You are on page 1of 16

National Conference on Water Resources & Flood Management with special reference to Flood Modelling

October 14-15, 2016 SVNIT Surat

HYDRODYNAMIC CHARACTERISTICS OF FLOWS IN A TWO-


LAYERED COMPOUND OPEN CHANNELS BY USING DYNAMIC
SGS MODEL

J. Sinha1*, P. L. Patel2, Samir K. Das3 and B. K. Samtani4


1
Scientist-D, Central Water and Power Research Station, Pune - 411024; Email :
jsinha_cwprs@yahoo.com
2
Professor, Department of Civil Engineering, Sardar Vallabhbhai National Institute of Technology,
Surat – 395007; Email : plpatel@ced.svnit.ac.in
3
Professor, Department of Applied Mathematics, Defence Institute of Advanced Technology, Pune –
411025; E-mail: samirkumar_d@yahoo.com
4
Professor, Department of Civil Engineering, Sardar Vallabhbhai National Institute of Technology,
Surat – 395007; Email : bks@ced.svnit.ac.in

ABSTRACT

In this paper, hydrodynamic characteristics due to uniform turbulent flow in compound open-
channels are simulated using two-layered two-dimensional model with dynamic subgrid scale (SGS)
scheme under large eddy simulation (LES) technique. The model has been validated using
experimental results of a wide symmetric and a narrow asymmetric compound open-channels
conveying shallow flows ( rh ≥ 3) and deep flows ( rh ≤ 2) respectively. The comparison has been
made for stream wise flow velocity in the channels. The developed model is used to explore the roles
of flow variables at horizontal interface in transfer of mass and momentum from lower layer to upper
layer near the junctions of the main channel and the floodplains.

Keywords: Upper and lower layers, deep and shallow flows, predictor-corrector approach, dynamic
subgrid scale, finite volume method

1. INTRODUCTION

Natural streams are like a compound section having a deep sub-section to convey low flows
and one or two shallower sub-section(s) either on one or both sides of the deeper section to
pass high flows during floods. A strong reaction in terms of velocity difference between
deeper and shallower sub-sections occurs, giving rise to generation of secondary flows in the
interface, which is responsible for transfer of lot of momentum in these regions. Until
recently, many researchers have been working in laboratory experiments (Tominaga and
Nezu, 1991; Joung and Choi 2008; Stocchino et al., 2011) and using numerical models (Naot
et al., 1993; Cater and Williams, 2008; Xie et al., 2013) on hydraulics of compound sections
along the vertical or inclined interface at the two sub sections whereas a little attention has
been paid to the flow characteristics along the horizontal interface.

A number of experimental (using Pitot static tube, Preston tube, hot film anemometers, two-
component laser Doppler anemometers, dual layer PIV system) and numerical model (using
linear/ non-linear k–ε, SKM method, RSM) studies reported in the past in compound open-
channel flows, revealed that intensive mass and momentum exchange takes place between
two sub-sections (main channel and floodplains) depending on their relative velocities, and
cross-sectional areas. Combination of velocity differences and complex cross-sectional area
develop natural shear instability in the vicinity of the junction between the main channel and
floodplains resulting into the formation of 3-D turbulent structures and, subsequently,
induced inclined secondary currents. Existing 2-D and 3-D numerical models have their own

WRF-24-1
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

advantages and disadvantages in prediction of secondary flows in compound open-channels.

Recently developed two-layered 2-D hydrodynamic model for predicting uniform turbulent
flow in compound open channels (Sinha et al., 2014) has been further extended for dynamic
subgrid scale (DSGS) turbulent closure scheme under large eddy simulation (LES) technique
to investigate characteristics of flow variables along the horizontal interface between the
lower main channel and, the upper main channel and floodplain taken together. The LES is
used in turbulent flow computation where the large-scale motions laden with high energy are
resolved, and the small-scale residual components are modelled. In this paper, the
Smagorinsky coefficient, varying in both space and time, has been treated as a dynamic
coefficient (Germano et al., 1991). The model has been used to simulate the laboratory
experimental results reported by Fraselle et al. (2008) and, Shiono and Feng (2003) for a
wide symmetric and a narrow asymmetric compound open-channel respectively.

2. TWO LAYER SHALLOW WATER FLOW EQUATIONS

A two-layered 2-D numerical model, equivalent to 3-D pseudo model, developed by Chau
and Jin (1995) for tidal flow in a boundary fitted orthogonal curvilinear coordinate system,
was extended by Sinha et al. (2014) for non-uniform rectangular grid system by incorporating
the DSGS closure scheme under LES approach. The model parameter of the scheme is
continuously updated to make the model feasible for studying the turbulent flow
characteristics at the horizontal interface of the lower main channel as lower layer and, the
upper main channel and floodplains taken together as upper layer in compound open channel.
In LES, spatially filtered layer-averaged governing equations of continuity and momentum
for lower and upper layers of compound channel, deduced from the filtered Navier-Stokes
equations, can be written (over bar is not shown) in the form of Eqs.(1-6). Pressure terms
appearing in hydrodynamic equations of both the layers are approximated with channel slope
and water depths in respective layers.

Equations for upper layer:

hu (hu uu ) (hu vu )


   wo (1)
t x y
(huuu ) (huuu2 ) (huuu vu ) h   hu xu    hu yxu   sx   ox (2)
   wouo  hu g u  hu gso     
t x y x x    y    
(huvu ) (huuu vu ) (huvu2 ) h   hu xyu    hu yu   s y   oy
   wovo  hu g u      (3)
t x y y x    y    

Equations for lower layer:


hl (hl ul ) (hl vl )
    wo (4)
t x y
(hl ul ) (hl ul2 ) (hl ul vl ) h   hl xl    hl yxl   ox   bx
    wouo  hl g u  hl gso      (5)
t x y x x    y    
(hl vl ) (hl ul vl ) (hl vl2 ) h   hl xyl    hl yl   o y   by
    wovo  hl g u      (6)
t x y y x    y    

Here, subscripts ‘u’ and ‘l’ designate the variables corresponding to upper and lower layers

WRF-24-2
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

respectively. In other subscripted variables, the symbols appearing as ‘o’, ‘s’ and ‘b’ denote
the position at the interface, surface and bed respectively. Here, hk, uk and vk represent filtered
water depth streamwise and spanwise layered average velocity respectively in kth layer;  ik is
layered average normal stress in ith direction in kth layer;  ijk is layered average filtered shear
stress at jth face in ith direction in kth layer;  si ,  oi and  bi are filtered shear stress in ith
direction at surface, interface and bed respectively; uo, vo and wo are filtered velocity at
interface in streamwise, spanwise and vertical directions respectively;  = mass density of
water; so= channel bed slope; and g = gravitational acceleration.

The bed shear stresses and both, shear stresses and flow velocities, at interface of the two
layers appearing in governing equations (Eqs.1-6) is expressed empirically which is similar to
that adopted in Chau and Jin (1995).

The filtered bed shear stresses are represented empirically as


 bx ul  by vl
 u*2 ;  u*2 (7)
 ul2  vl2  ul2  vl2
where u* is the friction velocity, and is defined as

 
1/ 2
u*  c f um2  vm2  (8)
 
hu uu  hl ul h v hv
where um  ; vm  u u l l (9)
h h
and h  hu  hl , representing total depth of water.
A horizontal interface is assumed at the top of lower layer, and c f is the coefficient of bed
resistance, determined using Eq.(10) as
n2 g
cf  (10)
h1/ 3
where n is the Manning’s roughness coefficient.

The filtered shear stresses at interface between two layers are approximated empirically as
 ox u  ul  oy v v
  mo u ;   mo u l
    (11)
where  mo  khl 1 
 ; and  is the mixing layer thickness at interface which can be
hl
 u*
  h
approximated as Prandtl mixing length, l0 , defined as
1/ 2
 hl 
  l0  khl 1  (12)
 h 
where k is the von-Karman constant, equals to 0.41.

The filtered surface shear stress terms, i.e.,  sx and  s y are not taken into account as the effect
of wind conditions have been assumed to be negligible in present model.

WRF-24-3
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

The filtered horizontal interfacial velocities (Chau and Jin, 1995) in streamwise and spanwise
directions can be expressed approximately as
h u hu h v hv
uo  u l l u ; vo  u l l u (13)
h h
The vertical interface velocity, w0 , resulting due to convective exchange between layers, is
considered positive upward and determined from layer-averaged continuity equation (Eq.1
or 4) of either of the layers after obtaining flow velocities and water surface levels of
respective layers.

3. DYNAMIC TURBULENCE MODEL

The filtered normal and shear stress terms are modeled using
 u u 
uu   tk  i  j 
 x 
(14)
 j xi k
i j k

where  tk is the SGS diffusivity which is modeled in present study using DSGS method for
each layer separately.

The SGS eddy viscosity is expressed as


t  C  2
k  2Sij Sij  k
(15)

1  u u 
where, Sijk   i  j  is the strain rate;    x.y  is filter width, representing the
1/ 2

2  x j xi k
characteristic length scale of the largest SGS eddies, and C is the model parameter specified
using recommendations of past studies (Germano et al., 1991). In this paper, the value of C is
determined using the dynamic method determined using resolved variables through filtering
of governing equations at two different scales. The value of C is determined using least
square minimization approach as
1 Lij M ij
C ( y, t )   k (16)
2 M ij M ij
k

where, < - > represents the spatial averaging in homogenous directions so as to stabilize the
computations following the summation convention of the repeated indices. In this paper,
streamwise direction, both in upper and lower layers of compound channel, is considered
homogenous as variation of flow velocities in this direction is analogous.

In DSGS method, the model parameter, C, is allowed to vary locally in the numerical solution
through internal calculations while adjusting the level of eddy viscosity (i.e.  tk  0 ) in the
flow field. The procedure is initiated by specifying a value of C, and subsequently, after each
time level, a new value of C for each grid cell is determined based on flow field. The
dynamic model has been developed by introduction of a test filter which is equal to two times
of the grid filter. The stresses obtained due to test filter are referred to as subtest-scale (STS)
stresses, and are modeled separately for lower and upper layers which is similar to that of
SGS stresses. The discrete trapezoidal filter approach has been adopted for an efficient
implementation of test-filtered quantities (Premnath et al., 2009).

In this paper, Lij and Mij are calculated separately for lower and upper layers at each grid cell
before taking their summative average in streamwise direction (see Eq.16). In order to keep

WRF-24-4
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

positive value of C,  tk  0 is imposed for lower and upper layers in model domains.
Subsequently, average of C both for lower and upper layers are calculated for each grid cell
along spanwise direction.

4. NUMERICAL METHOD

In present study, the governing equations have been discretised using the cell-centered finite
volume method on non-uniform Cartesian grid, and approximated by an explicit two step
predictor-corrector scheme to ensure second-order accuracy both in space and time. The
evaluation of convective and diffusive terms has been carried out separately for lower and
upper layers both in streamwise and spanwise directions. Convective flux are computed at
cell faces using linear interpolation from the information of flow variables available to the
left and right sides of the cell face under consideration (Singh and Bhallamudi, 1997). For
treatment of convective flux terms, slope limiter method ‘minmod’ has been adopted as it is
able to suppress spurious oscillation near the discontinuities. Diffusive flux having first order
derivative terms has been calculated in such a way that it remains consistent with convective
flux. Each component of source term in each layer is evaluated at the centre of grid cell under
consideration. Discretised form of the governing equations is solved using explicit two step
predictor-corrector approach. Numerical stability of the existing explicit scheme is ensured
using CFL conditions.

5. MODEL SETUP

5.1 Experimental data

Performance of the two-layered 2-D turbulent model have been assessed using experimental
data of symmetric and asymmetric compound open-channel flows reported by Fraselle et al.
(2008) and, Shiono and Feng (2003) respectively. Schematic diagrams of symmetric and
asymmetric compound cross-sections are included in Figure 1.

(a) Symmetric compound channel (Fraselle et al., 2008)

WRF-24-5
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

(b) Asymmetric compound channel (Shiono and Feng, 2003)


Figure 1: Schematic diagrams of prismatic compound open-channel

The experimental set up of symmetric compound channel consisted of a 10 m long tilted


flume with total width, B=1.2 m, width of either side flanked floodplain, bf=0.4 m, width of
main channel, bm=0.4 m, total depth of main channel, H=0.0756 and depth of channel in
floodplain, hf=0.025 m. In the experiment, measuring section was located along streamwise
direction at 0.6L downstream of the channel inlet, i.e., at a distance of 6 m from the inlet
section. For measurement of flow velocity in the channel, pitot tube was used (Fraselle et al.,
2008). In another experimental set up consisting of 20 m long tilted narrow asymmetric
compound open-channel (Shiono and Feng, 2003), flow depths of the main channel and
floodplain were H = 0.11m and h = 0.055m respectively with their respective widths of 0.1 m
and total channel width, B=0.2 m. The measuring section was located along the streamwise
direction at 14 m from the upstream section) near the free surface (i.e. at 0.108 m above the
channel bed).

5.2 Numerical experiments

The flow characteristics of numerical experiments used in present study are shown in Table
1. It is noted that hydraulic conditions of symmetric (Re=5.5x104 , Fr=0.43) and asymmetric
(Re=2.6x104, Fr=0.23) compound open channels in aforesaid experiments fall in the category
of shallow flows (rh≥3) and deep flows (rh≤2), respectively (Nezu et al., 1999).

In the numerical experiments, uniform grid spacing has been adopted in streamwise direction
while non-uniform grid spacing is chosen to capture flow conditions at the junction and near
the solid walls. In case of symmetric compound channel, 500x70 (i.e. streamwise x spanwise)
grid cells in upper layer and 500x26 grid cells in lower layer and, in case of narrow
asymmetric compound channel, 2000x40 grid cells in upper layer and 2000x20 grid cells in
lower layer are adopted following the grid dependence test.

Grid cells at the side walls of compound channels are refined in spanwise direction based on
the limiting criteria (y+>30) of log-law. In case of symmetric compound channel, Manning’s
coefficients for lower (nl) and upper (nu) layers are considered as 0.01 s/m1/3 and 0.013 s/m1/3,
respectively. However, in case of asymmetric compound channel, 0.012 s/m1/3 is considered
as Manning’s coefficients for both lower (nl) and upper (nu) layers.

WRF-24-6
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

Table 1: Hydraulic characteristics of numerical experiments in symmetric and


asymmetric compound open channels
UPPER LOWER MEAN DEPTH
CHANNEL LOWER LAYER BULK SLOP RATIO, FLOW
TYPE width depth width depth FLOW E rh [ (hu  hl ) ] TYPE
bu hu bl hl VEL. S0 hu

(m) (m) (m) (m) (M/S) X10-3


Symmetric 1.2 0.025 0.4 0.0506 0.369 1.0 3.0 shallow
(Fraselle et al.,
2008)
Asymmetric 0.2 0.055 0.1 0.055 0.237 0.5 2.0 deep
(Shiono and
Feng, 2003)

At upstream boundary, discharge and flow depth are specified and the remaining quantity is
extrapolated from the cells adjacent to boundary. The total discharge is taken in proportion of
area of each layer, and flow depths are calculated using Manning’s relationship. In addition,
the random perturbations up to 5% of mean bulk velocity were superimposed in the flow
velocities in both directions (Xie et al., 2013). At downstream boundary of each layer, flow
depth and velocity components are extrapolated using their first order derivative from the
interior grid cells. The side wall boundaries are treated using the wall function technique
(Sinha et al., 2014). The flow variables in each grid cell of both layers are set at time, t = 0 as
initial condition.

6. RESULTS AND DISCUSSIONS

In this section, model results in the form of depth/ layer averaged flow velocities, bed shear
stresses and flow characteristics along the horizontal interface between the two layers are
discussed. For hydraulic flow characteristics, longitudinal test sections have been considered
at 6.0 m and 14.0 m downstream of channel entrance in case of symmetric and asymmetric
compound channels respectively.

6.1 Mean flow velocities and shear stresses

Comparison of computed and measured depth averaged streamwise flow velocities is


included in Figure 2 for symmetric and asymmetric compound open-channels. Flow
velocities of both the channels are normalized with their respective mean friction velocity,
U*. In symmetric compound channel conveying shallow flows (rh≥3.0), it is seen that model
slightly over predicts the flow in the floodplains, whereas in asymmetric compound channel,
conveying deep flows (rh≤2.0), the model over predicts the flows in the main channel and
under predicts the same in the floodplain. The percentage difference between depth-averaged
streamwise flow velocities of upper and lower layers with reference to bulk mean flow
 u  uu 
velocity  l x100  in the same direction is shown in Figure 3 for both the channels and
 Ub 
the same has found to be of the order of 12-13% near the junction of the main channel and
the floodplain.

WRF-24-7
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

(a) Symmetric compound channel (Fp=Floodplain, Mc=Main channel)

(b) Asymmetric compound channel


Figure 2: Comparison of depth-averaged primary flow velocity

(a) Symmetric compound channel

WRF-24-8
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

(b) Asymmetric compound channel

Figure 3: Percentage difference between longitudinal flow velocity of two layers


with respect to mean bulk velocity in compound channels

Here, ul and uu represent layer averaged streamwise flow velocities in lower and upper layers
respectively, and Ub (=Q/A) represents the bulk mean flow velocity in the same direction.
Such decrease in streamwise flow velocity (velocity-dip) near the junction was also reported
by Tominaga and Nezu (1991). Figures 4(a) and 5(a) show the interface flow velocities in
spanwise and vertical directions along horizontal interface of lower and upper layers of both
the channels. In present model, the magnitude of secondary current (layer averaged)
generated in symmetric and asymmetric compound channels are of the order of 1.2% (see
Figure 4c) and 0.015% (see Figure 5c) of their mean bulk flow velocities (Ub) respectively. It
is evident from the figures that both the velocities attain their maximum values near the
junction region. In both the channels, spanwise interface velocity shows gradual rise in
magnitude before attaining its maximum value very close to the junction. It can be observed
that, in symmetric compound channel flow, secondary circulation pattern are opposite in
nature in both half of the lower layer with zero exactly at mid width of the channel. On the
other hand, in asymmetric channel, only counter clockwise circulation pattern is present near
the junction with zero magnitude at the far end of the channel. In both the channels, the
secondary interface velocities are directed towards the junction of the channel

spanwise (v0) Vertical (w0)


0.004
Velocity at Interface (m/s)

0.003 Mc
0.002
0.001
5E-18
-0.001
-0.002
-0.003
-0.004
0.4 0.5 0.6 0.7 0.8
Channel width at Interface (m)

WRF-24-9
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

Figure 4(a): Spanwise and vertical flow velocity profile at interface


in symmetric compound channel

6.05
Test section (m)

5.95
0.4 0.5 0.6 0.7 0.8
Channel width at Interface (m)

Figure 4(b): Contour plot of secondary currents at interface


in symmetric compound channel

Figure 4(c): Vector plot of secondary currents at interface


in symmetric compound channel

Spanwise (v0) Vertical (w0)

5E-05 Mc
Velocity at Interface (m/s)

4E-05
3E-05
2E-05
1E-05
0
0 0.02 0.04 0.06 0.08 0.1
Channel width at Interface (m)

Figure 5(a): Spanwise and vertical flow velocity profile at interface


in asymmetric compound channel

WRF-24-10
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

14.05

Test section (m)


14

13.95
0 0.02 0.04 0.06 0.08 0.1
Channel width at Interface (m)

Figure 5(b): Contour plot of secondary currents at interface


in asymmetric compound channel

Figure 5(c): Vector plot of secondary currents at interface


in asymmetric compound channel

From Figures 4(b) and 5(b), it is inferred that spanwise velocity at interface in both the
channels increases from zero to its maximum value at the junction of the channel. The
vertical interface velocity, in case of symmetric compound channel, is having negligibly
smaller value near the middle reach of the channel and rises to maximum value near the
junction of the channel. Such phenomenon can also be seen in vector plots [Figures 4(c) and
5(c)] of secondary current. This implies that transfer of momentum generally occurs towards
floodplains in both the channels.

Formation of complex structures at junction of the main channel and the floodplain in
compound open-channel flows can also be explained in terms of bed shear stresses using two
layered model developed in present study. Formation of vortices at the junction can be
observed for both the channels in bed shear profile, and contours plots as shown in Figures
6(a-b) and 7(a-b).

WRF-24-11
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

1.6
Fp Half of Mc
1.4
1.2
1
0.8
0.6
0.4
0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Half of channel width (m)

Figure 6(a): Bed shear stress profile in symmetric compound channel


Test section (m)

6.05

5.95
0 0.2 0.4 0.6
Channel width (m)

Figure 6(b): Contour plot of bed shear stress in symmetric compound channel

Figure 7(a): Bed shear stress profile in asymmetric compound channel

WRF-24-12
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

14.05

Test section (m)

14

13.95
0 0.05 0.1 0.15 0.2
Channel width (m)

Figure 7(b): Contour plot of bed shear stress in asymmetric compound channel

From the figures, it is seen that significant difference occurs in shear stress values at the
junction region in both the channels. Also, significant difference in shear stresses at
horizontal interface between the two layers of both the channels are observed, see Figures
8(a-b) and 9(a-b).

Figure 8(a): Shear stress profile at interface in symmetric compound channel

6.05
Test section (m)

5.95
0.4 0.5 0.6 0.7 0.8
Channel width at Interface (m)

Figure 8(b): Contour plot of shear stress at interface in symmetric compound channel

6.2 Spanwise variation of model parameter C

WRF-24-13
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

Figures 10(a-b) show spanwise variation of depth averaged model parameter C both in
symmetric and asymmetric compound channel flows which was determined locally during
the model simulation while adjusting the level of eddy viscosity in flow field.

Figure 9(a): Shear stress profile at interface in asymmetric compound channel

14.05
Test section (m)

14

13.95
0 0.02 0.04 0.06 0.08 0.1
Channel width at Interface (m)

Figure 9(b): Contour plot of shear stress at interface in asymmetric compound channel

WRF-24-14
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

(a) Symmetric compound channel

(b) Asymmetric compound channel


Figure 10: Dynamic coefficient profile in compound channel

From the figures, it is seen that, at sidewalls of the channel, the value of C reduces to the
minimum as eddy viscosity is minimum close to the walls due to occurrence of anisotropy
turbulence (Frohlich and Rodi, 2002). Further, away from the wall in floodplain (Fp), its
value rises steeply in symmetric channel reaching a maximum (0.138), and then decreases
slowly before attaining its minimum value (0.02) near the wall of lower layer. In main
channel (Mc) of symmetric compound channel, the value of model parameter rises again and
regains to its second peak at mid of the channel. The parameter varies within the range of
0.02 to 0.138 throughout the channel during the model simulations (Figure 10a) after
attaining steady state flow condition. The model parameter in asymmetric compound channel
rises away from the wall, and attains its maximum value (0.145) in main channel. The value
of C reduces rapidly to a lower value (0.028) near the solid wall of lower layer. Its spanwise
variation has been found to be in the range of 0.028 to 0.145, see in Figure 10(b).

7. CONCLUSIONS

A two-layer 2-D model for compound open channel flows is used to simulate turbulent flows
in a wide symmetric and a narrow asymmetric compound open-channel. Following are the
key conclusions drawn from the above studies:
1. The developed model has been able to simulate well depth average primary velocity in
compound open channels. The model is able to reproduce one inflection point in case
of shallow flows in wide symmetric compound channel while two inflection points in
case of deep flows in narrow asymmetric compound channel.
2. The significant difference in layered average streamwise velocities between lower and
upper layers, i.e., of the order of 10-12% of bulk mean velocity (Ub), has been noticed
near the junction of the channel.
3. The maximum secondary currents are generated near the junction of the main channel
and the floodplain; the direction of the currents indicates the possible trend of mass
and momentum transfer from lower layer to upper layer. Horizontal contour plots of

WRF-24-15
National Conference on Water Resources & Flood Management with special reference to Flood Modelling
October 14-15, 2016 SVNIT Surat

secondary currents, bed and interfacial shear stresses indicate the formation of vortices
near the junction of the main channel and the floodplains.
4. The variation of depth averaged dynamic model parameter, C, in the compound
channels has been found in the range of 0.02 to 0.145.

REFERENCES

Cater, J.E., and Williams, J.J.R. (2008) “Large eddy simulation of a long asymmetric compound open channel”,
J. Hyd. Res., 46(4): 445-453.

Chau, K.W., Jin, H.S. (1995) “Numerical solution of two-layer, two-dimensional tidal flow in a boundary fitted
orthogonal curvilinear co-ordinate system” Int. J. Num. Meths. Fluids 21(11): 1087-1107.

Fraselle, Q., Arnould, T., Lissoir, X., Bousmar, D. and Zech, Y. (2008) “Investigating diffusion and dispersion
in compound channels using low-cost tracer”, Proceeding river flow conference on fluvial hydraulics,
Cesme, Izmir, Turkey, 1: 529-538.

Frohlich, J. and Rodi, W. (2002) “Introduction to large eddy simulation of turbulent flows” Closure strategies
for turbulent and transitional flows, Ed. B.E. Launder and N.D. Sandham, Cambridge University
Press, Chapter 2: 267-298.

Germano, M., Piomelli, U., Moin, P., and Cabot, W.H. (1991) “A dynamic subgrid-scale eddy viscosity model”,
Physics of Fluids, A3(7): 1760-1765.

Joung, Y. and Choi, S. U. (2008) “Investigation of twin vortices near the interface in turbulent compound open-
channel flows using DNS data” J. Hydraulic Engineering, ASCE, 134(12): 1744-1756.

Naot, D., Nezu, I., and Nakagawa, H. (1993) “Hydrodynamic behaviour of compound rectangular open
channels”, J. Hydraulic Engg., ASCE, 119(3): 390-408.

Nezu, I., Onitsuka, K., Sagara, Y. and Iketani, K. (1999). Secondary currents and bed shear stress in compound
open-channel flows with shallow flood plain. Proc. 28th IAHR Congr., Graz, Austria

Premnath, K.N., Pattison, M.J. and Banerjee, S. (2009) “Dynamic Subgrid Scale Modeling of Turbulent Flows
using Lattice-Boltzmann Method”, Elsevier, January 2009.

Shiono, K. and Feng, T. (2003). “Turbulence Measurements of Dye Concentration and Effects of Secondary
Flow on Distribution in Open Channel Flows”, J. Hydraulic Engineering, 129(5): 373-384.

Singh, V., Bhallamudi, S.M. (1997) “Hydrodynamic modelling of basin irrigation”, J. Irrig. Drainage Eng.
123(6): 407-414.

Sinha, J., Das, Samir K., Patel, P.L. and Samtani, B.K. (2014) “Development of Two-layered model for
compound open-channel flow”, ISH Journal of Hydraulic Engineering, 20(3): 250-262.

Stocchino, A., Besio, G., Angiolani, S., and Brocchini, M. (2011). “Lagrangian mixing in straight compound
channels”, J. Fluid Mech., 675: 168-198.

Tominaga, A. and Nezu, I. (1991) ‘‘Turbulent structure in compound channel flows.’’ J. Hydraulic. Eng.,
117(1): 21–41.

Xie, Z., Lin, B. and Falconer, R.A. (2013) “Large-eddy simulation of the turbulent structure in compound open-
channel flows” Advances in Water Resources, 53: 66-75.

WRF-24-16

You might also like