You are on page 1of 12

International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


journal homepage: www.elsevier.com/locate/ijhmt

Large eddy simulation of density current on sloping beds


Saurabh Chawdhary a, Ali Khosronejad b, George Christodoulou c, Fotis Sotiropoulos b,⇑
a
Argonne National Laboratory, 9700 S. Cass Avenue, Lemont, IL 60439, USA
b
Department of Civil Engineering, Stony Brook University, Stony Brook, NY 11794, USA
c
Department of Civil Engineering, National Technical University of Athens, Zografou 15780, Greece

a r t i c l e i n f o a b s t r a c t

Article history: We carry out high-resolution large-eddy simulation (LES) of stratified flows developing as dense under-
Received 22 August 2017 flows over two different bed slopes of 5 and 15 in an laboratory-scale tank. Simulations for the corre-
Accepted 14 December 2017 sponding unstratified flows are also carried out to help elucidate the effects of stratification on the
Available online xxxx
dynamics of the flow and overall evolution of the effluent front. The simulated density current flows have
also been investigated experimentally by continuously releasing down a sloping plane a fluid that is
Keywords: slightly denser than the ambient fluid. The Navier Stokes equations with the Boussinesq approximation,
Density current
to account for the stratification due to density difference, are solved numerically using a second-order
Gravity current
Stratified flow
accurate finite-difference method (Khosronejad and Sotiropoulos, 2014). The simulated structure of the
Stratified jet density currents at quasi-equilibrium are shown to be in reasonable agreement with the experimental
Large-eddy simulation observations. The variation of the half-width of the density current plume with downstream length is
Immersed boundary method shown to exhibit different power law regimes on a log scale. Spectral analysis reveal existence of large
scale structures in the flow. Comparative analysis of the dynamics of the coherent structures in the strat-
ified and non-stratified flows reveal that stratification has a profound effect on the large-scale features of
the flow.
Ó 2017 Published by Elsevier Ltd.

1. Introduction mental flows and understanding of the conditions leading to their


formation in nature.
Stratified flows, in the form of density (or gravity) currents, are The mixing due to density currents and its dynamics have been
capable of transporting material across long distances in natural widely studied at laboratory scale under various conditions. These
geophysical flows. Such currents are formed as a result of mixing studies can be classified based on whether the density current
of fluids with differing density. The source of density difference develops on a horizontal plane or a sloping bed. Most commonly,
could be either temperature difference, salinity, presence of sus- density current is studied on a horizontal plane as a lock exchange
pended sediments or a combination of these factors. Examples of problem, emulating a dam break wherein a vertical lock between
naturally occurring phenomena due to density currents include two fluids of different densities is instantaneously removed. There
katabatic winds on the mountains, haboobs or dust storms, cold are three separate phases of mixing following the lock release
weather fronts under warmer air mass and thunderstorm outflow which are [28] (i) slumping phase, during which dense fluid
boundary [47]. Density current plumes on sloping surfaces can slumps and moves with almost constant front velocity; (ii) inertial
arise in various natural and man-made situations. Currents of phase, during which flow is dominated by inertial and buoyancy
heavier bottom fluid commonly develop underwater on open forces; and (iii) viscous phase, during which flow is dominated
slopes, spreading mostly laterally and sometimes vertically. They by viscous and buoyancy forces. Hacker et al. [22] performed lab-
can also occur in rivers flowing into the ocean or industrial efflu- oratory experiments with different aspect ratios of the lock and
ents merging in a river or lake, and advancement of saline under- identified the density structures formed during the different
currents in estuaries. Therefore, the study of density currents is phases of mixing. Breaking of Kelvin-Helholtz (KH) billows were
important in evaluating impact of underflows on natural environ- deemed responsible for the spreading of stratified fluid from the
density current head. Parsons and Garcia [45] explored the similar-
ity in density distribution in flow with higher Reynolds number.
⇑ Corresponding author. There are some early works which studied the two-dimensional
E-mail address: fotis.sotiropoulos@stonybrook.edu (F. Sotiropoulos). (2D) gravity current on a slope (by continuously releasing a ‘‘line

https://doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
0017-9310/Ó 2017 Published by Elsevier Ltd.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
2 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

source” of dense fluid) on long gently sloping fresh-water flumes problems. Härtel et al. were the first to perform 3D direct numer-
[15,58,6]. This action forms a dense underflow of gravity current ical simulation (DNS) of the lock- release problem. The flow topol-
spreading in vertical and streamwise direction, but is confined in ogy at the head of the dense current was analyzed to understand
the lateral direction. However, dense fluid released from a finite the lobe-and-cleft instability. Cantero et al. [8] analyzed the
source on an open slope results in three-dimensional (3D) spread- propagation of the front of 2D and cylindrical gravity currents
ing, giving rise to a negatively buoyant plume, where lateral using highly resolved 2D and 3D simulations. Tokyay et al.
spreading is much more important than vertical (see Fig. 1). Exper- [50,52,49,51] and Gonzalez-Juez et al. [20,21] employed large-
imental study on 3D spreading of density currents on slope was eddy simulation (LES) to investigate the interaction of gravity cur-
first performed by Hauenstein and Dracos [25] by releasing saline rent past obstacles. Other noTable 2D and 3D numerical works on
water from a point source on the sloping floor immersed in a fresh- lock-release gravity currents on a flat bed are [43,5,4,56,24,23,41,
water tank. Alavian [1] studied velocity and density distribution 7,42,7,13], respectively. Paik et al. [44] 3D unsteady Reynolds Aver-
and the steady state plume-shape for different angles of the slop- aged Navier-Stokes (RANS) simulations were also able to capture
ing bottom bed. Several important conclusions regarding the struc- the intricate 3D flow structures such as the KH billows, their break-
ture of the plume were drawn most notable being the fact that the down and lobe-and-cleft instabilities. They noted that 2D approx-
plume developed rapidly for some initial distance, and then imation of density currents tend to under-predict the propagation
became relatively constant in breadth. On mild slopes (< 1=10) of the dense front during the buoyancy-inertia regime. Tokyay and
the entrainment of ambient fluid was negligible and has well- García [53] used LES to investigate the effect of inflow conditions
defined plume boundaries. This study was followed by further on the propagation of density currents on mild (5%) slopes. RANS
experimental investigations of plumes on steep slopes only [55] modeling was used by Firoozabadi et al. [17] to simulate the den-
wherein the behavior of density currents on an incline with respect sity current plumes on extremely mild slopes. They concluded that
to different parameters was systematically quantified. It was found the inlet Richardson number has a profound effect on the structure
that, unlike 2D gravity plumes, spreading of a 3D plume front is of the density current. Other numerical studies on the density cur-
nonlinear. The spreading was highly dependent upon the buoyancy rents (on a slope) formed by a finite volume of fluid release instead
flux of dense fluid at the inlet, but only loosely dependent on inlet of a continuous release are in references [39,14,46,3,40]. However,
momentum flux. For similar inlet conditions, the spreading rate of these are not discussed here in detail since the focus of our work is
the dense plume decreased with slope. With the help of experi- density current formation due to continuous release (constant flux)
mental data and theoretical considerations, analytical expressions of dense fluid from a finite source (as opposed to line source).
for the longitudinal and lateral spreading were developed. Choi In this work we carry out LES of flow and density characteristics
and Garcia [10] developed a relationship for spreading rate and of 3D density currents down a sloping bed. We use the curvilinear
verified it with their earlier experimental results (for slopes from framework which has been already extensively validated for buoy-
2 to 10 ). Christodoulou and Tzachou [11] performed experiments ant and non-buoyant [36,33,32,34,31,35,37,9,30,29] flows. High-
for steep slopes to study unconfined 3D gravity currents at suffi- resolution LES in this work resolves intricate flow features and
ciently high Reynolds numbers. The dense plume in all cases enable us to study the vortex formation at the genesis of a dense
attained a ‘‘normal” width at some longitudinal distance after front near the inlet. This genesis is contrasted with an unstratified
which the dimensions of dense current changed only marginally. submerged jet to identify the role played by stratification on
This normal width correlated well with the buoyancy flux and under-flows. Frequency of large-scale structures is also identified
reduced gravity. Using more experiments, Christodoulou [12] via spectral analysis.
attempted to formulate the spreading laws relating the lateral This paper is organized as follows. In Sections 2 and 3, we
and longitudinal spreading of plume in different regions of the describe the numerical method and experimental setup of the
dense plume. It was argued that the spreading laws do not depend problem, respectively. This is followed by the description of the
on slopes for the range of slopes investigated (2 to 15 ). numerical procedure employed for the simulation in Section 4.
Laboratory experiments of density currents often require very The results obtained from the simulation are subsequently dis-
large setups and sophisticated data acquisition equipment, cussed. Finally, we conclude the paper with a summary and discuss
whereas field measurements of 3D gravity currents are difficult the areas of future research.
to carry out because of their unexpected occurrence. Hence,
numerical simulation is popular, yet challenging, tool for such 2. Numerical method

We solve spatially filtered continuity and Navier Stokes equa-


tions with Bousinesq approximation to account for stratification.
These governing equations for this case [36] are shown in non-
dimensional form as follows

@U j
J ¼0 ð1Þ
@n j
( ! !  )
j
1 @U i nil @ @ Gjk @ul @ nl p @ slj di3
¼  j ðU j ul Þ þ j Re   þ RiC
J @t J @n @n J @nk @n j J @n j J
ð2Þ
 1 2 3 
where J ¼ @ðn ; n ; n Þ=@ðx1 ; x2 ; x3 Þ is the Jacobian of the geometric
transformation, x3 is the vertical coordinate, nil ¼ @ni =@xl are the
transformation metrics, ui is the ith Cartesian velocity component,
U i ¼ ðnim =JÞum is the contravariant volume flux, Gjk ¼ nlj nkl are the
Fig. 1. Schematic of a 3D spreading of negatively buoyant plume on a slope. components of the contravariant metric tensor, p is the pressure,
Reproduced from Tsihrintzis and Alavian [55]. sij is the sub-grid stress (SGS) tensor arising when LES filter is
Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx 3

applied [29], d is the Kronecker delta. Additionally, C is the concen- cylindrical tank of 0.8 m diameter and then pumped to a constant
tration of dense fluid in volume fraction defined as [33]: head tank placed about 3 m above the channel. The flow-rate of
incoming of dense fluid was controlled with the help of 2 valves
q  qo
C¼ ð3Þ and a specially designed orifice meter. The colored dye used for
qs  qo
visualization consisted of potassium permanganate solution. Pho-
where qs is the density of the dense fluid at inlet, qo is the ambient tos of the spreading plume were taken manually at selected time
fluid density (the density of water at 20  C in the present case) and intervals by means of a 35 mm camera placed at an elevated plat-
q is the density of the mixed fluid. All quantities in the equations form about 6 m above the tank. More detailed description of the
are appropriately non-dimensionalized using qo ; uin , the bulk experimental setup and findings of the laboratory experiments
velocity at the inlet and h0 , the height of the inlet. The non- have been published previously in [11,12]. It was observed that
dimensional numbers in the equations are Richardson (Ri) and the density currents show different characteristics depending on
Reynolds (Re) numbers, defined as: the slope and on the initial Richardson (Ri) and Reynolds (Re)
numbers at the source. The most notable observation was that
uin h0 ðqs  qo Þ gh0
Re ¼ ; Ri ¼ ð4Þ large-scale disturbances resembling roll waves appeared for cases
m qo u2in with low slope, low Re and high Ri. This is in agreement with sim-
where m is the kinematic viscosity of water at 20  C and g is accel- ilar observations by Tsihrintzis [54]. Two experiments were chosen
eration due to gravity. The concentration field of the dense fluid is for numerical simulation and are described in the following
obtained by solving the following non-dimensional convection- section.
diffusion equation:
!   4. Computational details
1 @C @ Gjk @C @ CU j
¼ j Sc  ð5Þ
J @t @n J @ni @n jJ We simulate two test cases denoted as cases A and B which cor-
respond to experiments on slopes 5 and 15 to the horizontal,
In the above equation, Sc is the Schmidt number of the dense fluid.
respectively. The computational domain consists of a shallow con-
Saline solution at 20 °C was reported as Sc ¼ 700 [38]. The SGS ten-
tainer with sloped bed (L  W) inclined at h degrees to the horizon-
sor (sij ), which appears after applying the spatial filter to the turbu-
tal. Fig. 2 shows a schematic of the computational domain. As
lence, is modeled using the Smagorinsky model [48]:
shown in Fig. 2, the width and height of the rectangular opening
1 are b0 ð¼ 5 cmÞ and h0 . Table 1 shows values of physical parameters
sij  skk dij ¼ 2lt Sij ; ð6Þ
for the two cases. Initially the domain is filled with ambient den-
3
sity fluid (water at 20 °C). At t ¼ 0, fluid that is slightly denser (sal-
where the eddy viscosity (lt ) was further modeled by Smagorinsky
ine water) is released from a rectangular inlet.
as
Body-fitted curvilinear grids were used in both the simulations.
lt ¼ C s D2 jSj ð7Þ Table 2 gives details about the grids used in the simulations and
Fig. 3 shows the grids. The grid lines follow the sloping bed in a
In Eqs. (6) and (7), over-bar denotes spatial filtering operation, Sij is diverging fashion to contain the density current plume to be gen-
the filtered strain-rate tensor, dij is Kronecker delta, C s is the erated. The horizontal portion of the tank follows at the end of
Smagorinsky constant, D is the filter size (cube root of the grid cell the slope skewing the cells near the bottom of the slope. Grid
qffiffiffiffiffiffiffiffiffiffiffiffi
volume in the present method) and jSj ¼ 2Sij Sij . Smagorinsky con- points were distributed non-uniformly to cluster more points (in
X and Y directions) near the inlet in order to capture the small
stant C s is dynamically calculated using the method of Germano
scales produced as the buoyant flow enters the domain. The grid
et al. [19]. More details of the C s calculation in the context of our
spacing was kept uniform in Z direction.
numerical method can be found in Kang et al. [29].
The inlet was fed with a uniform flow of volume flux Q intro-
The governing equations are discretized in space using a second-
duced at t ¼ 0. The top free surface boundary was treated as free
order central finite difference scheme and advanced in time using a
slip. At the exit outflow boundary condition was used. The two side
second-order fractional step method [18,29]. Iterative solvers
walls and bottom walls were treated as no-slip boundaries.
implemented in Portable, Extensible Toolkit for Scientific Computa-
Simulations were run until time t ¼ T, when the heavy density
tion (PETSc) library are used for solving the discretized equations.
plume reached the far edge of the slope. At this time, a passive tra-
Generalized Minimal Residual (GMRES) method is used to solve
cer was released, along with the denser fluid from the inlet. It was
the linear system for the pressure-correction Poisson equation.
observed in the experiments that the plume shape of the tracer
Algebraic multigrid (AMG) is used as a preconditioner for the
reaches quasi-equilibrium state (‘‘steady state”) after an approxi-
GMRES method to accelerate the convergence. The nonlinear dis-
mate time t ¼ 3T. Similar procedure was followed in the simula-
crete momentum equation is solved using matrix-free Newton–
tions to obtain the quasi-equilibrium shape of the tracer plume.
Krylov method. The inner iterations of the Newton-Krylov solvers
That is, the simulations were carried out up until t ¼ 3T with the
also use GMRES method, but without preconditioning. The code is
passive tracer introduced into the flow at t ¼ T.
efficiently parallelized using PETSc library and Message Passing
Interface (MPI) to exploit massively parallel computer clusters.

3. Experiments

Laboratory experiments have been carried out in the Hydraulics


Laboratory of the National Technical University of Athens, Greece,
using a bed at various slopes in a shallow 5 m  7 m tank. The tank
was initially filled with tap water (ambient fluid) at t ¼ 0. A dense
saline fluid was released continuously for t > 0 at the top of the
sloping surface through a 1 m long and 5 cm wide channel. It con- Fig. 2. Schematic of the simulated cases. b0 ¼ 5 cm and h0 ; L and h are listed in
sisted of common salt-water solution, which was well mixed in a Table 1.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
4 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

Table 1
Details of the two different cases in the present study. hin represents the opening height of the inlet.

Case h h0 (cm) uin (m/s) Q ðcm3 =sÞ Dq=q L (m) W (m) Re Ri

A 15° 1.42 0.14 99.5 0.005 2 2.84 1989 0.03


B 5° 1.17 0.085 50 0.030 4 3.51 1000 0.47

Table 2
Details of the computational grid systems. Nx, Ny, Nz are the number of grid nodes in streamwise, spanwise, and vertical directions, respectively. u and uin are the shear velocity
and mean inlet velocity. Dzþ is the grid size in wall units in vertical directions.

Case Nx Ny Nz Dz at inlet u =uin Dzþ at inlet


A 1301 841 51 0:040hin 0.25 19.9
B 1501 501 71 0:0143hin 0.47 6.72

mately the size of the plume width. Due to this apparent difference
in the range of scales, the density currents are often classified as
either turbulent (case A) or laminar (case B) density currents.
[11,26]. This important difference between the two cases will be
discussed further in subsequent sections of this paper.

(a) Case A (b) Case B


Fig. 3. 3D views of the computational grid systems for (a) case A and (b) case B. The
grid view is stretched such that in streamwise to spanwise to vertical directions is
1:1:3.

5. Results and discussion

5.1. Plume shape at quasi-equilibrium state

Animations in Movies S1 and S2 show the computed instanta-


neous contours of tracer dye concentration released after quasi-
Movie S1. Case A: Instantaneous contours of tracer dye concentration released
equilibrium state of the dense plume is reached for cases A and after quasi-equilibrium state of the dense plume is reached.
B, respectively. The tracer contour levels in Fig. 4 show the com-
puted shape of the density current plume on the slope for both
cases. The dashed lines denote the visually identified boundaries
from the corresponding experimental photographs. When compar-
ing the plume with experiments, it should be noted that an accu-
rate comparison of contour levels with experimental images are
not possible because of issues related to reproduction of corre-
sponding dye visual. The plume shape and dye intensity in the
experimental image is affected by the quality of ambient light,
camera angle, and location. The color pixel at any point in the
image is representative of the superimposition of all fluid elements
from free surface to the bed. In order to plot the contours in simu-
lation results, we obtained maximum dye concentration in the ver-
tical direction for every point on the bed. In spite of the
aforementioned limitations and approximations, for both cases,
the simulated plume follows the experimental plume shape very
well initially. However, for distances far from the inlet, the plume
spreading is somewhat over-predicted. We also note that the
experimental plume shape is slightly asymmetric in the Y-
direction. This could be due to asymmetries or other spurious dis-
turbances inherent in the experimental setup. The over-prediction
of the current width away from the inlet was noted in previous Movie S2. Case B: Instantaneous contours of tracer dye concentration released
numerical simulations with similar Richardson and Reynolds num- after quasi-equilibrium state of the dense plume is reached.
bers (Re ¼ 1024; Ri ¼ 0:28) by Venetsanos et al. [57]. One striking
difference observed between the A and B cases is in the range of
scales of motion present in each case and visualized by the dye tra- In Fig. 5, the tracer concentration contours are plotted on the
cer. The case A appears to have multiple range of length scales, Y ¼ 0 plane passing through the center of the inlet opening. The
whereas case B shows prominent large scales which are approxi- laminar case B shows multiple current heads which can be

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx 5

Fig. 4. Comparison of computed (top) and measured (bottom) plumes at steady state for the case A (left) and B (right). Case A and B refer to the stratified flow on slopes of 15
and 5 angle, respectively. Contours are levels of tracers released form the inlet. Dotted lines mark the shape of plume in the corresponding experiments. The contours in the
computations denote the concentration of the dye in volume fraction. Flow is from top to bottom.

Fig. 5. Tracer concentration contours in a cross-sectional y ¼ 0 plane showing the structure of the current head t ¼ 200b0 =uin after the tracer release was started for case A
with 15 slope (top) and case B with 5 slope (bottom). The contours denote the concentration of the dye in volume fraction. Contours in (B) are scaled 3 times in vertical
direction for visibility. Flow is from right to left.

associated with the large-scale structures generated near the inlet respectively. Comparison between the measured and simulated
(see Section 5.2), and are compatible with the experimental obser- values show that the computed mean lobe spacing is within 15%
vations mentioned in Section 3. of the measured value and the computed lobe sizes fall in compa-
Focusing on the large-scale structures or the lobes in the case B rable range with measured lobe size (within 12%).
at quasi-equilibrium state, we measured the spacing between the
lobes as well as their sizes. First, we note that the lobes are small- 5.2. Effect of stratification on coherent structures
est near the inlet and evolve into bigger structures moving down-
stream. The lobes range from 19 to 58 cm in the computed plume To help us understand the effect of stratification on submerged
whereas the lobes in the instantaneous image of experimental jet, we look at the dynamics of the vortex formation close to the
plume range from 20 to 52 cm. The mean spacing among the lobes inlet immediately after the stratified jet is released. This is juxta-
in experiment and computation were 33 cm (ranging between 20 posed with a separate simulation with a similar setup (as Case B
cm and 55 cm) and 38 cm (ranging between 19 cm and 66 cm), but without stratification). Fig. 6 shows the iso-surfaces of the

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
6 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

second scalar invariant of the velocity gradient tensor (q-criterion) flow in this region. Further downstream, these organized struc-
visualizing the vortical coherent structures in the flow near the tures start to break up as flow undergoes transition to turbulence.
inlet region. At t ¼ tuin =h  1, in both cases, this value of q shows A swarm of hairpin-like vortices can be seen dominating the flow
a single vortex which roughly follows the shape of the inlet. In a downstream of this region. For unstratified case, the flow acceler-
region immediately next to the jet inlet, regularly repeating, orga- ation is slow, owing to the absence of buoyancy. For 3 < t < 30,
nized spanwise structures exist, revealing the laminar nature of the this effect is seen in the longitudinal propagation of plume (see

Fig. 6. Iso-surfaces of second scalar invariant of the velocity gradient tensor (q-criterion) visualizing vortical structures in stratified (left) and unstratified (right) submerged
jet at non dimensional time t  ¼ tuin =h between 0 and 30 for case B with 5 slope. Flow is from top-left to the bottom-right.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx 7

Movie S4. Submerged jet without stratification: Iso-surfaces of q-criteria struc-


tures near the inlet from time T ¼ tuin =h ¼ 0 when the release of water of ambient
density was started. The geometrical details and other conditions are same as case B.

Fig. 7. Iso-surfaces of second scalar invariant of the velocity gradient tensor (q-
criterion) on the slope after the plume reached quasi-equilibrium state for case B
(5 slope). Three different large-scale C-shaped structures are visible. Flow is from
top to the bottom.

Movie S4). A striking feature of the stratified case, as evident from


these figures, is the appearance of large-scale, C-shaped structures,
which are found to emanate aperiodically from the inlet and sweep
through the entire plume (see Movies S3 and S5). One such large-
scale structure can be clearly seen in Fig. 6 at t ¼ 30. Fig. 7 and
Movie S5 show several of these structures on the slope from top
view long after the plume shape reached equilibrium.

Movie S5. Iso-surfaces of q-criteria structures on the slope (top view) after the
quasi-equilibrium state of the plume is established for case B. Several C-shaped
structures can be seen emanating from near the inlet.

5.3. Spreading of dense plume

As the dense fluid travels from the finite source (inlet opening)
down the slope it entrains the lighter ambient fluid and spreads
longitudinally as well as laterally while losing momentum. The
nature of the spreading governing the equilibrium shape of the
plume is dependent on several factors, such as bed slope angle
(h), Ri, Re, inlet buoyancy flux, and inlet momentum flux. The
extent of lateral and longitudinal spreading of the dense plume
on slope has been the subject of several studies [16,12,25,55]. Tsih-
rintzis and Alavian [55] identified different spreading regimes of
the flow of density currents on a slope based on the type of dom-
inating forces in the flow - inertia-viscous, buoyancy-viscous and
gravity-viscous for mild slopes; and inertia-viscous and gravity-
Movie S3. Submerged jet with stratification: Iso-surfaces of q-criteria structures viscous for steep slopes.
near the inlet from time T ¼ tuin =h ¼ 0 when the release of saline fluid was started Herein, we use our simulation results to obtain the shape of the
in case B. plume and quantify its spreading. In Fig. 8, the computed plume

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
8 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

Fig. 8. Spreading law for lateral spreading of the steady state dense plume. Left: Case A with 15 slope and lower buoyancy; right: Case B with 5 slope and higher buoyancy.

half-width (b1=2 ¼ b=2) is plotted as a function of the streamwise Horsch [27]. Yet, it is considerably smaller than the experimental
distance from the origin (l) in log-log scale. The lengths in Fig. 8 value for low l=lB < 4. Another regime (l=lB > 28), with n ¼ 0:30 is
are normalized by the buoyancy length scale (lB ), which reads as identifiable at large distances, which is compatible with the sug-
follows: gested value of 0:33 [12].
The discrepancies between simulation and experiments at very
1=5
lB ¼ ðQ 3 =BÞ ð8Þ low values of l=lB , i.e. close to the source, can be attributed to small
differences in the specification of idealized boundary conditions,
where B ¼ gQ Dq=q is the buoyancy flux at the inlet of the jet. For such as constant velocity through the inlet as compared to non-
the conditions of the experiments A and B, the values of lB are uniform inlet velocity profile in the experiment. At large distances,
4.58 cm and 2.43 cm, respectively. Fig. 8 also includes dashed lines the experimentally observed smaller rate of growth (which is con-
outlining the two plumes from the experimental results as shown in sistent with an earlier numerical study [57]) may be due to a more
Fig. 4. Researchers have tried to extract a power law equation for predominant effect of bottom friction compared to the smooth
the variation of plume’s width versus its length, which reads as fol- wall assumed in the simulation, or even to inaccuracies in defining
lows in dimensionless form: the visual boundaries at very low concentrations in the experi-
 n ment. Furthermore, the procedure of defining the plume half-
b1=2 l
¼A ð9Þ width can also be responsible for some discrepancies. In order to
lB lB
obtain half-width, first, an outline boundary that contains the
On a log-log scale, linear regions are identifiable in the plots, indi- plume, as visualized by the dye, is drawn on the bed slope. At
cating the aforementioned power law. Based on the results of a any longitudinal distance (X), the half-width is then estimated by
large number of experiments on several slopes, Christodoulou averaging the extent of plume in lateral direction (Y). As can be
[12] concluded that (i) at l=lB  10, the plume half-width reaches seen from the contours in Fig. 4, the outline is not strictly defined.
a value b1=2 =lB  10 irrespective of slope and source conditions; The problem is further exacerbated in the case of experiments
because: (i) the plume is not symmetrical; (ii) the image is a per-
(ii) the power law indices expressing the lateral spreading, as per
spective view distorting the distances and (iii) only single time
Eq. (9), are estimated as n  1=3 for l=lB > 10 and n  1=2 for
instant is available from experimental results for calculation of
2 < l=lB < 10, being higher for larger slope. Using similarity argu-
plume width (unlike the numerical simulation where averaging
ments Hauenstein and Dracos [25] proposed a proportional rela-
over multiple time-instants was done).
tionship (l  b) between longitudinal and lateral spreading at
steady state. However, the similarity solution has been invalidated
theoretically as well as experimentally [10]. By means of theoretical 5.4. Spectral analysis
considerations Horsch [27] showed that the spanwise width of a
laminar density current flowing down an incline in the gravity- From the animation visualizing the vortical structures in Sec-
viscous dominated regime scales as the 7/9th power of the longitu- tion 5.2, the shedding of large-scale C-shaped structures from the
dinal distance. inlet region was observed. In order to further investigate the coher-
As seen in Fig. 8, the spreading results deduced from the simu- ent structures in the flow, velocity time series data was collected
lation are in reasonable agreement with the respective experi- for both cases on the plane of symmetry at several locations near
ments. In particular, in both cases the normalized half-width at the bed. Figs. 9 and 10 show streamwise velocity time series and
l=lB ¼ 10 is very close to 10, i.e. the value suggested in Christodou- its mean power spectra of velocity fluctuations for these locations.
lou [12]. The power law index n ¼ 0:55 obtained from the simula- At all four locations, the low frequency range show a constant
tion for case B is close to the experimentally infered value of decay of energy. For higher intermediate frequencies, the mean
5=3
n ¼ 0:5 for l=lB < 10 in Christodoulou [12]. Additionally, the same power spectra of velocity fluctuations exhibit f scaling, indicat-
value of n ¼ 0:55 is found to hold for a dominant part of the range ing the well-known inertial subrange, followed by sharp decay in
tested, including l=lB > 10. On the other hand, the value n ¼ 0:82 the viscous dissipation region. The stretch of the inertial subrange
deduced from the case A simulation over most of the region region decreases as we move downstream. For mean power spectra
(2 < l=lB < 28) is quite close to the 7=9 power law suggested in in case B, the slow frequencies of large-scale eddies can be seen

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx 9

0.15

0.14

0.13

Velocity (m/s) 0.12

0.11

0.1

0.09

0.08
(a) 0 10 20 30 40 50 60
Time (s)

0.16

0.14
Velocity (m/s)

0.12

0.1

0.08

0.06

(b) 0 10 20 30 40 50 60
Time (s)

0.14

0.12

0.1
Velocity (m/s)

0.08

0.06

0.04

0.02
0 10 20 30 40 50 60
(c) Time (s)

0.08
Velocity (m/s)

0.06

0.04

0.02

(d) 0 10 20 30 40 50 60
Time (s)

Fig. 9. Time series and mean power spectra of velocity fluctuations at (a) (6:3h0 ; 0:37h0 ), (b) (13h0 ; 0:54h0 ), (c) (22h0 ; 0:78h0 ), and (d) (34h0 ; 1:1h0 ) for case A with 15 slope
and lower buoyancy. The pairs denote streamwise distance and height above the bed.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
10 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

Fig. 10. Time series and mean power spectra of velocity fluctuations at (a) (12:5h0 ; 0:22h0 ), (b) (33h0 ; 0:35h0 ), (c) (65h0 ; 0:55h0 ) and (d) (102h0 ; 0:78h0 ) for case B with 5 slope
and higher buoyancy. The pairs denote streamwise distance and height above bed.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx 11

5=3
the well known f decay in the inertial subrange of the mean
power spectra of velocity fluctuations was also found. Further-
more, comparison of vortical structures of stratified case with
unstratified cases highlighted the role of stratification in the for-
mation of density current plume.
The spreading of the simulated plumes exhibit power law scal-
ing with exponents comparable to those determined by earlier
researchers. The plume shape as well as spreading compared well
in the region away from the inlet, but not close to the exit bound-
ary. Close to the far end of the slope the comparison of the plume
shape is difficult because of the ill-defined boundary of the plume
in this region for case B. Given that only two experiments are sim-
ulated owing to computational limitations, more research is
needed to further study the effect of different bed slopes and
Reynolds and Richardson numbers on the spreading laws. Compu-
tational paradigms such as local mesh-refinement (as in Angelidis
Fig. 11. Pre-multiplied mean power spectra of velocity fluctuations for four points
et al. [2]) are needed to perform faster high-resolution calculations.
in the dense plume showing prominent frequency of large-scale structures at
St ¼ 0:0122 for case B with 5 slope.
Conflict of interest

with distinctly larger amplitude compared to the inertial sub- I wish to confirm that there are no known conflicts of interest
range. Unlike this, for case A one can barely find such distinct fre- associated with our article.
quency in the spectra.
For case B, there is an identifiable peak in all of the spectra close Acknowledgment
to frequency f ¼ 0:05 Hz. This peak in frequency is better seen in
the following Fig. 11. In this figure, we plot the pre-multiplied Computational resources for this work were provided by the
spectra i.e. frequency times mean power spectral density (PSD) of Minnesota Supercomputing Institute, University of Minnesota.
the streamwise velocity fluctuations at four points for case B in the
plane of symmetry, away from the bed but close enough to lie in
the dense current. The streamwise distance and height above bed References
for the four points are - (21h0 ; 1:7h0 ), (33h0 ; 2:0h0 ), (47h0 ; 2:6h0 ),
[1] V. Alavian, Behavior of density currents on an incline, J. Hydraul. Eng. 112 (1)
and (102h0 ; 4:7h0 ). In all four of the spectra considered here, the (1986) 27–42.
lowest frequency, corresponding to the largest structures, are the [2] D. Angelidis, S. Chawdhary, F. Sotiropoulos, Unstructured cartesian refinement
same (i.e. St ¼ 0:0122). These structures are shed intermittently with sharp interface immersed boundary method for 3d unsteady
incompressible flows, J. Comput. Phys. 325 (2016) 272–300.
from the inlet, as could be seen in the Movies S3 and S5, which [3] V. Birman, B. Battandier, E. Meiburg, P. Linden, Lock-exchange flows in sloping
visualize iso-surface of second scalar invariant of the velocity gra- channels, J. Fluid Mech. 577 (2007) 53–77.
2 [4] T. Bonometti, S. Balachandar, J. Magnaudet, Wall effects in non-Boussinesq
dient tensor (at value of q ¼ 0:01u2in =h0 ). One such large-scale
density currents, J. Fluid Mech. 616 (2008) 445.
structure is shown in Fig. 6 at t  ¼ 30 for stratified case. In conclu- [5] T. Bonometti, M. Ungarish, S. Balachandar, A numerical investigation of high-
sion, the spectral analysis corroborates the existence large-scale reynolds-number constant-volume non-Boussinesq density currents in deep
ambient, J. Fluid Mech. 673 (2011) 574–602.
structures in case B which are not present in case A.
[6] R. Britter, P. Linden, The motion of the front of a gravity current travelling
down an incline, J. Fluid Mech. 99 (03) (1980) 531–543.
[7] M.I. Cantero, S. Balachandar, M.H. Garcia, High-resolution simulations of
6. Conclusion cylindrical density currents, J. Fluid Mech. 590 (2007) 437–469.
[8] M.I. Cantero, J. Lee, S. Balachandar, M.H. Garcia, On the front velocity of gravity
currents, J. Fluid Mech. 586 (2007) 1–39.
We employed the LES module of the VFS-Geophysics model to [9] S. Chawdhary, C. Hill, X. Yang, M. Guala, D. Corren, Jonathan Colby, F.
simulate a laboratory-scale turbulent stratified flow of a dense wall Sotiropoulos, Wake characteristics of a triframe of axial-flow hydrokinetic
jet flow over two sloping beds of 5 and 15 , which were also turbines, Renew. Energy 109 (2017) 332–345.
[10] S.-U. Choi, M.H. Garcia, Spreading of gravity plumes on an incline, Coast. Eng. J.
experimentally studied by Christodoulou [12]. 43 (04) (2001) 221–237.
The stratified jet resulted in formation of density current [11] G. Christodoulou, F. Tzachou, Experiments on 3-D turbulent density currents,
plumes on the slope. The numerical simulations are compared with in: Preprints of 4th Intern. Symposium on Stratified Flows, Grenoble, France,
1994.
experimental results and shown to be in good qualitative agree- [12] G.C. Christodoulou, Lateral growth of 3-D density currents, in: Proceedings of
ment in terms of plume shape at quasi-equilibrium state. Visual- the 29th Congress-International Association for Hydraulic Research, Beijing,
izations of the simulated coherent structures shed new striking China, vol. B, 2001, pp. 29–34.
[13] G. Constantinescu, Les of lock-exchange compositional gravity currents: a brief
insights into the physics of such flows, revealing rich dynamics review of some recent results, Environ. Fluid Mech. 14 (2) (2014) 295–317.
and large-scale instabilities. For the low slope angle case, very- [14] A. Dai, C. Ozdemir, M. Cantero, S. Balachandar, Gravity currents from
large-scale C-shaped structures appear to be emanating from the instantaneous sources down a slope, J. Hydraul. Eng. 138 (3) (2011) 237–246.
[15] T. Ellison, J. Turner, Turbulent entrainment in stratified flows, J. Fluid Mech. 6
inlet region. These structures appear aperiodically and travel down (03) (1959) 423–448.
the slope. A closer look at the image of the experimental plume [16] T.R. Fietz, I.R. Wood, Three-dimensional density current, J. Hydraul. Div. 93 (6)
reveals regions of high dye concentrations in large C-shapes, indi- (1967) 1–24.
[17] B. Firoozabadi, H. Afshin, E. Aram, Three-dimensional modeling of density
cating the presence of double C-shaped (or W-shaped) structures
current in a straight channel, J. Hydraul. Eng. 135 (5) (2009) 393–402.
instead of single. This difference could be due to any slight uncer- [18] L. Ge, F. Sotiropoulos, A numerical method for solving the 3D unsteady
tainty and/or asymmetry associated with the experimental condi- incompressible Navier-Stokes equations in curvilinear domains with complex
tions which are known to effect the flow regimes in the plume immersed boundaries, J. Comput. Phys. 225 (2) (2007) 1782–1809.
[19] M. Germano, U. Piomelli, P. Moin, W.H. Cabot, A dynamic subgrid-scale eddy
[11,12]. A spectral analysis of velocity in the plume shows the fre- viscosity model, Phys. Fluids A: Fluid Dyn. (1989–1993) 3 (7) (1991) 1760–
quency of the large-scale structures as St ¼ 0:0122. Existence of 1765.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063
12 S. Chawdhary et al. / International Journal of Heat and Mass Transfer xxx (2017) xxx–xxx

[20] E. Gonzalez-Juez, E. Meiburg, G. Constantinescu, The interaction of a gravity [39] T. Maxworthy, R. Nokes, Experiments on gravity currents propagating down
current with a circular cylinder mounted above a wall: effect of the gap size, J. slopes. Part 1. The release of a fixed volume of heavy fluid from an enclosed
Fluids Struct. 25 (4) (2009) 629–640. lock into an open channel, J. Fluid Mech. 584 (2007) 433–454.
[21] E. Gonzalez-Juez, E. Meiburg, T. Tokyay, G. Constantinescu, Gravity current [40] A. Ooi, N. Zgheib, S. Balachandar, Direct numerical simulation of three-
flow past a circular cylinder: forces, wall shear stresses and implications for dimensional gravity current on a uniform slope, Proc. Eng. 126 (2015) 372–
scour, J. Fluid Mech. 649 (2010) 69–102. 376.
[22] J. Hacker, P. Linden, S. Dalziel, Mixing in lock-release gravity currents, Dyn. [41] S.K. Ooi, G. Constantinescu, L. Weber, A numerical study of intrusive
Atmosph. Oceans 24 (1) (1996) 183–195. compositional gravity currents, Phys. Fluids (1994–present) 19 (7) (2007)
[23] C. Härtel, F. Carlsson, M. Thunblom, Analysis and direct numerical simulation 076602.
of the flow at a gravity-current head. Part 2. The lobe-and-cleft instability, J. [42] S.K. Ooi, G. Constantinescu, L. Weber, Numerical simulations of lock-exchange
Fluid Mech. 418 (2000) 213–229. compositional gravity current, J. Fluid Mech. 635 (2009) 361–388.
[24] C. Härtel, E. Meiburg, F. Necker, Analysis and direct numerical simulation of [43] S.K. Ooi, G. Constantinescu, L.J. Weber, 2D large-eddy simulation of lock-
the flow at a gravity-current head. Part 1. Flow topology and front speed for exchange gravity current flows at high Grashof numbers, J. Hydraul. Eng. 133
slip and no-slip boundaries, J. Fluid Mech. 418 (2000) 189–212. (9) (2007) 1037–1047.
[25] W. Hauenstein, T. Dracos, Investigation of plunging density currents generated [44] J. Paik, A. Eghbalzadeh, F. Sotiropoulos, Three-dimensional unsteady rans
by inflows in lakes, J. Hydraul. Res. 22 (3) (1984) 157–179. modeling of discontinuous gravity currents in rectangular domains, J. Hydraul.
[26] G.M. Horsch, The structure of two-dimensional, steady, miscible laminar Eng. 135 (6) (2009) 505–521.
density currents flowing down an incline, J. Hydraul. Res. 42 (2) (2004) 173– [45] J.D. Parsons, M.H. Garcıa, Similarity of gravity current fronts, Phys. Fluids
181. (1994–present) 10 (12) (1998) 3209–3213.
[27] G.M. Horsch, Scaling of three-dimensional, miscible, laminar density currents [46] A.N. Ross, P. Linden, S.B. Dalziel, A study of three-dimensional gravity currents
flowing down an incline, in: E-Proceedings of the VIII International Conference on a uniform slope, J. Fluid Mech. 453 (2002) 239–261.
on Protection and Restoration of the Environment, Chania, Greece, Chania, [47] J.E. Simpson, Gravity Currents: In the Environment and the Laboratory,
Greece, 2006. Cambridge University Press, 1999.
[28] H.E. Huppert, J.E. Simpson, The slumping of gravity currents, J. Fluid Mech. 99 [48] J. Smagorinsky, General circulation experiments with the primitive equations:
(04) (1980) 785–799. I. The basic experiment, Monthly Weather Rev. 91 (3) (1963) 99–164.
[29] S. Kang, A. Lightbody, C. Hill, F. Sotiropoulos, High-resolution numerical [49] T. Tokyay, G. Constantinescu, E. Gonzalez-Juez, E. Meiburg, Gravity currents
simulation of turbulence in natural waterways, Adv. Water Resour. 34 (1) propagating over periodic arrays of blunt obstacles: effect of the obstacle size,
(2011) 98–113. J. Fluids Struct. 27 (5) (2011) 798–806.
[30] S. Kang, X. Yang, F. Sotiropoulos, On the onset of wake meandering for an axial [50] T. Tokyay, G. Constantinescu, E. Meiburg, Lock-exchange gravity currents with
flow turbine in a turbulent open channel flow, J. Fluid Mech. 744 (2014) 376– a high volume of release propagating over a periodic array of obstacles, J. Fluid
403. Mech. 672 (2011) 570–605.
[31] A. Khosronejad, A. Hansen, J. Kozarek, K. Guentzel, M. Hondzo, M. Guala, P. [51] T. Tokyay, G. Constantinescu, E. Meiburg, Tail structure and bed friction
Wilcock, J. Finlay, F. Sotiropoulos, Large eddy simulation of turbulence and velocity distribution of gravity currents propagating over an array of obstacles,
solute transport in a forested headwater stream, J. Geophys. Res.: Earth Surface J. Fluid Mech. 694 (2012) 252–291.
121 (1) (2016) 146–167. [52] T. Tokyay, G. Constantinescu, E. Meiburg, Lock-exchange gravity currents with
[32] A. Khosronejad, C. Hill, S. Kang, F. Sotiropoulos, Computational and a low volume of release propagating over an array of obstacles, J. Geophys.
experimental investigation of scour past laboratory models of stream Res.: Oceans 119 (5) (2014) 2752–2768.
restoration rock structures, Adv. Water Resour. 54 (2013) 191–207. [53] T.E. Tokyay, M.H. García, Effect of initial excess density and discharge on
[33] A. Khosronejad, J. Kozarek, A. Hansen, K. Guentzel, M. Hondzo, P. Wilcock, M. constant flux gravity currents propagating on a slope, Environ. Fluid Mech. 14
Guala, J. Finlay, F. Sotiropoulos, Data-driven LES of turbulence and solute (2) (2014) 409–429.
transport in a natural stream, Bull. Am. Phys. Soc. 59 (2014). [54] V.A. Tsihrintzis, Theoretical and Experimental Investigation of Three-
[34] A. Khosronejad, J. Kozarek, F. Sotiropoulos, Simulation-based approach for Dimensional Boundary-Attached Density Currents, Ph.D. thesis, Department
stream restoration structure design: model development and validation, J. of Civil Engineering, University of Illinois, Urbana, IL, USA, 1998.
Hydraul. Eng. 140 (9) (2014) 04014042. [55] V.A. Tsihrintzis, V. Alavian, Spreading of three-dimensional inclined gravity
[35] A. Khosronejad, T. Le, P. DeWall, N. Bartelt, S. Woldeamlak, X. Yang, F. plumes, J. Hydraul. Res. 34 (5) (1996) 695–711.
Sotiropoulos, High-fidelity numerical modeling of the upper Mississippi river [56] M. Ungarish, Intrusive gravity currents in a stratified ambient: shallow-water
under extreme flood condition, Adv. Water Resour. 98 (2016) 97–113. theory and numerical results, J. Fluid Mech. 535 (2005) 287–323.
[36] A. Khosronejad, F. Sotiropoulos, Numerical simulation of sand waves in a [57] A. Venetsanos, G. Horsch, G. Christodoulou, Assesment of turbulence
turbulent open channel flow, J. Fluid Mech. 753 (2014) 150–216. modelling of density currents developing three dimensionally on a slope, J.
[37] A. Khosronejad, F. Sotiropoulos, On the genesis and evolution of barchan Mar. Environ. Eng. 8 (2) (2005) 147–154.
dunes: morphodynamics, J. Fluid Mech. 815 (2017) 117–148. [58] D. Wilkinson, I. Wood, Some observations on the motion of the head of a
[38] H. Lim, Y. Yu, J. Glimm, D.H. Sharp, Mathematical, physical and numerical density current, J. Hydraul. Res. 10 (3) (1972) 305–324.
principles essential for models of turbulent mixing, in: Nonlinear Conservation
Laws and Applications, Springer, 2011, pp. 405–413.

Please cite this article in press as: S. Chawdhary et al., Large eddy simulation of density current on sloping beds, Int. J. Heat Mass Transfer (2017), https://
doi.org/10.1016/j.ijheatmasstransfer.2017.12.063

You might also like