Vous êtes sur la page 1sur 15

Experimental and Numerical Study of the Flow Structure

around Two Partially Buried Objects on a Deformed Bed


Yovanni A. Cataño-Lopera1; Blake J. Landry2; Jorge D. Abad3; and Marcelo H. García, M.ASCE4

Abstract: The flow characteristics around two partially buried objects, a short cylinder (SC) and a truncated cone (TC) were examined
experimentally and numerically. Experiments were conducted in a large tilting flume with vertical walls made of acrylic plastic and steel floor
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

covered by a 24-m long sand pit. For each experiment, the object was placed on a flat sandy bed, the flume was filled to a water depth of
41 cm, and a steady unidirectional current was established at high Reynolds numbers while sinking of the object was tracked over time.
Velocity records were taken with an acoustic Doppler velocimeter (ADV) probe at the final stage of burial and scour hole evolution. The flow
structure around each object was further investigated using a model that solves the Reynolds-averaged Navier-Stokes (RANS) equations
coupled with a κ-ε turbulence closure model, and results were compared to the experimental measurements. Measurements and numerical
simulations allowed identification of regions of high and low velocity and turbulent kinetic energy accounting for zones of potential sediment
erosion and deposition that dictate the burial of the object. The κ-ε turbulence closure model was shown to reproduce the experimental mean
flow velocity and turbulent kinetic energy with acceptable accuracy. DOI: 10.1061/(ASCE)HY.1943-7900.0000619. © 2013 American
Society of Civil Engineers.
CE Database subject headings: Experimentation; Numerical analysis; Scour; Sediment transport; Turbulent flow; Flow rates; Cylinders.
Author keywords: ADV; CFD; Experiments; Object burial; Short cylinder; Scour hole; Sediment transport; Truncated cones; Turbulent
flows.

Introduction A significant number of experimental investigations have been


made on the scour and burial of two-dimensional (2D) objects such
To adequately predict scouring and burial of three-dimensional as pipelines in the field (Sumer and Fredsøe 2002) and in the labo-
(3D) objects placed on river and marine floors, the properties of ratory (Cevik and Yuksel 1999) under shoaling waves. Investiga-
the flow field around the object and the morphologic characteristics tions dealing with 3D objects on the sea floor were carried out by
of the bed must be adequately described. Under unidirectional McCarthy and Sabol (2000), Mayer et al. (2007), Traykovski et al.
flows, several investigations have focused on the flow field around (2007), and Guyonic et al. (2007), and in laboratory conditions
surface mounted obstacles: cubes, rectangular prisms, pyramids under oscillatory flows by Voropayev et al. (2003) and Cataño-
(Schofield and Logan 1990; Hussein and Martinuzzi 1996; Lopera et al. (2007), among others. Truelsen et al. (2005) studied
Martinuzzi and Tropea 1993; Martinuzzi and AbuOmar 2003), the scour and self-burial of mounted spheres under both steady
and spherical bodies (Thomschke 1971; Bagnold 1974; Willets currents and wave conditions. Testik et al. (2005) investigated
and Murray 1981). However, few investigations have been dedi- the flow field around short cylinders of different aspect ratios
cated to 3D objects on a movable bed subjected to scouring, either resting on a fixed bed, using particle image velocimetry (PIV)
under steady currents or oscillatory flows. and streak photography, under steady and oscillatory flows.
The use of acoustic technology has increased over the last de-
cade to measure the flow field around objects. Acoustic Doppler
1
Hydrodynamic Modeler, H2MHILL Inc., Boston, MA 02108; velocity profiler (ADVP) measurements were conducted to inves-
formerly, Visiting Research Associate, Ven Te Chow Hydrosystems tigate the mean velocity components and turbulent Reynolds
Laboratory, Univ. of Illinois at Urbana-Champaign, 205 N. Mathews stresses around a circular pier at equilibrium scour conditions
Ave., Urbana, IL 61801 (corresponding author). E-mail: yovanni100@
by Graf and Istiarto (2002). More recently, detailed ADV measure-
yahoo.com
2
Research Associate, Ven Te Chow Hydrosystems Laboratory, Univ. of
ments have been used by Raikar and Dey (2008) to characterize the
Illinois at Urbana-Champaign, 205 N. Mathews Ave., Urbana, IL 61801. flow field around a vertical square cylinder at intermediate and final
E-mail: blandry2@illinois.edu stages of scour hole evolution. Other studies in which ADVs
3 have been used include those of Song and Chiew (2001), García
Assistant Professor, Dept. of Civil and Environmental Engineering,
Univ. of Pittsburgh, 3700 O’Hara St., Benedum 943, Pittsburgh, et al. (2005), and Abad and García (2009) in the laboratory, and
PA 15261. E-mail: jabad@pitt.edu those of Nikora and Goring (2000) and Strom and Papanicolaou
4
Chester and Helen Siess Professor, Ven Te Chow Hydrosystems (2007) in natural streams. However, to date, no extensive experi-
Laboratory, Univ. of Illinois at Urbana-Champaign, 205 N. Mathews mental investigations have been performed with velocity measure-
Ave., Urbana, IL 61801. E-mail: mhgarcia@illinois.edu
ments around partially buried objects such as truncated cones and
Note. This manuscript was submitted on November 15, 2010; approved
on July 12, 2012; published online on February 15, 2013. Discussion per- short cylinders on a movable bed.
iod open until August 1, 2013; separate discussions must be submitted for Likewise, the development of numerical schemes and the use of
individual papers. This paper is part of the Journal of Hydraulic Engineer- computational fluid dynamic (CFD) models to investigate the scour
ing, Vol. 139, No. 3, March 1, 2013. © ASCE, ISSN 0733-9429/2013/3- mechanisms around 2D and 3D objects are rapidly increasing.
269-283/$25.00. Early efforts have focused on the hydrodynamics of the flow

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 269

J. Hydraul. Eng. 2013.139:269-283.


around objects mounted on a flat immobile bed, whereas more Rodi (1997) of the flow around a surface-mounted cube using
recent investigations have been extended to the study of the any of the κ-ε formulations underpredicted the upstream separation
flow structure over scoured regions around a variety of objects. and downstream reattachment lengths.
Over time, the numerical models have become more sophisticated Numerical modeling of conditions similar to those of the experi-
to deal with more complex flow domains, and the availability of ments documented by Testik et al. (2005) and Traykovski et al.
experimental data has increased. (2007) were conducted by Smith and Foster (2007) and Hatton
In the case of 2D objects, Liang and Cheng (1999, 2005) re- et al. (2007), respectively. Hatton et al. (2007) investigated the flow
ported numerical simulations of local scour and lee-wake scouring field around a 10% partially buried short cylinder on an immobile
of offshore pipelines. Li and Cheng (2001) and Liang and Cheng flat bed and compared the results to the field observations of
(2005) conducted numerical simulations of the scour under off- Traykovski et al. (2007). Smith and Foster (2005) used FLOW-
shore pipelines under steady currents. Salaheldin et al. (2004) used 3D to simulate the scouring underneath pipelines under oscillatory
the commercial software FLUENT to simulate the flow field flows using both RANS and LES approaches. The numerical
around circular piers on a scoured bed using Reynolds-averaged results were in reasonably good agreement with the experimental
Navier-Stokes (RANS) equations with a κ-ε turbulence closure observations. Jenkins et al. (2007) and Imman and Jenkins (2002)
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

model and validated their model with flume data from Ahmed presented a comprehensive literature review of numerical investi-
and Rajaratnam (1998). Other RANS-based studies of the flow gations dealing with simulated partially buried objects on the sea
around piers, such as that of Roulund et al. (2005), which used floor.
a κ-ω turbulence closure model, investigated the steady and The present study advances the understanding of the hydrody-
unsteady features of the flow field. Modeling the evolution of namics around partially buried objects through combined experi-
the scour hole has also been attempted by other researchers with mental and numerical investigations. An ADV probe was used
relative success in reproducing the main scour pattern characteris- to characterize the mean and turbulent characteristics of the flow
tics, including the studies of Wei et al. (1997), Olsen and Kjellesvig around a partially buried object on a scoured bed under unidirec-
(1998), Chen (2002), Nagata et al. (2005), and Zhao et al. (2010). tional flow. The numerical simulations were validated against the
Richardson and Panchang (1998) used RANS with the renormal- experimental results and showed fair agreement. The numerical
ization group theory (RNG) model as a turbulence closure scheme simulations allowed further exploration of the flow field, particu-
to simulate the flow within a scour hole at the base of a cylindrical larly in areas closer to both the object and bed that were not directly
bridge pier using the software FLOW-3D and compared their accessible through velocity measurements. The flow field was not
results with the laboratory observations of Melville and Raudkivi only modified by the object, but also by the shape and size of
(1977), obtaining acceptable agreement. Other numerical tech- the scour pattern defined by holes, ridges, and the surrounding
niques have been explored in recent years, especially those dedi- small-scale ripples. These results highlight the importance of the
cated to the unsteadiness associated with the coherent flow mean flow characteristics and turbulent intensities on the mobility
structures around obstacles. Unsteady RANS (URANS), along of sediments around the object, which impact the extent and
with a κ-ω turbulence closure scheme, was used by Chrisohoides development of the scour hole. This study may be used to predict
et al. (2003) to study the vortices induced by a vertical bridge abut- the behavior of mantalike and cylinderlike sea antitank mines
ment on a flat bed. Koken and Constantinescu (2008) used large placed in the continental shelf and objects resting on river beds.
eddy simulations (LES) to characterize the flow around a vertical The present findings will allow further development of mathemati-
spur dike in a straight channel with equilibrium scour conditions cal models of the local scour and burial mechanisms of 3D objects.
and showed that large-scale bimodal oscillations of the primary Furthermore, this experimental data set is of importance for vali-
necklace vortex are the mechanism mainly responsible for the am- dation of present and future computer models dealing with the
plification of the turbulence inside the scour hole. More recently, complex interplay of fluid, sediment, and object resting on a river
Kirkil and Constantinescu (2010) studied the unsteady dynamics of or ocean floor.
coherent structures around an in-stream vertical cylinder of rectan-
gular section. Other techniques, such as detached eddy simulations
(DES), have successfully captured the broad range of eddies asso- Experimental Setup and Test Conditions
ciated with the vortex shedding in the wake of a bridge abutment on
a flat bed, as shown by Paik and Sotiropoulos (2005). Paik et al. The flume used in this study was 49 m long, 1.8 m wide, and 1.2 m
(2007) used DES to study the bimodal dynamics of the horseshoe deep, with a steel floor and acrylic plastic sidewalls. The 31-cm-
vortex (HV) at the base of a wing-body junction. Although these deep sand pit spanned 24 m and covered the entire width of the
simulations reproduced the main characteristics of the flow field flume. The sediment was a well-sorted size distribution of silica
observed in the experiments reported by Devenport and Simpson sand with median diameter of d50 ¼ 0.25 mm, internal angle of
(1990) fairly well, the model was unable to accurately reproduce repose of φ ¼ 32.5°, and sediment porosity of λp ¼ 0.3. The
the location of the HV system. The model results placed the HV too longitudinal slope of the flume was set to zero, and the steady flow
far upstream of the object compared to the experimentally observed was set to a depth of h ¼ 41 cm above the sandy bed. The object
case. The location of the HV seems to be influenced by whether or was placed at the center of the flume at a distance of 13 m from the
not random fluctuations are added to the inflow of the numerical origin of the sand pit, adequate for flow development.
domain. The experimental truncated cone (TC) was made of a combina-
Rodi (1997) compared several κ-ε based RANS and LES mod- tion of foam and concrete, to reproduce a specific gravity similar to
els with laboratory observations of the flow around a free rectan- prototype manta mines with γ c ¼ 2.0 and center of mass set nearer
gular cylinder and a surface-mounted cube from Martinuzzi and the base, whereas the short cylinder (SC) was made of concrete and
Tropea (1993). Both RANS- and LES-based models have been had a specific gravity of 2.3.
shown to reproduce the experimental length of the vortex shedding The water temperature was measured for each experiment, and
in the wake of the object with accuracy. However, the variation corrections to water viscosity and density were made accordingly.
in the amplitude of shedding is not well captured when using Before running each experiment, the object was gently placed over
the standard κ-ε closure scheme. The simulations conducted by an initial flat bed. Next, to prevent the scouring of the bed around

270 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


the object, the water level was raised slowly 41 cm above the sand that was corrected by temperature T for each experiment, 18.5 and
bed. The initial burial depth was estimated to be Bdo ¼ 0.2 and 19.1°C in the SC and TC cases, respectively. The object Reynolds
0.3 cm for the TC and SC, respectively. number was calculated in the SC case as Rd ¼ UD=v, where D is
A 3D SonTek 10 MHz ADV probe was employed to measure the cylinder diameter, whereas in the TC case was calculated as
the underlying velocity field. The instrument consisted of a sound Rd ¼ UDm =v, where Dm is the average between top and bottom
transmitter, three sound receivers, and a signal processing module diameters of the cone. The Shields parameter was computed using
(Sontek Inc. 2001). Scattering of the signal occurred within a finite the formula θ ¼ u2 =½ðγ s − 1Þgd50 , where γ s is the specific gravity
water volume of approximately 0.1 cm3 and located approximately of the sediment, g is the acceleration attributable to gravity, and the
5 cm below the tip of probe (García et al. 2005). The user sampling shear velocity u is calculated after curve fitting of the velocity pro-
rate was set to 25 Hz and the ADV velocity range was set to file measured 2 m upstream of the object by using the expression
30 cm=s, based on past experimental investigations (Sontek Inc. uðzÞ ¼ ðu =kv Þ lnðz=zo Þ, where kv ¼ 0.41 is the von Kármán con-
2001; Voulgaris and Trowbridge 1998). Velocity measurements stant and zo is the roughness length. The total bed shear stress is
were recorded during 2 min in regions away from the object related to the shear velocity that depends on the total roughness
and during 3 min in regions near the object characterized by length zo , which is the summation of the components attributable
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

relatively higher shear stresses and flow reversals. to skin friction zos , form drag zof , and near-bed sediment transport
The ADV was attached to two Velmex BiSlide frame assemblies zot . For simplicity, only skin friction height, estimated as
that enabled vertical (z-axis) and transversal (y-axis) displace- zos ¼ 2.5d50 , is considered for the computation of u in Table 1.
ments. The frame assemblies were mounted on a movable carriage Other variables are defined in the “Notation” section.
allowing for motion along the longitudinal (x-axis) direction of the The critical Shields parameter, above which sediment was en-
flume. Figs. 1(a and b) depict the spatial locations of the recorded trained into suspension because of the action of the flow, was ap-
velocity profiles with respect to the object and flow direction for the proximately θcr ¼ 0.041, as estimated from the Shields diagram
SC and TC, respectively. In the case of the SC, a total of 72 velocity (García 2008). The Froude numbers associated pffiffiffiffiffi with the meanpflow
ffiffiffiffiffiffi
profiles were taken with ADV, whereas in the case of the TC, a total and the object are calculated using F ¼ U= gh and Fd ¼ U= gD,
of 65 velocity profiles were recorded. In each case, an additional respectively. The Strouhal number was computed as Sh ¼ D=UT s ,
velocity profile, located 2 m upstream of the object at y ¼ 0 cm, where T s is the period of the vortex shedding oscillation. Fast
was recorded to characterize the approaching velocity field with Fourier transform (FFT) analysis of the ADV velocity signals per-
average velocities of U ¼ 21.8 and 19.4 cm=s for the SC and mitted identifying not a fully periodic but rather an aperiodic vortex
TC, respectively. shedding downstream in the wake of the object for which the
Each vertical velocity profile was composed of 12 velocity point Strouhal number was computed to be 0.41 and 0.36 for the SC
records, with the lowest point taken at approximately 1.7 cm above and TC, respectively (Table 1). These values are comparable to
the local bed level at equilibrium morphodynamic conditions experimental values in the case of a short cylinder under a unidi-
(EMC). In this paper, EMC refers to the final stages of scour evo- rectional current (Testik et al. 2005) and to those obtained using
lution in which, on average, sediment entrainment and deposition, CFD modeling using a LES turbulent closure scheme (Smith
triggered by localized near-bed turbulent bursting phenomena, are and Foster 2007).
in equilibrium, and any further object burial is negligible. A Canon
electro-optical system (EOS) 20D digital single-lens reflex (SLR)
camera was used to capture intermediate and final configurations of ADV Data Quality Analysis
the scour region around the object. The ADV probe was also used
to monitor the burial of the specimen over time. On average, burial To define whether or not the recorded velocity time series are of
equilibrium conditions were attained after approximately 90 min adequate quality to describe mean flow velocities and turbulent
from the start of the run. On the presumption of flow symmetry parameters, the velocity signals acquired with the ADV system
around the object, velocities were recorded on a single side of the were analyzed. The first step was to determine if the length of
specimen, starting from the center, as shown in Figs. 1(a and b). the signal was long enough to compute the statistical moments as-
The shape of the scour pattern around the object assured the sym- sociated with the flow (García et al. 2005; Abad and García 2009).
metric condition presumption was correct. Once the bed configu- A running mean analysis for each recorded time series was per-
ration and the object reached equilibrium conditions, velocity formed to verify that flow velocities are computed within a 95%
measurements around the object were taken. confidence interval. For the case of x-direction velocities (u),
Bathymetry in the vicinity of the objects was recorded using two the convergence time within a 95% confidence level is generally
optical methods. A Keyence laser displacement sensor (model several tens of seconds. In the case of the vertical (w) and transverse
LB-301) with a spot diameter of 1.2 × 2.5 mm, maximum response (v) components, the convergence time is typically much longer
frequency of 915 Hz, and vertical resolution of 50 μm was em- than in the x-direction (Abad and García 2009). The analysis re-
ployed to acquire detailed measurements of the final configuration vealed that for most cases, the convergence times are smaller than
of the bottom topography around the SC and TC. Figs. 2(a and b) the total signal sample time (e.g., of at least 160 s for each position),
show the recorded bathymetries at EMC for the SC and TC, respec- demonstrating that the recorded signals are all adequate for mean
tively. The laser sensor was temporarily attached to the same frame and turbulent flow characterization. The analysis also revealed that
as the ADV, after the water in the tank was slowly drained to avoid the length of the recorded signals was less demanding for the TC
disturbing the bed morphology. Laser scans were performed with than for the SC, because shear and turbulent intensities are in
grid resolution of 0.5 × 1 cm in the transverse and longitudinal general larger in the latter case. The flow distortions induced by
directions, respectively. the sharp edges of the SC, as the cylinder rests perpendicular to
A list containing the dimensionless flow parameters used in the the current, are more intense than those generated by TC.
present study is shown in Table 1. All listed parameters were cal- The second step, to define the quality of the measured data,
culated using the approaching depth-averaged flow velocity U demands filtering of the recorded velocity signals. The filtering
measured approximately 2 m upstream of the object. The Reynolds process involves despiking of aliased points within the signal
number was computed as R ¼ Uh=v, where v is the water viscosity (Goring and Nikora 2002) and removing points of low correlation

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 271

J. Hydraul. Eng. 2013.139:269-283.


Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 1. Schematic plan view of the location of vertical velocity measurements around (a) SC; (b) TC; note: dimensions are in cm

272 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 2. (Color) Laser scan image of the scour pattern at equilibrium conditions around (a) the SC; (b) the TC; note: object dimensions and flow
characteristics are presented in Table 1

Table 1. Summary of Object and Flow Properties by García et al. (2005) by means of the parameter F ¼ LS f R =U c,
Variable/parameter Short cylinder Truncated cone where Ls is an energy containing length scale, fR at 25 Hz is sam-
pling frequency of the ADV, and U c is the convective velocity.
h (cm) 41 41
In the experiments, global values of F ¼ 45.8 and 51.8 were cal-
Uðcm=sÞ 21.8 19.4
u ðcm=sÞ 1.4 1.5
culated for both the SC and TC after setting Ls equal to the mean
Red 21,520 28,579 water depth of 41 cm and U c ¼ 21.8 and 19.4 cm=s, respectively.
Re 86,081 76,209.3 The value F > 20 implies that the ADV, a priori, describes turbu-
Sh 0.41 0.36 lent flow characteristics with good accuracy (García et al. 2005;
Fd 0.19 0.16 Abad and García 2009), as long as the filtered signals are also
F 0.11 0.10 of good quality. The value of F was further investigated locally
F 45.8 51.8 as a requirement for turbulence representation. In the latter case,
θ 0.049 0.053 the convective velocity was assumed to be U c ¼ u 0 , the local aver-
θcr 0.041 0.041 age longitudinal velocity (García et al. 2005). Because a SonTek
hc (cm) — 10
ADV probe with fR ¼ 25 Hz with a sampling time between
Lc 20 —
Dbc (cm) — 20 160 and 120 s was used, F was larger than 20 for most velocity
Dtc (cm) — 10 records.
Bdo (cm) 0.3 0.2 In the case of mean flow characterization, past studies have
D (cm) 20 — shown that mean ADV velocities are typically accurate within
Bd =D, Bd =hc 0.15 0.05 1% (Voulgaris and Trowbridge 1998; Abad and García 2009).
γc 2.3 2.0 As recommended by García et al. (2005), an additional step con-
ks (mm) 0.3 0.3 sisted in removing the noise from the recorded signals to further
d50 (mm) 0.25 0.25 assure that the velocity signals were of enough quality to properly
T (°C) 18.5 19.1 describe mean and turbulent flow characteristics. Henceforth, the
σdg 1.28 1.28
contribution of noise to the signal was removed by integrating
φ 32.5° 32.5°
λp 0.3 0.3 the average value of the power spectrum at the noise plateau
across the range of frequencies (f) defined by 0 < f ≤ fR =2 follow-
ing the methodology proposed by Voulgaris and Trowbridge
(1998). In addition, an inverse Fast Fourier transform of the cor-
coefficient (COR) and low signal-to-noise ratio (SNR) (Sontek Inc. rected power spectrum was performed to estimate the autocorrela-
2001). The threshold values of COR ¼ 70 dB and SNR ¼ 15 dB tion function, convective velocity, length, and time scales (García
are recommended by the manufacturer of the ADV (Sontek Inc. et al. 2005).
2001), and should not be taken as universal but rather as guidelines
for analysis, because the value of COR depends on the local flow
conditions (Strom and Papanicolaou 2007). A low COR value is Numerical Model Setup
not necessarily indicative of a low quality point, but could be
the result of sampling a flow with strong local fluctuations asso- The numerical simulations of the flow field around the two objects
ciated with higher turbulence intensities. In this study, the majority were performed with the three-dimensional and nonhydrostatic
of the velocity records are characterized by SNR > 15 dB, largely model FLOW-3D (Flow Science Inc. 2009), which solves the tran-
because of the presence of a large quantity of fine particles within sient 3D Navier—Stokes and continuity equations. The model
the flow, whereas most velocity records exhibit COR values greater resolves fluid—fluid and fluid—air interfaces with a nonboundary
than 60. fitted rectangular grid and volume of fluid (VOF) approach that
Next, for turbulent flow description, dependence of bad resolves the grid cells into separate fractional fluid components
data points was investigated following the procedure proposed containing the fraction of fluid and nonfluid fraction in the cell

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 273

J. Hydraul. Eng. 2013.139:269-283.


(Hirt and Nichols 1981). The fluid fraction function Ff is defined indicating that a statistically steady state was attained. Therefore,
to be equal to 1.0 in the fluid and 0.0 outside fluid, i.e., in the mean flow values of velocity and TKE were computed between 70
void (Flow Science Inc. 2009). Similarly, a fractional area-volume and 80 s.
obstacle representation (FAVOR) approach is used to parameterize The simulations considered clear water conditions in which
the flow within cells, which contain fluid-obstacle boundaries. The interaction of sediment particles with the flow coherent structures
FAVOR approach uses three-dimensional quadratic equations to is neglected, and no morphodynamic changes are considered. This
define complex obstacle geometries. These geometries are speci- assumption is supported by the findings of studies that showed
fied separately from the numerical mesh (Richardson and Panchang the influence of sediment particles on the flow turbulence in the
1998). Details on the κ-ε turbulent closure scheme used in this vicinity of the object and within the scour hole is negligible for
study are presented in Smith and Foster (2007). Wilcox (2000) low sediment concentrations (Kirkil et al. 2008). Low sediment
provides the closure equations for the turbulent kinetic energy (k) concentrations are characteristic at EMC, during which the ADV
and the dissipation rate of turbulent energy (ε) with standard co- measurements were conducted and were corroborated by visual
efficients. Additional applications of FLOW-3D in morphodynamic inspection. The FLOW-3D model uses a variable time step to
settings are described in Abad et al. (2008). preserve numerical stability at all times using the Courant number
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

The numerical setup consists of the measured bathymetry, criterion. Each simulation was performed using eight cores on one
which includes the sand bed and the object at equilibrium dual quad-core Intel Xeon E5354 and took approximately 300 h of
conditions embedded in a Eulerian rectangular structured mesh. wall-clock time.
Figs. 2(a and b) show the recorded bathymetries at EMC for the
SC and TC, respectively. The mesh consisted of cells of variable
size, transitioning from finer cells in the object vicinity to linearly Analysis of the Flow Field Structure around the
increasing coarser cells toward the boundary of the domain. Objects
Initially, two mesh sizes were tested. The first mesh consisted of
0.5-cm-sided regular cubic cells spanning approximately five ob- Average Flow Velocities
ject diameters around the object in the streamwise x-direction, three
diameters in the transverse y-direction, and two diameters in the Fig. 3 shows contour plots of measured and modeled dimensionless
vertical z-direction. A second mesh consisted of 0.3-cm-sided regu- mean flow streamwise velocities along the center plane of each
lar cubic cells covering a similar spatial domain. In both cases, the object on the xz plane. The vectors in the left frames represent
cell sizes increased linearly toward the edges of the flow domain. the resultant of the three measured velocity components, and their
The finer mesh spacing in the vicinity of the object was found tails represent the measurement location. Mean dimensionless
necessary to better resolve the complex flow. For both objects, the velocities (U nd , V nd , W nd ) are computed by dividing each time
coarser and finer grids used a total of 2,730,000 and 4,080,000 averaged velocity component by the mean streamwise velocity U
cells, respectively. The simulations showed that the mean flow (Table 1), as
structure, mean velocities, and turbulent kinetic energy were in U nd ¼ ū=U; V nd ¼ v̄=U; W nd ¼ w̄=U ð1Þ
better agreement with the experimental data for the finer mesh
cases. The analysis presented subsequently corresponds to the case where in the case of the experiments, ū ¼ ð1=NÞΣNi¼1 ui is the time
of simulations conducted with the finer mesh. averaged velocity in the streamwise direction; and N = total number
The boundary conditions (BC) were set in the following manner. of instantaneous ADV measurements.
A free-surface boundary condition was imposed at the top of the Experiments and numerical simulations identified several simi-
domain for a mean water depth of 41 cm, measured from the initial larities and differences of the flow structure around the two objects.
bed level. In FLOW-3D, free surfaces are modeled with the VOF Fig. 3 exhibits contours of measured (left frames) and simulated
technique (Hirt and Nichols 1981). A symmetry boundary condi- (right frames) dimensionless streamwise velocities along the center
tion was imposed at the lateral boundaries of the model domain. plane of the objects. Fig. 3 shows that U nd exceeds by approxi-
The bottom of the domain was set to a nonslip boundary condition mately 30 and 50% the average velocity U immediately above
and consisted of the measured bathymetry for which a roughness the SC and TC, respectively. Conversely, significant flow deceler-
length of κs ¼ 0.254 mm, equivalent to the mean grain size of the ation and reversal occurs upstream and downstream of the object.
sand bed, was used. This roughness length was also set for the The numerical results shown in the right frames of Fig. 3 reveal a
bathymetric region corresponding to the object because the con- smoother transition between contour zones when compared to the
crete surface of the two objects had a mean roughness length of experimental frames because of the much larger density of velocity
0.3 mm. A continuative BC was specified at the downstream points offered by the numerical mesh. For both objects, the most
end of the computational domain where the specific experimental significant distortion of the longitudinal velocity occurs along the
flow characteristics are unknown. In this type of boundary condi- center plane and decays in the transverse direction. Recirculation
tion, the gradients of all flow parameters in the direction normal to zones, indicated by the reversal of the velocity vectors, emerge
the flow are set to zero (Flow Science Inc. 2009). A uniform flow immediately downstream (front) and upstream (back) of the object.
field with mean velocity equal to the mean flow velocity was set as Sediment transported from the back of the object deposits along
initial conditions (IC). Streamwise velocities and turbulent inten- the downstream recirculation zone that is formed as the flow passes
sities obtained from a preliminary run of a fully developed flow the object. Although the experimental measurements reveal regions
over a flat bed were set at the inflow boundary. The other BCs of flow deceleration downstream of both obstacles as shown on
in the preliminary run were set similar to the aforementioned the left frames of Fig. 3, there are regions that lack velocity mea-
description, using uniform mean velocity profiles of 21.80 and surements very close to the bottom. The numerical simulations,
19.38 cm=s as upstream BC for the SC and TC, respectively. however, overestimate the size of the downstream recirculation
The total simulation time, including the preliminary runs, was zones. Accumulation of sediment forming the ridge downstream
near 80 s, much longer than the flow-through period of approxi- of the object helps stabilize the object and prevents it from rolling
mately 6 s. Analysis of contour plots of velocity and turbulent backward into the scour hole. The combination of vertical and
kinetic energy (TKE) showed negligible variations after 70 s, lateral recirculation zones, along with vortex structures around

274 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


the object, are responsible for the accumulation of sediment that modeled pointwise streamwise dimensionless velocities at selected
forms the central sand ridge [Figs. 2(a and b)]. This sand ridge ex- locations in Figs. 1(a and b). Mean and RMS errors between mea-
tends for approximately one and two object diameters downstream sured and modeled velocities at several representative locations
of the TC and SC, respectively [Figs. 2(a and b)]. around each object were tabulated, although are not shown here
The center of the recirculation zone is located near the bed for simplicity. The locations chosen for analysis are mainly along
and on the upstream face of the scour hole (right frames in transects y ¼ 0 cm and x ¼ 0 cm, characterized by the most
Fig. 3). A region of accelerated flow atop the object, characterized dramatic flow distortions exhibited in Fig. 3. The analysis shows
by low pressure that in turn originates a lift force that competes that mean and RMS errors of streamwise velocities fall within 20%,
against the weight of the object, is shown in Fig. 3. This whereas errors are larger for the vertical velocities and largest for
region, shown in the numerical results frames, is also evidenced the transversal velocities. The best agreement, between model and
by the larger vectors nearer the top of the object on the left experimental values, occurs upstream of the object and outside of
frames, particularly in the case of the TC. For both objects, the scour hole. In the vicinity of the object, discrepancies increase
a good qualitative agreement between measured and modeled toward the solid boundaries, e.g., at locations P69 and P26 in the
dimensionless velocities is observed away from the solid SC case, and at P55, P44, and P03 in the TC case (Fig. 4). Better
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

boundaries. agreement in streamwise velocities is observed in both object cases


At the final stages of scour hole evolution, recirculating flow for z > 5 cm. An 8% difference is observed in the measured and
velocities within the scour hole are no longer energetic enough modeled streamwise velocities at location P01 in the SC case. In the
to induce significant additional erosion. However, at this stage, region above 5 cm off the bed, the maximum differences are
near-bed instantaneous velocities may induce bed shear stresses observed at the free surface. In the SC case, the measured and
larger than the local critical shear stress, leading to local entrain- modeled velocities differ by 12% near the free surface at location
ment of sediment into suspension. P26. The differences observed at the free surface could be attrib-
Contour plots of dimensionless vertical velocities, W nd , not utable to the mesh resolution. The largest percent mean and RMS
shown here for simplicity, indicate that the most significant distor- errors are for the transversal velocity (V nd ), attributable mostly to
tion in the vertical velocity occurs along y ¼ 0 cm and decays the fact that such velocities are very small relative to streamwise
away from the object’s center plane. Fig. 3 also reveals that the velocities, and therefore induce larger relative errors. Furthermore,
RANS κ-ε turbulence model can reproduce the main experimental errors are larger downstream of the object because of well docu-
mean flow structures from a qualitative standpoint. Nevertheless, mented difficulties of RANS-based models to reproduce separated
to assess the ability of the model to reproduce measured velocities, flows (Rodi 1997; Smith and Foster 2007).
a direct comparison of pointwise velocity profiles was conducted. As previously noted, relatively high differences between
Fig. 4 shows a comparison between profiles of measured and measured and modeled velocities are observed near the bottom.

Fig. 3. (Color) Measured (left frames) and simulated (right frames) dimensionless mean streamwise velocities along y ðcmÞ ¼ 0 for the SC (top) and
TC (bottom); arrows in left frames indicate mean flow velocity

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 275

J. Hydraul. Eng. 2013.139:269-283.


Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 4. Comparison between measured and modeled profiles of mean local average velocity (u 0 ) along y ¼ 0 cm for SC (left) and TC (right); filled
and open symbols correspond to measured and modeled values, respectively

The discrepancy could be attributable to the bottom roughness hydrodynamic model to reproduce the main flow structure features
height used in the simulations, which may not account accurately revealed by the experimental observations.
for the experimental conditions. As previously stated, the bottom
roughness height used in the numerical simulations was 0.254 mm,
equivalent to the mean grain size of the sand bed. Further inves- Average Flow Turbulent Characteristics
tigations should be dedicated to using the bottom roughness as
a tuning parameter to improve the match between modeled The temporal evolution and final configuration of the tandem
and recorded velocities, particularly in the near bed region. Never- system, consisting of object and scour hole, are not solely deter-
theless, this is beyond the scope of the present work, because one mined by the mean flow configuration, but also by the turbulent
of the main goals of the study was to evaluate the ability of the flow structure. Turbulence is known for directly affecting and being

276 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


modulated by sediment entrained into suspension (Abad and turbulent kinetic energy TKE_ND is obtained by dividing Eq. (2)
García 2009). Although the mean flow may dictate general burial by one half of the square of the mean bottom shear velocity u,
and scour characteristics around the object through the mean or equivalently,
flow-sediment interactions, turbulent fluctuations may be more
TKE
responsible for the events that dictate sediment entrainment and re- TKE ND ¼ ð3Þ
suspension in the vicinity of boundaries, in particular those regions 0.5u2
of fluid-object-sediment interaction (Kirkil et al. 2008). Because of
the experimental conditions of the present study, in which point
Production of turbulent kinetic energy is generated by shear and
velocity measurements were taken around each object at distinctive
intensifies across the shear layers between the outer flow and the
times, instantaneous snapshots of the flow field are not available to
recirculation zones in the front and back of the objects. Fig. 5 shows
describe instantaneous flow field evolution (e.g., vortex cores).
the contours of TKE in a vertical xz-plane and illustrates areas of
Techniques that allow such analysis include PIV and streak photog-
higher TKE residing within the scour hole and along the central
raphy (Testik et al. 2005). In this paper, however, the mean flow
sand ridge. Fig. 5 also shows that the streamwise extension of
field turbulence characterization is performed in terms of time-
higher simulated TKE (right column) is in qualitative agreement
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

averaged signal values. This section analyzes the turbulent structure


with the measurements (left column) for both objects; yet, compar-
around each object and its implications on the final configuration
ing the two cases, the vertical extension is larger in the TC case.
of the surrounding scour hole.
This can be attributed to the blunter shape of the TC compared to
the SC. Conversely, contour plots of TKE in a horizontal xy-plane,
TKE not shown here, reveal that turbulence energy production is higher
in the SC induced by its sharp edges as opposed to the smooth side
Fig. 5 shows the measured and modeled normalized TKE along faces of the TC. A second region of high TKE develops inside the
the center plane y ¼ 0 cm of the two objects. The experimental scour hole upstream of the obstacle, though of lesser strength as
TKE was calculated as compared to the downstream side. Upstream of the object, ripples
1 f are large enough to affect the flow and induce relatively high levels
TKE ¼ ðu 0 2 þ v 0 2 þ w 0 2 Þ − R ðNLx þ NLy þ NLz Þ ð2Þ of TKE at their crests, which interact with the turbulent structures
2 2
(e.g., vortex cores) upstream of the object. Regions of high TKE
where u 0 , v 0 , and w 0 = mean of the velocity fluctuations in extend downstream for approximately 5D in the SC case and
the streamwise, transverse, and vertical directions, respectively; 6Dm in the TC case. As the flow passes the object, energy is ex-
and NLx , NLy , and NLz = noise levels in the three directions tracted from the mean flow, and high levels of TKE intensity are
(García et al. 2005; Abad and García 2009). The normalized induced primarily by the alternate shedding of large vortices in the

Fig. 5. (Color) Measured (left frames) and κ-ε simulated (right frames) mean dimensionless TKE field along transect y ðcmÞ ¼ 0 for the SC (top) and
the TC (bottom); arrows in left frames indicate mean flow velocity

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 277

J. Hydraul. Eng. 2013.139:269-283.


near wake. The TKE is enhanced by the process of time averaging hole. This HV does not have sufficient energy to continue scouring
of the flow field, which results in large variations of the velocity and at the equilibrium stage.
thus turbulent kinetic energy at certain locations. The vortex struc- Coherent vortex structures detach from the sides of the object,
tures generated atop and along edges of the object are convected with higher intensity in the case of the SC than in the TC case, as
downstream by the mean flow and gradually lose coherency as the shown in the left frames in Fig. 6(a). The κ-ε model overpredicts
vortices increase in size and their circulation decreases. Eventually, the length of the recirculation zone, because it is unable to predict
several diameters downstream of the object, TKE levels recover the vortex shedding periodic motion because of its inability to re-
to the far-field characteristic levels, where the disturbance by the solve fluid exchange in the near wake of the object (Rodi 1997;
object is negligible. The best agreement between modeled and Smith and Foster 2007). As displayed in Fig. 6(a), the length of
measured TKE occurs upstream of the object, as in the case of the upstream separation zone extends for approximately one object
streamwise velocities. diameter, but is still located within the scour hole. Smith and Foster
(2007) showed that the length of the separation zone upstream
of the object tends to constant values of approximately four and
Vortex Structures five object diameters for κ-ε and LES turbulence models, respec-
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

tively. In the present study, simulations of the two objects on a flat


The flow structure around 3D obstacles resting freely upon a immobile bed revealed a length of the upstream separation zone
deformed bed is more complex than the flow passing around a extending between two and three object diameters. The recircula-
short cylinder on a flat bed. In the present case, diverse 3D vortex tion length, defined as the distance between the center of the object
structures of various scales generated by the object, scour hole, and and the location where velocity path lines converge in the horizon-
surrounding bedforms coexist. Here, attention is focused on the tal xy-plane along y ¼ 0 cm, is approximately four times the diam-
average turbulent characteristics of the flow. eter of the object, which is consistent with the observations of
The complete 3D vortex structure is difficult to discern with the Testik et al. (2005) and Smith and Foster (2007). The latter study
vorticity field, because the vorticity not only includes the rotational showed that for object Reynolds numbers larger than 1,500, both
field components, but also signals from the boundary generated the upstream separation length and the downstream recirculation
shear (Smith and Foster 2007). Regions of high boundary shear length reached asymptotic constant values.
generate from the sides and top of the object and the surrounding Notice in Fig. 6(a) (left frames) that both HV and AV are not
scour hole. The Q-criterion (Dubief and Delcayre 2000) is used to completely symmetrical. The longitudinal extensions of the HV
show instantaneous snapshots of the 3D coherent structures around are slightly more elongated on the right side of the SC, because
the two objects in Fig. 6(a) (left frames) at EMC stages. The main the cylinder is not completely perpendicular to the direction of flow
coherent structures are the HV core located within the scour hole, propagation. As a result, the deposition and scour patterns are not
and the small vortex tubes formed past the crest of the surrounding fully symmetrical [Fig. 2(a)]. Likewise, the scour and deposition
ripples. The HV is primarily responsible for mobilizing sediment patterns around the TC are slightly asymmetrical, though to a lesser
near the bed and promoting the formation and growth of the scour extent than in the case of the SC [Fig. 2(b)].
hole upstream of the object. Notice an arch vortex (AV) detaching
from the top of the object, though relatively less coherent than
HV. Contours of the y-component of the vorticity are displayed Wall Shear Stress Distribution
in the right frames of Fig. 6(a). Fig. 6(b) shows contours of the
y-component of the vorticity along transect y ¼ 26 cm. Regions The presence of the object substantially deforms the approaching
of higher vorticity coincide with the location of the vortex tubes flow, causing it to accelerate and increase the bottom shear stresses
forming past the crests of far-field ripples shown in the left frames near the object compared to the far flow field. The relation between
of Fig. 6(a). the flow strength and the shape of the object influences the amount
The HV originates upstream of the object and helps form the of sediment entrained into suspension and the manner in which the
scour hole by entraining sediment as it moves up and down and sediment is shed away from its vicinity. If the instantaneous bed
sheds sediment toward the edges and downstream of the obstacle. shear stresses are higher than the critical bed shear stresses needed
This coherent structure mostly persists within the scour hole in to entrain sediment from the bottom, scouring of the bed occurs
the back and sides of the object. At a distance of approximately provided that the incoming bed load rate is less than the entrain-
a half object diameter downstream of the object, the structure loses ment rate. This ultimately leads to the burial of the object, given the
coherence and dissipates as it interacts with the mean flow. Observe removal of sediment is large enough to deplete the span shoulder.
that the y-component of the vorticity signature shown in Fig. 6(a) Fig. 7 shows the mean Shields parameter around the SC and TC
extends for approximately the same distance downstream of the (left frames: object on a flat bed; right frames: object on the scoured
two objects, yet the vertical extension is larger in the case of bed) simulated with the κ-ε model. Observe that at initial condi-
the TC than in the SC. This agrees with the vertical distribution tions, i.e., object on a flat bed, the Shields parameter θ is more than
of TKE shown in Fig. 5. Observe also that the vortex tubes atop twice the critical Shields parameter θcr ð¼ 0.041Þ at the lateral
the crests of the upstream ripples exhibit larger coherence upstream sides of the object. Here, 0.041 was calculated for a flat bed; how-
of the object than in its close vicinity. These structures recover their ever, in reality it should be adjusted for gravitational effects because
coherence once turbulent eddies detached from the object are no of a variable bathymetry (Kirkil et al. 2008). Note that the shear
longer energetic enough to disrupt the ripple-generated vortices. stress is higher at the edges of the SC than in the case of the
The resulting interaction of vorticity regions generated by object TC. At equilibrium scour conditions (right frames in Fig. 7), the
and ripples give rise to a more complex flow structure when com- Shields parameter is generally smaller than the critical value,
pared to the simpler case of an object lying on a flat bed (Smith i.e., ranging from 0.01–0.04. However, the local Shields parameter
and Foster 2005, 2007; Hatton et al. 2007; Testik et al. 2005). Note remains slightly larger than the critical Shields value at several
that upstream of the object several other vortical features are ob- locations, such as at the top of both objects and crests of far field
served. A streak of high vorticity, resulting from the upstream ripples, by approximately 70 and 30%, respectively (Fig. 7). In the
boundary layer (BL) interacts with the main HV within the scour latter case, this leads to continuous sediment transport at the ripple

278 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 6. (Color) Simulated coherent structures around the two objects (a) left frames: instantaneous three-dimensional coherent structures around SC
(top) and TC (bottom) visualized using the Q-criterion (Q ¼ 1.0); right frames: mean y component of vorticity signatures along y ¼ 0 cm for the SC
(top) and TC (bottom); (b) mean y component of vorticity signatures along transect y ðcmÞ ¼ 26 in the far field for SC (top) and TC (bottom)

crest, inducing ripple migration away from the object. In FLOW- The unsteadiness of the flow field, though not captured in the
3D, the shear velocity, u , is solved iteratively at no-slip boundaries current κ-ε model simulations, can induce shear velocities in excess
with the combined smooth and rough logarithmic law of the wall of two to three times mean values, leading to local erosion within
equation. As previously stated, velocity and bathymetry mea- the scour and deposition regions, as in the cases of flow around
surements were conducted once the scouring process in the short cylinders (Smith and Foster 2007) and vertical cylinders
vicinity of the object ceased, as indicated by visual inspections. (Kirkil et al. 2008). Modeling relying on the computation of a time-
Even though the scouring around the object and its sinking have averaged shear stress, as in this study, generally results in under-
ceased, ripples in the far field maintained a very regular two- prediction of the scour hole downstream of the object, particularly
dimensional structure and steady migration rates on the order for higher Reynolds numbers (Li and Cheng 2001; Sumer et al.
of 1 cm= min. 1988). Fig. 7 shows that at flat bed conditions, i.e., at the beginning

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 279

J. Hydraul. Eng. 2013.139:269-283.


Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

Fig. 7. (Color) Modeled mean Shields parameter around the object (a-SC, b-TC) on a flat bed (left) and deformed beds (right) at stationary flow
conditions

of the run, there is a significant difference in Shields stress between a RANS κ-ε turbulence model. Simulations with a dense
regions exposed to potential sediment deposition in the SC and TC computational mesh in the vicinity of the object generate results
cases. The Shields parameter can easily exceed the critical Shields in fair agreement with measured average velocities and turbulent
parameter in almost 80 and 60% at the lateral edges for the SC and kinetic energy. The model overestimates the recirculation length
TC, respectively. because it fails to adequately solve flow separation. However,
Observe also in Fig. 7 the regions of very small shear stresses the mean flow features and flow velocities away from the
approximately one diameter upstream and over a larger region obstacle are well reproduced.
downstream of the object. This is in agreement with the model ob- • The experiments and RANS simulations show that the accelera-
servations by Smith and Foster (2007) for short cylinders of similar tion near the sides of the object attributable to the obstruction
aspect ratio as the one used in the present study. Initially, sediment and adverse pressure gradients induced by the object increases
tends to deposit in these low shear regions, but as scour and the level of turbulence and bed shear stresses compared to the
deposition around the object progress over time, the location of far field values. This in turn leads to scouring of the bed and
such regions do not necessarily coincide with the scour and dep- subsequent sinking of the object. Under similar flow conditions,
ositional patterns indicated at the beginning of the experiment the flow field around the truncated cone is less energetic and
under flat bed conditions. This difference can be attributed to therefore induces smaller shear stresses than in the case of
the complex interaction between the evolving bed topography, the the short cylinder lying horizontally on the floor. Smaller shear
incoming flow, and the surrounding ripples not present at the stresses are associated with a smaller sediment transport capa-
beginning of the run. The scour and deposition regions around city of the flow resulting in less sediment entrainment, leading
the object, both upstream and downstream, coincide with the re- ultimately to less burial of the object.
gions in which the incoming flow is highly disturbed because of • Regions of high turbulent kinetic energy generated downstream
the blockage of the object. This is in accordance with the previous and atop the objects are observed in both experiments and nu-
discussion regarding the extension of the vortex shedding and its merical simulations. Regions of higher TKE intensities extend
influence on the sediment transport. for several diameters downstream of the object. On the upstream
side of the object, relatively high values of TKE are observed,
yet are smaller in magnitude when compared to their down-
Conclusions stream counterparts.
• The numerical simulations identified two main coherent struc-
• A set of ADV velocity measurements were conducted around tures near the base of the objects at equilibrium morphodynamic
two 3D objects, a short cylinder and a truncated cone, on a conditions: a necklace horseshoe vortex embedded within the
scoured bed at equilibrium morphodynamic conditions under scour hole, and an arch vortex detaching from the top of the
unidirectional flows. object. The turbulent kinetic energy associated with both obsta-
• Simulations of the flow structure around the partially buried cles is much higher than that associated with the approaching
3D objects were performed for high Reynolds numbers using boundary layer flow.

280 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


• The numerical study used standard values of turbulence para- kv = von Kármán constant;
meters in the commercial fluid dynamics model. Further study Lc = length of the short cylinder;
of turbulence parameter sensitivity and the effect of bottom Ls = energy containing length scale;
roughness are needed to improve the agreement between numer- NLx , NLy , and NLz = noise levels in the three spatial directions;
ical and experimental results. R = Reynolds number;
• Bedforms surrounding the object modify the turbulent shear Rd = object Reynolds number;
stresses considerably. Significant differences are observed be- Sh = Strouhal number;
tween the object mounted on the flat bed and the partially buried T = water temperature;
object. In the case of flat bed, maximum shear stresses appear T s = period of the vortex shedding oscillation;
close to the lateral edges of the object. Higher shear stresses are TKE ND = normalized TKE;
observed in the case of the short cylinder than in the truncated U = mean flow velocity of the current;
cone case. As a result, the extent and depth of the scour is larger U nd , V nd , W nd = dimensionless velocities;
at equilibrium conditions for the short cylinder case. u = bed shear velocity;
• Comparison of the flow structures around the short cylinder and u 0 = local average longitudinal velocity;
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

truncated cone in the flat bed mounted and partially buried u, v, w = instantaneous velocities in the three spatial
cases suggests that at the beginning of the experiment, when coordinates;
the object is mounted on a flat bed, TKE levels are much higher u 0 , v 0 , w 0 = turbulent velocity fluctuations around the mean in the
near the bed level when compared to the case at EMC. This can three spatial coordinates;
be attributed to the fact that at the mature equilibrium condition, ū, v̄, w̄ = mean flow velocities in the three spatial coordinates;
mean and turbulent intensities are in equilibrium with the en- x, y, z = spatial coordinates;
trainment and deposition sediment rates. zo = total roughness length;
• Sand ridges, formed by the supply of sediment scoured from zof = form drag;
behind and along the sides of the object, extend downstream zos = skin friction;
along the center plane of the object for approximately 2Lc in zot = near-bed sediment transport;
the case of the short cylinder, and approximately 2Dm in the γ c = specific gravity of object;
case of the truncated cone. This turns out to be a self-protecting γ s = specific gravity of sediment;
mechanism against further sinking. ε = dissipation rate of turbulent kinetic energy;
θ = Shields parameter;
θcr = critical Shields shear stress;
Acknowledgments λp = sediment porosity;
ν = kinematic viscosity of fluid; and
The writers are thankful for the financial support of the University φ = internal angle of repose of the sand.
of Illinois at Urbana-Champaign and the Coastal Geosciences Pro-
gram of the U.S. Office of Naval Research, Grant N00014-01-1-
0337. The writers thank David Waterman and Dr. Talia Tokyay
for their valuable comments and suggestions on the manuscript.
References
Gratitude is also extended to Dr. Carlos M. García for his input
during ADV data quality analysis. The participation of Dr. Jorge Abad, J. D., and García, M. H. (2009). “Experiments in a high-amplitude
D. Abad was possible because of startup funding from the Depart- Kinoshita meandering channel: 1. Implications of bend orientation on
ment of Civil and Environmental Engineering, University of mean and turbulent flow structure.” Water Resour. Res., 45, W02401.
Pittsburgh. The writers would like to thank the anonymous re- Abad, J., Rhoads, B., Güneralp, İ., and García, M. (2008). “Flow structure
viewers and the journal’s associated editor for their positive at different stages in a meander-bend with bendway weirs.” J. Hydraul.
feedback and suggestions, which were very helpful to improve Eng., 134(8), 1052–1063.
the manuscript. Ahmed, F., and Rajaratnam, N. (1998). “Flow around bridge piers.” J. Hy-
draul. Eng., 124(3), 288–299.
Bagnold, R. A. (1974). “Fluid forces on a body in shear flow; experimental
Notation use of stationary flow.” Proc., R. Soc. London, Ser. A, 340(1621),
147–171.
The following symbols are used in this paper: Cataño-Lopera, Y., Demir, S. T., and García, M. H. (2007). “Self-burial of
short cylinders under oscillatory flows and combined waves plus
Bdo = initial burial depth;
currents.” IEEE J. Ocean. Eng., 32(1), 191–203.
D = diameter of the SC; Cevik, E., and Yuksel, Y. (1999). “Scour under submarine pipelines in
Dm = mean diameter of the TC; waves in shoaling conditions.” J. Waterw., Port, Coastal, Ocean
d50 = median sediment size; Eng., 125(1), 9–19.
F = parameter describing accuracy of ADV Chen, H. C. (2002). “Numerical simulation of scour around complex piers
measurements; in cohesive soil.” Proc., First Int. Conf. on Scour of Foundations, Texas
F = Froude number associated to mean flow; A&M Univ., College Station, TX, 14–33.
Fd = Froude number associated to the object; Chrisohoides, A., Sotiropoulos, F., and Sturm, T. (2003). “Coherent struc-
Ff = fluid fraction function as defined in FLOW-3D; tures in flat-bed abutment flow: Computational fluid dynamics simula-
fR = user defined ADV sampling frequency; tions and experiments.” J. Hydraul. Eng., 129(3), 177–186.
Devenport, W. J., and Simpson, R. L. (1990). “Time-dependent and time
g = acceleration attributable to gravity;
averaged turbulence structure near the nose of a wing-body junction.”
h = average water depth; J. Fluid Mech., 210, 23–55.
hc = height of the TC; Dubief, Y., and Delcayre, F. (2000). “On coherent vortex identification in
k = turbulent kinetic energy TKE; turbulence.” J. Turbul., 1(1), 011.
ks = surface mean roughness height of the object and Flow Science Inc. (2009). FLOW-3D V 9.4 user’s manual, Flowscience
sandbed; Inc., Santa Fe, NM.

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 281

J. Hydraul. Eng. 2013.139:269-283.


García, M. H. (2008). “Sedimentation engineering: Processes, measure- Nikora, V., and Goring, D. (2000). “Flow turbulence over fixed and weakly
ments, modeling and practice; Manual of practice 110.” Chapter 2, mobile gravel beds.” J. Hydraul. Eng., 126(9), 679–690.
Sediment transport and morphodynamics, ASCE. Olsen, N. R. B., and Kjellesvig, H. M. (1998). “Three-dimensional numeri-
García, C. M., Cantero, M. I., Niño, Y., and García, M. H. (2005). cal flow modeling for estimation of maximum local scour depth.”
“Turbulence measurements with acoustic Doppler velocimeters.” J. Hydraul. Res., 36(4), 579–590.
J. Hydraul. Eng., 131(12), 1062–1073. Paik, J., Escauriaza, C., and Sotiropoulos, F. (2007). “On the bimodal
Goring, D. G., and Nikora, V. I. (2002). “Despiking acoustic Doppler dynamics of the turbulent horseshoe vortex system in a wing body
velocimeter data.” J. Hydraul. Eng., 128(1), 117–126. junction.” Phys. Fluids, 19(4), 045107.
Graf, W. H., and Istiarto, I. (2002). “Flow pattern in the scour hole around a Paik, J., and Sotiropoulos, F. (2005). “Coherent structure dynamics
cylinder.” J. Hydraul. Res., 40(1), 13–19. upstream of a long rectangular block at the side of a large aspect
Guyonic, S., Mory, M., Wever, T. F., Ardhuin, F., and Garlan, T. (2007). ratio channel.” Phys. Fluids, 17(11), 115104.
“Full-scale mine burial experiments in wave and current environ- Raikar, R. V., and Dey, S. (2008). “Kinematics of horseshoe vortex devel-
ments and comparison with models.” IEEE J. Ocean. Eng., 32(1), opment in an evolving scour hole at a square cylinder.” J. Hydraul.
119–132. Res., 46(2), 247–264.
Hatton, K. A., Foster, D. L., Traykovski, P., and Smith, H. D. Richardson, J. E., and Panchang, V. G. (1998). “Three-dimensional simu-
(2007). “Numerical simulations of the flow and sediment transport
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

lation of scour-inducing flow at bridge piers.” J. Hydraul. Eng., 124(5),


regimes surrounding a short cylinder.” IEEE J. Ocean. Eng., 32(1), 530–540.
249–259. Rodi, W. (1997). “Comparison of LES and RANS calculations of the flow
Hirt, C. W., and Nichols, B. D. (1981). “Volume of fluid (VOF) method around bluff bodies.” J. Wind. Eng. Ind. Aerodyn., 69–71, 55–75.
for the dynamics of free boundaries.” J. Comput. Phys., 39(1), Roulund, A., Sumer, B. M., Fredsoe, J., and Michelsen, J. (2005). “Numeri-
201–225. cal and experimental investigation of flow and scour around a circular
Hussein, H. J., and Martinuzzi, R. J. (1996). “Energy balance for turbulent pile.” J. Fluid Mech., 534, 351–401.
flow around a surface mounted cube placed in a channel.” Phys. Fluids, Salaheldin, T. M., Imran, J., and Chaudhry, M. H. (2004). “Numerical mod-
8(3), 764–780. eling of three-dimensional flow field around circular piers.” J. Hydraul.
Imman, D. L., and Jenkins, S. A. (2002). “Scour and burial of bottom Eng., 130(2), 91–100.
mines: A primer for fleet use.” Scripps Institution of Oceanography Schofield, W. H., and Logan, E. (1990). “Turbulent shear flow over surface
(SIO) series 02-8, Univ. of San Diego, San Diego, CA. mounted obstacles.” J. Fluids Eng., 112(4), 376–385.
Jenkins, S. A., Inman, D. L., Richardson, M. D., Wever, T. F., and Wasyl, J. Smith, H. D., and Foster, D. L. (2005). “Modeling of flow around a cylinder
(2007). “Scour and burial mechanics of objects in the nearshore.” IEEE
over a scoured bed.” J. Waterw., Port, Coastal, Ocean Eng., 131(1),
J. Ocean. Eng., 32(1), 78–90.
14–24.
Kirkil, G., and Constantinescu, G. (2010). “Flow and turbulence structure
Smith, H. D., and Foster, D. L. (2007). “Three-dimensional flow around
around an in-stream rectangular cylinder with scour hole.” Water
a bottom-mounted short cylinder.” J. Hydraul. Eng., 133(5),
Resour. Res., 46, W11549.
534–544.
Kirkil, G., Constantinescu, S. G., and Eterna, R. (2008). “Coherent struc-
Song, T., and Chiew, Y. M. (2001). “Turbulence measurements in nonuni-
tures in the flow field around a circular cylinder with scour hole.”
form open-channel flow using acoustic Doppler velocimeter (ADV).”
J. Hydraul. Eng., 134(5), 572–587.
J. Eng. Mech., 127(3), 219–232.
Koken, M., and Constantinescu, G. (2008). “An investigation of the flow
Sontek Inc. (2001). ADV operation manual, 1st Ed., Sontek Inc.,
and scour mechanisms around isolated spur dikes in a shallow open
San Diego, CA.
channel: 2. Conditions corresponding to the final stages of the erosion
and deposition process.” Water Resour. Res., 44(8), W08407. Strom, K. B., and Papanicolaou, A. N. (2007). “ADV measurements around
Li, F., and Cheng, L. (2001). “Prediction of lee-wake scouring of pipe- a cluster microform in a shallow mountain stream.” J. Hydraul. Eng.,
lines in currents.” J. Waterw., Port, Coastal, Ocean Eng., 127(2), 133(12), 1379–1389.
106–112. Sumer, B. M., and Fredsøe, J. (2002). The mechanics of scour in the marine
Liang, D., and Cheng, L. (1999). “Numerical model for local scour under environment, advanced series on ocean engineering, Vol. 17, World
offshore pipelines.” J. Hydraul. Eng., 125(4), 400–406. Scientific, Teaneck, NJ.
Liang, D., and Cheng, L. (2005). “Numerical model for wave-induced Sumer, B. M., Jensen, H. R., Mao, Y., and Fredsøe, J. (1988). “Effect of
scour below a submarine pipeline.” J. Waterw., Port, Coastal, Ocean lee-wake on scour below pipelines in current.” J. Waterw., Port,
Eng, 131(5), 193–202. Coastal, Ocean Eng., 114(5), 599–614.
Liang, D., and Cheng, L. (2005). “Numerical model for wave-induced Testik, F. Y., Voropayev, S. I., and Fernando, H. J. S. (2005). “Flow around
scour below a submarine pipeline.” J. Waterway Port Coastal Ocean a short horizontal bottom cylinder under steady and oscillatory flows.”
Eng., 131(5), 193–202. Phys. Fluids, 17(4), 047103.
Martinuzzi, R. J., and AbuOmar, M. (2003). “Study of the flow around Thomschke, H. (1971). “Experimentelle untersuchung der stationären um-
surface-mounted pyramids.” Exp. Fluids, 34(3), 379–389. strömung von kugel und zylinder in wandnähe.” Ph.D. thesis, Karlsruhe
Martinuzzi, R., and Tropea, C. (1993). “The flow around surface mounted, Univ., Karlsruhe, Germany.
prismatic obstacles placed in a fully developed channel flow.” J. Fluids Traykovski, P., Richardson, M. D., Mayer, L. A., and Irish, J. D. (2007).
Eng., 115(1), 85–92. “Mine burial experiments at the Martha’s Vineyard Coastal Observa-
Mayer, L. A., Raymond, R., Glang, G., Richardson, M. D., Traykovski, P., tory.” IEEE J. Ocean. Eng., 32(1), 150–166.
and Trembanis, A. C. (2007). “High-resolution mapping of mines and Truelsen, C., Sumer, B. M., and Fredsøe, J. (2005). “Scour around spherical
ripples at the Martha’s Vineyard coastal observatory.” IEEE J. Ocean. bodies and self-burial.” J. Waterw., Port, Coast. Ocean Eng., 131(1),
Eng., 32(1), 133–149. 1–13.
McCarthy, E. M., and Sabol, B. (2000). “Acoustic characterization of Voropayev, S. I., Testik, F. Y., Fernando, H. J. S., and Boyer, D. L. (2003).
submerged aquatic vegetation: Military and environmental monitoring “Burial and scour around short cylinder under progressive shoaling
applications.” OCEANS MTS/IEEE Conf. and Exhibition, Providence, waves.” Ocean Eng., 30(13), 1647–1667.
RI, Vol. 3, 1957–1961. Voulgaris, B., and Trowbridge, J. H. (1998). “Evaluation of the acoustic
Melville, B. W., and Raudkivi, A. J. (1977). “Flow characteristics in local Doppler velocimeter (ADV) for turbulence measurements.” J. Atmos.
scour at bridge piers.” J. Hydraul. Res., 15(4), 373–380. Ocean. Technol., 15, 272–289.
Nagata, N., Hosoda, T., Nakato, T., and Muramoto, Y. (2005). “Three- Wei, G., Chen, H. C., Ting, F., Briaud, J. L., Gudavalli, S. R., and Perugu,
dimensional numerical model for flow and bed deformation S. (1997). “Numerical simulation to study scour rate in cohesive soils.”
around river hydraulic structures.” J. Hydraul. Eng., 131(12), Research Rep. to the Texas Dept. of Transportation, Dept. of Civil
1074–1087. Engineering, Texas A&M Univ., College Station, TX.

282 / JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013

J. Hydraul. Eng. 2013.139:269-283.


Wilcox, D. C. (2000). Turbulence modeling for CFD, 2nd Ed., DCW Zhao, M., Cheng, L., and Zang, Z. (2010). “Experimental and
Industries, La Cañada, CA. numerical investigation of local scour around a submerged vertical
Willetts, B. B., and Murray, C. G. (1981). “Lift exerted on stationary circular cylinder in steady currents.” Coastal Eng., 57(8),
spheres in turbulent flow.” J. Fluid Mech., 105, 487–505. 709–721.
Downloaded from ascelibrary.org by Eng Periodicals 09 on 02/28/13. Copyright ASCE. For personal use only; all rights reserved.

JOURNAL OF HYDRAULIC ENGINEERING © ASCE / MARCH 2013 / 283

J. Hydraul. Eng. 2013.139:269-283.

Vous aimerez peut-être aussi