You are on page 1of 35

Environ Fluid Mech (2016) 16:659693

DOI 10.1007/s10652-016-9449-0

ORIGINAL ARTICLE

Variations of bed elevations due to turbulence around


submerged cylinder in sand beds

K. Sarkar1 C. Chakraborty2 B. S. Mazumder1,3

Received: 15 January 2015 / Accepted: 25 March 2016 / Published online: 8 April 2016
 Springer Science+Business Media Dordrecht 2016

Abstract This paper presents the spatio-temporal variations in bed elevations and the
near-bed turbulence statistics over the deformed bed generated around the submerged
cylindrical piers embedded vertically on loose sediment bed at a constant flow discharge.
Experiments were carried out in a laboratory flume for three blockage ratios in the range of
0.040.06 using three different sizes of submerged cylinders individually placed vertically
at the centerline of the flume. Clear-water experimental conditions were maintained over
the smooth sediment bed surface with a constant flow discharge (Q 0:015 m3 =sec),
thereby giving three different cylinder Reynolds numbers ReDc UmmDc (=10200, 12750,
15300) away from the cylinder locations, where Um is the maximum mean velocity, Dc is
the cylinder diameter and m is the kinematic viscosity of fluid. Instantaneous sand bed
elevations around the cylinders were recorded using a SeaTek 5MHz ultrasonic ranging
system of net 24 transducers to estimate bed form migration, and the near-bed velocity data
at transducer locations over the stable deformed bed around the pier-like structures were
collected using down-looking three-dimensional (3D) Micro-acoustic Doppler velocimeter
to estimate the bottom Reynolds shear stresses and the contributions of bursting events to
the dominant shear stress component. The flow perturbation generated due to relatively
lower flow blockage ratio favored to achieve the stable bed condition more rapidly than the
others, and larger upstream scour-depth and deformed areas were noticed for greater flow

& B. S. Mazumder
bijoy@isical.ac.in
K. Sarkar
soundofphysiks@gmail.com
C. Chakraborty
chandan@isical.ac.in
1
Fluvial Mechanics Laboratory, Physics and Applied Mathematics Unit, Indian Statistical Institute,
Kolkata 700 108, India
2
Geological Studies Unit, Indian Statistical Institute, Kolkata 700 108, India
3
Present Address: Department of Applied Mechanics, Indian Institute of Engineering Science and
Technology (IIEST), Shibpur, Howrah 711103, India

123
660 Environ Fluid Mech (2016) 16:659693

blockage ratio due to larger cylinder diameter. For larger blockage ratio in the upstream of
scour-hole near the bed, occurrences of probabilities of both boundary-ward interactions
(Q1 and Q3) were the dominant; whereas in the downstream of the scoured region,
occurrences of probabilities of second and third quadrant events (Q2 and Q4) were
dominant. On the other hand, for the lower blockage ratio, quadrant (Q2) was dominant
over Q4 in the downstream of scour-hole, and in the upstream of scour-hole, quadrant Q4
was the dominant.

Keywords Bed elevation  Cylindrical pier  Blockage ratio  Exceedance probability 


Bottom shear stress  Spectral analysis  Quadrant analysis

1 Introduction

Local scour is a natural phenomenon caused by the erosive action of flowing water over
sand beds in rivers and streams. An immovable obstruction like a bridge pier restricts the
flow area, leading to an increase in velocity around the sides accompanied by various kinds
of vortices, such as, horseshoe vortices, wake vortices, and trailing vortices developed at
the top of the obstructions for the submerged case. When the scoured region develops at
the upstream of obstruction, the scoured bed material may be deposited downstream of the
obstruction.
Controlling the size of scour-hole is of prime importance to assure bridge safety and to
bind the variations of river morphology [16, 65]. Local scour around such engineering
constructions changes the morphology of the river/sea bed and affects aquatic ecosystems
and fish habitats quite sensitively [4, 5, 8, 73]. In fact, a lot of fish eggs and infant fishes
can easily be found in a scour hole, especially located near a shore line. Some kind of
fishes such as Biwia zezera can only be seen and regenerating in a scour hole in the
downstream of the Kizu River [76, 77]. Hence the suitable knowledge about the mecha-
nism of scouring process and the hydrodynamics of flow structures around the bridge piers
in rivers is of great importance to produce a preferable habitat by utilizing the interaction
between the structure and hydro-morphological processes.
Extensive research work has been performed on scour at sand-embedded piers and bed
sills to estimate the temporal variations of pier scour depth [8, 18, 24, 44, 46, 49], to
quantify numerically the maximum scour-depth at bridge piers in alluvial sand beds [3, 33,
53, 69, 73] and to examine the dependence of scouring on different shape factors like
circular, rectangular, square shaped objects [45, 50, 58]. Ballio and Radice [7] used
underwater fibre optics for continuous measurement of local scour, and it was successfully
used for bottom profiling during mobile bed. Link et al. [37] performed experiments using
a laser distance sensor (LDS) to study the spatio-temporal variation of geometric properties
of scour-holes due to 0.2 m diameter circular cylinder embedded on a sand bed. Muzza-
mmil and Gangadhariah [46] developed an expression for the maximum equilibrium scour
depth from the vortex velocity distribution inside the scour-hole, which gave a better
results compared to the results of well-known models. Gaudio and Marion [24] and
Dargahi [18] proposed different time scales for the simulation of the temporal variations of
the maximum scour depth. A 3-D numerical FLUENT model was used by Salaheldin et al.
[53] to simulate the separated flow around vertical circular piers of various size, shape and
dimension in clear water. Their computations were performed using different turbulence

123
Environ Fluid Mech (2016) 16:659693 661

models to study the scour initiation processes around the piers. Zhao and Fernando [80]
employed an Eulerian two-phase model using FLUENT software to simulate scour around
pipelines. The aim was to evaluate the model efficiency using available benchmark data
and to use the model to find important information on flow dynamics. The flow-particle and
particle-particle interactions were considered in the model formulation. Large-eddy sim-
ulation (LES) and laboratory observations were performed by Kirkil et al. [33] to examine
the coherent structures in the flow field around a circular cylinder. Tregnaghi et al. [67, 68]
investigated the effect of flash floods on the scour at sequence sills with symmetrically
triangular-shaped and long flood recession hydrographs, respectively. They provided
methodology to predict the final scour depth developed downstream of a sill after the
occurrence of a flood. Numerical simulations were performed by Tregnaghi et al. [69] to
predict the local scour under time varying flow conditions. The simulation was based on
the characteristics of the flow hydrograph, temporal evolution of scour and the geometry of
the scour hole. Their approach was general and could be applied to a wide range of scour
environments (e.g. bed sills, step-pool systems). Catano-Lopera et al. [13] investigated
experimentally as well as numerically the flow characteristics around two partially buried
objectsone is a short cylinder and other one a truncated cone.
Scour around submerged bridge piers is considerably important in the field of offshore
and coastal engineering constructions like submerged cylindrical breakwaters (Cheng et al.
[15]) used for coastal protection in trapping of water waves in riverine environment. It may
be noted that the unfinished bridge pier constructions, or pier-like structures especially
constructed for the ecological purpose would act as submerged object in flood situation.
The problems of scour around and subsequent burial of short cylinders placed on a sandy
bed under progressive waves were studied by Voropayev et al. [72]. The motivation of
their study was to observe possible scour and burial effects of cylindrical objects in coastal
environments, where non-linear progressive waves played a significant role. The geometry
of the scour hole and the flow structure around short cylinders under the action of wave
alone and combined wave and current was experimentally studied by Catano-Lopera and
Garcia [11]. Their results showed that both width and length of the scour hole due to the
short cylinder primarily depends on the KeuleganCarpenter number KC Um Tw =Dc
and the aspect ratio (L/Dc), where Um is the maximum velocity, Tw is the wave period, Dc
is the cylinder diameter and L is the length of the cylinder. Their study was aimed at
understanding the dynamics of pipelines and short cylinders placed on a sandy bed under
the action of oscillatory flows, which occurred in shallow water regions of coastal area.
Dey et al. [20] introduced a submerged factor in determining the scour depth and its
relation to the horseshoe vortex circulations. The shape of the scour-hole in case of
submerged cylinder is similar to that in the emerged condition. Thus the maximum scour
depth appears at the upstream of pier and as the height of the pier decreases, i.e. sub-
mergence ratio decreases, the maximum scour-depth reduces. Young and Testik [78]
provided a scaling analysis of the effects of submerged vertical and semicircular break-
waters on the scour development, which accounts for sediment scour and wave field
characteristics. In their work, the scour was classified into two regimes: one was attached
scour when the Keulegan-Carpenter number KC value was less than or equal to p, while the
detached scour when the KC value was greater than p. Attached scour occurred when the
scour hole was connected to the onshore breakwater face and the detached scour occurred
when the scour hole was not connected to the onshore face of the breakwater. Subse-
quently, Young and Testik [79] investigated in a laboratory tank the effects of submerged
vertical and semicircular breakwaters on the local wave field to estimate the wave
reflection coefficient. Catano-Lopera et al. [12] investigated the scour and burial of conical

123
662 Environ Fluid Mech (2016) 16:659693

frustums placed on a sandy bed under waves alone (WA) and combined flows (CF)
conditions. The observations indicated that equilibrium burial depth was smaller than
burial of other objects such as short cylinders lying on sand bed under equivalent
hydrodynamic conditions. Euler and Herget [21, 22] developed a simple analytical
approach based on the obstacle Reynolds and Froude numbers to determine the local scour
around submerged obstacles with various submerged factors in the fluvial environment.
Maity and Mazumder [41] described the turbulence statistics of flow above and within the
crescentic scour-hole generated at the upstream of a horizontal short cylinder placed over
the sandy bed transverse to the flow. Recently, Sarkar et al. [52] studied the space-time
dynamics of scour-hole structures around circular and non-circular submerged cylinders,
like circular, elliptical, square and triangular, associated with turbulence properties. It was
interesting to note that the upstream scour depth was maximum for circular and minimum
for elliptical cylinder; and side-wise scouring rate was also significant for the elliptical one.
In spite of extensive experimental works on local scour around piers, only few inves-
tigations have been made to understand the coherent structures of turbulence and in par-
ticular the bursting phenomena of turbulent flow around pier-like obstacles [28, 29, 31, 36].
Izadinia et al. [29] investigated the turbulent flow and sediment entrainment around a
bridge pier, and determined the contributions of each quadrant to the Reynolds shear stress.
Link et al. [36] recently investigated in the laboratory the coherent structures and sediment
particle motion around a cylindrical pier to register the scour hole structure in a fine sand
bed. Termini [62], Termini and Lo [63] and Termini and Sammartano [64] investigated the
relation between the horizontal turbulent intensity and the bed form generation with the
influence of turbulence structures and bursting events. Moreover, investigations have been
made on turbulence characteristics of flow within the scour hole obtained with different
geometrical schemes [27, 65]. However, these studies were not concerned with the evo-
lution of bed topography of different scales and geometries around the submerged vertical
piers of variable cylindrical diameters, and the associated turbulence. The implications of
horizontal turbulent intensity on the scour geometry around submerged cylinders have a
significant importance in high flood situations, when water flows over unfinished bridge
pier constructions, flow over a caisson placed underwater at the bottom around the piers for
scour protection etc. In fact, any pier like objects, placed near the river bank would act as a
submerged object in flood situation.
The objective of the present paper is to study the spatial and temporal changes of bed
form structure around a submerged cylindrical pier of different diameters embedded ver-
tically in a sand bed and the associated near-bed bursting phenomena at a constant flow
discharge. More precisely, an attempt has been made to examine the stochastic nature of
bed form features, the susceptible regions for sediment entrainment and deposition due to
the near-bed bursting events around the cylindrical pier; and how these are linked with the
geomorphic properties of scour holes. The novelty of this work is to experimentally
investigate the space-temporal evolution of bed topography with different scales and
geometry around a submerged pier of variable sizes using ultrasonic ranging system
(URS). In addition, the power spectra of velocity components and the contributions of
bursting events to the bottom Reynolds shear stresses relevant to the stable bed form
features are investigated. The results described in this paper are of great interest to the civil
engineers, and earth and environmental scientists, who consider the dynamics of bed form
migration around the bridge piers embedded on river-beds, sea-beds and shallow water
regions of coastal areas [60, 61] and also who examine the formations of bedform elevation
features in a recent stream to determine palaeo-hydraulic conditions [34, 55, 56].

123
Environ Fluid Mech (2016) 16:659693 663

2 Experimentation

2.1 Test channel

Experiments are conducted in a re-circulating closed circuit laboratory flume (Mazumder


et al. [42]; Mazumder et al. [43]) specially designed at the Fluvial Mechanics Laboratory
(FML) of the Physics and Earth Sciences Division, Indian Statistical Institute, Calcutta.
Both experimental and re-circulating channels of the flume are of same dimension (10 m
length  0.50 m width  0.50 m height), which looks like elliptical in shape (Fig. 1ab).
The experimental walls of the flume are made of Perspex windows of a length of 8 m,
affording clear view of the flow. Two non-clogging types of centrifugal pumps providing
the discharge are located outside of its main body of the flume. The intake and outlet pipes
are freely suspended from the overhead structure to allow tilting the flume. The outlet pipes
(Pump-1 and 2) are fitted with by-pass pipes and valves, so that the discharges are adjusted
for the desired velocity. The electromagnetic discharge meters are fitted with the outlet
pipes to facilitate its continuous monitoring of the flow. The upstream bend of the channel
is divided into three sub-channels of equal dimensions. Moreover, two honeycomb cages
are placed at the back end of the sub-channels in front of the jets of high velocities coming

Fig. 1 Schematic diagram of a plan view and b front view of the flume (after Fig.1 of Mazumder et al. [42])

123
664 Environ Fluid Mech (2016) 16:659693

out from the outlets, and one honeycomb cage is placed at the other end of the sub-
channels in order to ensure the laminarize and vortex free flow of water through the
experimental channel. The positions of honeycomb cages in front of the jets and at the
upstream end of the experimental channel substantiate the curvature effect free flow from
the measurements.

2.2 Experimental set-up and measurement techniques

To perform the test, a loose sand bed of thickness h0 (= 5.5 cm) and 7 m long covering the
entire 50 cm width of the flume was laid down at the bottom surface. The sand had a
median particle diameter d50 of 0.25 mm, a standard geometric deviation rg of 0.685 and a
specific gravity of 2.65. Water depth was kept constant at d = 25 cm for all experiments.
The effective water depth was h = d  h0 = 19.5 cm. The flow discharge (Q = 0.015
m3 =sec) was chosen in such a way that the dimensionless bottom shear stress was below
the threshold value for the initiation of sediment movement at the undisturbed plane sand
bed of grain size d50 = 0.25 mm, i.e. when there was no sand transport at the bed. For the
present experiments, only the pump-2 was operated for the low water discharge in the
flume. The vertical velocity data were collected at three different horizontal locations at the
centreline of the flume using 3-D micro-acoustics Doppler velocimeter (ADV) developed
by SonTek, for 180 sec at a sampling rate 40 Hz to ensure the uniform laminarized flow
over the plane bed surface. The sampling volume of ADV was located 5 cm below the
transmitter probe and the precise distance depends on the individual probe geometry. The
velocity data were collected at several vertical positions starting from the lowest level
0.50 cm above the bed surface and the highest level at about 13.5 cm for each profile for
the flow Reynolds number Red Umm d  63700 and the Froude number Fr p Um 0:163,
gd
where Um (= 25.5 cm/s) is the maximum fluid velocity and g is the acceleration due to
gravity. Special care was taken to collect data with correlation [70 % and signal to noise
ratio (SNR)[15. However, very near to the bed, slight deviation (5 %) in the correlation
was observed. The bottom shear stress is characterized by the dimensionless Shields
parameter s0  and can be approximated as
s0
s0  1
qs  qgd50
where s0 qghR s with hR is the hydraulic radius, s the hydraulic slope of the flume is of
order 0.0001, q the fluid density, qs the sediment density and g is the acceleration due to
gravity. No measurements near the free surface were made in the present study. The ADV
data were processed to remove noise using a phase space threshold de-spiking technique
described by Goring and Nikora [25]. The velocity data were analyzed for all three hor-
izontal locations and found almost no change amongst the results of those velocity data.
The dimensionless stream-wise mean velocity profile u over the plane bed surface
exhibited a log-law up to flow depth about 13 cm from the sand bed (Fig. 2).
According to Long et al. [39], Tachie et al. [66], Yang et al. [75] and recently Absi [1], a
two-dimensional flow in the central portion of the flume was achieved with width/depth
ratio ( 2). Since in the present work the width/depth ratio was  3:7 with the occurrence
of maximum velocity at the flow depth 13.5 cm, it could be ascertained the two-di-
mensional flow at the central line of the flume, and hence the effect of secondary currents
on the flow is negligible. Therefore, it was ensured that the flow was free from the 3-D

123
Environ Fluid Mech (2016) 16:659693 665

Fig. 2 Mean velocity profiles


over plane sand bed

effect at the experimental channel. The quantitative evidence of lateral and vertical mean
velocity profiles (v) and (w) throughout the depth is shown in Fig. 2.
Here the velocity data at different frequency cutoffs (e.g. 20, 10, 5 and 2 Hz) were
filtered to calculate the spectra. The noise at high frequencies created an aliasing effect
from frequencies greater than the Nyquist frequency (here, fn 20 Hz) that were folded
into the lower frequencies. By applying a low-pass filter the data at 20 Hz (or less), this
noise was removed. Actually, the accuracy of the individual time series velocity data of all
three components was verified by observing the slope of the curve in the inertial sub-range
of the energy density plot against frequency in loglog scale. For the present study, the
energy density versus frequency plot of de-spiked and low-pass filtered velocity data of all
components with different cutoff frequencies are shown in Fig. 3ac. A clear 5/3 slope at
the inertial sub-range is observed from the figure. Hence, it is confirmed the accuracy of the
measurement of velocity samples at three directions i.e. u, v and w velocity data.
A series of experiments was conducted at the laboratory with varying blockage ratio to
the flow caused by the rigid circular cylindrical pier-like obstacles of different diameters
embedded vertically in the sand bed at the flume centreline. The blockage ratio was defined
as the ratio between the area of flow blocked by the cylinder and the total flow area in the
undisturbed channel. Overall three different blockage ratios were used in a range of
0.040.06 using cylinder diameters of 4, 5 and 6 cm and a common height 15 cm with
submergence ratio of 0.6. The selected blockage ratios were similar to the actual prototype
conditions as compiled in hydrologic investigations (USGS Atlases [71]) and laboratory
measurements of Atabay and Knight [6] and Seckin et al. [54]. Moreover, the cylinder
Reynolds number was defined by ReDc UmmDc , where Um is the maximum mean velocity,
Dc is the cylinder diameter and m is the kinematic viscosity of fluid. The values of ReDc are
10200, 12750 and 15300 for respective cylinder diameters Dc = 4, 5, 6 cm. The effective
cylinder height was 9.5 cm at the initial time for each of the cases. For each experiment, a
single cylinder was mounted vertically at a distance 6 m downstream from the channel

123
666 Environ Fluid Mech (2016) 16:659693

Fig. 3 Energy density versus frequency plot of velocity data for a u, b v and c w velocity

inlet. Henceforth, three different blockage ratios are referred to as the cylinder diameters of
4, 5 and 6 cm respectively throughout the paper.
For all three cases, the flow discharge was identical. The investigations were made on
the scour geometry at the vicinity and the bed form elevations around the different cylinder
diameters embedded on the sediment bed. Though the flow discharge (Q = 0.015 m3/sec)
was not enough to reach the threshold value of the bottom shear stress on the plane bed
surface, at the vicinity of the cylinder it was sufficiently strong to cause sediment motion.
No bed forms, such as, ripples and dunes were observed in the stream-wise direction on the
sediment bed surface except the occurrence of some local deformed area (scour and
deposition) in the vicinity around the submerged piers. The discharge was kept undisturbed
for a longer time to develop a nearly stable scour structure around the pier.
During the process of bed evolution around the cylindrical piers, instantaneous mea-
surements of the sand bed elevations (z) were made using a SeaTek 5 MHz URS composed
of net 24 transducers. The URS was mounted on an aluminum trolley just submerged on
the free surface level over the cylindrical obstacle. The aluminium trolley with transducers
as usual induced some disturbance just on the water surface layer about 0.5 cm thick. The
effect of additional disturbance propagated maximum 2 cm below the water surface, and it

123
Environ Fluid Mech (2016) 16:659693 667

was so negligible that could not penetrate to the bed surface. Fig. 4 shows the schematic
diagram of the experimental set-up with all 24 transducers mounted on the aluminium
trolley, which covered the area over the embedded cylinder. Each transducer had a high
precision of 0.01 cm and was capable of 4 Hz sampling rate. Fig. 5ac show the alu-
minum trolley with all transducers placed on it, the experimental set up and a schematic of
the 24 transducer locations respectively. The transducers are referred to as Tr-1, Tr-2Tr-
24. Here the transducer Tr-12 placed over the centre of the cylindrical pier assured that the
pier remained stable. The bed evolution due to local scour and deposition at different
locations around the cylinder was recorded continuously by URS up to 200 mins for each
of the three experiments. During the experimental time of 200 mins, i.e., up to the
stable bed form, continuous video recording of bed form evolution through the glass side of
the flume was made. The video recording was stopped nearly at the time of stable bed
condition, because the main motivation was to investigate the spatiotemporal evolution of
scouring and the bed deformation pattern due to cylindrical obstacles of different sizes.
The shape of the scouring zone and the rate of upstream scour depth became negligible
after about 200 mins of run for all the cases. This was confirmed by test runs, critically
measuring the scoured bed form structures continuously for a longer period of time (about
300 mins), when less than 0.15 cm scour was recorded per hour at the upstream of the
cylinder (at the location of Tr-11). In this study, this condition was chosen to be the
stable bed condition similar to Sarkar et al. [52].
When the stable deformed bed conditions were achieved, the URS trolley was removed
and near-bed instantaneous velocity data of all three components at the level 0.50 cm
above the sand bed were collected at the transducer locations using micro-ADV for
180 sec at a sampling rate 40 Hz, where the particle-particle interactions were very low.
Below the level 0.50 cm, the measurements of velocity data were undetected and erro-
neous; perhaps the particleparticle interactions were very high. During the collection of
velocity data, no changes in scour and bed form were observed. The ADV sampling
volume is 9  108 m3 and is approximately cylindrical oriented along the transmitter
beam axis. Since the fluid contained a lot of particulate matter of size of the order of
microns, the transmitted pulses from the ADV were back-scattered by the particulate
matter in water within the transmitter acoustic beams, and hence the ADV measured the
fluid velocity with fluctuations. The 3-D Micro ADV is a high precision instrument, which
operates using a Doppler back-scattering technique, known as the pulse-to-pulse coherent

Fig. 4 Schematic of experimental set-up: Water channel, 5 MHz ultrasonic ranging system (URS), sand
bed, cylindrical pier (after Fig. 1 of Sarkar et al. [52])

123
668 Environ Fluid Mech (2016) 16:659693

Fig. 5 a Transducers placed at different arrays on the trolley, b photograph of the whole experimental set-
up, c schematic of the locations of different transducers over the cylindrical pier (after Fig. 2ac of Sarkar
et al. [52])

Doppler technique. In this technique, the instrument sends two pluses of sound separated
by a time lag; then it measures the phase of the return signal from each pulse. The change
in the phase divided by the time between pluses is directly proportional to the velocity of
the particles in the water. Pulse coherent process is used because it provides the best
possible temporal and spatial resolution. The sampling volume is located approximately
5 cm below the transmitter probe tip; the precise distance depends on individual probe
geometry. It has a diameter equal to that of the 0.6 cm ceramic of the transmitter (Lohr-
mann et al. [38]; SonTek Inc.[59]), and the vertical length is 0.32 cm. Factory calibration
of the ADV is specified to be  1.0 % of the measured velocity (i.e., an accuracy of 
1 cm/sec is on a measured velocity of 100 cm/sec). The ADV has been used in a variety of
applications for the turbulence measurements, such as over dunes (Ojha and Mazumder
[48]), circular cylinder (Debnath et. al. [19]), cube (Lacey and Rennie [35]) and many
more.
In order to check the uncertainty associated with the ADV data, one data series of 7200
samples for duration of 3 mins have been collected at a location 5 mm above the bed.
Uncertainty in a single measurement in terms of standard deviation and uncertainty in the
mean of the sample in terms of standard error were calculated and presented in Table 1. In
addition, the standard deviation and error associated with the measurement and the mean of
bed elevation data by the URS were also calculated using one data series of 1200 samples
for duration of 5 mins collected at Tr-12 (i.e. the fluctuation in the elevation of cylinder
head during the experiment) and also presented in Table 1.

123
Environ Fluid Mech (2016) 16:659693 669

Table 1 Uncertainty associated with the ADV and URS data


Uncertainty in a measurement (Standard deviation) Uncertainty in the mean (SE)

u 2.74 cm/sec U 0.03 cm/sec


v 1.76 cm/sec V 0.02 cm/sec
w 0.61 cm/sec W 0.00 cm/sec
qu0 v0 5.98 dyne/cm2 qu0 v0 0.07 dyne/cm2
2
qv w0 0
1.33 dyne/cm qv0 w0 0.01 dyne/cm2
2
qu w0 0
2.06 dyne/cm qu0 w0 0.02 dyne/cm2
z(t) 0.02 cm Z 0.00 cm

3 Experimental observations

3.1 Bed topography around cylinders

For all three different cylinder diameters (Dc = 4, 5 and 6 cm) or three different
cylinder Reynolds numbers (ReDc UmmDc 10200; 12750; 15300), the initial movement of
sediment bed around the embedded cylinder was similar. The local bed surface area around
the cylinder was significantly deformed due to the interaction of flow with the submerged
cylinders. Figure 6ac shows the photographs of stable deformed bed form features
developed around the three cylinder diameters (Dc = 4, 5, and 6 cm) respectively. In fact,

Fig. 6 Photograph of the deformed surface after achievement of stable state around the cylindrical piers of
diameter a 4 cm, b 5 cm and c 6 cm. The flow direction is from left to right

123
670 Environ Fluid Mech (2016) 16:659693

Table 2 Measurement of the deformed beds


Properties Case I. 4 cm dia Case II. 5 cm dia Case III. 6 cm dia

Maximum scour-depth 3.4 4.6 5.1


(cm) at Tr-11
Scour radius (cm) Along flow 8.8 9.5 10.3
from Tr-12
Transverse to 8.7 9.3 10.5
the flow
Deposition peaks distance 1st major 13 15.4 17.9
(cm) from Tr-12 deposition
2nd major 24.1 27.4 32.6
deposition
Affected deformed zone Length-wise 52 65.8 97.5
(cm)
Breadth-wise 21 26.4 29
(cm)

Table 2 shows that scour dimension is a function of cylinder diameter. It is observed that
the scoured dimensions and the downstream bed form structures about the centreline
depend on the diameter of the cylinders. It is interesting to note that a ridge is observed at
downstream along the centreline for the cylinder of Dc = 4 cm diameter (Fig. 6a), an
asymmetric bed form structure with some ripples in one side for Dc = 5 cm diameter
(Fig. 6b), and a symmetric deformed bed about centerline with several ripples in both sides
along the downstream direction for Dc = 6 cm diameter (Fig. 6c). In fact, scour develops
at the upstream of the cylinder due to the formation of horse-shoe vortices and the
deformed bed at the downstream of the cylinders due to the shedding wake vortices
generated by the separation of flow (Tsutsui [70]). These wake vortices act like small
tornadoes, lifting up sediment particles and transporting them toward downstream. Espe-
cially for the case of 5 and 6 cm cylinder diameters, lunate type bedforms developed along
two sides of centreline deposition region with a scoured zone at the end of stoss-side in the
upstream direction. At a certain distance downstream of this zone, linguoid bed forms were
generated especially in the case of 6 cm diameter cylinder. Moreover, a long sediment
ridge along the centreline was observed just downstream of the peak region, which
increased in both length and width with cylinder size.

3.2 URS data analysis over deformed beds

The temporal changes in bed elevations at some intervals of time obtained from the URS
data along Row0 are shown in Fig. 7ac and along Row1 in Fig. 7df for 4, 5, 6 cm
diameter cylinders respectively. Along Row0 (centre line of the flume) bed deformation
was started immediately after a few minutes and scoured bed materials were deposited just
downstream of the cylinder base. In this study, the scoured bed refers the bed height to be
smaller than 5.5 cm (initial bed height) and above the 5.5 cm bed height, it is referred to as
deposition. Changes along Row1 (transducer locations near the right side of the cylinder)
could be seen only after 89 mins of the experiment. Scour-hole was slowly spreading at
upstream of the cylinder centreline and the deposition took place just downstream of it.
Major erosion and deposition were observed along Row0 (centre line of the embedded
pier). The location of the transducer Tr-11 just upstream of the cylinder at Tr-12 showed

123
Environ Fluid Mech (2016) 16:659693 671

Fig. 7 Temporal changes in bed elevation at some time intervals along Row0 (shown in ac) and along
Row1(shown in df) for 4, 5 and 6 cm diameter cylinder respectively

continuous scouring due to the horse-shoe vortices developed at the frontal area of the
cylinder until it reached to stable state. After a certain minutes, significant accumulation of
eroded sediment was identified at the downstream vicinity of the cylinder at Tr-13 in each
of the cases. At around 69 mins it attained a maximum deposition (where bed
height = 9.86 cm) for the 6 cm diameter case, which was approximately about 1.36 and
1.05 times of the maximum deposition recorded for 4 and 5 cm cylinder diameters
respectively. It is interesting to note that flip-flops were observed from the bed elevations
recorded by the transducers Tr-13 and Tr-14 at downstream of the cylinder at Tr-12. The
location of Tr-13 was observed to be peak deposition in 10 mins, and gradually the peak
shifted to the location of Tr-14 nearly at 15 mins of run, and then a continuous decrease in
elevation was observed at Tr-13 due to turbulent wakes. Continuous scouring initiated to
occur (from the initial bed height of 5.5 cm) just at the downstream (Tr-13) after 27 mins
for the smallest cylinder and for the cylinder of diameters Dc = 5 and 6 cm, the scouring
began at 36 mins. In fact, at 9 mins, sediments got accumulated heavily at Tr-13 location
(just downstream of cylinder along centreline) coming mostly from eroded Tr-11 location.
As time goes on, these sediments eroded and got accumulated at further downstream
location (Tr-14) and also extended towards the side wall creating smaller sloped bed forms.
Thus, at 9 mins it was deposition at the Tr-13, and at 15 min (Fig. 7ac), the bed heights at
the same location were smaller than that at 9 mins.
From Fig. 7df along the Row1 (transducers were placed on the left side of the
cylinder), S-shaped bed profiles were observed with time, which was mostly like a

123
672 Environ Fluid Mech (2016) 16:659693

trough and crest profiles. Interestingly, at the transducer locations along the Row1, the
scour started at 108 mins of the experiment at the downstream of the smallest cylinder of
4 cm diameter case for cylinder Reynolds number ReDc 10200, and at 72 and 51 mins
for the case of 5 cm cylinder diameter (where ReDc 12750) and the diameter 6 cm (for
ReDc 15300) respectively. At the initial stage, the flow velocity was so low that the
dimensionless bottom shear stress was below the threshold value for a grain-size move-
ment, i.e. when there was no sand transport at the bed (Eq. 1). But eventually after a long
time, due to the smallest blockage ratio where the change of flow velocity happened to be
negligible, the starting time of scour or erosion was observed delay at the downstream of
the cylinder because the probability of occurrence turbulent fluctuations responsible for
dislodge the sand sizes was less, whereas the increase in blockage ratio causes to generate a
little larger turbulence length scale corresponds to the higher flow velocity between the
wall and the cylinder boundary, where the probability of occurrence turbulent fluctuations
responsible for dislodge the sand sizes was high enough, and so the scour started to occur
faster than the earlier. Hence, the increase in diameter of the pier (blockage ratio or
cylinder Reynolds number) leads to be faster the starting time of scouring at the down-
stream. As the time passes, the sand material deposits and spreads at both sides. It was
found that scour formation reached nearly a stable state for the cylinder of smallest
diameter (Dc = 4 cm) after approximately 70 mins from the starting time. The scour
around the 5 and 6 cm diameter cylinders took 108 and 170 mins respectively to achieve
nearly a stable state from their starting times of scouring. It is observed that the
stable states for individual cylinder diameters (4, 5, 6 cm) achieved at respective total times
around 178, 180, 221 mins. The maximum scour-depth was invariably at the transducer
location just upstream of the cylinders. The upstream as well as the downstream scoured
regions were observed to be the largest for the 6 cm diameter and the smallest for the 4 cm
diameter case. As the cylinder diameter (blockage ratio) increases, a small amount of bed
deformation can be observed along the width-wise transducer locations at later time after
150 mins of experiment, especially for the case of the 6 cm cylinder (largest blockage
ratio).
The surface contour plots of bed form elevations (z) at the end of each run are shown in
Fig. 8ac for three different cylinder cases respectively. Contour maps of this paper were
generated using a kriging procedure in the OriginLab contouring package. Kriging tech-
nique employs a weighted moving average interpolation (or/and extrapolation) method that
minimizes the estimated variance of a predicted point from the weighted average of its
neighbors (Best and Kostaschuk [9]). The weighted value is determined by the spatial
correlation structure of the original data. It is observed from the figures that the upstream
scour depth was maximum about 5.09 cm for the 6 cm diameter cylinder, which was
approximately 1.4 and 1.1 times of that for 4 and 5 cm cases respectively. Sediment ridge
portions were observed along the centreline at approximately 6, 7.5 and 9.5 cm down-
stream of 4, 5 and 6 cm diameter cylinders respectively.
The upstream slope of the scour hole determines the stability of the upstream part of the
scour hole and adjacent bed protection. Now, to investigate the variations in gradients or
slopes of the deformed bed around and to make the comparative study, the bed slopes at the
end of the runs along the flow are shown in Fig. 9ac, and transverse to the flow in Fig. 9d
f for all three cases. The bed gradient was measured as a quotient equals to the rise divided
by run, where rise of a bed profile between two points is the difference between the
elevation of the bed at those two points and the run is the horizontal distance of those two
points as used by Sarkar et al. [52]. The bed slope along the flow at the cylinder upstream

123
Environ Fluid Mech (2016) 16:659693 673

Fig. 8 Surface contour plots of bed elevation (z) at the end of each run for a 4 cm, b 5 cm and c 6 cm
diameter cylinder respectively. The flow direction is shown by black colored arrow

location was observed to be maximum at the deepest part of the scour-hole, approximately
48 for 6 cm diameter and minimum about 39 for 4 cm diameter, which are comparable to
the findings of Link et al. [37], where average slope of scour-hole sides were 45 in the
middle part of upstream scour-hole. These steeper slopes at the deeper parts of the scour
hole is generated by horseshoe vortex. The average upstream scour slope (i.e. average
slope at the upper and deeper parts of the scour hole) was estimated to be equal to the angle
of repose (Hoffmans and Verheij [26]). Sharply peaked deposition region was captured for
4 cm diameter case at about 7 cm downstream of it with a gradient of 0.64. Interestingly,
downstream of 6 cm diameter cylinder, two sediment heaps were generated at two sides of
the flume centreline. In the transverse direction, the bed slopes along the transverse
direction were smaller than that along the flow with an average slope of 31 . Fig. 9df
showed that the deformed bed forms generated around the cylinders were of symmetrical
in nature with respect to the flume centreline.

123
674 Environ Fluid Mech (2016) 16:659693

Fig. 9 Bed gradient contour plots at the end of each run: ac along the flow direction and df transverse to
the flow direction for 4, 5 and 6 cm diameter cylinder respectively. The flow direction is shown by black
colored arrow

3.3 Statistical characteristics of surface evolution

The statistical analysis of bed evolution pattern at some selected locations, such as Tr-1,
Tr-7, Tr-11, Tr-13, and Tr-14 for the total 200 mins of experimental time, is performed.
These locations are chosen to represent differences in bed evolution. At a particular
location, we define the change in bed elevation by
dzt zt Dt  zt 2
where z(t) denotes the bed elevation at time t at that location and Dt is the temporal
resolution of the experimental data (here Dt 0.25 s). This leads to three cases to be
considered for the surface evolution processes [23, 57, 58] the positive values of elevation
increments dzt [ 0 correspond to deposition (D jdztj), the negative values of
dzt\0 erosion events (E jdztj) and the zero values of dzt0 correspond to
inactivity at that time and location. Each of these three processes, deposition, erosion and
inactivity, has a characteristic time scale of operation.

123
Environ Fluid Mech (2016) 16:659693 675

The exceedance probability of the series of deposition (D), for the deposition magnitude
Dj is defined as
j
PD [ Dj 3
m1
where j corresponds to the rank of the deposition in the sorted series of D in descending
order, e.g. j 1 for the largest deposition magnitude in the series of D and m is the length
of the D series, i.e. equal to the maximum value of j. Similarly, the exceedance probability
PE [ Ej of the series of erosion (E) for the erosion magnitude Ej is calculated.
Here, PD [ Dj and PE [ Ej for all three cases of cylinder diameters (4, 5, and 6
cm) are plotted in Fig. 10 for five selected transducer locations (Tr-1, Tr-7, Tr-11, Tr-13,
and Tr-14) in loglog scale. In fact, the exceedance probability is the probability that a
certain depositional height or scour depth recorded in particular time will be exceeded
during the whole experimental time at certain location. It may be mentioned here that the
transducer (Tr-1 and Tr-7) are located in one side of the pier and the transducers (Tr-11,
Tr-13 , and Tr-14) are on the centreline of the pier, where T-11 is on the upstream and T-13
and T-14 are placed along the downstream of the pier. From the figure, it is clearly
observed that the overall minimum exceedance probabilities of deposition PD [ Dj and
erosion PE [ Ej events for all three cylinder diameters are observed at Tr-1 with slightly
greater probability for largest diameter size. Plots of the exceedance probability at the
locations Tr-7 and Tr-14 clearly indicate the least or negligible effect in depositional or
erosional events. The transducers Tr-11 and Tr-13 located respectively at the upstream and
downstream of the pier show significant changes in the bed elevation or scour. It is
interesting to note that the exceedance probabilities of deposition and erosion events at the
upstream and downstream of the pier are identical. Moreover, PD [ Dj and PE [ Ej
possessed smaller values at the upstream location Tr-11 for magnitude equal to 1 mm than
that at the downstream location Tr-13 of the pier. At the location Tr-11 upstream of the
pier, where the scour depth is maximum, highest probabilities of exceedance are observed
for largest diameter of cylinder above depositional magnitude 0.5 mm for D and above
erosional magnitude 1.0 mm for E. Comparing the probabilities of exceedance PD [ Dj
and PE [ Ej at Tr-11 with that at Tr-7 and Tr-1, it is clearly observed that the probability
of exceedance is always greater for any magnitude of D or E at Tr-11 than that at Tr-7 and
Tr-1. Further downstream location Tr-14, it is observed that the probabilities of exceedance
PD [ Dj and PE [ Ej are almost identical for all the embedded cylinder diameters.
Interestingly, PD [ Dj and PE [ Ej possessed quite smaller values for magnitude
equal to 1 mm at Tr-1 relative to other transducers locations around the pier. As the bed
elevation increment (deposition or erosion) process is random in nature, the loglog plots
of exceedance probability of deposition and erosion help to estimate the chance of a
natural, inherent, or hydrological risk of bridge failure due to unwanted major bed
deformation events.

4 Near-bed velocity measurements

4.1 Near-bed Reynolds shear stresses

In turbulent flow, the instantaneous velocity components (u,v,w) in the Cartesian coordi-
nate system (x,y,z) are given by,

123
676 Environ Fluid Mech (2016) 16:659693

Fig. 10 Exceedance probabilities of deposition and erosion events i.e. PD [ Dj and PE [ Ej for all
three cases: 4 cm (thin continuous lines), 5 cm (thick continuous lines), and 6 cm (thick dashed lines)
respectively at five selected locations Tr-1, Tr-7, Tr-11, Tr-13, and Tr-14

u U u0 ; v V v0 ; w W w0 4
where U, V, W are the time-averaged velocities in (x,y,z)-directions and u0 , v0 , w0 are the
fluctuating components. Here x-axis is along the flow, y-axis transverse to the flow and z-
axis normal to the flow; origin (0, 0, 0) is taken at the center of the pier at the centreline of
the flume. The near-bed ADV velocity data were filtered to remove noises using the phase-
space threshold de-spiking method described by Goring and Nikora [25] for all locations of
the transducers. At least 90 % of the raw data remained after using the de-spiking method.
Excluded signals were replaced using a cubic polynomial interpolation. The time-averaged
local Reynolds shear stress components along three planes (xy, yz and xz) are determined
by:

suv qu0 v0 ; svw qv0 w0 ; suw qu0 w0 5


Here, the total 7200 velocity data collected using ADV at 40Hz rate were considered for
averaging.
The contour plots of stable bed elevation increments (z) and the corresponding plots of
time-averaged Reynolds shear stresses suv , svw and suw along three different planes (xy, yz
and xz) around the different cylinder diameters (4, 5 and 6 cm) are shown in Fig. 11a1a3,
b1b3, c1c3 and d1d3 respectively. Effectively, from the figures it is observed that

123
Environ Fluid Mech (2016) 16:659693 677

Fig. 11 Contour plots of stable bed elevations for all three cases are shown in subfigures (a1a3). Contour
plots of Reynolds shear stress components suv , svw and suw for all three cylinder sizes in subfigures (b1b3),
(c1c3), and (d1d3) respectively. The flow direction is shown by black colored arrow

suv [ suw at the downstream of all the cylinder diameters, which correspond to the higher
bed elevation increments at the downstream of the cylinders, indicating that depositional
increments lead to increase shear stress suv along xy-plane. It is readily observed that the
overall magnitude of suv is maximum and that of suw is minimum. The increase in cylinder
diameter (blockage ratio) leads to increase in both shear stress components suv and suw in
the downstream with a maximum effect of suv for 6 cm diameter cylinder and the shear
stress svw was negative and minimum.
The loglog plots of exceedance probabilities for the positive tails of instantaneous
near-bed Reynolds shear stress distributions along the three planes are presented in
Fig. 12a1a4, b1b4, and c1c4 for 4, 5, and 6 cm cylinder diameters respectively at
selected transducer locations Tr-7, Tr-11, Tr-13, and Tr-14. These plots show the proba-
bility that a certain instantaneous Reynolds shear stress value recorded in a particular time
is exceeded during the 3 mins of velocity measurement at certain location. Interesting to
note that the flip-flops were observed between the exceedance probabilities of shear
stresses svw and suw . At the side location Tr-7 of cylinders, the probability of exceedance
for svw surpasses the suw . A similar phenomenon was noticed at the upstream location Tr-
11, but the profiles were overlapped at some portion. However, at the downstream location
Tr-13, the exceedance probability of suw showed beyond that of svw for lower instantaneous

123
678 Environ Fluid Mech (2016) 16:659693

Fig. 12 Exceedance probabilities for the positive tail of instantaneous Reynolds shear stresses at three
planes (i.e. of suv with dashed lines, svw with dotted lines and suw with continuous lines) for all three
cylinders: 4 cm (subfigure a1a4), 5 cm (subfigure b1b4), and 6 cm (subfigure c1c4) respectively for four
selected locations Tr-7, Tr-11, Tr-13, and Tr-14

shear stress values and overlapped each other at higher shear stress values (Fig. 12a3c3).
At further downstream location Tr-14, a similar phenomenon like Tr-13 is observed except
for the largest cylinder diameter, where the shear stresses suw and svw overlapped each
other at higher shear stress values (Fig. 12c4). Overall it is observed from the Fig. 12 that
the exceedance probabilities for the positive tails of instantaneous near-bed Reynolds shear
stress distribution suv along xy-plane (horizontal plane) is maximum in the both upstream
and downstream of the cylinders. In fact, in the near-bed region, the Reynolds shear stress
suv along the xy-plane is the dominant shear stress, which was due to the generation of
tremendous secondary currents, especially v-velocity component, caused by the kind of
eddy generated around and at the wake of the obstacles. The increased dominance of
vortex-shedding and the strength of vorticity for the span-wise velocity (v) in comparison
to that of the other two velocity components for the deformed bed condition can be verified
through a critical watch over the power spectra plots of these velocity components.

123
Environ Fluid Mech (2016) 16:659693 679

4.2 Power spectra analysis

The power spectrum of the de-trended velocity data at different points at the upstream and
downstream, in the wake of the pier as well as inside the scour hole was conducted in order
to find the dominant vortex-shedding frequency of the large scale coherent structures. The
power spectrum at each point was calculated using Fast Fourier Transform (FFT) of auto-
covariance function of velocity time-series data. The resultant power spectra P(f) at dif-
ferent points around each pier for plane bed condition were calculated as, Pf
Pu f 2 Pv f 2 Pw f 2 0:5 cm2 =s2 and presented in Fig. 13. Here, Pu f , Pv f and
Pw f are the power spectra for stream-wise, transverse and vertical velocity components
respectively. In order to know how much energy is contained in each frequency, we
multiplied each of the spectra by the corresponding frequency which gives the spectra in
variance preserving form. Distinct frequency peaks in the power spectrum plots indicate
the frequencies of coherent eddies that are shed from the cylinder boundary layer. For 4 cm
diameter cylinder, the peak power spectra at sides and wake of the cylinder were found at
exactly the same frequency of 0.703 Hz. With increase in the cylinder diameter, the
dominant vortex-shedding frequency decreased to 0.586 and 0.508 Hz for 5 and 6 cm
diameter cylinder respectively, which in terms of the Strouhal number, St (fDc =U0 ) is
approximately 0.14, 0.15 and 0.16 respectively, where f is the vortex-shedding frequency,
U0 is the depth-averaged incoming flow velocity = 19.5 cm/sec.
The power associated with the peak frequency for the power spectrum distribution
indicates the strength of vorticity of different kind of vortices formed by the cylinder. In
turn, the strength of wake vorticity expresses the capacity of the wake vortices to entrain
and move sediment bed from the flanks and rear of each cylinder. From the figure, it was
evident that the vorticity strength at the flume centerline (along Row0) and at the near wake
region (Tr-9 and Tr-10) just down-stream of the pier, was maximum. It confirmed the
fastest scour initiation process for the case of lowest size cylinder. In the just upstream of
cylinder location (Tr-11), the maximum vortex-shedding frequency with maximum asso-
ciated power was observed for the largest cylinder (f = 0.625 Hz), and consequently the
both were minimum for the smallest cylinder (f = 0.156 Hz). Again it is interesting to note

Fig. 13 Resultant power spectra fP(f) at different points (Tr-1 to Tr-14) around each pier: 4 cm (thin
continuous line), 5 cm (thick continuous line) and 6 cm (thick dashed line) for plane bed condition

123
680 Environ Fluid Mech (2016) 16:659693

that comparatively the strength of the vortices at Tr-3 and Tr-4 away from the near-wake
region was maximum for the largest cylinder and minimum for the smallest cylinder,
which in turn is responsible for the generation of larger secondary bed form on mobile
sandy bed for the larger cylinders.
The power spectra of u, v and w-velocity data at different transducers locations for the
cylinder of 4 cm diameter are plotted in Fig. 14 for plane bed case and in Fig. 15 for
scoured bed condition. It was evident that the strength of wake vortex was maximum for
span-wise velocity (v-component) especially along the cylinder downstream centerline
location for plane bed condition and interestingly, for the scour bed condition the corre-
sponding maximum value was achieved at side of the cylinder base (Tr-7) and just
downstream of the pier location (Tr-13), thus in turn it confirmed the tremendous sec-
ondary current generation at near the bed at these locations. For both the conditions, the
vortex strength was overall prominent for the stream-wise (u) and span-wise (v) velocity
components, that contribute largely to the suv component to be dominant than the other two
Reynolds shear stress components.
Contrary to the flat-bed case (Fig. 14), it was also noticed that the power spectra in the
scoured-bed case (Fig. 15) became very weak and several dominant frequencies were
present there with considerable decrease in associated power. This situation indicated that
when the equilibrium scour hole and the overall stable bed form were achieved, the
capacity of wake-vortices to entrain and move the sediment bed from the sides and
upstream of the cylinder considerably decreased. It was also observed that inside the
upstream scour hole (Tr-11), the strength of vortices was lower than those outside the scour
hole. It can be argued from the Figs. 14,15 that for the scour bed condition, considerable
increase in the power associated with the span-wise velocity data (v) took place in scoured
region at the flanks and rear of the cylinder when compared to the plane bed condition.

Fig. 14 The power spectra plots of u, v and w-velocity data: f  Pu f (thin circled black line), f  Pv f (thin
starred black line), f  Pw f (thick dashed ash line) at different locations for plane bed condition for 4 cm
diameter pier

123
Environ Fluid Mech (2016) 16:659693 681

Fig. 15 The power spectra plots of u, v and w-velocity data: f  Pu f (thin circled black line), f  Pv f (thin
starred black line), f  Pw f (thick dashed ash line) at different locations for scoured bed condition for 4 cm
diameter pier

4.3 Quadrant analysis of near-bed Reynolds shear stress

Coherent structures with large flux events had been proposed to explain the bursting
phenomena responsible for resistance to the motion, sediment transport process, turbulence
production, hence the mixing [2, 29, 31]. The turbulence over the deformed bed structure
around the piers was examined through quadrant analysis [40, 47, 74] to estimate the major
turbulent events characterizing the coherent structures. This study considers the Reynolds
shear stress component suv qu0 v0 along xy-plane, because suv is the most prominent
near-bed shear stress over the scoured surface generated by three different cylinder
diameters observed from the exceedance probability (Fig. 12) and according to the dis-
cussion in previous section.
Quadrant analysis was originally devised to sort out the contributions to the total shear
stress s from each quadrant events by Corino and Brodkey [17]. The idea of quadrant
analysis is used in the present work to determine the contributions of turbulent events to the
total shear stress component suv qu0 v0 on the u0 v0 -plane. Thus, in this technique one
considers the frequency of occurrence and contribution to Reynolds shear stress from the
four quadrant events defined by: (1) left boundary-ward interaction Q1i
1; u0 [ 0; v0 [ 0 occurring when the high-speed fluid moves towards left, (2) ejection
event Q2i 2; u0 \0; v0 [ 0 characterizing the low-speed fluid towards left boundary,
(3) right boundary-ward interaction Q3i 3; u0 \0; v0 \0 occurring when low-speed
fluid moves towards right, and (4) sweep event Q4i 4; u0 [ 0; v0 \0 occurring high-
speed fluid towards right boundary. The more significant event is ejection (Q2) which
transports low-momentum fluid along positive v-direction (left boundary), and sweep (Q4)
which transports high-momentum fluid along negative v-direction. It may be noted that
both the ejection and sweep events (Q2 and Q4) produce turbulent energy; and the both
wall-ward interactions (Q1 and Q3) establish energy dissipation along the plane (Pope
[51]). Therefore, bursting events on this u0 v0 -plane have a significant influence on the
entrainment of sand particles towards the boundaries into the flowing water.

123
682 Environ Fluid Mech (2016) 16:659693

The existing quadrant analysis is used to analyze the deformed scoured bed generated
around vertically embedded circular cylinders influenced by the various turbulent events
near the bed. Each velocity pair from the ADV data is investigated either through
examining the entire signal data or filtering those data above a threshold value H. A
hyperbolic hole H is defined, excluding from the analysis a region of instantaneous values
of ju0 v0 j that are greater than Hju0 v0 j. At any point in a turbulent flow, the contribution of
the total shear stress from quadrant i excluding the region H, is defined as,
Z
1 n 0
\u0 v0 [ i;H lim u tv0 tIi;H u0 ; v0 dt 6
n!1 n 0

where \ [ denotes a conditional average and n is the total number of observations. The
indicator function Ii;H is defined as,
8 0 0
< 1; if u ; v is in ith quadrant and
>
0 0
Ii;H u ; v if ju0 v0 j
Hju0 v0 j 7
>
:
0; otherwise
Here, H is the threshold parameter in the Reynolds stress by which one can extract the
more intense values of u0 v0 that exceed Hju0 v0 j. The stress fraction [39, 46] by i-th quadrant
is defined as,
hu0 v0 ii;H
Si;H 8
u 0 v0
which gives the Reynolds shear stress fraction associated with each of the turbulent events.
By definition, for H = 0, S1;0 S2;0 S3;0 S4;0 1.
Figure 16ac depicts the contributions of all four quadrant events to the Reynolds shear
stress suv i.e. stress fraction jSi;H j for i = 1 to 4 for threshold parameter H for all three
cylinders of 4, 5, and 6 cm diameter respectively. Extensive influence of cylinder sizes or
blockage ratios and spatial locations of transducers on stress fraction jSi;H j of an individual
turbulent event can be readily observed.
The space fraction or occurrence of probability of each of the quadrant events are
computed [28, 37, 69], which gives the fraction of total observations (n) contributing to the
instantaneous Reynolds shear stress by each of the quadrant events for a given H, using
P
0 0 Ii;H
Ni;H u ; v 9
n
where Ii;H is the indicator function (Eq. 7). The space fractions Ni;H u0 ; v0 given by Eq. (9)
occupied by all four quadrants (i = 1 to 4) for H ranging from 0 to 5 are shown in Fig. 17a
c for all three cylinders of 4, 5, and 6 cm diameter respectively.
As occurrence probability does not consider the history of occurring events, it cannot
define the circumstances of a bursting event by stating information about the movements of
one event to another event or occurrence of remaining in the same event for consecutive
time intervals. Thus, to find the transition probability of the movement of an event at time t
to another event at time t Dt is defined [31, 32] as,
ni!j
Pi!j ; i; j 1; 2; ::4 10
ni

123
Environ Fluid Mech (2016) 16:659693 683

Fig. 16 Stress fraction jSi;H j for each quadrant against H for all three cylinders of diameter a 4 cm, b 5 cm,
and c 6 cm at four selected locations : Tr-7 (with continuous lines), Tr-11 (with thin dashed lines), Tr-13
(with thick dashed-dotted lines), and Tr-14 (with thick dashed lines)

where, Pi!j is the transition probability or probability of movement from one event at time
t to another event at time t Dt (denoted by P11, P12 etc. hereafter), ni!j is the number
of such transitions and ni is the number of occurrence of bursting event of quadrant i.
Considering the lateral or marginal movements between two adjacent quadrants, e.g. n1!2 ,
n3!4 , cross movements between crossed quadrants, e.g. n1!3 , n4!2 and stable movements,
i.e. when the same event occurs in the time step t and t 1, e.g. n1!1 , n2!2 , we analyze the
dependence of occurring a quadrant event at time step t Dt on the occurrence of some
quadrant event at previous time step t.

123
684 Environ Fluid Mech (2016) 16:659693

Fig. 17 Space fraction Ni;H  100 % for each quadrant for H ranging from 0 to 5 for all three cylinders of
diameter a 4 cm, b 5 cm, and c 6 cm at four selected locations : Tr-7 (with continuous lines), Tr-11 (with
thin dashed lines), Tr-13 (with thick dashed-dotted lines), and Tr-14 (with thick dashed lines)

5 Discussions

Bridge piers, or pier-like structures i.e. obstacles of different sizes embedded vertically in
river bed can be used as a tool for river restoration / rehabilitation by adding morphological
undulation on river geometry. A left pier after a removed/renewed bridge, or an unfinished

123
Environ Fluid Mech (2016) 16:659693 685

bridge pier construction, or a pier-like structure especially constructed for the ecological
purpose would act as a submerged object in flood situation. The above considerations
motivate the need to investigate the spatial and temporal changes of bed elevations and
associated near-bed turbulence phenomena due to submerged cylindrical piers at a constant
discharge. Study of flow associated with these structures in this fluvial environment is
inherently difficult since it is usually hard to measure the near-bed flow structures and to
quantify the nature of turbulence and bottom Reynolds shear stresses over these structures.
This paper presents the high-resolution experimental data-set and subsequent results
relating to the development of scour features with time around cylindrical obstacles
embedded in a sand bed.
From the stable bed form structures developed around the cylinders, it was noticed that
the submerged vertical cylinders significantly changed the local bed surface area, indi-
cating the scour dimension as a function of cylinder diameter or blockage factor. More-
over, a long sediment ridge along the centreline was observed just downstream direction of
the peak deposited region, which increased in both length and width with cylinder size.
It is noticed that along Row0 (centreline of the flume) from the transducers data,
continuous scouring initiated at the downstream location (Tr  13) of the cylinder of 4 cm
diameter after 27 mins and it was about 36 mins for the larger cylinder of 5 and 6 cm
diameters. It may be referred here that scouring means below the original bed thickness 5.5
cm and deposition means above the thickness 5.5 cm. Greater flow perturbations at the
upstream base of the larger cylinder generated larger amount of bed material to be eroded
and got accumulated at the just downstream of the cylinder around the flume centreline.
Thus, for the case of larger cylinder, it takes longer time for continuous downstream scour
to take place. It is interesting to note that S-shaped bed profiles were observed to be
generated with time along Row1 (transducers Tr-15 to Tr-20 at the right side of the
cylinder), where the scour occurred at the downstream of the cylinder of smallest diameter
(Dc 4 cm) at 108 min of the experiment, and for the case of larger blocking diameters
(fir Dc 5 and 6 cm), the starting times of scouring were 72 and 51 mins respectively.
Therefore, unlike the centreline scouring, the bed deformation in the transverse direction
with continuous bed erosion was delayed to take place as cylinder diameter decreases. It
was found that the stable state of scoured region around the cylinder of smallest diameter
achieved earlier than that for the larger diameter of cylinders. So the reaching of stable bed
condition depends on flow perturbation. The bed slope along the flow at the upstream was
maximum, = 48 for Dc = 6 cm cylinder and minimum, 39 for Dc 4 cm cylinder,
whereas the bed slope towards the transverse direction was smaller than that along the
flow, which is an average slope about 31 .
The statistical analysis of bed evolution at some selected locations reveals that at
upstream of the cylinder, where the scour-depth was maximum, the highest probabilities of
exceedance for deposition and erosional events were observed for largest diameter of
cylinder above jdztj 0.5 mm for deposition D and above jdztj 1.0 mm for erosion
E, whereas at the immediate downstream of cylinder, the highest exceedance probability
values of deposition and erosion were found for the largest diameter of cylinder above
jdztj 0.4 mm for both D and E. At the side location of the cylinder diameters, the
exceedance probability profiles for all three cylinder diameters overlapped with slightly
greater value for largest size of cylinder above jdztj 0.7 mm, and at further down-
stream location (Tr-14), all three profiles were overlapped with each other above jdztj
0.8 mm for both D and E. It is important to note that the depositional and erosional events
are the random variables which define the thickness of the preserved stratigraphic
sequences (Kolmogorov [34]).

123
686 Environ Fluid Mech (2016) 16:659693

From the contour plots of time-averaged bottom Reynolds shear stresses suv , svw and suw
for different cylinder diameters (blockage ratios), it is observed that the overall magnitude
of suv was maximum and that of suw was minimum. Especially with increase of cylinder
diameter, positive value of suv was observed at right side just downstream of the cylinder
and the negative value of suv was observed at left side of the cylinder base. The fluid
velocity, which passed along the right side of the cylinder base, changed the flow direction
towards the left wall just downstream of the cylinder, caused the higher magnitude of suv
(Keshavarzi et al. [31]). Similarly, reverse phenomena occurred from the left side of the
cylinder base, resulting higher negative values of suv . At further downstream of the
cylinder, where the magnitude of suv was smaller, higher bed elevation occurred [22]. For
svw component, due to the dominance of intermittent flow reversal of secondary currents,
the negative shear stress was observed at just downstream and around right side of the
cylinder indicating the flux of momentum along the left boundary [29]. However, at the
upstream of the cylinder and around the left side, positive svw was observed [29] indicating
the flux of momentum along the right boundary. The magnitudes of svw at the upstream of
the cylinder were quite smaller than that at the downstream. Note that at the upstream of
maximum scoured region, maximum shear stress was shown by svw component. However,
the suw component showed negative value at just upstream of the cylinder [29] due to the
return flow, whose magnitude increased with the cylinder diameter. It is to be noted that the
minimum of suw was observed at the downstream of the cylinder, where maximum sedi-
ment accumulation was generated and hence the location of minimum suw moved to further
downstream with the increase of cylinder diameter.
Exceedance probabilities for the positive tails of instantaneous Reynolds shear stresses
at the selected locations for all three cylinder diameters showed that for all locations, the
exceedance probability of suv was the highest among the three shear stresses with highest
tails at the downstream location Tr-13 and lowest tails at upstream location Tr-11.
Moreover, two profiles of suw and svw overlapped each other for higher instantaneous shear
stress values, which increased with cylinder diameter (30 dyne/cm2 for diameter Dc 4
cm, 60 dyne/cm2 for diameter Dc 5 cm and 100 dyne/cm2 for diameter Dc 6 cm). At
further downstream location Tr-14, the exceedance probabilities of suw always exceeded
those of svw .
The influence of cylinder sizes or blockage ratios and spatial locations of transducers on
stress fraction jSi;H j of an individual turbulent event are readily seen from the Fig. 16ac.
Just downstream of the cylinder at transducer location Tr-13 (the location of downstream
scour, where the scour depth increased as cylinder size increased), contributions of ejection
event (Q2) decreased with increase in cylinder diameters (blocking ratio) for all threshold
parameter H. The sweeping event (Q4) increased with increase in cylinder diameters 4 and
5 cm, except for 6 cm diameter for all threshold parameter H. However, at further
downstream location Tr-14, the contributions of all the events due to Dc 4 and 5 cm
were greater than that due to 6 cm diameter, and these were equally important for both 4
and 5 cm diameters except the largest diameter. Furthermore, the contributions to jSi;H j at
Tr-14 from all the events were greater than that at Tr-13 for all Dc 4, 5, 6 cm diameters
except for sweeping event due to Dc 6 cm cylinder diameter. For example, at the near-
bed region, downstream location Tr-13 of the cylinder diameter Dc 4 cm for fixed H =
10, the contributions from the quadrant events are S1;10 0:150 S2;10 0:945, S3;10
0:273 S4;10 0:695, sum of them is 0.762, which is 76 % of average shear stress s; for
diameter Dc 5 cm for H = 10, the contributions from the quadrants are S1;10 0:184,
S2;10 0:482, S3;10 0:244, S4;10 0.542, sum of them is 0.596, which is 60 % of

123
Environ Fluid Mech (2016) 16:659693 687

average shear stress s; and for Dc 6 cm for H = 10, the contributions are S1;10 0:078,
S2;10 0:303, S3;10 0:103, S4;10 0:449, sum is 0.572, which is 57 % of average
shear stress s. Similarly, at further downstream Tr-14 due to the cylinder diameter
Dc 4 cm for H = 10, the contributions from the quadrants are S1;10 2:744,
S2;10 2:298, S3;10 2:680, S4;10 2:114, sum of them is 1.013, which is 101 % of
average shear stress s; for diameter Dc 5 cm for H = 10, the contributions from the
quadrants are S1;10 1:753, S2;10 2:284, S3;10 1; 540 , S4;10 1:938 , sum is
0..928, which is 93 % of average shear stress s; and for Dc 6 cm for H = 10, the
contributions are S1;10 0:137, S2;10 0:475, S3;10 0:129, S4;10 0:379, sum is
0.589, which is 59 % of average shear stress s. Therefore, it is noted that at both down-
stream locations Tr-13 and Tr-14 the sum of the contributions of all quadrant events to the
stress fraction jSi;H j decreased with increase in cylinder diameter (blocking ratio) for all
threshold parameter H.
However, at the upstream location Tr-11, increase of cylinder diameter or blocking ratio
led to decrease the values of all the quadrant events in magnitude for all threshold
parameter H, but the contributions of all events due to Dc 4; 5 cm were greater than that
due to Dc 6 cm at the same location. For upstream scour location Tr-11 of Dc = 4 cm for
H = 10, the contributions of quadrant events are S1;10 = 23.730, S2;10 = -19.339, S3;10 =
20.190, S4;10 = -23.607, sum of them is 0.974, which is 97 % of average shear stress s; for
Dc = 5 cm for H = 10, the contributions are S1;10 = -20.142, S2;10 = 17.011, S3;10 =
-14.756, S4;10 = 18.896, sum is 1.008, which is 100 % of average shear stress s; and for
Dc = 6 cm for H = 10, the contributions are S1;10 = 0.740, S2;10 = -0.379, S3;10 = 0.736,
S4;10 = -0.435, sum is 0.662, which is 66 % of average shear stress s. Therefore, at the
upstream scour region Tr-11, sum of the contributions from all four events to jSi;H j
increased for Dc = 4 and 5 cm and then decreased for Dc = 6 cm for all threshold parameter
value H.
From the Fig. 6ac it is noted that the contributions of all quadrant events to the total
shear stress jSi;H j at the upstream location Tr-11 were much greater than that at both Tr-13
and Tr-14 in the downstream locations for all the cylinder diameters or blockage ratios. For
Dc = 4 and 5 cm, all the events at the scoured location Tr-11 were much greater than that at
the location Tr-13, but for diameter Dc = 6 cm, contributions from all the four events at the
side location Tr-7 were greater than the others. At the upstream location Tr-11 of all the
cylinders Dc = 4, 5, 6 cm, where the scour occurs, contributions from all the quadrant
events were almost identical except for Dc = 6 cm, where interaction (Q1) was greater than
the interaction (Q3). On the other hand, just at the downstream location Tr-13 of Dc = 4
cm, ejection (Q2) [[ sweep (Q4) and left wall-ward interaction (Q1)  right wall-ward
interaction (Q3); for Dc = 5 cm, Q4 [[ Q2 and Q3 [ Q1; for Dc = 6 cm, Q4  Q2 and
Q1  Q3 for all threshold parameter H.
The occurrence of probability Ni;H u0 ; v0 plot shows that at the upstream of scour hole
(Tr-11) near the bed, occurrence of probability was higher for interaction (Q1) and sweep
(Q4) for 4 cm diameter cylinder, considering higher H-values. With increase in cylinder
diameter (blockage ratio), occurrence of sweep events was outshined by outward inter-
actions (Q1). For 6 cm diameter cylinder, highest occurrence of probabilities was observed
by the two interaction events [30]. In this maximum-scoured region, minimum velocities
were generated, which did not have enough energy to transport sediments to further
downstream; hence even if it initiated the sediment movements by lifting smaller sediment
grains up, eventually it failed to progress further downstream and as a consequence sed-
imentation occurred in this region [30]. Interestingly, at side of the cylinder at Tr-7, where

123
688 Environ Fluid Mech (2016) 16:659693

the rate of scour was slightly slower than that in upstream location (Tr-11), the interaction
event (Q3) showed highest occurrence of probability throughout for any H-value. In the
downstream scoured location (Tr-13), where the scour-depth increased with increase in
cylinder diameter, near the bed, ejection (Q2) was the mostly occurring event for the 4 cm
diameter cylinder [30]. Interestingly, occurrence probability of ejections (Q2) was
exceeded by the sweep events (Q4) with increase in cylinder size, considering higher H-
value. The other two events, (Q1 and Q3), possessed nearly the same occurrence proba-
bilities for all three cases. So, in another word, these two events contributed significantly to
balance the sediment entrainment and deposition in a water channel by giving off its space
for occurrence of ejections and sweeps [14, 29]. Outside the scour hole near the bed at
further downstream location (Tr-14), ejections (Q2) and sweeps (Q4) were the mostly
occurring events. Moreover, with the increase in cylinder diameter, probability of occur-
rence of Q2 and Q4 also increased.
The computed transitional probabilities of bursting events analyzed from velocity data
are shown in Table -3 for four locations, viz. side (Tr-7), just upstream (Tr-11), just
downstream (Tr-13) and further downstream (Tr-14) of cylinders respectively for all cases.
It was observed that irrespective of spatial variation, the transitional probability of

Table 3 The transition probability matrix of four bursting events for three cylinder sizes (blockage ratio) at
four selected locations
Location Q1jtDt Q2jtDt Q3jtDt Q4jtDt

4 cm 5 cm 6 cm 4 cm 5 cm 6 cm 4 cm 5 cm 6 cm 4 cm 5 cm 6 cm

Tr-7
Q1jt 52.17 40.24 45.85 21.77 22.47 26.75 10.05 16.47 11.73 16.01 20.82 15.86
Q2jt 26.29 28.83 29.80 39.85 33.84 33.95 22.89 20.22 22.27 10.96 17.10 13.98
Q3jt 11.17 15.59 12.77 18.47 19.94 20.37 45.23 38.55 41.00 25.12 25.91 25.86
Q4jt 22.68 22.13 21.02 12.21 15.64 14.55 25.77 27.89 29.55 39.34 34.33 34.88
Tr-11
Q1jt 39.33 39.52 39.28 22.30 19.43 18.88 14.27 12.96 17.61 24.09 28.09 24.22
Q2jt 23.53 22.10 24.68 41.68 40.56 39.45 23.32 23.56 25.55 11.46 13.77 10.31
Q3jt 13.97 13.05 19.05 29.13 28.41 22.28 39.58 41.03 42.06 17.32 17.51 16.60
Q4jt 23.28 22.03 29.94 13.64 14.59 13.78 24.47 24.85 24.89 38.61 38.52 31.38
Tr-13
Q1jt 46.54 44.51 47.31 27.92 29.41 29.89 8.51 8.69 9.27 16.97 17.39 13.53
Q2jt 25.86 27.54 27.07 47.15 47.30 49.56 14.92 14.37 13.31 12.07 10.78 10.06
Q3jt 10.15 11.42 11.35 20.10 21.98 25.37 42.94 38.11 33.44 26.81 28.49 29.83
Q4jt 16.04 14.99 13.51 11.81 12.65 11.73 22.76 20.79 20.72 49.39 51.56 54.04
Tr-14
Q1jt 51.36 44.68 35.58 24.47 24.43 28.45 8.69 11.03 12.31 15.48 19.85 23.65
Q2jt 29.81 24.91 21.77 40.39 40.23 41.73 19.40 21.26 20.67 10.41 13.60 15.83
Q3jt 8.70 10.55 12.28 18.97 24.28 28.27 45.59 37.89 33.12 26.74 27.28 27.33
Q4jt 17.43 17.77 17.36 11.21 14.71 17.46 24.38 23.42 20.78 46.97 44.09 44.39

123
Environ Fluid Mech (2016) 16:659693 689

stable movements was higher than marginal and cross movements for any cylinder size
[30, 31]. This indicated that the probability of continuous occurrence of any event was
higher than fluctuating between the quadrants. At the scoured region (Tr-7), transitional
probabilities of P11 were maximum for all three cylinder diameters and that of P22 was
minimum. Hence, effective sediment entrainment did not take place towards the down-
stream from there. In the upstream maximum scoured location (Tr-11), maximum transi-
tion probability was shown by P22 for 4 cm diameter cylinder. However, with the increase
of cylinder size, P33 exceeded the transition probability of P22 in this location (Tr-11),
where horse-shoe vortex occurred. Interestingly, P44 gradually decreased with increase in
cylinder size. In the downstream scoured region (Tr-13), interestingly P22 and P44 showed
maximum transition probability [31], which also progressively increased with cylinder size
and lowest transition probability, was shown by P33. This was due to the shedding wake
vortices created just downstream of the cylinder [10, 65]. It is noteworthy that the sum of
the magnitude of P22 and P44 in the downstream was much greater than that of P11 and
P33 in the upstream of the cylinder of any size. Considering further downstream location
(Tr-14), highest transition probability was showed by P11 for smallest cylinder and by P44
for largest cylinder. Clearly, it was seen that with increase in cylinder size, occurrence of
P11 and P33 tremendously reduced. With increase in cylinder size, the location Tr-14
gradually comes closer to the cylinder center and the downstream scoured area, and hence
it plays an important role in suspension and entrainment of sediment particles there to
further downstream location through rolling, sliding, and saltation of sediment particles.

6 Conclusions

A new experimental technique for high-resolution measurement of spatiotemporal scour-


hole and bed form structures around the submerged cylindrical pier-like obstacles due to
turbulent flow was described. The amount of flow perturbation generated due to flow
blockage for the smallest cylinder diameter favored the stable bed condition to achieve
more rapidly than the others, and larger upstream scour-depth and deformed area were
noticed for greater flow blockage due larger cylinder diameters. It was interesting to note
that at upstream of the cylinder, where scour depth was maximum, highest probabilities of
exceedance were observed for largest diameter of cylinder above jdztj = 0.5 mm for D
and above jdztj = 1.0 mm for E. However, at the upstream location of the cylinder, the
exceedance probability of instantaneous Reynolds shear stress was lowest for all three
components of shear stress, but probability for suv was the highest within all other shear
stress components. For suw at Tr-7 and Tr-14, exceedance probabilities were the maximum.
Power spectra of velocity data were calculated at different locations at both upstream
and downstream for both plane and scoured bed conditions to find the dominant vortex-
shedding frequency and the corresponding strength of different kind of vortices. For the
smallest cylinder, the peak power spectra at sides and wakes of the cylinder were found at
the same frequency of 0.703 Hz. With increase in the cylinder diameter, the dominant
vortex-shedding frequency decreased to 0.586 Hz for 5 cm diameter and 0.508 Hz for
6 cm diameter cylinder. The vorticity strength at the flume centerline (along Row0) and at
the near wake region (Tr-9 and Tr-10) downstream of the pier, was maximum. It confirmed
the fastest scour initiation process for the case of smallest cylinder. It was shown that the
strength of wake vortex was maximum for span-wise velocity (v-component) especially
along the cylinder downstream centerline location for plane bed condition, and

123
690 Environ Fluid Mech (2016) 16:659693

interestingly, for the scour bed condition the corresponding maximum value was achieved
at side of the cylinder base (Tr-7) and just downstream of the pier location (Tr-13), thus in
turn it confirmed the tremendous secondary current generation at near-bed region at these
locations.
Quadrant analysis of Reynolds shear stress showed that in the upstream scour hole (Tr-
11) near the bed, occurrence probabilities of left boundary-ward interaction (Q1) and
sweep (Q4) were higher for 4 cm cylinder. But, with increase in cylinder diameter, the
occurrence of sweep event was outshined by two interaction events (prominent for 6 cm
diameter cylinder). However, in the downstream scoured location (Tr-13) near the bed,
ejections (Q2) were the mostly occurring event for the 4 cm diameter cylinder, but
interestingly, with increase in cylinder size, occurrence probabilities of ejection (Q2) were
exceeded by the sweep event (Q4).
The transition probability of bursting events showed that transition probability for
stable movements was higher than marginal and cross movements for any cylinder size at
any location around the pier. In the upstream maximum scoured location (Tr-11), maxi-
mum transition probability was showed by P22 for 4 cm diameter cylinder. However, as
cylinder size increases, P33 exceeded the transition probability of P22 in this location (Tr-
11). In the downstream scoured region (Tr-13), interestingly P22 and P44 showed maxi-
mum transition probability, which also progressively increased with cylinder size and
lowest transition probability was shown by P33.
Here only three experiments using three different cylinder diameters with a common
submergence height were concentrated to study the bed form dynamics. The interactions
between the flow and cylinders embedded on the sandy bed determined the bed form
characteristics and scour-hole structures because of the flow blockage ratio which included
the dimensions of the cylinder. It is important to consider the potential effects of several
flow blockage ratios on the scour hole and bed elevation problems. Therefore, a detailed
study is required considering these factors to quantify the role of these important param-
eters. It is of interest to perform more experiments to improve our understanding of the bed
form structures using different cylinder diameters and submergence ratios; this will be
studied in a separate paper to come out with generalized results.
The present study would be helpful to the civil and mechanical engineers to modify
various pier-like submerged hydraulic system designs, such as cylindrical breakwaters used
for coastal protection, a caisson placed underwater around the pier bottom for scour
protection, and an unfinished bridge pier construction during possible high-flood situation
due to scouring with associated turbulence. Moreover, the information could be utilized as
a database for testing numerical turbulence modeling of sedimentation process around
submerged cylindrical bridge pier-like objects. Erosion changes the morphology of the
river/sea bed and affects the aquatic ecosystems and fish habitats quite sensitively. This
investigation on the hydrodynamics of turbulent flow structure over scoured bed around
such submerged hydraulic constructions contributes to the field of environmental fluid
dynamics.

Acknowledgments Authors would like to express their sincere thanks to the Department of Science and
Technology (DST), New Delhi for approving the Project (No. SERB/ S4/ ES-127/2004) with a financial
support to conduct this experimental research work at the Fluvial Mechanics Laboratory, Indian Statistical
Institute, Kolkata. Authors also like to express their thanks to the Editor-in-Chief, and two anonymous
reviewers for their critical comments and suggestions to improve the paper.

123
Environ Fluid Mech (2016) 16:659693 691

References
1. Absi R (2011) An ordinary differential equation for velocity distribution and dip-phenomenon in open
channel flows. J Hydraul Res 49:8289
2. Afzalimehr H, Moghbel R, Gallichand J, Sui J (2011) Investigation of turbulence characteristics in
channel with dense vegetation. Int J sediment Res 26(3):269282
3. Ali KHM, Karim O (2002) Simulation of flow around piers. J Hydraul Res 40(2):161174
4. Arlinghaus R, Engelhardt C, Sukhodolov A, Wolter C (2002) Fish recruitment in a canal with intensive
navigation, implications for ecosystem management. J Fish Biol 61:13861402
5. Armstrong JD, Kemp PS, Kennedy GJA, Ladle M, Milner NJ (2003) Habitat requirements of Atlantic
salmon and brown trout in rivers and streams. Fish Res 62:143170
6. Atabay S, Knight DW (2002) Bridge afflux experiments in compound channels, R & D project record
W5A061/PR6 (Afflux at bridges and culverts Review of current knowledge and practice, Annex 6).
The Environment Agency, Bristol
7. Ballio F, Radice A (2003) A non-touch sensor for local scour measurements. J Hydraul Res 41:105108
8. Barkdoll B, Huckins C (2012) The role of bridge scour in relation to stream restoration. World Environ
Water Resour Congr 25462555: doi:10.1061/9780784412312.255
9. Best J, Kostaschuk R (2002) An experimental study of turbulent flow over a low-angle dune. J Geo-
physical Res 107(C9):1819. doi:10.1029/2000JC000294
10. Cantwell BJ, Coles D (1983) An experimental study of entrainment and transport in the turbulent near
wake of a circular cylinder. J Fluid Mech 135:321374
11. Catano-Lopera YA, Garcia MH (2007) Geometry of scour hole around, and the influence of the angle
angle of attack on the burial of finite cylinders under combined flows. Ocean Eng 34:856869
12. Catano-Lopera YA, Landry BJ, Garca MH (2011) Scour and burial mechanics of conical frustums on a
sandy bed under combined flow conditions. Ocean Eng 38(10):12561268
13. Catano-Lopera YA, Landry BJ, Abad J, Garca MH (2013) Experimental and numerical study of the
flow structure around two partially buried objects on a deformed bed. J Hydraul Eng 139(3):269283
14. Cellino M, Lemmin U (2004) Influence of coherent flow structures on the dynamics of suspended
sediment transport in open channel flow. J Hydraul Eng (ASCE) 130(11):10771088
15. Cheng S, Liu S, Zheng Y (2003) Application study on submerged breakwaters used for coastal pro-
tection. In: International conference on Estuaries and Coasts, Hangzhou
16. Comiti F, Andreoli A, Lenzi MA (2005) Morphological effects of local scouring in step-pool streams.
Earth Surf Process Landf 30:15671581. doi:10.1002/esp.1217
17. Corino ER, Brodkey RS (1969) A visual investigation of the wall region in turbulent flow. J Fluid Mech
37(1):130
18. Dargahi B (2003) Scour mechanism downstream of a spillway. J Hydraul Res 41(4):417426
19. Debnath K, Manik MK, Mazumder BS (2012) Turbulence statistic of flow over scoured cohesive
sediment bed around circular cylinder. Adv Water Resour 41:1828
20. Dey S, Raikar R, Roy A (2008) Scour at submerged cylindrical obstacles under steady flow. J Hydraul
Eng 134(1):105109
21. Euler T, Herget J (2011) Obstacle-Reynolds-number based analysis of local scour at submerged
cylinders. J Hydraul Res 49(2):267271. doi:10.1080/00221686.2010.547719
22. Euler T, Herget J (2012) Controls on local scour and deposition induced by obstacles in fluvial envi-
ronments. Catena 91:3546
23. Ganti V, Straub KM, Foufoula-Georgiou E, Paola C (2011) Space-time dynamics of depositional
systems: experimental evidence and theoretical modeling of heavy-tailed statistics. J Geophys Res
116:F02011. doi:10.1029/2010JF001893
24. Gaudio R, Marion A (2003) Time evolution of scouring downstream of bed sills. J Hydraul Res
41(3):271284. doi:10.1080/00221680309499972
25. Goring DG, Nikora VI (2002) Despiking acoustic Doppler velocimeter data. J Hydraul Eng
128(1):117126
26. Hoffmans GJCM, Verheij HJ (1997) Scour manual. CRC Press, Boca Raton
27. Hopfinger EJ, Kurniawan A, Graf WH, Lemmin U (2004) Sediment erosion by Gortler vortices: the
scour-hole problem. J Fluid Mech 520:327342
28. Istiarto I (2001) Flow around a cylinder in a scoured channel bed. PhD-thesis Nr 2368, Ecole
Polytechnique Federale de Lausanne
29. Izadinia E, Heidarpour M, Schleiss AJ (2013) Investigation of turbulence flow and sediment entrain-
ment around a bridge pier. Stoch Environ Res Risk Assess 27:13031314. doi:10.1007/s00477-012-
0666-x

123
692 Environ Fluid Mech (2016) 16:659693

30. Keshavarzi A, Gheisi AR (2006) Stochastic nature of three-dimensional bursting events and sediment
entrainment in vortex chamber. J Stochast Environ Res Risk Assess 21(1):7587
31. Keshavarzi A, Melville B, Ball J (2014) Three-dimensional analysis of coherent turbulent flow structure
around a single circular bridge pier. Environmental Fluid Mech 14(4):821847. doi:10.1007/s10652-
013-9332-1
32. Keshavarzi A, Shirvani A (2002) Probability analyses of instantaneous shear stress and entrained
particles from the bed. In: CSCE/EWRI of ASCE Environmental Engineering Conference, Niagara
Falls
33. Kirkil G, Constastantinescu G, Ettema R (2008) Coherent structures in the flow field around a circular
cylinder with scour hole. J Hydraul Eng ASCE 134(5):572587
34. Kolmogorov AN (1951) Solution of a problem in probability theory connected with the problem of the
mechanism of stratification. Trans Am Math Soc 53:171177
35. Lacey RWJ, Rennie DC (2012) Laboratory investigation of turbulent flow structure around a bed-
mounted cube at multiple flow stages. J Hydraul Eng 138(1):7183
36. Link O, Gonzalez C, Maldonado M, Escauriaza C (2012) Coherent structure dynamics and sediment
particle motion around a cylindrical pier in developing scour holes. Acta Geophys 60(6):16891719.
doi:10.2478/s11600-012-0068-y
37. Link O, Pfleger F, Zanke U (2008) Characteristics of developing scour-holes at a sand-embedded
cylinder. Int J Sediment Res 23(3):258266
38. Lohrmann A, Cabrera R, Kraus NC (1994) Acoustic-Doppler velocimeter (ADV) for laboratory use. In:
Pugh CA (ed) Fundamental and advancements in hydraulic measurements and experimentation. ASCE,
Reston, pp 351365
39. Long D, Steffler PM, Rajaratnam N (1990) Study of flow structure in submerged hydraulic jump.
J Hydraul Res 28:437460
40. Lu SS, Willmarth WW (1973) Measurements of the structure of the Reynolds stress in a turbulent
boundary layer. J Fluid Mech 60:481512
41. Maity H, Mazumder BS (2014) Experimental investigation of the impacts of coherent flow structures
upon turbulence properties in regions of crescentic scour. Earth Surf process Landf 39(8):9951013.
doi:10.1002/esp.3496
42. Mazumder BS, Ray RN, Dalal DC (2005) Size distributions of suspended particles in open-channel flow
over sediment beds. Environmetrics 16:149165
43. Mazumder BS, Pal DK, Ghoshal K, Ojha SP (2009) Turbulence statistics of flow over isolated scalene
and isosceles triangular-shaped bedforms. J Hydraul Res 47(5):626637
44. Melville B, Chiew Y (1999) Time scale for local scour at bridge piers. J Hydraul Eng 125(1):5965
45. Mohammed TA, Noor MJMM, Ghazali AH, Yusuf B, Saed K (2007) Physical modeling of local
scouring around bridge piers in erodable bed. J King Saud Univ Engg Sci 19(2):195207
46. Muzzammil M, Gangadhariah T (2003) The mean characteristics of horseshoe vortex at a cylindrical
pier. J Hydraul Res 41(3):285297
47. Nakagawa H, Nezu I (1981) Structure of space-time correlations of bursting phenomena in an open-
channel flow. J Fluid Mech 104:143
48. Ojha SP, Mazumder BS (2008) Turbulence characteristics of flow region over a series of 2D dune
shaped structures. Adv Water Resour 31:561576
49. Oliveto G, Hager W (2005) Further results to time-dependent local scour at bridge elements. J Hydraul
Eng 131(2):97105
50. Parola A, Mahavadi S, Brown B, El Khoury A (1996) Effects of rectangular foundation geometry on
local pier scour. J Hydraul Eng 122(1):3540
51. Pope SB (2000) Turbulent flows. Cambridge University Press, Cambridge
52. Sarkar K, Chakraborty C, Mazumder BS (2015) Spacetime dynamics of bed forms due to turbulence
around submerged bridge piers. Stoch Environ Res Risk Assess 29(3):9951017. doi:10.1007/s00477-
014-0961-9
53. Salaheldin T, Imran J, Chaudhry H (2004) Numerical modeling of three-dimensional flow field around
circular piers. J Hydraul Eng ASCE 130(2):91100
54. Seckin G, Knight DW, Atabay S, Seckin N (2004) Bridge afflux experiments in compound channels,
Unpublished technical paper presented for JBA consulting Engineers & Scientists and the Environment
Agency
55. Sengupta S (1966) Studies on orientation and imbrication of pebbles with respect to cross-stratification.
J Sediment Petrol 36(2):362369
56. Sengupta S (2007) Introduction to sedimentology. CBS Publications and Distributors, New Delhi

123
Environ Fluid Mech (2016) 16:659693 693

57. Singh A, Lanzoni S, Wilcock PR, Foufoula-Georgiou E (2011) Multiscale statistical characterization of
migrating bed forms in gravel and sand bed rivers. Water Resour Res 47:W12526. doi:10.1029/
2010WR010122
58. Singh A, Foufoula-Georgiou E, Porte-Agel F, Wilcock PR (2012) Coupled dynamics of the co-evo-
lution of gravel bed topography, flow turbulence and sediment transport in an experimental channel.
J Geophys Res 117:F04016. doi:10.1029/2011JF002323
59. SonTek Inc. (2001) ADV principles of operation, technical document, San Diego
60. Sumer BM, Fredsoe J (2002a) The mechanics of scour in the marine environment. World Scientific,
River Edge, p 552
61. Sumer BM, Fredsoe J (2002b) Time scale of scour around a large vertical cylinder in waves. In:
Proceedings of the 12th International Offshore Polar Engineering Conference, pp 5560
62. Termini D (2005) Experimental investigation on the horizontal turbulence and the bed deformation:
preliminary results. In: Proceedings of the international symposium stochastic hydraulics, IAHR
Congress, Nijmegen
63. Termini D, Lo Re C (2006) Analysis of the relation between the flow horizontal turbulence and the
bed deformation. In: Proceeding international symposium on sediment dynmics hydromorphology
fluvial systems, Dundee
64. Termini D, Sammartano V (2008) Experimental analysis of relations between coherent turbulent
structures and formation of bed forms. Arch Hydro-Engg Environ Mech 55(34):125143
65. Termini D, Sammartano V (2012) Morphodynamic processes downstream of man-made structural
interventions: experimental investigation of the role of turbulent flow structures in the prediction of
scour downstream of a rigid bed. Phys Chem Earth 49:1831. doi:10.1016/j.pce.2011.12.006
66. Tachie MF, Balachandar R, Bergstrom DJ (2004) Roughness effects on turbulent plane wall jets in an
open channel. Exp Fluids 37:281292
67. Tregnaghi M, Marion A, Coleman S (2009) Scouring at bed sills as a response to flash floods. J Hydraul
Eng 135(6):466475
68. Tregnaghi M, Marion A, Coleman S, Tait S (2010) Effect of flood recession on scouring at bed sills.
J Hydraul Eng 136(4):204213
69. Tregnaghi M, Marion A, Bottacin-Busolin A, Tait SJ (2011) Modelling time varying scouring at bed
sills. Earth Surf Process Landf 36:17611769. doi:10.1002/esp.2198
70. Tsutsui T. (2008) Fluid force acting on a cylindrical pier standing in a scour. In: Bluff bodies aero-
dynamics & applications, Milano, BBAA VI international colloquium, pp 2024
71. USGS (1978) Hydrologic Investigation atlases HA591HA611. Department of the Interior, Denver
72. Voropayev SI, Testik FY, Fernando HJS, Boyer DL (2003) Burial and scour around a short cylinder
under progressive shoaling waves. Ocean Eng 30:16471667
73. Wallerstein NP (2003) Dynamic model for constriction scour caused by large woody debris. Earth Surf
Process Landf 28:4968. doi:10.1002/esp.426
74. Wu Y, Christensen KT (2006) Reynolds-stress enhancement associated with a short fetch of roughness
in wall turbulence. AIAA J 44(12):30983106
75. Yang SQ, Tan SK, Lim SY (2004) Velocity distribution and dip phenomenon in smooth uniform open
channel flows. J Hydraul Eng ASCE 130(12):11791186
76. Yano H (2003) Masters thesis (in Japanese). Osaka Kyoiku University, Kashiwara
77. Yasunori M (2008) Local scour around a submerged cylindrical pier. In: 4th international conference on
scour and erosion (A-13)
78. Young DM, Testik FY (2009) Onshore scour characteristics around submerged vertical and semicircular
breakwaters. Coast Eng 56:856876
79. Young DM, Testik FY (2011) Wave reflection by submerged vertical and semicircular breakwaters.
Ocean Eng 28:12691276
80. Zhao Z, Fernando HJS (2007) Numerical simulation of scour around pipelines using an EulerEuler
coupled two-phase model. Environ Fluid Mech 7(2):121142

123

You might also like