You are on page 1of 9

Flow Patterns and Turbulence Structures in a Scour

Hole Downstream of a Submerged Weir


Dawei Guan 1; Bruce W. Melville, M.ASCE 2; and Heide Friedrich 3
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

Abstract: Scouring downstream of submerged weirs is a common problem resulting from the interaction of the three-dimensional turbulent
flow field around the structures and the mobile channel bed. This paper presents the distributions of flow patterns, bed shear stresses, and
turbulence structures in the approach flow and the scour hole downstream of a submerged weir. The experiments were conducted under the
clear-water scour condition for an equilibrium scour hole. The experimental results show that the flow structures are considerably changed by
the presence of the structure. A large recirculation zone and a flow reattachment region are formed downstream of the submerged weir.
Strongly paired cellular secondary flows are observed in the scour hole. The turbulence structures ahead of the recirculation zone govern
the dimensions of the scour hole. DOI: 10.1061/(ASCE)HY.1943-7900.0000803. 2014 American Society of Civil Engineers.
Author keywords: Submerged weir; Scour; Flow pattern; Bed shear stress; Turbulence; Secondary flows.

Introduction the first investigators to detect the three-dimensional turbulence


structure downstream of a W-weir in a meandering channel.
Submerged weirs (or dams) are low-head hydraulic structures con- In order to precisely predict the scouring downstream of sub-
structed in the channel of a waterway for the purpose of limiting merged weirs, it is important to develop a good understanding
excessive channel-bed degradation, raising the upstream water of the turbulence flow structures around such hydraulic structures.
level, and reducing the flow velocity. The flow depth on the weir The present study aims to obtain information on flow patterns,
crest is deep enough for barges to get through, even during dry boundary shear stresses, turbulence intensities, and Reynolds shear
seasons. Thus, once built, such weirs will improve river navigation stresses in the scour zone. The experiments were confined to the
conditions. In sloping, straight channels, several consecutive clear-water scour condition and the scour hole downstream of one
submerged weirs can be constructed, if necessary. The effect of single submerged weir at the equilibrium phase.
a submerged weir is to suddenly change the channel-bed elevation.
This sudden change of bed elevation at a submerged weir not
only influences the flow pattern, but also results in local scour Experiments
downstream of the structures.
For practical purposes, the most important scour parameters
are the scour-hole dimensions (i.e., maximum scour depth ds and Experimental Set-up
length ls ) at the equilibrium phase. Therefore, maximum scour The experimental work was conducted in a 12-m-long, 0.38-m-
depth and length have been widely studied, providing a selection deep, and 0.44-m-wide glass-sided, tilting flume (Fig. 1) in the
of empirical equations (Bormann and Julien 1991; DAgostino and Hydraulic Laboratory of the University of Auckland. At the up-
Ferro 2004; Marion et al. 2004; Chen et al. 2005; Marion et al. stream end of the flume, the water is fed into a mixing chamber
2006) to be applied to the design of submerged weirs. However, and enters the flume through a honeycomb flow straightener, which
there are only a few studies on the flow structure in the scour hole effectively eliminates any rotational flow component induced in the
downstream of submerged weirs. Ben Meftah and Mossa (2006) return pipelines, so that uniform flow is obtained. At the down-
studied flow turbulence in an equilibrium scour hole downstream stream end, sediment from the scour hole is trapped in a separated
of one weir in a sequence of weirs. Bhuiyan et al. (2007) were hopper-like sump, from where pumps return the flow to the inlet
end of the flume.
The sediment used in the experiments was coarse sand, with
1 median diameter d50 0.85 mm and relative submerged particle
Ph.D. Student, Dept. of Civil and Environmental Engineering, Univ. of
Auckland, Private Bag 92019, Auckland 1142, New Zealand (correspond- density 1.65. The sediment size distribution was near uniform,
ing author). E-mail: dgua324@aucklanduni.ac.nz with a standard deviation g 1.3. The weir used in the experi-
2
Professor, Dept. of Civil and Environmental Engineering, Univ. of ments was a 10-mm-thick rectangular plastic plate, with the same
Auckland, Private Bag 92019, Auckland 1142, New Zealand. E-mail: width as the flume. In the experiments, the weir was inserted into
b.melville@auckland.ac.nz the bed with a 40-mm protrusion from the initial flat bed, and was
3
Lecturer, Dept. of Civil and Environmental Engineering, Univ. of located 4.5 m from the outlet of the flume. During the test, the
Auckland, Private Bag 92019, Auckland 1142, New Zealand. E-mail: approach flow depth y and tail water depth yt were maintained
h.friedrich@auckland.ac.nz
at 150 mm.
Note. This manuscript was submitted on August 9, 2012; approved on
The upstream flow was fully developed in the experiment. The
July 16, 2013; published online on December 16, 2013. Discussion period
open until June 1, 2014; separate discussions must be submitted for indi- average approach flow velocity U 0 for this experiment was esti-
vidual papers. This paper is part of the Journal of Hydraulic Engineering, mated from the vertical distribution of approach flow velocities
Vol. 140, No. 1, January 1, 2014. ASCE, ISSN 0733-9429/2014/1-68- on the centerline of the flume in front of the weir, when the uniform
76/$25.00. flow was achieved as shown in Fig. 2. The approach average flow

68 / JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014

J. Hydraul. Eng., 2014, 140(1): 68-76


Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Schematic display of the flume

1000

water surface level 150mm


Flow Depth (mm)

100

10

1
0.22 0.24 0.26 0.28 0.3 0.32 0.34
Time-averaged velocity (m/s)

Fig. 2. Approach flow velocity profile (semilogarithmic coordinate


Fig. 3. Sensor arrangement for measuring water surface and bed
system)
profiles

velocity U 0 0.296 m=s was determined as the velocity at 0.368y


grid (see Fig. 3) on a carriage that can be moved along the top rail
(Yalin 1992). The average approach flow shear velocity, u
of the flume. The system was operated at 5 Hz and allowed meas-
0.014 m=s, was estimated from the logarithmic form of the velocity
urement of the whole scour region in about 1 min. Taking into ac-
profile (Fig. 2); the average approach flow critical shear velocity
uc 0.021 m=s was determined using the Shields diagram for count the detection of suspended sediment particles, the outliers of
the respective particle size (Melville 1997). Thus, the ratio of the raw bed profile data were filtered in the data postprocessing.
bed shear to critical shear velocity for the approach flow (u =uc The procedure of the program was to use 3 as the threshold
0.67) was calculated. The corresponding Reynolds number (R for the outlier detection (well known as the 3- rule), where
p is the standard deviation derived from the original data set. The
Uy=) and Froude number (F U= gy) were 44,400 and 0.24,
respectively. filtered data were then analyzed by spline interpolation procedures.
The final resolution of each processed bed profile is 10 10 mm.
At the start of the experiment, the sediment bed was levelled
Bed Profile and Velocity Measurement with a scraper after setting the weir. The flume was then filled
Theoretically, for clear-water scour, the equilibrium of the scour to the desired water depth. The filling process took place slowly
process should be defined as the condition when the dimensions to avoid disturbance around the weir before the actual experiment.
of the scour hole do not grow with time. However, even in small- Water temperature was measured in order to set the initial exper-
scale laboratory experiments, it may take several days or weeks imental parameters for the MTAs. After starting both pumps with
to attain equilibrium conditions (Melville and Chiew 1999). Thus the required settings, the water depth and the slope of the flume
it is important to conduct continuing bed-profile measurements to were adjusted to get uniform flow for the approach flow upstream
understand the scour process and ensure the scour equilibrium of the weir, while the flow depth downstream of the weir was con-
phase is obtained. The three-dimensional scour geometry down- trolled by adjusting the location of an overflow pipe in the sump.
stream of the weir was measured throughout the experiment using When uniform flow was obtained, bed profiles and water surface
Seateks multiple transducer arrays (MTAs) (SeaTek Instrumenta- were measured with 27 MTAs sensors as a function of time. The
tion, Florida) as a function of time. This instrument is an ultrasonic scour process lasted around 23 days until the scour hole reached
ranging system, comprising 32 transducers, which can detect the equilibrium. The time evolution of maximum scour depth and final
distance from the sensors to reflective objects. The measuring ac- scour geometry are shown in Figs. 4 and 5. As shown in Fig. 4, the
curacy of the system is approximately 1 mm. A detailed descrip- temporal development of the scour hole experiences three stages.
tion of this device can be found in Friedrich et al. (2005). During The maximum scour depth ds develops very quickly during the first
the test, only 27 transducers were employed, among which two day, then progresses at a decreasing rate over the following 19 days,
transducers were used for water surface measurements, with a as it approaches the equilibrium stage. During the final stage, the
125-mm interval. The transducers were mounted in a rectangular values of ds fluctuated around an average value of 151 mm, which

JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014 / 69

J. Hydraul. Eng., 2014, 140(1): 68-76


160
The closest measuring location to the bed was always 5 mm from
the bed surface. At each measurement point, 2-min samples were
dse collected. Throughout the experiments, the signal-to-noise (SNR)
120 ratio for each beam was maintained above 15. After the experi-
ds (mm) ments were completed, the output data from the velocimeter were
80 filtered using WinADV software (Wahl 2000). The filter was set to
remove spikes [using the phase-space threshold method of Goring
40 and Nikora (2002)] and data with low correlation (Minimum
COR < 70). Although the best configurations for the velocimeter
were carefully chosen during the experiments, the results for some
0
0 5 10 15 20 25 locations at 8.5 to 10 cm above the bed in the centerline longitu-
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

t (day) dinal section still had relatively high noise and low correlations.
These locations are called velocimeter weak spots or velocity holes,
Fig. 4. Temporal development of the maximum scour depth and are mainly caused by the return signal interference from the
boundary (Martin et al. 2002). After filtering, around 55% of data
for these weak spots were of good quality and therefore retained.
For all other measurement points, more than 80% of the data were
is taken as the maximum scour depth at the equilibrium phase (dse ) retained after filtering. The velocity power spectrum for the filtered
in this study. data points was examined with Kolmogorovs 5=3 law, conform-
After the scour hole reached equilibrium, the flow field ing that the data presented in this paper are of high quality.
was measured using a three-component, downward-facing Nortek
Vectrino+ acoustic velocimeter (Nortek AS, Rud, Norway). The
probe measures the velocities 50 mm beneath the acoustic transmit- Results and Discussion
ter, which must be submerged, and consequently velocities within
the first 55-mm depth beneath the water surface were not measured.
Measurements were taken along the centerline longitudinal section Two-Dimensional Velocity Distribution
and on three other transverse cross sections. The velocimeter was Fig. 6 shows the distribution of time-averaged velocity vectors
used with a sampling rate of 200 Hz. The sampling volume was on the centerline longitudinal section. The velocity vectors are
cylindrical, having a 6-mm diameter and an adjustable height vary- determined from the average values of the streamwise and vertical
ing from 1 to 7 mm. As suggested by Dey et al. (2011), the sam- velocity components. It can be seen that the upstream flow is quite
pling height was set as 12.5 mm in the near-bed zone to avoid uniform, even at the equilibrium stage. When this uniform flow
interfering with sediment particles; a 4-mm sampling height was approaches the submerged weir, the flow pattern is altered by the
used in the upper flow zone of the centerline longitudinal section, sudden change of bed elevation. The approach flow is accelerated
and a 7-mm height for the three other transverse cross sections. at the crest of the weir, and a weak back flow is created immediately

Section U Section M Section D


Transverse Direction (mm)

200
0

40

100
0
-10

-140
0

-80

-60
-40

-40

0
-120

20
0

-100

0
20

-2
-80
0

-20

-1

0
-60

20
-20

-100
0

-140
-200
-0.5 0 0.5 1 1.5 2 2.5 3
Longitudinal Direction (m)

Fig. 5. Scour geometry at the equilibrium phase, contours in millimeters

200
Vertical distance (mm)

100

0
Recirculation zone
-100 Flow reattachment region 0.20 m/s

-0.5 0 0.5 1 1.5 2 2.5


Longitudinal distance (m)

Fig. 6. Velocity vectors distribution in the centerline longitudinal section

70 / JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014

J. Hydraul. Eng., 2014, 140(1): 68-76


upstream of the weir. At the upstream base of the weir, a small scour hole is 84% (127 mm=151 mm), which implies that the effect of
hole (around 20-mm depth) was observed, which was produced by secondary flows cannot be ignored when studying clear water scour
weak vortices, generated by the interaction of the approach flow at submerged weirs in a relatively deep flow. It should be noted that
and the associated back flow. Downstream of the weir, a large the flow aspect ratio (flume width/flow depth) at section M is
recirculation zone developed (Fig. 6). Immediately downstream of around 1.5, which means side wall effects also contribute to the
the weir, vortices can be clearly observed, indicated by a recirculat- formation of secondary flows in the scour hole and to the scour-
ing movement of sediment close to the weir. For other parts of this hole geometry. In a large river, the flow aspect ratio is larger, which
zone, the occasionally random movement of sediment particles is might increase the number of secondary flow cells and change the
seen at the equilibrium phase of scouring. At the rear of the recir- scour-hole geometry.
culation zone, the main flow reattaches to the bed, creating the
flow reattachment region (Fig. 6). Inside this region, the flow is
Bed Shear Stress
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

considerably turbulent, and the velocities are quite small. The


observed maximum scour-depth point (1.06 m away from the weir) At the equilibrium stage of clear-water scour, the stability of sedi-
is located close to the end of this region. When the flow passes the ment particles in the scour hole is based on the equilibrium con-
maximum scour-depth cross section, no obvious reverse velocities ditions of the forces acting on them, which can be simplified as a
can be seen on the centerline longitudinal section, and a relatively balance of flow drag force FD , lift force FL , and submerged weight
uniform flow is redeveloped downstream of the end of the of sediment particles FG . Accordingly, the critical shear stress c0 of
scour hole. sediment particles resting on a bed, sloping in the streamwise
The time-averaged velocity vector distributions in three trans- direction, can be defined by the following equation:
verse cross sections (here taken as sections U, M, and D, respec-  
c0 tan
tively; see Fig. 5), which are located at x 0.50 m, 1.06 m, and cos 1 1
2.75 m, respectively, are presented in Fig. 7. These vectors are c tan
determined from the mean velocities of the transverse and vertical
where c0 = critical shear stress on a sloping bed; c = critical shear
velocity components. As indicated in Fig. 7, secondary flows de- stress on a horizontal bed (calculated as u2c 0.45 Pa); = bed
velop at all three cross sections. For the U and D cross sections [see slope (measured from an horizontal datum); and = submerged
Figs. 7(a and c)], especially for the former one, the magnitude of angle of repose of sediment (taken as 36, as measured in this
the velocity vectors is relatively small compared with those in the study). A complete analysis of incipient sediment motion on non-
maximum depth cross section M [Fig. 7(b)]. This is driven by horizontal slopes can be found in Chiew and Parker (1994). Inside
cross-sectional anisotropic turbulence (Prandtl 1952); the strong the equilibrium scour hole, reversed velocity vectors are observed
secondary flows observed in the maximum depth cross section M on the upstream slope (Fig. 6).
[Fig. 7(b)], being categorized as Prandtl second kind (driven by According to past research (Kim et al. 2000; Biron et al. 2004;
turbulence). The secondary flows are characterized by paired cir- Pope et al. 2006), four common methods can be used for the es-
cular flow cells, which are quasi-symmetrically located at both timation of bed shear stress with experimental data: (1) reach aver-
sides of the centerline sand ridge. A similar flow pattern can be age method, (2) current velocity profiles (law of the wall, or log
seen in Fig. 7(c); this figure also shows nonsymmetry in flows, this profile method), (3) Reynolds stress measurement, and (4) TKE
being related to the nonsymmetrical bed surface. The pattern of (turbulence kinetic energy) method. The assumptions, suitability,
cellular secondary flows and the associated observed sand ridge and limitations of these four methods have been critically reviewed
in this research is consistent with the observations and theory of by Kim et al. (2000) and Biron et al. (2004). Considering the appli-
Nezu and Nakagawa (1984) and Nezu et al. (1988). These secon- cability of these four methods, two of them are employed in this
dary flows have a significant effect upon the development of study, namely, the Reynolds shear stress measurement and the TKE
the scour hole and the final bed geometry. They also account for method. The Reynolds shear stresses, x;z , are defined as u 0 w 0 .
the formation of the centerline sand ridge and help to explain the The turbulent kinetic energy density, E, is calculated from
deepest point of scour hole being found close to the side wall, rather
1
than on the centerline of the flume. The ratio of the maximum scour E u02 v02 w02 2
depth on the centerline to the maximum scour depth in the scour 2

150 150
(a) 0.02 m/s (b) 0.02 m/s
100
Vertical distance (mm)
Vertical distance (mm)

100

50 50

0 0
150
(c) 0.02 m/s -50
100
-100
50
-150
-200 -100 0 100 200 -200 -100 0 100 200
Transversal distance (mm) Transversal distance (mm)

Fig. 7. Velocity vectors distribution in the U, M, and D cross sections

JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014 / 71

J. Hydraul. Eng., 2014, 140(1): 68-76


where u 0 , v 0 , w 0 are fluctuation velocity components along frequent sediment recirculating movements were observed during
the downstream, transverse and vertical directions, respectively. the experiment, including at equilibrium, but they did not result
A simple relationship between TKE and bed shear stress has been in deepening of the scour hole. Elsewhere within the scour hole,
formulated as 0 CE (Soulsby 1981), where 0 is bed shear measured Reynolds shear stresses are around or below the thresh-
stress and C is an empirical coefficient. The empirical factor C old and calculated bed shear stresses are slightly greater than the
was found to be 0.20 (Soulsby 1981), while 0.19 has been adopted threshold values. Considering the form of Eq. (2), the TKE method
by others (Stapleton and Huntley 1995; Thompson et al. 2003; takes into account the three-dimensional velocity fluctuations, thus
Pope et al. 2006) and has been found to apply to a complex flow it is less applicable when used for two-dimensional estimation,
fields (Biron et al. 2004). Therefore, C 0.19 has been used in especially when secondary flows exist. This may account for the
this study. The critical question of obtaining bed shear stress esti- overestimation in the scour hole. On the downstream slope of the
mates using single point measurements is how to determine the scour hole, near-bed velocity accelerates as flow depth decreases,
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

appropriate measurement height above the bed. According to the which causes a reduction to the velocity gradient. As a result, the
recommendation of Biron et al. (2004), the best option for using measured Reynolds shear stresses are very close to zero and are
single-point measurements to estimate bed shear stress is to posi- below the threshold values.
tion the instrument at around 10% of the flow depth. Then, it is
above the thickness of the roughness layer and is less affected
Turbulence Characteristics
by unexpected increases in SNR or Doppler noise that may occur
closer to the bed (Finelli et al. 1999; Kim et al. 2000). The mea-
sured points for estimating bed shear stresses in this study were Turbulence Intensities
all set at 10 mm above the bed. The measured and calculated values Contours of the turbulence intensity distributions for the longitu-
for these near-bed points were used for direct estimation of bed dinal direction and three transverse cross sections (U, M, and D)
shear stress. are presented in Figs. 9 and 10, respectively. The contour values
Threshold bed stresses, measured Reynolds shear stresses, and for the downstream, transverse and vertical directions are calcu-
calculated bed stresses were obtained, and a comparison of the ex- lated from
perimental bed shear stresses and local threshold bed shear stresses q q q
obtained from Eq. (1) is presented in Fig. 8. It should be noted that u 0 2 v 0 2 w 0 2
the values of critical bed shear stresses on the upstream slope of the TI u ; TI v ; TI w 3
U0 U0 U0
scour hole are negative, which corresponds to the direction of the
bottom reverse flow, while in Fig. 8 only absolute values are used where U 0 is the average approach flow velocity. For the centerline
for comparison. It can be seen that for the approach flow and near longitudinal section, the turbulence intensities for all three direc-
the end of the scour hole, bed stresses obtained from the observed tions show a very similar distribution [Figs. 9(ac)]. More specifi-
Reynolds shear stresses and from the TKE method do not exceed cally, the values of TI u , TI v , and TI w upstream of the weir are
the threshold. This is consistent with the experiment being con- rather small compared with those in the scour hole. The peak values
ducted under conditions of no general sediment transport. Although are found immediately downstream of the weir and above the origi-
reverse flows are observed on the upstream slope of the scour hole, nal bed level. Downstream of the locations of the peak values, the
the values of measured Reynolds shear stresses are still positive in turbulence intensities are damped, as the distance from the weir
this region, which is consistent with the velocity gradients still increases. It is important to note that the positions where the peak
being positive close to the bed (Figs. 6 and 8). For this case, values of turbulence intensities occur are found at the upstream end
estimation of bed shear stress from near bed Reynolds shear stress of the recirculation zone. Furthermore, the measurements of the
measurements may be unreliable, because the measurement equip- turbulence intensities in the centerline longitudinal section show
ment is incapable of acquiring data in the negative velocity gradient the turbulent flow to be anisotropic, with u 0 1.2v 0 1.7w 0 .
layer, which is very thin (less than 10 mm) and just above the bed. Fig. 10 shows turbulence intensity distributions in three trans-
At the upstream end of the recirculation zone, the experimental bed versal cross sections (U, M, and D). For section U [Figs. 10(ac)],
stresses considerably exceed the absolute threshold values, which although the turbulence intensities are very small compared with
is in agreement with our experimental observations. In this area, those in the scour hole, a trend can be observed. The areas of very

200
Flow
(Pa)

1.5
Vertical distance (mm)

100 1
Bed Shear stress

0.5

0 0
Original Bed Line
-0.5

Observed Reynolds Shear Stress


100 Scour Hole
Calculated Bed Shear Stress (TKE)
Threshold Bed Shear Stress

-200
-0.5 0 0.5 1 1.5 2 2.5
Longitudinal distance (m)

Fig. 8. A comparison of estimated shear stress and threshold shear stress along the centerline upstream and downstream of the submerged weir

72 / JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014

J. Hydraul. Eng., 2014, 140(1): 68-76


200
Vertical distance (mm)
100
0.07 0.15 0.4
0.45

5
0.2
0.1 0.4 0.3 5

0.1
0 0.2
0.3 0.3
0.25 0.25
0.2
-100
(a)
200
Vertical distance (mm)

100 0.06 0.25


Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

0.15 0.3 0.15


0.08 0.4 0.35 0.2
0 0.3
0.25 0.25 0.25
0.2
-100 5 0.2
(b)
200
Vertical distance (mm)

100
0.04 0 5 0.2 0.15 0.08
0. 0. 3 0.25 0.2
0.04 0.25 0.08
0 0.2
0.2 0.15 0.1
-100 0.15
(c)
-0.5 0 0.5 1 1.5 2 2.5
Longitudinal distance (m)

Fig. 9. Turbulence intensity distributions in the centreline longitudinal section: (a) downstream; (b) transverse; (c) vertical directions, respectively

TIu TIv TIw


Vertical distance (mm)

150
(a) (b) (c)

Section U
100 0.06 0.0
5

0.07
0.03

0.07 0.07 0.07 0.07 5 0.043


4
0.06 0 . 03

0.04
0.08 0.08 0.08 0 .0 7 0.035
50 8
9 0.09 0.0 0.0
8 0.09 0 .0 3
0.0 0.1 9 0.08 0.08 0.0 0.04 0.04 0.035
0.11
0.1 0.09 0.09 8 0.038
0
150
(d) (e) (f)
Vertical distance (mm)

100 0.18
0.36 0.16
0.26
6

0.26 0.2 0.18


0.3

0.22
50 0.34 0.28

Section M
0.28 0.2
0.34 0.28
32
0.

0
28 0.26
0. 0.1
0.2
0.24 6
0.

-50
18

0.2 0.3 0.28


6
0.24 0.16 0.18
0.22 26 0 .2 4
-100 0.
0.

0.18 0.12 0.16


24

0.18 0.22 0.1


Vertical distance (mm)

-150
150
(g) (h) (i)
Section D

100 85 0.1 0.1 0.12


0.0 0.1 0.1 0.15 0.05
0.1

0.15 0.085 0.12 0.1 0.06


0.1 0.1 0.1 0.05
50 2 0.1 0.12 0. 1 0.18 0.05 0.0 05
0. 0.04
0 .1 0.15 0.1 2 0.1
0.18 2 0.1 0.04 0.035 4

0
-150 -100 -50 0 50 100 150 -150 -100 -50 0 50 100 150 -150 -100 -50 0 50 100 150
Transversal distance (mm) Transversal distance (mm) Transversal distance (mm)

Fig. 10. Turbulence intensity distributions in the U, M, and D cross sections

low turbulence intensities correspond with areas of high streamwise the bed decreases. The decrease of turbulence intensities closer to
velocities. The same can be observed in straight natural rivers. For the bed, which is in line with the dissipating trend of upstream tur-
section M [Figs. 10(df)], turbulence intensities are highest around bulence intensities, is caused by the damping effect of bed boun-
5 cm above the original bed level and reduce as the distance from daries. With respect to section D [see Figs. 10(gi)], turbulence

JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014 / 73

J. Hydraul. Eng., 2014, 140(1): 68-76


200

Vertical distance (mm) 100


0.006 0.15 0.02
0.22
0.01 0.2 0.12 0.1 8 0.04
0.15 0.0
0 0.06
0.1
0.06 0.08
-100 0.06

-0.5 0 0.5 1 1.5 2 2.5


Longitudinal distance (m)

Fig. 11. TKE distribution in the centerline longitudinal section


Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

200
Vertical distance (mm)

100 0.3 0.3 10 10 1 .5 0.5 0.3


24 20 15
0.5
10 5 3
0 10
5 1.5 0.5
3 5 3
3
-100 3 0.3

-0.5 0 0.5 1 1.5 2 2.5


Longitudinal distance (m)

Fig. 12. Normalized Reynolds shear stress distribution in the centerline longitudinal section

uw uv vw
Vertical distance (mm)

150
(a) (b) (c)

Section U
100
0.08 0.2 0
-0.1

0. 0.06

0
0. 0.08 0
-0.1
0
0.1

-0.3

0.3
1

03
0

0.5 0.7 0.03 6


-0.
0 .3

-0.05
0.5 0 0
0.3

50 -0.3 0 0.0
0.5

0.7 0.1 0.0 0.03 -0.08


0.7 3
0.8 0.5 0.3 0
0

-0.05
0
150
(d) (e) (f)
100
Vertical distance (mm)

6 10 10 0 1 -2
1
12
2

8 2 2 0
0

0
-2

14
-2

50 12 10
0
2

12 0
Section M
0
6

12 2
2

10 0
6
4

10 8 1
0

0 -1
10
8 -1 0
0

8 -6 10 2 3 1
-50 -4
6 -2 0 1 1
6 8 2
2

-1

4 2 4 0
-100 4 6 4 0
2 -2
1

4
Vertical distance (mm)

-150
150
(g) (h) (i)
Section D

100
2 0
0. 0
0.

0.6
0

0.2
0.2

0.5
15

1
0.2

-0.15

0.6
11.5
0.6 0.2 0 -0.15
0

50 0.8 0.5
0

1 0.8 0
0

1 -0.5 0.3 -0.3

0
-150 -100 -50 0 50 100 150 -150 -100 -50 0 50 100 150 -150 -100 -50 0 50 100 150
Transversal distance (mm) Transversal distance (mm) Transversal distance (mm)

Fig. 13. Normalized Reynolds shear stresses distributions in the U, M, and D cross sections

74 / JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014

J. Hydraul. Eng., 2014, 140(1): 68-76


intensities in all directions are damped, resulting in considerably longitudinal direction, a recirculation zone and a flow reattachment
smaller values, less than those in the scour hole, while still exceed- region are developed. The turbulence structures at the upstream end
ing values observed in the upstream cross section U. The distribu- of the recirculation zone govern the dimensions of the scour hole,
tions of TI u , TI v , and TI w show a certain degree of irregularity, as indicated by the observed maximum turbulence intensities, TKE,
compared to the distributions in the upstream cross section. and Reynolds shear stresses on the upstream slope of the scour
The distribution of normalized TKE, which is calculated from hole. The location of maximum scour depth is found at the rear of
E=U 20 , in the centerline longitudinal section, is shown in Fig. 11, the flow reattachment region and close to the left flume glass wall.
and highlights a similar pattern to that observed for turbulence The observed Reynolds shear stress near the bed and the calculated
intensities. As previously reported (Bradshaw et al. 1967), TKE bed shear stresses from TKE method are larger than absolute values
reflects the energy extracted from the mean flow by the motion of of critical bed shear stresses immediately downstream of the weir,
the turbulent eddies. Thus it is possible to conclude that the strong- and smaller than critical bed shear stresses elsewhere in the scour
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

est and largest eddies are developed at the upstream end of the hole and further downstream.
recirculation zone. For the transverse direction, strongly paired cellular secondary
flows are observed in the scour hole. A certain degree of symmetry
Reynolds Shear Stress of Reynolds shear stress uv distributions at cross sections are
Figs. 12 and 13 present the distributions of normalized Reynolds observed, which directly account for the formation of secondary
shear stresses for the centerline longitudinal section and three trans- flows. These secondary flows have a significant effect upon the
verse cross sections (U, M, and D). The Reynolds stresses values development of the scour hole and the final bed geometry. Their
here are calculated from effect should be considered in the study of scour at low-head
structures in a relatively deep flow.
u 0 w 0 u 0 v 0 v 0 w 0
uw ; uv ; vw 4
u2 u2 u2
Acknowledgments
where u = average approach flow shear velocity. Fig. 12 shows
that the largest Reynolds shear stresses uw occur immediately The authors would like to thank China Scholarship Council (CSC)
downstream of the weir, with values dissipating in the scour for the financial support of this research.
hole and further downstream. As supported by the distribution of
TKE (Fig. 11) and the distribution of uw (Fig. 12) in the centerline
longitudinal section, it is possible to infer that the large magnitude Notation
of turbulence structure on the upstream slope of the scour hole
The following symbols are used in this paper:
governs the scour hole size (maximum scour depth and length).
C = empirical factor used in TKE method for calculating bed
This is in agreement with work undertaken by Ben Meftah and
shear stress;
Mossa (2006).
ds = maximum scour depth;
As seen in Figs. 13(a, d, and g), Reynolds shear stresses uw
dse = maximum scour depth at the equilibrium phase
and uv are dominant, while vw values are relatively small in
d50 = median diameter;
all three cross sections. It also can be seen that the highest Reynolds
E = turbulence kinetic energy density;
shear stress values are found near the bottom of the sections, for
FD = flow drag force exert on a sediment particle;
both U and D, while for section M they are observed around the
FG = submerged weight of a sediment particle;
original bed level. Thus bottom friction at sections U and D was the
FL = lift force on a sediment particle;
dominant factor to account for shear stress distributions, but for
g = gravity;
section M the distributions of Reynolds shear stresses are strongly
ls = maximum scour length;
dependent on upstream dissipating shear stresses.
As discussed above, secondary flows are observed at all TI u , TI v , TI w = turbulence intensities along the downstream,
three sections (U, M, and D; see Fig. 7). The values of uv in transverse, and vertical directions, respectively;
Figs. 13(b, e, and h) also reveal secondary flows. Negative uv t = scour time;
values are found on the left side of the flume centerline, while pos- U 0 = average approach flow velocity;
itive values are observed on the right side, as seen in Fig. 13(e), u, v, w = mean velocity components along the downstream,
showing a certain degree of symmetry. Similar patterns can be seen transverse, and vertical directions, respectively;
in Figs 13(b and h). Since the values of uv and vw should be zero u 0 , v 0 , w 0 = fluctuation velocity components along the downstream,
when no secondary flows exist, the Reynolds shear stress values not transverse, and vertical directions, respectively;
only reveal the concentration of turbulence, but also indicate the u = average approach flow shear velocity;
intensities of secondary flows. uc = average approach flow critical shear velocity;
y = approach flow depth;
yt = tail water depth;
Conclusions c = critical shear stress on a horizontal bed;
c0 = critical shear stress on a sloping bed;
The results of an experimental study of flow patterns, bed shear uw , uv , vw = normalized Reynolds shear stresses;
stresses, and turbulence structures in the approach flow towards 0 = bed shear stress;
a submerged weir, and the resulting scour hole are presented. = relative submerged particle density;
The experiments were undertaken in clear-water scour conditions = bed slope;
in a laboratory flume. The equilibrium scour-hole condition was = submerged angle of repose of sediment;
obtained. The three-dimensional flow-field data was obtained by = water density;
a Nortek Vectrino+ acoustic velocimeter. s = sediment density;
The results show that the presence of a submerged weir con- g = standard deviation; and
siderably changed the flow structure. Along the flume centerline = kinematic viscosity of fluid, considered as 1 106 m2 =s.

JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014 / 75

J. Hydraul. Eng., 2014, 140(1): 68-76


References Marion, A., et al. (2004). Effect of sill spacing and sediment size grading
Ben Meftah, M., and Mossa, M. (2006). Scour holes downstream of bed on scouring at grade-control structures. Earth Surf. Processes
sills in low-gradient channels. J. Hydraul. Res., 44(4), 497509. Landforms, 29(8), 983993.
Bhuiyan, F., et al. (2007). Hydraulic Evaluation of W-Weir for River Marion, A., et al. (2006). Sediment supply and local scouring at bed sills in
Restoration. J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(2007)133: high-gradient streams. Water Resour. Res., 42(6), W06416.
6(596), 596609. Martin, V., et al. (2002). ADV data analysis for turbulent flows: Low
Biron, P. M., et al. (2004). Comparing different methods of bed shear correlation problem. Proc., Hydraulic Measurements and Experimen-
stress estimates in simple and complex flow fields. Earth Surf. Proc- tal Methods 2002, ASCE, CO, 110.
esses Landforms, 29(11), 14031415. Melville, B. W. (1997). Pier and abutment scour: Integrated approach.
Bormann, N. E., and Julien, P. Y. (1991). Scour downstream of grade- J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(1997)123:2(125),
control structures. J. Hydraul. Eng., 10.1061/(ASCE)0733-9429 125136.
(1991)117:5(579), 579594. Melville, B., and Chiew, Y. (1999). Time scale for local scour at bridge
Downloaded from ascelibrary.org by National Institute Technology - Silchar on 11/04/16. Copyright ASCE. For personal use only; all rights reserved.

Bradshaw, P., et al. (1967). Calculation of boundary-layer development piers. J. Hydraul. Eng., 125(1), 5965.
using the turbulent energy equation. J. Fluid Mech., 28(03), 593616. Nezu, I., and Nakagawa, H. (1984). Cellular secondary currents in straight
Chen, Z., et al. (2005). Experimental study on the upstream water level conduit. J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(1984)110:2
rise and downstream scour length of a submerged dam. J. Hydraul. (173), 173193.
Res., 43(6), 703709. Nezu, I., Nakagawa, H., and Kawashima, N. (1988). Cellular secondary
Chiew, Y.-M., and Parker, G. (1994). Incipient sediment motion on non- currents and sand ribbons in fluvial channel flows. Proc., 6th APD-
horizontal slopes. J. Hydraul. Res., 32(5), 649660. IAHR Congress, APD-IAHR, Kyoto, Japan, 2(1), 5158.
DAgostino, V., and Ferro, V. (2004). Scour on alluvial bed downstream of Pope, N. D., et al. (2006). Estimation of bed shear stress using the turbu-
grade-control structures. J. Hydraul. Eng., 10.1061/(ASCE)0733-9429
lent kinetic energy approachA comparison of annular flume and field
(2004)130:1(24), 2437.
data. Cont. Shelf Res., 26(8), 959970.
Dey, S., et al. (2011). Near-bed turbulence characteristics at the entrain-
Prandtl, L. (1952). Essentials of fluid dynamics, Blackie and Son,
ment threshold of sediment beds. J. Hydraul. Eng., 10.1061/(ASCE)
London.
HY.1943-7900.0000396, 945958.
Finelli, C. M., et al. (1999). Evaluating the spatial resolution of an acoustic Soulsby, R. L. (1981). Measurements of the Reynolds stress components
doppler velocimeter and the consequences for measuring near-bed close to a marine sand bank. Marine Geol., 42(14), 3547.
flows. Limnol. Oceanogr., 44(7), 17931801. Stapleton, K. R., and Huntley, D. A. (1995). Seabed stress determinations
Friedrich, H., et al. (2005). Three-dimensional measurement of laboratory using the inertial dissipation method and the turbulent kinetic energy
submerged bed forms using moving probes. Proc., XXXI IAHR method. Earth Surf. Processes Landforms, 20(9), 807815.
Congress, Korea Water Resources Association, Seoul, 396404. Thompson, C. E. L., et al. (2003). The manifestation of fluid-transmitted
Goring, D. G., and Nikora, V. I. (2002). Despiking acoustic doppler bed shear stress in a smooth annular flumeA comparison of meth-
velocimeter data. J. Hydraul. Eng., 10.1061/(ASCE)0733-9429(2002) ods. J. Coastal Res., 19(4), 10941103.
128:1(117), 117126. Wahl, T. L. (2000). Analyzing ADV data using WinADV. Proc., 2000
Kim, S. C., et al. (2000). Estimating bottom stress in tidal boundary layer Joint Conf. on Water Resources Engineering and Water Resources
from acoustic doppler velocimeter data. J. Hydraul Eng., 126(6), Planning and Management, ASCE, Reston, VA.
399406. Yalin, M. S. (1992). River mechanics, Pergamon Press, Oxford, U.K.

76 / JOURNAL OF HYDRAULIC ENGINEERING ASCE / JANUARY 2014

J. Hydraul. Eng., 2014, 140(1): 68-76

You might also like