interface. This fact can be explained by the transition effect between regions 1 and 3 (see
Figure 2), where the jet-like diffusion is accompanied by high levels of flow-turbulence intensities and
large kinetic energy production, as also observed in a previous study by Ben Meftah and Mossa [20].
Water 2020, 12, 186 9 of 17

At x/xs > 1.33, k undergoes a sharp decrease, showing values O(10−2 ), it also shows a very gradual
reduction going downwards. At equilibrium scour condition, k shows very small values, ranging
between 0.005 and 0.013, near the bed-flow interface for the different downstream positions x/xs . In the
scour hole, at equilibrium condition, the stability of the sediment particles is due to the significant
decrease of turbulent energy production at the bed-flow interface.

Figure 6. Vertical profiles of normalized turbulent kinetic energy, k/Uc 2 , in scour hole and at different
downstream positions x/xs along the plane of flow symmetry (y = 0).

3.3. Turbulent Length Scales


Determination of the eddies scales in a turbulent flow is of crucial importance for experimental and
numerical investigations, defining suitable domain-dimensions (area or volume) for computation [20].
Since the condition of incipient movement of the sediment particles is significantly influenced by the
size of turbulent eddies, in this section, we try to experimentally determine the characteristic eddy
length scales of the turbulent flow in the scour hole1at equilibrium condition. The integral length scale
Li (= Lx , Lz ) is simply calculated as the product of the integral time scale Ti and the local time-averaged
velocity Ui (= U, W), where Ti (= Tx , Tz ) is computed integrating the autocorrelation of the measured
instantaneous flow velocity ui (t) [= u(t), w(t)]. The autocorrelation of the measured instantaneous
flow velocity was determined after a spectral analysis of the flow velocity fluctuation at the different
measurement points.
Figures 7 and 8 depict the vertical profiles of the integral length scales Lx and Lz , respectively,
normalized by the mean average diameter of sediments, d50 , at different downstream positions x/xs .
The processed data and the downstream positions x/xs are the same as those covered in the previous
sections. Herein, Lx is the integral length scale in the x-direction and Lz in the z-direction, calculated by
means of the variables (U, Tx ) and (W, Tz ), respectively.
Contrary to what observed in Figures 3–6 for the flow turbulence properties in the scour hole,
Figure 7 shows that larger values of Lx occur almost at the location of lower turbulence levels.
At x/xs = 0.33 and 0.64, i.e., the positions of the jet-like flow and jet diffusion in regions 1 and 2 (Figure 2),
Lx has a size of the order of 1 ÷ 5d50 . At x/xs = 1, Lx is considerably increased, compared to the values
at the upstream positions x/xs = 0.33 and 0.64. It vertically decreases, going down towards the scour
bed, from value O(10d50 ) to value O(d50 ) near the sediment bed. At x/xs = 1.33, Lx shows a maximum
value O(27d50 ) and it monotonically decreases with increasing depth, reaching a value O(7d50 ) close to
Water 2020, 12, 186 10 of 17

the scour bed. At x/xs > 1.33, a sharp increase of Lx can be clearly noted. It shows an average size of
O(40d50 ), and it ranges between a minimum and maximum of 20d50 and 54d50 , respectively. Contrary
to the region of high turbulence levels (x/xs < 1), at x/xs > 1.33, the Lx -magnitudes rapidly decrease
with increasing flow depth, yielding maximum gradient values along the vertical. The integral length
scale distribution in the scour hole at equilibrium provides an integrated hydrodynamic picture on the
formation and relative macroscopic scales of the turbulent eddies.

Figure 7. Vertical profiles of normalized turbulent length scales, Lx /d50 , at different downstream
positions x/xs at the plane of flow symmetry (y = 0).

Figure 8. Vertical profiles of normalized turbulent length scales, Lz /d50 , at different downstream
positions x/xs at the plane of flow symmetry (y = 0).

2
Water 2020, 12, 186 11 of 17

Figure 8 illustrates the dimensionless integral length scale Lz /d50 versus the dimensionless vertical
position Z/zs in the scour hole at equilibrium phase. Contrary to what is shown in Figure 7 with Lx ,
Figure 8 indicates that the larger values of Lz appear at the location of higher turbulence levels. At x/xs
≤ 1.33, Lz shows both the largest values and the highest gradient along the vertical. It attains maximum
values of almost 3d50 at x/xs = 1 and 1.33. These values significantly decrease near the scour bed to an
order of 0.1d50 to 0.4d50 . This behavior could play an important role in increasing the vertical lifting
force to move the sediment particles before reaching an equilibrium condition. At x/xs = 0.33 and 1.33,
Lz slightly decreases to values O(1d50 ). This distribution of Lz at these positions seems reasonable, as it
is strongly influenced by the incoming inclined flow-jet in the scour hole, which increases the vertical
velocity component. At x/xs > 1.33, Lz shows very small values, less than 1d50 . This is explained by the
smooth outflow from region 3, as observed in Figure 2, where a streamwise velocity dominance over
the vertical component occurs.
The results obtained from the distribution of the integral length scales Lx and Lz seem to indicate
that, in addition to the drag force, the erosion capacity of the flow increases with the increase of the
vertical lifting force acting on the sediment particles. The vertical lifting force, essential for moving the
Water 2020, x, x FOR PEER REVIEW 11 of 16
sediment particles, is a direct result of an appropriate magnitude of the vertical velocity components.
Figure
Figure 99 displays
displays the the normalized
normalized Kolmogorov’s
Kolmogorov's micro-scale
micro-scale η/d η/d5050 versus
versus the
the normalized
normalized vertical
vertical
coordinate Z/z at different distances from the bed sills x/x .
coordinate Z/zss at different distances from the bed sills x/xs.s The main observation fromThe main observation from Figure
Figure 9
9 is
is
thethe expectedsignificant
expected significantreduction
reductionofofηηcompared
comparedto to LLxxand
andLLz,z ,by
by anan average
average (over
(over all
all measured
measured
points)
points) factor of 340 and 25, respectively. Through the scour hole, η shows values ranging between
factor of 340 and 25, respectively. Through the scour hole, η shows values ranging between
0.017d
0.017d50 to0.044d
50 to 0.044d5050 , an
, an equivalent
equivalent of 0.04
of 0.04 to 0.15
to 0.15 mm.mm.
For the Forpositions
the positions x/xs η
x/xs ≤ 1.33, ≤ shows
1.33, ηtheshows
smallestthe
smallest values, which are almost of constant magnitude along
values, which are almost of constant magnitude along the vertical and are very comparable, the vertical and are very comparable,
indicating
indicating aa kind kind of of local
local isotropy
isotropy of of the
the flow
flow turbulence
turbulence at at these
thesescales.
scales. For x/xss >
For x/x > 1.33,
1.33, ηη shows
shows an an
increase
increase by by an
an almost
almostfactor
factorofof1.5
1.5at x/xss =
atx/x 2, presenting
= 2, presenting an average length
an average length scale
scale of
of the
the order
order of of 0.56d
0.56d50 50,,
and then continues to increase monotonically reaching an average value
and then continues to increase monotonically reaching an average value of the order of 0.73d50, quite of the order of 0.73d 50 , quite
constant
constant at at both positions x/x
both positions x/xss = 2.33 and
= 2.33 and 2.67.
2.67. Figure
Figure 99 indicates
indicates that
that for
for all
all profiles,
profiles, regardless
regardless of of the
the
positions x/x
positions x/xss,, η/d
η/d50 is invariant
50 is invariantalong
alongthe thevertical
verticaldirection.
direction.Figure
Figure9 9also
alsopoints
pointsout outthat
thatthethesize
sizeofofη
ηis is influenced
influenced bybythethe level
level of of
thethe flow
flow turbulence
turbulence intensity.
intensity. In areas
In areas of high
of high turbulence,
turbulence, suchsuch
as atasx/xats
x/x s ≤ 1.33,
≤ 1.33, η considerably
η considerably decreases
decreases to smaller
to smaller values.
values. This This
may may be explained
be explained by thebydiffusion
the diffusion
of theofinlet
the
inlet jet-like flow in the scour hole, which induces the formation of small
jet-like flow in the scour hole, which induces the formation of small eddy scales that increase velocity eddy scales that increase
velocity
fluctuation. fluctuation.

/d50
0.00 0.05 0.10 0.15 0.20
0.0
-0.1
-0.2 x/xs
0.33
-0.3
0.67
Z/zs

-0.4 1.00
-0.5 1.33
-0.6 2.00

-0.7 2.33
2.67
-0.8
-0.9
-1.0

Figure 9. Vertical profiles of normalized Kolmogorov’s micro-scale, η/d50 , at different downstream


Figure 9. x/x
positions Vertical
s at theprofiles
plane ofofflow
normalized
symmetry (y = 0).
Kolmogorov’s micro-scale, η/d50, at different downstream
positions x/xs at the plane of flow symmetry (y = 0).

3.4. Scaling of the Maximum Scour Depth at Equilibrium


In this section we focus on the scaling of the maximum scour depth at the equilibrium condition,
derived from the phenomenological theory of turbulence. At equilibrium, the scour hole between
two consecutive bed sills is shown schematically in Figure 10. The scour hole is caused by the effect
of a sort of an entering jet flow of both initial thickness and initial velocity comparable, respectively,
U c hc U s hs U s  zs  ht 
U c hs  zs  ht  (6)
 
U s hc hc
From12,Equation
Water 2020, 186 (2), we can obtain a scaling of the mean velocity Us as: 12 of 17

1 
1 
d  4
1 1
3.4. Scaling of the Maximum gd50 at f  D   gd50  50 
U s U bScourDepth f 2  D
4Equilibrium
2 (7)
 hs 
In this section we focus on the scaling of the maximum scour depth at the equilibrium condition,
derived from the Equation
Substituting phenomenological theory of
(7) into Equation (6)turbulence.
produces: At equilibrium, the scour hole between
two consecutive bed sills is shown schematically 1 
in Figure 10. The scour hole is caused by the effect
 1 comparable,
  2
 3   d 50 
4 2
zs  ht
of a sort of anhentering jet flow of both 
initial 3 
thickness
 and initial 4velocity  respectively,
Uc, D   Fthe qisDthe
 unit water
 3   3   3 
s
to the flow depth,  hc , anditsFdcorresponding
c   velocity, f over d c sill.Inc Figure f 10,
3  (8)
hc hc
discharge, ls is the equilibrium scour  clength
h  and ht is the tailwater depth at the position xs of the
maximum scour depth at equilibrium,
Since for rough turbulent flow f(D) tends to defined asbe thea vertical
constantdistancefunction, between
Equation the(8)initial
can bebedreduced
profile
and the free-surface
to the following form: flow (at xs the total flow depth is h s = z s + ht ). The effective parameter of the scour
hole, illustrated in Figure 10, can be expressed as follows:
4 1 
hs zs  ht  3 

 3 
hs = zs + ht = f (hc , UFcd, cd50 , L,S c o , g, ρw , ρs , ν)
(9)
(3)
hc hc

Original bed
Free-surface flow
ls
q hc ht

zs
Sill

Sill
z
x
xs
Figure 10. Definitional sketch of scour hole downstream of bed sills, Uo indicates the initial jet velocity.
Figure 10. Definitional sketch of scour hole downstream of bed sills, Uo indicates the initial jet
velocity.
The application of dimensional analysis to the variables of Equation (3) leads to the following
expression:
It should be mentioned here hs that the
zs +scaling
ht law
 of the maximum
LSo
 scour depth at equilibrium,
=
investigated applying the phenomenological = f
theoryλ , , F dc ,
c of turbulence, Re c as shown by Equation (9),(4) is
hc hc hc
explicitly expressed as a function of only the dimensionless parameters Fdc and λ, also appearing in
For fully turbulent flow, Reg > 70 [31,32], the dependence upon the Reynolds number, Rec , could
Equation (5) through dimensional analysis. It should be noted that, contrary to dimensional analysis,
be neglected, and therefore Equation (4) can be finally expressed as:
the application of the phenomenological theory of turbulence establishes a unique and complete
(with defined exponents) relationshiphs between
zs + ht some of the

LScharacteristic
o
 variables of the scour hole.
= = f λc , , Fdc (5)
Since the flow is fully turbulent within
hc thehcscour hole and hc the bed sediment is characterized by a
relative roughness in the range of 10−4 < λc <10−1 [30], the scaling law is validated for σ = −5/3, in the
The application of the continuity equation between the section at x = 0 (over the bed sill) and the
section at xs , position of maximum scour, yields:

Uc hc = Us hs = Us (zs + ht )
Uc hs ( zs + ht ) (6)
Us = hc = hc

From Equation (2), we can obtain a scaling of the mean velocity Us as:

! (1+σ)
p (1+σ) 1 p d50 4 1
Us ∝ Ub ∼ ∆gd50 λ 4 f (D) =
2 ∆gd50 f 2 (D) (7)
hs
Water 2020, 12, 186 13 of 17

Substituting Equation (7) into Equation (6) produces:

!− ( 1 + σ ) (1+σ)
hs zs + ht 4 d50 (3−σ) 2
− (3−σ 4 − (3−σ) 2
− (3−σ
= ∝ Fdc (3−σ) f ) (D) = Fdc (3−σ) λc f ) (D) (8)
hc hc hc

Since for rough turbulent flow f (D) tends to be a constant function, Equation (8) can be reduced to
the following form:
(1+σ)
hs zs + ht 4 −
= ∝ Fdc (3−σ) λc (3−σ) (9)
hc hc
It should be mentioned here that the scaling law of the maximum scour depth at equilibrium,
investigated applying the phenomenological theory of turbulence, as shown by Equation (9), is explicitly
expressed as a function of only the dimensionless parameters Fdc and λ, also appearing in Equation (5)
through dimensional analysis. It should be noted that, contrary to dimensional analysis, the application
of the phenomenological theory of turbulence establishes a unique and complete (with defined
exponents) relationship between some of the characteristic variables of the scour hole. Since the flow is
fully turbulent within the scour hole and the bed sediment is characterized by a relative roughness
in the range of 10−4 < λc <10−1 [30], the scaling law is validated for σ = −5/3, in the energy inertial
subrange, and thus Fdc (4/(3−σ) λc (−(1+σ)/(3−σ) = Fdc (6/7) λc (1/7) . In order to experimentally evaluate the
validity of this scaling law, shown in Equation (9), in Figure 11 we plot measured values of (zs + ht )/hc
as a function of FPEER(6/7) λ (1/7) . The experimental data used for this scope are those illustrated in Table 1,
Water 2020, x, x FOR dc c
REVIEW 14 of 16
acquired by Ben Meftah and Mossa [3].

4.5

4.0

3.5

3.0

2.5
(zs + ht)/hc

2.0

1.5 Eq. (10)


1.0 Deviation 15%
0.5 Ben Meftah and Mossa [3]
0.0
1.6 1.8 2.0 2.2 2.4 2.6

Fdc6/7c1/7

Figure 11. Validation of the scaling law of the maximum scour depth downstream of bed sills
Figure 11. Validation of the scaling law of the maximum scour depth downstream of bed sills obtained
obtained by applying the fundamental laws of turbulent energy spectrum with σ = −5/3, and thus
by (4/(3−σ)
applying(−(1+σ)/(3−σ)
the fundamental laws of turbulent energy spectrum with σ = −5/3, and thus
Fdc(4/(3−σ) (−(1+σ)/(3−σ)
λc = Fdc (6/7) λc (1/7) . The data of Ben Meftah and Mossa [3] were obtained
Fdc λc = Fdc(6/7)λc(1/7). The data of Ben Meftah and Mossa [3] were obtained downstream of
downstream of bed sills in alluvial channels and with a bed sediment size of d50 = 1.8 mm.
bed sills in alluvial channels and with a bed sediment size of d50 = 1.8 mm.
Figure 11 clearly shows that the data of the normalized maximum scour depth (zs + ht )/hc plotted
law Fdc (6/7) λcBen
versus the scaling25.0 (1/7) tend to collapse into a single curve. From the interpolation of the
Meftah and Mossa [3]
data shown in Figure 11 the following
Bormannexpression,
and Julien [33]predicting the maximum scour depth at equilibrium,
is obtained: 20.0 Eq. (10)
hs 15%zs + ht 6 1
Deviation = = 1.43Fdc 7 λc 7 (10)
h
Eq. (11)c h c
15.0
It is worth mentioning thatDeviation
Equation 50%(10) is the regression fitting of data collected downstream of
(zs + ht)/hc

bed sills in alluvial channels. As reported in Ben Meftah and Mossa [3], the distance L between sills
varies between 1 and
10.0 4 m and the non-cohesive bed sediment consists of very coarse sand particles of

5.0

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Water 2020, 12, 186 14 of 17

mean average
Water 2020, size
x, x FOR PEER = 1.8 mm. Despite the non-presence of the parameter LS0 /hc in Equation
d50REVIEW (10),
14 of 16
with respect to Equation (5), which takes into consideration the effects of both the distance and the
slope between sills, Equation (10) suitably predicts the maximum scour depth downstream of bed sills
at equilibrium condition.4.5 This is proved by the low deviation of the data, ±15%, from the best fit line
of Equation (10). 4.0
To give more validity
3.5 to this approach, in addition to the data obtained by Ben Meftah and
Mossa [3], in Figure 12 we also plot data previously obtained by Bormann and Julien [4]. The scour
3.0
data of Bormann and Julien [4] were collected downstream of a grade control structure of different
face slopes (18◦ , 45◦2.5 , and 90◦ with respect to the horizontal) in a sand-bed channel of d50 equal to
(zs + ht)/hc

0.3 mm and 0.45 mm. 2.0 Due to the complexity of the phenomenon and the wide variety of conditions
(including different 1.5 characteristics of the sediment particles, different geometries of the grade control
Eq. (10)
structure, and different entering jet typologies), the data by Bormann and Julien [4], plotted according
1.0
to the scaling law of Equation (9), show more scattering thanDeviation those by15%Ben Meftah and Mossa [3].
0.5
As compared to previous studies [4,33], when data were fitted by Benother scaling
Meftah laws,
and Mossa [3] the scaling law of
Equation (9) predicts 0.0with more accuracy the maximum equilibrium scour depth, as shown by the
50%-deviation from the best fit line of Equation (11). In fact, the normalized scour depth (zs + ht )/hc
1.6 1.8 2.0 2.2 2.4 2.6
(6/7) λ (1/7) . A linear
by Bormann and Julien [4] shows an overall increasing Fdc6/7c1/7 trend as a function of Fdc c
regression analysis of these data leads to the following expression for predicting the equilibrium scour
depthFigure
at grade control structures
11. Validation (under
of the scaling Bormann
law of and Julien’s
the maximum [4] downstream
scour depth conditions):of bed sills obtained
by applying the fundamental laws of turbulent energy spectrum with σ = −5/3, and thus
hs zs + ht 6 1
= of Ben Meftah
Fdc(4/(3−σ)λc(−(1+σ)/(3−σ) = Fdc(6/7)λc(1/7). The data = 0.53F
and dc 7 λc 7 [3] were obtained downstream of (11)
Mossa
hc hc
bed sills in alluvial channels and with a bed sediment size of d50 = 1.8 mm.

25.0
Ben Meftah and Mossa [3]
Bormann and Julien [33]
20.0 Eq. (10)
Deviation 15%
Eq. (11)
15.0 Deviation 50%
(zs + ht)/hc

10.0

5.0

0.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Fdc6/7c1/7

Figure 12. Validation of the scaling law of the maximum scour depth using data obtained by Bormann
Figure 12. Validation of the scaling law of the maximum scour depth using data obtained by Bormann
and Julien [4] downstream of grade control structures of different features and with sediment sizes of
and Julien [4] downstream of grade control structures of different features and with sediment sizes of
d50 ranging between 0.30 and 0.45 mm, Reg ranging between 396 and 2093 and with relative roughness
d50the
in ranging
rangebetween
of 10−4 <0.30 and
λc <10 −20.45
. Themm, Reg ranging
scaling betweenfor
law is validated 396
σ=and 2093
−5/3, in and with relative
the energy inertialroughness
subrange.
in the range of 10 < λc <10 . The scaling law is validated for σ = −5/3, in the energy inertial subrange.
−4 −2

It is worth noting that the scaling approach introduced by Equation (9) to predict the equilibrium
4. Conclusions
scour depth is only expressed as a function of the densimetric Froude, Fdc , and the relative roughness,
λc , with exponents that
Considering function
an of σ. It does not
exhaustive involve all
prediction of the
thecharacteristic parameters
scour development is affecting the scour
fundamental for
features, i.e., the parameter
structural stability, we examined LS /h
0 c that appears in Equation (5) through dimensional
experimentally the local scouring processes downstream analysis. In any
of
case, its application shows reasonable accuracy in predicting the equilibrium
hydraulic grade control structures. This study focuses on the analysis of flow hydrodynamic scour depth in different
hydraulic
structuresstructures
within theforscour
sand-bed
hole rivers. The limitation
at equilibrium and onof this
the approach
applicationin the present
of the form, however,
phenomenological
is that it
theory ofdoes not exhibit
turbulence a general
to find a scalingproportional
law predictingrelationship for all scour
the equilibrium hydraulic structure typologies,
depth.
Specifically, by means of velocity measurements, we observed that at equilibrium conditions the
flow in the scour hole is fundamentally characterized by three distinct regions: (i) A first region
consists of a free entering jet flow, plunging from the crest of bed sill into the scour hole and strongly
eroding the bed sediments; (ii) a second region located near the scour bottom, extending upstream
Water 2020, 12, 186 15 of 17

as shown in Equations (10) and (11) by the coefficients 1.43 and 0.53. This implies that a further better
combination between this scaling approach and other characteristic parameters, obtained through
dimensional analysis, can lead to finding a general expression to predict the maximum equilibrium
scour depth.

4. Conclusions
Considering that an exhaustive prediction of the scour development is fundamental for structural
stability, we examined experimentally the local scouring processes downstream of hydraulic grade
control structures. This study focuses on the analysis of flow hydrodynamic structures within the
scour hole at equilibrium and on the application of the phenomenological theory of turbulence to find
a scaling law predicting the equilibrium scour depth.
Specifically, by means of velocity measurements, we observed that at equilibrium conditions
the flow in the scour hole is fundamentally characterized by three distinct regions: (i) A first region
consists of a free entering jet flow, plunging from the crest of bed sill into the scour hole and strongly
eroding the bed sediments; (ii) a second region located near the scour bottom, extending upstream
due to vortex formations (eddies) generated by the jet diffusion, allowing to reach the equilibrium
condition; and (iii) a third less-turbulent region, localized downstream of the first and the second
regions and characterized by an almost unidirectional flow in the x-direction.
The detailed analysis of the flow turbulence characteristics in the equilibrium scour hole showed
that both the Reynolds shear stresses and turbulence intensities (the streamwise and the vertical ones)
exhibit their greatest values in the first region, where the approaching jet has effects, while a sharp
reduction of them occurs in the third region. Moreover, these turbulent properties (Reynolds shear
stresses and turbulence intensities) show negative vertical gradient going down towards the scour
bed. A heterogeneous spatial distribution of these turbulent features is especially evident in the first
and second regions, where the jet effect is quite strong. The analysis of the turbulent kinetic energy
shows that the region of jet-like flow, region 1, is a location of maximum turbulent energy production.
At equilibrium condition, the stability of sediment particles is due to a significant decrease of turbulent
energy production at the bed-flow interface.
The distribution of the longitudinal and vertical integral length scales, in the plane of flow
symmetry, was also analyzed. The larger values of Lx occur almost at the location of lower turbulence
levels. On the contrary, Lz shows maximum values at the regions of higher turbulence levels.
This distribution of Lx and Lz reflects the role of the flow turbulent eddies in the incipient motion of
the sediment particles. The increase of Lz , related to the increase of the vertical velocity component,
enhances the vertical lifting force, giving more possibilities for particles to move.
Finally, the phenomenological theory of turbulence was applied to the scour hole, deducing a
new scaling law of the maximum scour depth at equilibrium in non-cohesive bed rivers. Once fixed
the spectral exponent of the turbulent energy spectrum, the scaling low becomes a simple function of
the densimetric Froude number, Fdc , and the relative roughness, λc . This scaling law was evaluated
and validated using laboratory measured scouring data of sand-bed with a relative roughness in the
range of 10−4 < λc <10−1 . It showed a quite reasonable accuracy in predicting the equilibrium scour
depth in different hydraulic structures.

Author Contributions: M.B.M. performed the experiments, analyzed the data, designed the study and wrote the
paper; F.D.S., D.D.P. and M.M. contributed suggestions, discussions and reviewed the manuscript. All authors
have read and agreed to the published version of the manuscript.
Funding: This research recevied no external funding.
Acknowledgments: The experiments were carried out at the Hydraulic Laboratory of the Mediterranean
Agronomic Institute of Bari (Italy).
Conflicts of Interest: The authors declare that they have no conflicts of interest.
Water 2020, 12, 186 16 of 17

References
1. Adduce, C.; Sciortino, G. Scour due to a horizontal turbulent jet: Numerical and experimental investigation.
J. Hydraul. Res. 2006, 44, 663–673. [CrossRef]
2. Balachandar, R.; Kells, J.A. Local channel in scour in uniformly graded sediments: The time-scale problem.
Can. J. Civ. Eng. 1997, 24, 799–807. [CrossRef]
3. Ben, M.M.; Mossa, M. Scour holes downstream of bed sills in low-gradient channels. J. Hydraul. Res. 2006,
44, 497–509.
4. Bormann, N.E.; Julien, P.Y. Scour downstream of grade-control structures. J. Hydraul. Eng. 1991, 117, 579–594.
[CrossRef]
5. Carstens, M.R. Similarity laws for localized scour. J. Hydraul. Div. 1966, 92, 13–36.
6. D’Agostino, V.; Ferro, V. Scour on alluvial bed downstream of grade-control structures. J. Hydraul. Eng. 2004,
130, 24–37. [CrossRef]
7. Espa, P.; Sibilla, S. Experimental study of the scour regimes downstream of an apron for intermediate
tailwater depths. J. Appl. Fluid Mech. 2014, 7, 611–624.
8. Gaudio, R.; Marion, A.; Bovolin, V. Morphological effects of bed sills in degrading rivers. J. Hydrau. Res.
2000, 38, 89–96. [CrossRef]
9. Lenzi, M.A.; Marion, A.; Comiti, F. Local scouring at grade-control structures in alluvial mountain rivers.
Water Resour. Res. 2003, 39, 1176. [CrossRef]
10. Mason, P.J.; Arumugam, K. Free jet scour below dams and flip buckets. J. Hydraul. Eng. 1985, 111, 220–235.
[CrossRef]
11. Pagliara, S.; Amidei, M.; Hager, W.H. Hydraulics of 3D plunge pool scour. J. Hydraul. Eng. 2008, 134,
1275–1284. [CrossRef]
12. Papanicolaou, A.N.; Bressan, F.; Fox, J.; Kramer, C.; Kjos, L. Role of structure submergence on scour evolution
in gravel bed rivers: Application to slope-crested structures. J. Hydraul. Eng. 2018, 144, 1087–1093. [CrossRef]
13. Wang, L.; Melville, B.W.; Guan, D.; Whittaker, C.N. Local scour at downstream sloped submerged weirs.
J. Hydraul. Eng. 2018, 144, 04018044. [CrossRef]
14. Tregnaghi, M.; Marion, A.; Gaudio, R. Affinity and similarity of local scour holes at bed sills. Water Resour.
Res. 2007, 43, W11417. [CrossRef]
15. Lu, J.Y.; Hong, J.H.; Chang, K.P.; Lu, T.F. Evolution of scouring process downstream of grade-control
structures under steady and unsteady flows. Hydrol. Process. 2013, 27, 2699–2709. [CrossRef]
16. Manes, C.; Brocchini, M. Local scour around structures and the phenomenology of turbulence. J. Fluid Mech.
2015, 779, 309–324. [CrossRef]
17. Gioia, G.; Bombardelli, F.A. Localized turbulent flows on scouring granular beds. Phys. Rev. Lett. 2005, 95,
014501. [CrossRef]
18. Bombardelli, F.A.; Gioia, G. Scouring of granular beds by jet-driven axisymmetric turbulent cauldrons.
Phys. Fluids 2006, 18, 088101. [CrossRef]
19. Ali, S.Z.; Dey, S. Impact of phenomenological theory of turbulence on pragmatic approach to fluvial
hydraulics. Phys. Fluids. 2018, 30, 045105. [CrossRef]
20. Meftah, M.B.; Mossa, M. Turbulence measurement of vertical dense jets in crossflow. Water 2018, 10, 286.
[CrossRef]
21. Meftah, M.B.; Malcangio, D.; De Serio, F.; Mossa, M. Vertical dense jet in flowing current. Environ. Fluid
Mech. 2018, 18, 75–96. [CrossRef]
22. Meftah, M.B.; De Serio, F.; Mossa, M.; Pollio, A. Experimental study of recirculating flows generated by
lateral shock waves in very large channels. Environ. Fluid Mech. 2008, 8, 215–238. [CrossRef]
23. Meftah, M.B.; Mossa, M. Partially obstructed channel: Contraction ratio effect on the flow hydrodynamic
structure and prediction of the transversal mean velocity profile. J. Hydrol. 2016, 542, 87–100. [CrossRef]
24. Meftah, M.B.; Mossa, M. A modified log-law of flow velocity distribution in partly obstructed open channels.
Environ. Fluid Mech. 2016, 16, 453–479. [CrossRef]
25. Ben Meftah, M.; De Serio, F.; Mossa, M. Hydrodynamic behavior in the outer shear layer of partly obstructed
open channels. Phys. Fluids. 2014, 26, 65102. [CrossRef]
26. Ben Meftah, M.; Mossa, M. Prediction of channel flow characteristics through square arrays of emergent
cylinders. Phys. Fluids. 2013, 25, 45102. [CrossRef]
Water 2020, 12, 186 17 of 17

27. Ben Meftah, M.; Mossa, M.; Pollio, A. Considerations on shock wave/boundary layer interaction in undular
hydraulic jumps in horizontal channels with a very high aspect ratio. Eur. J. Mech. B Fluids. 2010, 29, 415–429.
[CrossRef]
28. Scurlock, S.M.; Thornton, C.I.; Abt, S.R. Equilibrium scour downstream of three-dimensional grade control
structures. J. Hydraul. Eng. 2011, 138, 167–176. [CrossRef]
29. Ghodsian, M.; Mehraein, M.; Ranjbar, H.R. Local scour due to free fall jets in non-uniform sediment.
Sci. Iranica. 2012, 19, 1437–1444. [CrossRef]
30. Ali, S.Z.; Dey, S. Origin of the scaling laws of sediment transaport. Proc. R. Soc. A 2017, 473, 20160785.
[CrossRef]
31. De Vincenzo, A.; Brancati, F.; Pannone, M. An experimental analysis of bed load transport in gravel-bed
braided rivers with high grain Reynolds numbers. Adv. Water Res. 2016, 94, 160–173.
32. Mirauda, D.; De Vincenzo, A.; Pannone, M. Statistical characterization of flow field structure in evolving
braided gravel beds. Spat. Stat. 2019, 34, 100268. [CrossRef]
33. Meftah, M.B.; Mossa, M. New approach to predicting local scour downstream of grade-control structure.
J. Hydraul. Eng. 2019, 146, 04019058. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like