You are on page 1of 163

TWO LAYER NUMERICAL MODEL FOR

HYDRODYNAMIC AND SOLUTE TRANSPORT IN


COMPOUND OPEN CHANNELS
Thesis
Submitted in partial fulfillment of requirement for the degree of
DOCTOR OF PHILOSOPHY
in
CIVIL ENGINEERING

by
JIWESHWAR SINHA
(Reg. No.: D 09 CE 111)

Under Supervision of
Dr. P.L. Patel, Dr. B.K. Samtani & Dr. S.K. Das

CIVIL ENGINEERING DEPARTMENT


SARDAR VALLABHBHAI NATIONAL INSTITUTE OF TECHNOLOGY
SURAT 395 007, GUJARAT, INDIA
(December 2016)

II

III

Dedicated to Lord Shri Sainath,


Kul-devta
&
My father
Late H aribans N. Sinha (Siree)

IV

Acknowledgement
I would like to take this opportunity to first and foremost thank God
for being my strength and guide during this research work. Without Him,
I would not have had the wisdom and the physical ability to do so.
I am deeply grateful to all my research supervisors Dr. P. L. Patel, Dr.
S.K. Das and Dr. B.K. Samtani for their constant support and guidance.
Dr. Patels continuous encouragement helped me to complete this research
work in time. Dr. Das was always available for technical discussions
throughout this work.
I am thankful to Dr. D.C. Jinwala, Dr. J. Banerjee, Dr. S M Yadav
and Dr. V.L. Manekar for evaluating my work as examiners in Research
Progress & Pre-synopsis Seminar Committee and for their guidance to
improve the quality of thesis.
I would like to express my sincere thanks to Dr. M.K. Sinha,
Director, CWPRS & Shri M.D. Kudale, Additional Director, CWPRS for
providing the departmental infrastructure facilities during this research
work. I am also thankful to Smt. V.M. Bendre, Ex. Director, CWPRS for
permitting me to pursue higher studies.
My special thanks are due to Dr. Shailesh B. Patel and Shri Viraj
Loliyana for their help during the research work and writing of this thesis.
I would also like to thank Dr. Prafulkumar V. Timbadiya, Dr. U.C.
Roman, Dr. P. Chandra, Dr. A.K. Bagwan, S/Smt. Seema Shiyekar,

Shivani Sahu, Komal Vighe, Vaibhavi Roy, Miss Kavya, S/Shri V.B.
Sharma, S.N. Jha, Priyank Sharma, Nitin Bharadiya, Lalit Pal, Anil,
Akshay, Hasmukh Patel for extending all kinds of support during the
research work.
I would also like to express my gratitude towards my mother Smt.
Bachchi Sinha, In-laws Smt. Premlata Sinha & Shri Ayodhya P. Sinha,
Elder brother Shri Jitendra Sinha, and My wife Sudha, son Navneet &
daughter Smriti for providing me inspiration and enthusiasm at all level
to work hard. Sincere thanks to all my family members without their
support it is not possible to pursue higher studies.
Last but not least I would like to thank all known and unknown
persons for their kind co-operation, help and motivation during the
research.
Jiweshwar Sinha

VI

Abstract
In nature, rivers are made up of a centrally located main channel and floodplains on one
or both sides of it. The river systems are being overloaded with pollutants due to rapid
land use changes in their respective catchments. A thorough understanding of the flow
interaction between the main channel and floodplains is essentially required for river
management and related works. Also, the prediction of hydraulic characteristics in
natural river systems is required in assessment of their pollutant transport capacity along
the flow. The natural river can be represented as a two layered system in the form of
upper layer and lower layer, and a horizontal interface in between them where convective
exchange between the layers is allowed. A two-layer two-dimensional (2-D)
mathematical model is developed for hydrodynamics and solute transport in compound
open-channel using Cartesian equations for non-uniform grid system. Turbulent exchange
across the interface is treated empirically. Turbulent terms are dealt with turbulence
closure schemes under large eddy simulation (LES) approach, such as, standard
Smagorinsky (SS) scheme and dynamic subgrid scale (DSGS) scheme. In order to obtain
high accuracy and oscillation-free solutions, two step predictor-corrector explicit scheme,
second-order accurate in space and time, and slope limiter function minmod are
employed for solving the governing equations.

Performance of the developed hydrodynamic and solute transport models have been
assessed for turbulent flows in wide symmetric and narrow asymmetric compound
channels. A normal asymmetric compound channel conveying shallow and deep flows
was also considered for the aforesaid purpose. The comparison of primary velocity and
bed shear stress was performed based on past experimental data from literature using
statistical error measuring index, root-mean-square error (RMSE). A number of other
flow characteristics, such as secondary velocity, shear stress at horizontal interface,
subgrid scale (SGS) turbulent shear stresses in upper and lower layers, have been
simulated in aforesaid models. The simulated result on flow parameters like interfacial

VII

shear stress and vertical flow velocity gives insight about the overall momentum
exchange between the main channel and the floodplain of straight compound openchannel. The analysis of model results revealed the existence of horizontal vortices near
the junction of the main channel and the floodplain where momentum exchange takes
place. The dominance of shear stresses at the junction of the main channel and floodplain
is existed irrespective of channel cross-sections. The model reproduces the existence of
one inflection point in case of shallow flows and two inflection points in case of deep
flows. The significant difference in layered average primary velocities (in the range of 14% to +12% of Ub) between lower and upper layers is present near the junction of the
channel. Horizontal isolines plots of secondary currents (ranging between 0.02% of Ub to
1.1% of Ub); bed (varying between1.15

to 1.35 ) and interfacial (varying between

0.08 N/m2 to 0.4 N/m2) shear stresses indicated the formation of vortices near the
junction of the main channel and the floodplains. Variation of dynamic SGS coefficient,
C, (in the range of 0.02 to 0.145) for turbulent flows in wide symmetric and narrow
asymmetric compound channels, poses slight effects in flow characteristics in the
aforesaid channels.

Performance of solute transport model is also assessed using past experimental data on
symmetric and asymmetric compound open-channels. Transport processes are
investigated through injection of inert pollutant at various locations in compound openchannels. The model performs better for pollutant injected at other than shear layer
locations. The significant differences in distribution of solute concentration between the
computed and experimental results were observed due to under estimation of secondary
currents near the junction of main channel and floodplain. The performance of transport
model is better at downstream locations as compared to upstream locations due to better
mixing of pollutants. Further, the presence of vertical velocity component at horizontal
interface causes higher rate of distribution of concentrations vertically as compared to in
horizontal directions.

VIII

Contents

Acknowledgement

Abstract

VII

Contents

IX

List of Figures

XII

List of Tables

XVI

List of Notations and Abbreviations

XVII

Chapter

Name

Introduction

Page
No.
1-6

1.1

General

1.2

Brief literature review

1.3

Motivation for present investigation

1.4

Research objectives

1.5

Thesis organisation

Literature review

7-21

2
2.1

General

2.2

Previous studies in compound channel flows

Hydrodynamics of compound channel flows

2.2.1.1 Experimental investigations

2.2.1

2.2.2
2.3

2.2.1.2 Numerical investigations

14

Solute transport in compound open channels

18

Concluding remarks

21

IX

Theoretical consideration and modelling

22-52

3.1

General

22

3.2

Governing equations of two-layered flows

22

3.2.1

Hydrodynamic equations

23

3.2.2

Governing equations for solute transport

27

Method of solution

29

3.3.1

Finite volume method (FVM)

30

3.3.2

Discretisation of governing equations

33

3.3.3

Expansion of turbulent stress terms

38

Turbulence closure scheme

39

3.4.1

Large eddy simulation (LES)

40

3.4.2

Subgrid scale (SGS) modelling

40

3.5

Initial and boundary conditions

44

3.6

Predictor-corrector scheme

49

3.7

Stability conditions

50

3.8

Concluding remarks

52

3.3

3.4

Simulation of hydrodynamic characteristics:


Results and discussions

53-117

4.1

General

53

4.2

Experimental data

53

4.2.1

Symmetric compound channel

53

4.2.2

Asymmetric compound channel

54

Hydrodynamic model simulation results

56

Standard Smagorinsky model (SS Model)

56

4.3.1.1 Symmetric compound channel

56

4.3.1.2 Asymmetric compound channel

69

Dynamic subgrid scale model (DSGS model)

98

4.3.2.1 Symmetric compound channel

98

4.3
4.3.1

4.3.2

4.3.2.2 Asymmetric compound channel

107

4.4

Concluding remarks

116

5.1
5.2

Simulation of solute transport: Results and


discussions
General
Experimental data

118-131
118
118

5.2.1

Symmetric compound channel

118

5.2.2

Asymmetric compound channel

119

Transport model simulation results

120

5.3.1

Symmetric compound channel

121

5.3.2

Asymmetric compound channel

127

Concluding Remarks

131

5.3

5.4

Conclusions

132-135

6.1

General

132

6.2

Hydrodynamic model

132

6.3

Solute transport model

134

6.4

Future scope of works

134

References

136

Research publications

142

XI

List of Figures
Figure
No.

Title

Page
No.

3.1

Schematic of two layered compound open channel flows

23

3.2

Representation of physical domain in model

30

3.3

Non-orthogonal cell grid and its neighbours

31

3.4

Control volume of grid cell (i,j)

32

3.5

Left and right state at cell face [(i+1/2,j) & (i,j+1/2)]

35

3.6

Computational cells showing auxiliary control volume (dashed


lines) used for calculations of derivative terms of diffusive flux

36

3.7

Model grid along with ghost cells

46

3.8

Diagram of partial slip boundary condition

47

3.9

Flowchart of developed SS and DSGS models

51

4.1

Schematic of symmetric compound channel (Fraselle et al., 2008)

54

4.2

Schematic of asymmetric compound channel (Shiono and Feng,


2003)

55

4.3

Schematic of asymmetric compound channel (Tominaga and Nezu,


1991)

56

4.4(a)

Variation of primary flow velocity of symmetric compound


channel (SS model) with C (Alpha=0.5)

59

4.4(b)

Variation of primary flow velocity of symmetric compound


channel (SS model) with Alpha (C=0.1)

59

4.5

Computational grid cells and boundaries in symmetric compound


channel for SS model

60

4.6

Comparison of primary flow velocity in symmetric compound


channel (SS model)

61

4.7

Computed flow field in symmetric compound channel (SS model)

62

4.8

Percentage difference between primary flow velocities of two


layers with respect to mean bulk velocity of wide symmetric
compound channel (SS model)

62

XII

4.9

Spanwise bed shear stress in wide symmetric compound channel


(SS model)

64

4.10

Secondary flow velocities at horizontal interface in wide


symmetric compound channel (SS model)

65

4.11

Variation of shear stress at horizontal interface in wide symmetric


channel (SS model)

67

4.12

Variation of SGS turbulent shear stress in upper layer of wide


symmetric channel (SS model)

68

4.13

Variation of SGS turbulent shear stress in lower layer of wide


symmetric compound channel (SS model)

69

4.14

Comparison of primary flow velocity in narrow asymmetric


compound channel (SS model)

72

4.15

Percentage difference between primary flow velocities of two


layers with respect to mean bulk velocity of narrow asymmetric
compound channel (SS model)

73

4.16

Spanwise bed shear stress in narrow asymmetric compound


channel (SS model)

74

4.17

Secondary flow velocities at horizontal interface of narrow


asymmetric compound channel (SS model)

75-76

4.18

Variation of shear stress at horizontal interface of narrow


asymmetric compound channel (SS model)

4.19

Variation of SGS turbulent shear stress in upper layer of narrow


asymmetric compound channel (SS model)

78-79

4.20

Variation of SGS turbulent shear stress in lower layer of narrow


asymmetric compound channel (SS model)

80

4.21(a)

Variation of primary flow velocity of normal asymmetric


compound channel (Exp.S-2: SS model) with C (Alpha=0.5)

82

4.21(b)

Variation of primary flow velocity of normal asymmetric


compound channel (Exp.S-2: SS model) with Alpha, (C=0.1)

82

4.22

Comparison of primary flow velocity in normal asymmetric


compound channel (SS model)

84

4.23

Percentage difference between primary flow velocities of two


layers with respect to bulk mean velocity in normal asymmetric
compound channel (SS model)

85

4.24

Comparison of bed shear stresses in normal asymmetric compound


channel (SS model)

XIII

77

86-87

4.25

Secondary velocity at horizontal interface in normal asymmetric


compound channel (SS model)

88-91

4.26

Variation of shear stresses at horizontal interface in normal


asymmetric compound channel (SS model)

92-94

4.27

Variation of SGS turbulent shear stress in upper layer of normal


asymmetric compound channel (SS model)

94-95

4.28

Variation of SGS turbulent shear stress in lower layer of normal


asymmetric compound channel (SS model)

96-97

4.29

Variation of SGS coefficient across wide symmetric compound


channel (DSGS model)

99

4.30

Comparison of primary flow velocity in wide symmetric


compound channel (DSGS model)

100

4.31

Percentage difference between primary flow velocities of two


layers with respect to bulk mean velocity of wide symmetric
compound channel (DSGS model)

100

4.32

Variation of bed shear stress in wide symmetric compound channel


(DSGS model)

4.33

Secondary flow velocities at horizontal interface of wide


symmetric compound channel (DSGS model)

103

4.34

Variation of shear stress at horizontal interface of wide symmetric


compound channel (DSGS model)

104

4.35

Variation of SGS turbulent shear stress in upper layer of wide


symmetric compound channel (DSGS model)

105

4.36

Variation of SGS turbulent shear stress in lower layer of wide


symmetric compound channel (DSGS model)

106

4.37

SGS coefficient profile in narrow asymmetric compound channel


(DSGS model)

107

4.38

Comparison of primary flow velocity in narrow asymmetric


compound channel (DSGS model)

108

Percentage difference between primary flow velocities of two


layers with respect to bulk mean velocity of narrow asymmetric
compound channel (DSGS model)

108

4.39

101-102

4.40

Variation of bed shear stress in narrow asymmetric compound


channel (DSGS model)

109-110

4.41

Secondary flow velocities at horizontal interface of narrow


asymmetric compound channel (DSGS model)

111-112

XIV

4.42

Variation of shear stress at horizontal interface of narrow


asymmetric compound channel (DSGS model)

112-113

4.43

Variation of SGS turbulent shear stress in upper layer of narrow


asymmetric compound channel (DSGS model)

114

4.44

Variation of SGS turbulent shear stress in lower layer of narrow


asymmetric compound channel (DSGS model)

115

5.1

Comparison of depth-averaged solute concentration at different


locations downstream of inlet in wide symmetric compound
channel for Exp.I-1

122-123

5.2

Isolines plot of depth-averaged solute concentration (g/l) in layers

124

of wide symmetric compound channel for Exp.I-1


5.3

Comparison of depth-averaged solute concentration at different


locations downstream of inlet in wide symmetric compound
channel for Exp.I-4

5.4

Isolines plot of depth-averaged solute concentration (g/l) in layers

125-126

126

of wide symmetric compound channel for Exp.I-4


5.5(a)

Solute concentration profile in narrow asymmetric compound


channel for Exp.C-1

127

5.5(b)

Isolines plot of solute concentration (ppb) in layers of narrow


asymmetric compound channel for Exp.C-1

128

5.6(a)

Solute concentration profile in narrow asymmetric compound


channel for Exp.C-2

129

5.6(b)

Isolines plot of solute concentration (ppb) in layers of narrow


asymmetric compound channel for Exp.C-2

129

5.7(a)

Solute concentration profile in narrow asymmetric compound


channel for Exp.C-3

130

5.7(b)

Isolines plot of solute concentration (ppb) in layers of narrow


asymmetric compound channel for Exp.C-3

130

XV

List of Tables

Table
No.

Title

Page
No.

4.1

Compound channel geometries and hydraulic characteristics of


experimental data

55

4.2

Grid cells used in wide symmetric compound channel for grid


dependence test and verification of SS model

57

4.3

Combination of values of parameters C and

58

4.4

Grid cells used in narrow asymmetric compound channel for grid


dependence test and verification of SS model

70

4.5

Grid dependence test for normal asymmetric compound channel


under SS model

81

5.1

Solute injection and measurement locations in simulation


experiments in wide symmetric and narrow asymmetric compound
channels

XVI

120

Notations and Abbreviations


The following symbols are used in present study:

Latin Letters
A
A1
A

area of grid cell; [m2]


constant, in Eq.(3.61); [=5.3]
area of auxiliary grid cell, in Eq.(3.33); [m2]

bl
b
bu

lower layer width; [m]


total width of channel; [m]
upper layer width; [m]

C
cf
cl
cu

SGS model coefficient, in Eq.(3.49);


coefficient of bed resistance, in Eq.(3.4 & 3.6);
layer averaged concentration in lower layer; [kg/m3]
layer averaged concentration in upper layer; [kg/m3]

dr
ds
dv

length of face of grid cell, in Eq.(3.20); [m]


surface area of face of finite volume cell, in Eq.(3.18); [m2]
volume of finite volume cell, in Eq.(3.18); [m3]

Dx
Dy

diffusion coefficient in x-direction, in Eq.(3.11); [m2/s]


diffusion coefficient in y-direction, in Eq.(3.11); [m2/s]

E
Eh
Ec

x-component of total flux at grid cell, in Eq.(3.17)


x-component of hydrodynamic part of total flux at grid cell, in Eq.(3.10);
x-component of transport part of total flux at grid cell, in Eq.(3.16);

F
Fc
Fv

total flux at face of grid cell, in Eq.(3.18)


convective flux, in Eq.(3.32)
viscous flux, in Eq.(3.32)

g
G
Gh
Gc

acceleration due to gravity; [m/s2]


y-component of total flux at grid cell, in Eq.(3.17)
y-component of hydrodynamic part of total flux at grid cell, in Eq.(3.10)
y-component of transport part of total flux at grid cell, in Eq.(3.16)

hl

layer averaged water depth in lower layer; [m]

XVII

hu
h (=hu+hl)

layer averaged water depth in upper layer; [m]


total water depth; [m]

k
Ke

von-Karman constant, in Eq(3.8); [=0.41]


rate of exchange of concentration from one layer to other, in Eq.(3.14);
[per sec]

lo
Lij

Prandtl mixing length, in Eq.(3.8); [m]


resolved turbulent stresses, in Eq.(3.55)

Mij

substituted parameter for different STS and SGS stresses at test filter level,
in Eq.(3.56a);

n
nl
nu
nx
ny
Ni
Nj

Mannings roughness coefficient, in Eq.(3.6); [s/m1/3]


unit normal vector at face of grid cell of lower layer
unit normal vector at face of grid cell of upper layer
x-component of unit normal vector at face of grid cell
y-component of unit normal vector at face of grid cell
total number of grid point in ith direction, in Eq.(3.31)
total number of grid point in jth direction, in Eq.(3.31)

sl
so
su
S
Sc
Sh
Sij
Sx

quantity of inert solute in lower layer, in Eq.(3.15); [kg/m2/s)]


longitudinal slope of channel;
quantity of inert solute in upper layer, in Eq.(3.14); [kg/m2/s)]
source vector of combined governing equation, in Eq.(3.17);
Source vector of solute transport equation, in Eq.(3.16);
Source vector for hydrodynamic equation, in Eq.(3.10);
strain rate
x-component of face vector of auxiliary grid cell, in Eq.(3.34)

ul

layer averaged primary flow velocity along x-direction in lower layer;


[m/s]
layer averaged flow velocity component in x-direction, in Eq.(3.4); [m/s]
flow velocity component at horizontal interface along x-direction, in
Eq.(3.9); [m/s]
layer averaged primary flow velocity along x-direction in upper layer;
[m/s]
local friction flow velocity, in Eq.(3.3); [m/s]
fluctuation in flow velocity, in Eq.(3.38); [m/s]
State vector of combined governing equation, in Eq.(3.17);
mean bulk velocity; [m/s]
State vector of solute transport equation, in Eq.(3.16);
State vector for hydrodynamic equation, in Eq.(3.10);
global friction flow velocity; [m/s]

um
uo
uu
u*
u'
U
Ub
Uc
Uh
U*

XVIII

vl
vm
vo
vu

layer averaged spanwise flow velocity along y-direction in lower layer;


[m/s]
layer averaged flow velocity component in y-direction, in Eq.(3.5); [m/s]
flow velocity component at horizontal interface along y-direction, in
Eq.(3.9); [m/s]
layer averaged spanwise flow velocity along y-direction in upper layer;
[m/s]

wo

Vertical flow velocity at horizontal interface, in Eqs.(3.1 & 3.2); [m/s]

Zb

vertical position of channel bed with respect to datum; [m]

Greek Letters

density of water; [kg/m3]


flow variable at center of grid cell, in Eq.(3.27)
flow variable at center of auxiliary grid cell, in Eq.(3.33)

positive coefficient in flux approximation, in Eq.(3.28)


wave celerity, in Eq.(3.31); [m/s]
kinematic viscosity of water, in Eq.(3.61); [m2/s)
stress tensor, in Eq.(3.37)
shear stresses at horizontal interface in x-direction, in Eq.(3.1b); [N/m2]
shear stresses at horizontal interface in y-direction, in Eq.(3.1c); [N/m2]
bed shear stress in x-direction, in Eq.(3.2b); [N/m2]
bed shear stress in y-direction, in Eq.(3.2c); [N/m2]
shear stresses at free surface in x-direction, in Eq.(3.1b); [N/m2]
shear stresses at free surface in y-direction, in Eq.(3.1c); [N/m2]
SGS turbulent normal stress at x-face in x-direction in lower layer, in
Eq.(3.2b); [N/m2]
SGS turbulent normal stress at x-face in x-direction in upper layer, in
Eq.(3.1b); [N/m2]
SGS turbulent shear stress at x-face in y-direction in lower layer, in
Eq.(3.2c); [N/m2]
SGS turbulent shear stress at x-face in y-direction in upper layer, in
Eq.(3.1c); [N/m2]

XIX

SGS turbulent shear stress at y-face in x-direction in lower layer, in


Eq.(3.2b); [N/m2]
SGS turbulent normal stress at y-face in y-direction in lower layer, in
Eq.(3.2c); [N/m2]
SGS turbulent shear stress at y-face in x-direction in upper layer, in
Eq.(3.1b); [N/m2]
SGS turbulent normal stress at y-face in y-direction in upper layer, in
Eq.(3.1c); [N/m2]
shear stress at wall, in Eq.(3.62); (N/m2)
SGS eddy diffusivity in x-direction at lower layer, in Eq.(3.15); [m2/s]
SGS eddy diffusivity in x-direction at upper layer, in Eq.(3.14); [m2/s]
SGS eddy diffusivity in y-direction at lower layer, in Eq.(3.15); [m2/s]
SGS eddy diffusivity in y-direction at upper layer, in Eq.(3.14); [m2/s]

mixing layer thickness at interface, in Eqs.(3.7 & 3.8); [m]

area of finite volume cell, in Eq.(3.20); [m2]

Superscripts

p
*
p+1

auxiliary grid variables


predictor value of variables at known time, t
corrector value of variables at known time, t
value of variables at unknown time, t+dt

Subscripts
b
c
h
i
ij
k
l
l
L
m
o
p
r

at bed
convective grid cell
hydrodynamic part
in ith direction (i=x, y)
ith face and jth direction
kth layer (k=u, l)
lower layer
laminar flow
left side of face of grid cell
mean value
at horizontal interface
grid cell point
face of grid cell

XX

R
s
u
v

right side of face of grid cell


at free surface
upper layer
viscous grid cell

Abbreviations
AGC
CFL
DNS
DSGS
ENO
Exp.
FLDA
FVM
LDA
LES
MUSCL
RANS
RMSE
SGS
SS
STS
TVD

Auxiliary grid cell


Courant-Friedrichs-Lewy
Direct numerical simulation
Dynamic subgrid scale
Essentially non-oscillating
Experiment
Fiber-optic laser Doppler anemometer
Finite volume method
Laser Doppler anemometer
Large eddy simulation
Monotone upwind scheme for conservation laws
Reynolds averaged Navier-Stokes
Root mean square error
Subgrid scale
Standard Smagorinsky
Subtest scale
Total variation diminishing

XXI

Chapter 1

Introduction

1.1 GENERAL

The natural rivers are, invariably, compound in cross-sections, consisting of deep main channel
flanked by floodplains on either or both sides to accommodate high flows during the floods. The
rivers have become more prone to pollutant and sediment overloading due to extensive
urbanisation across the globe. The characterisation of hydraulic parameters is of immense
importance in developing transport models of organic and inorganic pollutants for selection of
water intake locations, and prediction of erosion and deposition of sediments in natural channels.
Better understanding and complete control over transport phenomena are the essentially required
for many industries such as water sector, oil and chemical sectors and thermal sector in the
interest of environmental and economic benefits.

The flow velocity in the floodplains is lower than main channel due to shallow water depth. As a
result of the velocity gradient, shearing occurs at the interface between the main channel and the
floodplains, leading to the flow patterns most of which are characterized by large-scale vertical
structures with vertical axes (macro-vortices). Mixing and transport processes of solutes in
compound channels are mainly controlled and regulated by the exchange of mass and
momentum between the main channel and the lateral floodplains. Coherent structures are
commonly recognized as major agents of mass transport, owing to their ability to trap mass and
inhibit exchanges among different flow regions.

Prediction of hydraulic characteristics and pollutant transport in open-channel is undertaken


using series of three dimensional (3-D) and two dimensional (2-D) models. The 3-D models,
1

though give the precise results, may have their limited applications while applying the same for
long reach of natural rivers due to their extensive computational requirements. The 2-D models,
under such circumstances, may be very useful for monitoring the flow and transport
characteristics of natural channels with large reaches. Brief description of recent work
undertaken in the foregoing theme is included in following paragraphs.

1.2 BRIEF LITERATURE REVIEW

Numerous works on straight compound channel flow have been reported in recent past using
analytical methods (Tang and Knight, 2008; Yang et al., 2013), laboratory experiments
(Rajaratnam and Ahmadi 1981; Knight and Hamed 1984; Tominaga and Nezu 1991; Nugroho
and Ikeda, 2007; Fernandes et al., 2014) and numerical simulations (Naot et al., 1993; Thomas
and Williams, 1995; Sofialidis and Prinos, 1998; Nugroho and Ikeda, 2007; Joung and Choi,
2008; Cater and Williams, 2008; Kara et al., 2012; Xie et al., 2013) in investigation of flow
characteristics and mechanism of lateral momentum transfer from main channel to floodplains.

With the advent of high-speed computers, the study on two and three-dimensional flows in open
channels had experienced surge of interest in recent years. A number of numerical studies on
compound open-channel hydraulics, mainly based on the Reynolds-averaged Navier-Stokes
(RANS) equations, have been reported in the literature. Naot et al. (1993), Pezzinga (1994), and
Sofialidis and Prinos (1998) solved three-dimensional form of equations and showed that their
models are capable of predicting reasonably well the time-averaged primary flow including bed
shear stress distributions for different compound open-channel geometries. Pezzinga (1994) and,
Sofialidis and Prinos (1998) were able to reproduce the primary and secondary flows quite well
using nonlinear turbulence closure schemes. However, the bed shear stress predictions were less
accurate (Pezzinga, 1994), due to implementation of wall functions. The large eddy simulation
(LES), an alternative to RANS modelling, resolves the large-scale turbulence, and hence, allows
an accurate computation of flow characteristics. Thomas and Williams (1995) were the first to
employ LES in three-dimensional flow to study the flow and turbulence characteristics in a
compound open-channel. Recently, Cater and Williams (2008) and Kara et al. (2012) also
2

successfully applied the LES using 3-D equations in the simulation of compound open-channel
flows. Practical applicability of commercially available 3-D hydrodynamic models in practical
engineering problems is due to their extensive computational efforts and generic utilisation.
Owing to such difficulties, simplified two-layer models, in which each layer can be considered
two-dimensional, vertically homogeneous, and depth-averaged with a horizontal interface, could
be very handy in resolving the practical issues of compound channels.

Chau and Jin (1995) developed a two-layered, two-dimensional model based on finite difference
approach using CFD-based numerically generated boundary fitted orthogonal coordinate system,
and used to simulate density stratified unsteady flows in a natural water-body with complex
topography. Later on, Jin et al. (1998) applied the model to simulate turbulent flows in a
sinusoidal meandering compound open channel. In both, above-mentioned numerical studies
(Chau and Jin, 1995; Jin et al., 1998), grid block technique was used wherein the boundary
fitted orthogonal curvilinear grids were generated in the physical plane, and the same was
transformed into rectangular grids in the computational plane resulting in complicated
mathematical formulations with additional transformation relations. Prandtls mixing length
approach was used for modelling turbulent shear stress terms.

The 2-D depth averaged numerical models have also been in use in past due to their less
computational efforts in prediction of inert pollutant transport with relevant turbulent closure
schemes in compound open channel flows. Djordjevic (1993) presented a two dimensional
depth-averaged numerical model using the operator-splitting approach for unit Schmidt number
for solute transport in both streamwise and lateral directions in steady open channel flows. Wood
and Liang (1989) developed a 2-D semi-analytical model using eigen-function expansion
technique for prediction of tracer concentration in compound open channel flows. Chatila and
Townsend (1998) developed a 2-D finite-difference based model for simulating inert pollutant
transport in compound open-channels using constant eddy-viscosity model, and accounted for
lateral momentum transfer (LMT) feature at horizontal interface of main channel and
floodplains. Fraselle et al. (2008) carried out experimental studies in a symmetric compound
laboratory flume for diffusion and dispersion of tracer. Fraselle et al. (2010) compared their

results with commercial CFD software TELEMAC-2D and showed the role of turbulent eddies
in solute transport.

Existing numerical models have their own advantages and disadvantages in prediction of
secondary flows and solute concentration in compound open-channels. Most of the reported
studies have focused their attention mainly on vertical or inclined interface at the junction of
main channel and floodplains whereas a little attention has been paid to the flow characteristics
along the horizontal interface. Also, the past studies did not analyze the influence of flow depth
ratio and cross-sectional channel characteristics (wide or narrow) on transverse mixing of
pollutants.

Very few publications deal with the transport of pollutants in rivers with compound sections. It
can be argued that a 3-D model is required for more precise simulation of the flow field.
However, the effort and cost in developing a 3-D model would outweigh the improvement
gained over a 2-D model simulation. Thus, a 2-D model is generally considered to be a
reasonable compromise between simplified 1-D simulation and difficult 3-D models.

1.3 MOTIVATION FOR PRESENT INVESTIGATION

Invariably, the river systems are compound in cross-sections, and being overloaded with
pollutants due to rapid land use changes in their respective catchments. The prediction of
hydraulic characteristics in natural river systems is required in assessment of their pollutant
transport capacity along the flow. Due to rapid advancement in computational fluid dynamics
(CFD) and its popularity vis--vis physical modelling, the writer is prompted to take up the
problem of hydrodynamics and solute transport in compound channel using two layer 2-D
models with LES closure scheme.

1.4 RESEARCH OBJECTIVES

Keeping view the research gap and suitability of two layer 2-D model in prediction of hydraulic
and transport characteristics in compound open-channel, present study has been planned with
following research objectives:
a) Development of two layer 2-D hydrodynamic numerical model using LES approach for
straight compound open-channels.
b) Validation of developed numerical model using past experimental data on symmetric and
asymmetric compound channels.
c) Development of two layer 2-D solute transport model using LES approach for straight
compound open-channel.
d) Validation of solute transport model using past experimental data on symmetric and
asymmetric compound open-channels.

1.5 THESIS ORGANISATION

The brief description of PhD Thesis, comprising six chapters, are described in following
paragraphs:

Chapter 1 provides the background information and brief literature review on hydrodynamics
and transport phenomena involved in compound channel flow. The motivation, research
objectives and thesis structure are described.

In Chapter 2, detailed literature review on hydrodynamics and transport phenomena in


compound open-channel flows, involving experimental, analytical and numerical studies, are
presented.

In Chapter 3, theoretical aspects dealing with two-layered two dimensional mathematical models
and their numerical solutions for hydrodynamics and transport of inert solute are presented. Two

types of turbulence schemes, viz., standard Smagorinsky (SS) and dynamic subgrid scale
(DSGS) closure schemes, under large eddy simulation (LES) technique, are discussed.

In Chapter 4, numerical model results on hydrodynamics are presented in both symmetric and
asymmetric compound open-channel flows. The results obtained from both standard
Smargorinsky (SS) and dynamic subgrid scale (DSGS) model are discussed in the chapter. The
performance of present numerical model is explored with respect to laboratory data from
literature. Further, flow properties at horizontal interface of the channels are also explored in
detail.

In Chapter 5, numerical model results are presented on inert solute transport in both symmetric
and asymmetric compound open-channel flows. The results obtained from dynamic subgrid scale
(DSGS) model coupled with hydrodynamic equations are presented in this chapter. The
performance of developed numerical model is explored with respect to laboratory data from
literature.

In Chapter 6, the key conclusions derived from foregoing study on hydrodynamics and transport
numerical models are presented. Also, the directions for future work on selected theme are
included.

Chapter 2

Literature Review

2.1

GENERAL

In nature, rivers are made up of a centrally located main channel and one or two floodplains
extending laterally away from it. A thorough understanding of the flow interaction between the
main channel and the floodplains is essentially required for river management and related works.
Since 1960s, compound channels have been studied extensively through both laboratory and
numerical model studies. Previous studies related with hydrodynamic and solute transport
phenomena in compound open-channel flows are presented in this chapter.

2.2

PREVIOUS STUDIES IN COMPOUND CHANNEL FLOWS

In this section, different phenomena associated with hydrodynamic and solute transport model of
straight compound open channels are reviewed.

2.2.1 Hydrodynamics of Compound Channel Flows

Prior to early sixties, very little attention was paid on complex flow patterns which exist between
the main channel and the floodplains in compound channels, but, later on developments have led
to a clearer understanding of hydraulic mechanism involved at the level of model studies. Prandtl
(1925) proposed a logarithmic velocity profile for velocity distribution in pipe flow using mixing
length hypothesis. Subsequently, Von Karman (1930) proposed a velocity distribution model
7

using similarity hypothesis. A general form of velocity distribution law was developed by
Prandtl (1932) which is generally considered as the Prandtl-Von Karman velocity law. Chow
(1959) discussed open channel hydraulics problems and related structures in his classical
literature that describe fundamental principles in open-channels. Numerous researchers have
studied in detail about the flow phenomenon in compound open channels both, experimentally
and numerically, afterwards. Further, in pursuance of analysing the complex flow features in
straight compound channel, various types of laboratory channels were used under different
conditions. Similarly, in order to fully analyse the flow phenomena of straight compound open
channels, numerous numerical methods were proposed. In furtherance of these researches, the
laboratory produced results have been compared with simulated results of straight compound
open-channel flows. Based on these researches, it is established that the flow structure in straight
compound open channels is extremely complex in nature which induces significant mass and
momentum exchange, and formation of vortices. One can obtain one-dimensional (1-D) SaintVenant equations by integrating along the depth and width from the Navier-Stokes equations (NS); and, similarly, by integrating along the depth and assuming hydrostatic pressure distribution
in vertical scale, two-dimensional (2-D) shallow water equations (SWE) are obtained. Since the
basic flow equations describing the open channel flows are non-linear and coupled and, as a
result, it is difficult to solve them analytically. However, only in small number of cases, it is
possible to obtain an analytical solution of the Navier-Stokes equations under certain
assumptions. In order to solve or close turbulent shear stress terms, several model are considered
and these are (i) Boussinesq eddy viscosity approach (ii) Mixing length hypothesis (iii) Zero
equation models (iv) One equation models (v) Two equation models (vi) Multi-equation models
(vii) Sub-grid-scale (SGS) models (viii) Direct numerical simulation (DNS) and extension
thereof.

The hydrodynamic behaviour of compound channel flow is composed of the averaged wall
shear stress and the continuous mass exchange between the floodplain and the main channel.
The intensity of vortex formed near the junction of the floodplain and the main channel
corresponds to continuous mass exchange between them. In addition to above, a fully
developed turbulent flow in compound open-channels is specified by the momentum transfer
from the faster flow in the main channel to the slower flow in the floodplain. The shear layer is
8

developed at interface of the main channel and floodplain, and the secondary currents are
generated at the corners of channel cross-sections due to bed-generated turbulence and freesurface effects. The presence of such co-existing phenomena results into a very complex flow
field in straight compound open channels.

2.2.1.1 Experimental investigations

The hydrodynamics of two stage open channels of symmetric type, was first investigated in the
laboratory by Sellin (1964) wherein momentum transfer mechanism between the deep main
channel and the shallow floodplains along with the presence of vortices and their effect on
velocity and discharge were demonstrated using flow visualisation technique. He described
phenomenon of formation of vertical vortices due to transfer of fluid from the main channel to
the floodplains with high momentum. The studies were performed in a 4.5 m long symmetrical
compound open-channel. He was not able to estimate overbank flow which depends upon
empirical weighing of roughness number requiring sizable amount of experimental data. He also
opined that numerical expression for the effect of momentum transfer mechanism depends upon
the effect of variations of scale, cross-sectional shape, and floodplain and channel roughness. He
inferred that lateral momentum exchange from the main channel to the flood plains was
significantly increased which, ultimately, caused reduction in the main channel conveyance and
increase in flow resistance in the channel. In addition to the vertical vortices due to the
anisotropy of turbulence, generation of secondary currents was observed in the longitudinal
direction which played an important role in the momentum exchange, especially near the
junction.

Zheleznyakov (1971) carried out investigations into the problem of main channel and floodplain
interaction using a 5.2 m wide laboratory flume having parabolic shaped flood channels of 0.45
m and 0.6 m width on either side. The tests were performed for discharges ranging from 0.0069
cumecs to 0.0514 cumecs under bed slopes ranging from 1/1000 to 1/2000. He reported the
effect of turbulent flow mixing between the main channel and the floodplains, and inferred that a
significant reduction of flow occurs in the main channel. The formation of vortices with vertical
axes and turbulent eddies at the junction of main channel and floodplain were also observed by
9

him. Based on his experiments, it was also inferred that interaction of flows between the main
channel and the floodplain is inversely dependent on the depth of floodplains as the interaction
of flow between the main channel and the floodplain becomes weaker with increasing water
depth. He named such a momentum-exchange phenomenon as the kinematic effect. Similar
observations were also reported by Myers and Elsaway (1975), and Myers (1978).

Ghosh and Jena (1971) investigated shear stress distribution for rough and smooth sidewalls in
symmetric compound flume of length 8.5 m, main channel width 0.203 m and floodplains width
0.076 m. They obtained boundary shear distribution in the flume for different flow depths using
combined technique of Preston tube and Patel calibration (Patel, 1965). They observed that
boundary shear in the flume gets redistributed due to roughening of sidewalls and bed of the
channel. Such redistribution of boundary shear was also reported by Ghosh and Mehta (1974)
who conducted laboratory experiments with different compound channels.

Rajaratnam and Ahmadi (1970, 1981) carried out laboratory tests in a symmetrical compound
channel of 18.29 metres long with main channel width of 0.2032 metres flanked by two
floodplains, each 0.508 metres wide to manifest interaction mechanism in the channel. They
developed equations based on their experimental results to express velocity and shear stress
profiles in main channel and floodplain. They also showed that mixing zone is extended by a
distance of 6 times the bankfull depth across main channel and floodplain.

Price (1975) showed that turbulent shear occurring between the main channel and the floodplain
is dependent on difference of mean velocity between them. Based on his experiments carried out
in a laboratory flume, he opined that shear layers should be considered as vertical and be
regarded as rough surfaces.

Myers (1978) measured shear stress distributions around the periphery of a complex channel,
consisting of a deep section and one flood plain and the results were used to quantify the
momentum transfer due to interaction of flows between the main channel and floodplain.

10

In order to explore the interaction of flows and shear forces at different levels between the main
channel and the floodplains, Knight and Demetriou (1983) conducted experimental studies on
symmetrical rectangular compound open channel flume of length 15 m, at a constant bed slope
of 9.66 x 10-4, having 0.152 m wide and 0.076 m deep centrally located channel along with 0.229
m wide and 0.076 m high two floodplains on either sides. Flow discharge and velocity were
measured using a differential type Venturi meter, and a Novar stream flow miniature propeller
current meter, respectively. Shear stresses were measured using a Preston tube. Percentage of
flows carried by different sections and percentages of shear force distributions at different
interfaces of the channel were calculated for different relative water depths [(H-h)/H=0.5, 0.4,
0.33, 0.25, 0.2, 0.1] and relative widths [B/b=4, 3, 2, 1] of the main channel and the floodplain.
Here, water depth in main channel, H is 0.152 m; water depth in floodplain, h is 0.076 m, total
width of channel, B is 0.305 m, and width of main channel, b is 0.076 m. Among different
interfaces, they also calculated apparent shear forces and discharges in terms of percentage along
the horizontal interface. The horizontal interface was considered by dividing the whole channel
section into two parts; one part as the lower main channel and the other part as the upper main
channel and floodplains taken together. It was evident from the experiments that, at low relative
water depths and high relative channel widths, values of shear force were positive which
indicated retardation of flow in the lower main channel due to the flow in upper region. At high
relative water depths, the values of shear forces turned out to be negative indicating that the flow
in the lower main channel is accelerated due to flow of the upper region.

Knight and Hamed (1984) carried out experimental studies and presented boundary shear stress
distributions at different sections of a symmetric compound channel to quantify momentum
transfer mechanism in terms of apparent shear force acting on channel interface. He studied the
progressive roughness of floodplains in six steps to observe the influence of differential
roughness between floodplains and main channel on lateral momentum transfer process.
Empirical equations were also presented for shear force on floodplains in terms of percentage of
total shear force. Supplementary relationships were established for apparent shear forces on
vertical, inclined and horizontal interfaces within the cross section. The influence of momentum
transfer between sub-areas on vertical and lateral distribution of longitudinal velocity was also
assessed.
11

Formation of vertical vortices and helical secondary currents at the junction of main channel and
floodplains was also observed by Shiono and Knight (1989) in their experiments carried out in
the SERC Flood Channel Facility (FCF) at Hydraulics Research Station, Wallingford, UK
wherein they had undertaken secondary current measurements using Laser Doppler Anemometer
(LDA). With experimental observations, they inferred the importance of geometry of wide
symmetrical compound channel in influencing the structure of secondary currents in main
channel. Shiono and Knight (1989) further showed that a relatively smaller secondary cell of
clockwise is formed near left sidewall of main channel apart from two larger secondary cells in
upper and corner region of main channel. The FCF, a large scale facility having channel length
and width of 60 m and 10 m, respectively, was constructed in 1986 for experimental
investigations of overbank flows in rivers. It was configured as rigid or mobile bed channel as
per requirement. Phase A of the facility mainly included on straight and skewed fixed bed
compound channels flows, and experimental results were presented by Ackers (1993) who
introduced the coherence concept in account of the interaction effect between the main channel
and the floodplains.

Tominaga and Nezu (1991) carried out experimental studies in a laboratory flume composed of
painted iron plates channel bed and smoother glass plates side walls to investigate secondary
current patterns, turbulent structure of flows such as turbulent intensities and Reynolds stresses
in straight asymmetric compound open-channel of length 12.5 m and width 0.4 m, under various
relative water depth ratios, and reported a 3-D data set using Fiber-optic Laser Doppler
anemometer (FLDA). The data set was used by several researchers (Naot et al., 1993; Kara et al.,
2012; Xie et al., 2013) for examining the validity of their numerical model of compound openchannel flows. They reported turbulent flow characteristics of the channels for different water
depth ratios, and showed the presence of strong inclined secondary current on momentum
transport generated near the junction of the main channel and the floodplain. They inferred that
secondary currents along with vertical vortices in the longitudinal direction, played significant
role in the momentum exchange near the junction region. They also reported the strength of
secondary current as about 4% of the maximum longitudinal velocity (Umax) whereas other
researchers (Naot et al., 1993; Nezu, 1996), reported it as 2-3% of the maximum longitudinal
12

velocity (Umax). They also inferred that component of turbulence intensity increases near the
junction of the main channel and floodplain.

Nezu and Nakayama (1997) experimentally obtained detailed information about 3-D flow
structures in compound open channel using both Particle Imaging Velocimetry (PIV) and LDA.
In their works, they observed presence of strong interaction between upward flows and
horizontal vortices at the junction of the main channel and floodplain.

Nezu et al. (2004) developed a new dual-layer PIV system to analyze the velocity distribution at
different elevations simultaneously, and investigated 3-D coherent properties of horizontal
vortices, convection mechanism and transverse momentum exchange between the main channel
and the floodplain for shallow and deep water depths.

Sanjou et al. (2010) conducted an experimental study on a large scale compound open channel
flume using acoustic Doppler velocimetry (ADV) to reveal space correlation properties and
velocity fluctuations between the main channel and floodplains. Based on their experimental
observations, they concluded that two layer model can be very useful to understand the spanwise
flow features and interaction between the upper and the lower layers in the main channel.

Stocchino and Brocchini (2010) presented a statistical analysis of the properties of quasi-2D
macro-vortices which form at the interface of a compound open channel under quasi uniform
flow condition, and showed that these structures remain constant in the streamwise direction.
Stocchino et al. (2011) conducted extensive experimental studies for showing Lagrangian mixing
process and quasi-2D macro-vortices properties in a uniform compound open channel flow by
assuming a shallow water conceptual model for varying physical flow parameters like flow depth
ratio rh [=H/(H-h)] and Froude number. Here, H=total flow depth and h=main channel flow
depth.

Fernandes et al. (2014) carried out experimental studies in laboratory compound channel to
investigate longitudinal flows and, corresponding mixing and boundary layers for six uniform
flows generated in different locations of the channel. They reported effects of shallowness and
13

floodplain roughness on longitudinal streamwise velocities and lateral shear stresses. They also
reported relationship of mixing layers developed with the growth rate for different roughness in
channel sections.

2.2.1.2 Numerical investigations

A number of numerical studies on compound open-channel flows were undertaken using


different turbulence models. Pezzinga (1994), Lin and Shiono (1995) and Hosoda et al. (1999)
used nonlinear k- model to numerically simulate compound open-channel flows. Hosoda et al.
(1999) predicted the relation between horizontal vortex and secondary currents. Similar studies
were carried out by Naot et al. (1993) and Cokljat and Younis (1995) who employed algebraic
stress model (ASM) and Reynold stress model (RSM) for turbulence closure respectively.
Thomas and Williams (1995) performed a large eddy simulation (LES) of turbulent flows in a
compound open channel for a Reynolds number (Re) of 42000, and presented detailed mean
velocity field, secondary circulation field, bed shear stress distribution, and lateral stress
distribution. Joung and Choi (2008) investigated about a twin vortices formed near the interface
for a turbulent flow of Re=3900 in a compound open-channel having width-to-depth ratio,
[(bmc+bfp)/(hu+hl)], of unity and flow depth ratio, rh [=(hu+hl)/hu], of two using DNS method,
and compared the results with published experimental works. Naot et al. (1993) investigated
hydrodynamic behaviour of the turbulent flow in a compound rectangular open channel to the
flood plain depth, roughness and symmetry where the free shear instability and boundary
generated turbulence is primarily responsible for complex secondary flows.

Hydrodynamic behaviour of flows in four straight compound channels was reported by Naot et
al. (1993) using 3-D steady state model in conjunction with k- turbulence model and algebraic
stress model developed by Naot and Rodi (1982). Special wall functions were considered in
treatment of channel walls. They inferred that smaller secondary currents influence flow
behaviours, such as velocity and boundary shear stress in channels. Simulated results, such as
streamwise velocity, secondary flow, shear stress and mass exchange between the main channel
and the floodplains were compared with experimental results of Tominaga and Nezu (1991).
They also inferred that large differences in Reynolds number of flows between the floodplain
14

and main channel were one of the reasons for discrepancies occurred between the model results
and the experimental observations.
Pender and Manson (1994) implemented a k- turbulence model in a set of 3-D flow equations
and tested the same against the results obtained from McKeogh and Kiely (1989). The model
was used for predicting water levels and flow velocities in a compound channel.

Smagorinsky (1963) derived an eddy viscosity subgrid scale (SGS) model based on a Boussinesq
approximation and obtained Smagorinsky coefficient. The classical Smagorinsky model is valid
only if scales of size on the order of filter width are within the inertial range, which are larger
than the dissipative scales, yet smaller than the scales that contain of the energy. Large eddy
simulation (LES) of classical Smagorinsky model along with the application of van-driest
damping function for solid walls was first applied by Thomas and Williams (1995) to simulate
complex turbulent structures in a straight compound channel under Reynolds numbers of 42,000
and 43,000. They reported reasonable match of time averaged streamwise flow velocity and
secondary flows with measured data of the SERC Flood Channel Facility at Hydraulics Research
Station, Wallingford, UK, but failed to represent surface depression effect in the channel. They
reported generation of bottom shear stresses, turbulence intensities and turbulent kinetic energy
of same magnitude as that of experimental results but locations of occurrence of their peaks were
different. They inferred that, due to isotropic nature of Smagorinsky (1963) model, the model
was not able to capture secondary flows in channels; and, due to coarse mesh resolution, the
model over-predicted mean velocities by about 8% near the mid of the main channel.

Chau and Jin (1995) reported development of a two-layer, two-dimensional unsteady


mathematical model using time-splitting finite difference method with two fractional steps based
on curvilinear orthogonal co-ordinates and grid block technique, and undertook simulation of
flows with density stratification in a natural water-body with complicated topography. The
governing equations were derived from N-S equations. The two layers were considered as
separated by a horizontal interface lying between them. Further, it was assumed that upper layer
only existed above the horizontal interface. Values of flow variables at the interface were
estimated based on flow variables at upper and lower layers using empirical relationship for eddy
15

viscosity. The turbulent exchange was treated empirically across the interface of the flow
domain. The hydrodynamic model was applied for Tolo Harbour, Hong Kong. They reported
velocity hodographs showing different patterns of tidal circulations at various positions in each
layer. Based on trends of computed Lagrangian pathline they inferred that tidal excursion was
dependent on types of tide in flow domain. In order to capture the interaction between turbulence
and mixing across the interface, they suggested for use of more sophisticated turbulence model.
They also suggested incorporation of transport equations in their model as future scope of work.
Their models were used by Jin et al. (1998) for flow simulation in a meandering compound open
channel.

Nadaoka and Yagi (1998) reported characteristic of large eddies in compound open-channel
under different flow depth conditions using Sub-Depth Scale, SDS-2DH model. In model, eddyviscosity approach, corresponding to Sub-Depth Scale turbulence, was used to model small-scale
turbulences and reproduce mechanism of eddy generation. Wavelength of vortex obtained from
model was compared with experimental results. They also predicted average flow velocity and
bed shear stress profiles in shear layer. It was reported that the model under predicted the flow
velocities near mid of the main channel for relative depths < 0.15 and over predicted the same for
relative water depths > 0.20. It was also reported that flow velocities on floodplains was underpredicted for high relative water depth conditions.

Sofialidis and Prinos (1998) modified low-Reynolds, nonlinear k- turbulence model developed
by Craft et al. (1993) to predict simple flows, and applied the same to study compound open
channel flow with low relative depths (the ratio of water depth on the floodplain to the total
depth). They reported the prediction of turbulent parameters such as turbulent shear stresses and
turbulent kinetic energy and, flow interaction between main channel and floodplain. The model
was not able to generate secondary currents of desired strength in open channel. They stressed
the use of nonlinear stress-strain relationship for representing anisotropy of Reynolds stresses at
the free surface and dissipation equation for prediction of turbulent kinetic energy. Sofialidis and
Prinos (1999) further reported development of k- model, and applied it in straight compound
channel with different relative flow depths. They applied redistribution mechanism for normal
turbulent stresses and simulated eddy viscosity, t, near free surface. They reported success in
16

reproduction of streamwise velocity, secondary currents and turbulent shear stresses from the
developed model but the model was unable to simulate depression of the maximum velocity near
free surface. They concluded that, with increase in relative flow depth, lateral shear at interface is
intensified for an increased range of flow depth in floodplain. Further, the presence of weaker
secondary currents causes the velocity-maxima below the free surface and turbulence levels are
enhanced in the region of interaction.

Jin et al. (1998) applied two-layer two-dimensional mathematical model of Chau and Jin (1995)
for unsteady flows in a sine-generated meandering compound channel, and evaluated the change
of flow patterns such as water level, velocity vector, Lagrangian pathline and friction factor in
terms of water depth ratio between the main channel and the floodplains. Empirical treatment
was proposed for approximation of turbulent exchange across the horizontal interface. They
inferred that shallower floodplain flow was affected severely by meandering flow of main
channel, but it was independent of water depth ratios greater than 0.9.

Shi et al. (2000) reported development of 3-D model for Large Eddy Simulation (LES) of
turbulent flow in symmetric straight compound channels using Series-A experimental
programme of the UK FCF, and compared the results with algebraic stress model. They reported
computational difficulties in using these models and inferred that the models are helpful in
replicating physical processes like intrinsic mechanisms of flow and accurately predicting the
flow features including discharge capacity in straight compound channels. They also inferred
that their models are capable of determining the impact of flow on other processes such as scour
and pollution transport.

Othman and Valentine (2006) developed 3-D model in Cartesian coordinate system for uniform
flow in compound open channel of the UK FCF using non-linear k- (NKE) turbulence model
and used Reynolds stress components for generation of secondary currents. The SIMPLE
technique was used to approximate pressure terms. They reported prediction of primary mean
velocity, secondary currents, bulging of contours at bottom corner of the main channel,
inclination of contours near free surface towards the channel centre and, depression of the

17

maximum velocity below free surface along with formation of vortices in the main channel and
floodplains.

Zarrati et al. (2008) derived semi-analytical equations for streamwise vorticity along with
secondary Reynolds stresses for shear stress distribution in straight open channels of different
cross-sections, such as rectangular, trapezoidal and compound cross-sections. They reported
evaluation of relative shear stress distribution equation along the width of different channel
cross-sections using experimental data, and investigated the effect of secondary flows due to
shear layer between the main channel and floodplains.

Kara et al. (2012) reported model results of large eddy simulation (LES) of turbulent flow in an
asymmetric compound open channel with deep and shallow flow depths and the results were
compared with experimental data of Tominaga and Nezu (1991). They also reported the effects
of water depth ratio of main channel and floodplain on primary and secondary flows, secondorder turbulence statistics and anisotropy of turbulence. They demonstrated that anticlockwise
rotated vortex pair along horizontal axes caused velocity bulge at interface of the main channel
and the floodplain. They also reported that resulting secondary currents as well as primary and
secondary Reynolds stresses contributed to the streamwise momentum transport.

Xie et al. (2013) employed LES approach with dynamic SGS model using partial cell treatment
technique in investigating turbulent structure, large-scale vortical structures and instantaneous
secondary flows in an asymmetric compound open-channel which proved to be significant
parameters in flow resistance and sediment transport fluxes. They reported comparison of mean
velocity and boundary shear stress distributions, secondary currents, and turbulence statistics
with experimental data of Tominaga and Nezu (1991). They inferred that vortical structures can
cause significant lateral exchange of mass and momentum in compound channels.

2.2.2 Solute Transport in Compound Open Channels

Flow characteristic of the rivers during the flood, and transport of sediment and pollutant in
natural systems are the sole reasons to study the channel with compound sections. Heat and mass
18

transfer induced transport phenomena, such as convection and diffusion play a vital role in
human life. Complete understanding of transport phenomena is essential for many industrial
applications like aerodynamic, flows in furnaces, heat exchangers, and chemical reactors. In
nature, both advection and diffusion are responsible for transport processes to occur in natural
rivers or channels with high water. Hydrodynamic processes in such rivers tend to be complex
due to complex geometry and bed roughness conditions. Numerous research works have been
undertaken till date to understand this complex hydrodynamic behaviour. However, a few works
were only undertaken to reveal the effect of hydrodynamic process on mass transport. The
traditional approach to investigate such processes is based on observations and experiments
which have limited scope of obtaining complete information about the processes. The complex
transport phenomena expressed in terms of complex mathematical equations can only be solved
using numerical methods. Research works related to solute transport in straight compound openchannels undertaken till date are reviewed and presented in following paragraphs.

Lin and Shiono (1995) developed 3-D model by solving Navier-Stokes equations in conjunction
with linear and nonlinear k- turbulent schemes to investigate transport processes of solute in
compound open channels. They reported comparison of flow velocity and the solute transport
rates with experimental works undertaken by Wood and Liang (1989). With the help of both
linear and non-linear k- turbulent approaches they showed influence of secondary currents on
the mixing processes in the compound open-channel.

Simoes and Wang (1997) reported three-dimensional numerical model in conjunction with two
algebraic eddy viscosity schemes to simulate time-dependent turbulent flows and transport of
dissolved materials in compound open channels. Accuracy along with advantages and drawbacks
of the two models were reported by them based on comparison of simulated results with
experimental data.

Chatila and Townsend (1998) presented 2-D finite-difference mathematical model using depthaveraged versions of 3-D Navier-Stokes equations in conjunction with advection-diffusion
equation and constant eddy-viscosity model to simulate inert pollutant transport in compound
open channels. They employed Platzman's space staggered scheme and double time step
19

operation in their numerical solution. They reported a comparison of model results with
laboratory data obtained from continuous-injection dye tracer experiments in asymmetrical
compound open channel at Hydraulics Laboratory, University of Ottawa.

Shiono et al. (2003) reported numerical model studies using two turbulence closure schemes, viz.
k- model and algebraic stress model (ASM) for prediction of solute transport injected at three
injection points near the water surface for deep flow depth in a narrow asymmetric compound
open channel of length 20 m and width 0.2 m. They put forward a comparison of simulated
results with experimental work of Shiono and Feng (2003) undertaken in laboratory flume using
sophisticated measuring equipments such as laser Doppler anemometer (LDA) and laser induced
fluorescence (LIF). Fluorescence dye (Rhodamine 6G) was used as solute material injected at 13
m from the channel entrance and measured at 1 m downstream of the injection point. They
reported a skewed distribution of solute on floodplain and inferred that algebraic stress model
(ASM) performed better than k- model. They investigated the cause of skewed distribution
through the variations of secondary flow and eddy diffusivity. They also inferred that, in deep
flow case, the effect of secondary currents on peak concentration in shear layer was significant.
They further reported the presence of secondary current cells of large magnitude in upper region
and smaller magnitude in left corner of the main channel. They demonstrated that, in solute
concentration distribution, the role of turbulent Schmidt number was significant in comparison to
eddy viscosity.

Laboratory experiments for solute transport in a wide symmetric compound open-channel were
undertaken by Fraselle et al. (2008). Total length and width of channel were 10 m and 1.2 m
respectively. Width of main channel and floodplains on either side was equal to 0.4 m each. A
constant discharge of 0.01855 m3/s in compound channel at longitudinal slope 0.001 resulted rise
of water level up to a height of 0.0756 m in main channel for a subcritical flow condition
(Fr=0.4). Flow measurements were undertaken using a pitot tube. Fraselle et al. (2010) reported
depth-averaged primary flow velocity profile of the experiments while comparing results of
numerical models of Telemac-2D software for two turbulence schemes, i.e., k- and Elder
schemes. Diluted common salt (NaCl) was used as solute material. Total four experiments were
undertaken for transport of solute by injecting solute material at different places in main channel
20

at 2 m downstream of channel entrance. Measurements were taken at four sections, viz. 1 m, 2


m, 4 m and 7 m downstream of the injection point along the length of the compound channel.
They reported the dispersion of solute near the junction of the channel due to the formation of
strong turbulent structures when injected in shear layer region. They also reported the role of
turbulent eddies in solute transport especially in shear zone near the interface where strong
vorticity was present.

2.3 CONCLUDING REMARKS

In previous numerical model studies on compound channels, a number of models were


developed based on 1-D, 2-D and 3-D equations, and most of the cases, dealt with using a bodyfitted coordinate system. For closure of turbulence, mostly, k- model had been adopted. Existing
numerical models have their own advantages and disadvantages in prediction of secondary flows
and solute concentration in compound open-channels. Most of the reported studies have focused
their attention on vertical or inclined interface at the junction of main channel and floodplains
whereas little attention has been paid to flow characteristics along the horizontal interface. Also,
the past studies did not analyze the influence of flow depth ratio and cross-sectional channel
characteristics (wide or narrow) on transverse mixing coefficient. The comprehensive
performance of dynamic SGS model is yet to be investigated for their applications in compound
open channel flows (Kawahara et al., 2008). In the light of such evidence, it is worth to develop
2-layered 2-D mathematical model for hydrodynamic and solute transport using Cartesian
equations for non-uniform grid system. It is also proposed to use turbulence closure schemes
using LES approach, such as, standard Smagorinsky (SS) model and dynamic subgrid scale
(DSGS) model.

21

Chapter 3

Theoretical Consideration and Modelling

3.1 GENERAL

In this chapter, the governing equations used in formulating hydrodynamic and transport flow
models in straight rectangular compound channel are presented. A rectangular compound
channel has been represented in the form of a two layer system separated along the horizontal
interface at junction of the main channel and the floodplains, considering lower main channel as
bottom or lower layer, and upper main channel and floodplains taken together as top or upper
layer. Basic governing equations for lower and upper layers are described separately.
Subsequently, the partial differential equations describing the transport of inert pollutants in the
compound channels are presented. Also, the numerical scheme and its stability and relevant
boundary conditions being used in solution of aforesaid equations are presented.

3.2 GOVERNING EQUATIONS OF TWO-LAYERED FLOWS

The turbulent flows in natural channels can be solved using simplified two-layer twodimensional models. Such considerations simplify the analysis and yields reasonably good
results. Two-dimensional shallow water flow equations are obtained by integrating threedimensional equations of motion along the depth individually for each layer. Derivation of twolayered two-dimensional equations for body-fitted coordinate system was presented by Chau and
Jin (1995) and Jin et al. (1998). These equations have been transformed into Cartesian coordinate
system and used in present study for simulation of hydraulic variables and inert pollutant

22

transport. The schematic of two-dimensional and two layered flows system of a typical
compound channel is shown in Fig. 3.1.

3.2.1 Hydrodynamic Equations

Conservation of mass and momentum of an incompressible fluid over depth of each layer can be
expressed in terms of partial differential equations, are derived in terms of depth-averaged
equations. For such two layered system, the position of interface is considered horizontally in
between upper and lower layers. In shallow water flows, where interface level is at or lower than
upper layer bed, the lower layer vanishes and only upper layer exists.

Fig. 3.1 Schematic of two layered compound open channel flows

The two layered two dimensional hydrodynamic equations for shallow water flows in Cartesian
co-ordinate system with consideration of mass and momentum exchange between the layers are
derived in the same way as suggested by Abbott (1979) for depth-averaged formulations. The
layer averaged continuity and momentum equations for each layer are described in following
paragraph. The detailed derivation of the equations is available elsewhere (Chau and Jin, 1995).

The governing equations are derived based on following key assumptions:


1. The pressure distribution is hydrostatic in each layer, i.e., convection and friction terms in
momentum equation in vertical direction are negligibly smaller than the pressure gradient
23

and gravitation. This assumption is valid for gradually varied flow. Although some
details are lost in the vicinity of sharp gradients if the shallow water equations are used.
The overall results are adequate for engineering purposes (Cunge, 1975).
2. Density in each layer is constant and equal, i.e., variation in density between the layers is
neglected.
3. Coriolis force is negligible.

Equations for upper and lower layers can be expressed using Eqs. (3.1) and (3.2) respectively as
(3.1a)

(3.1b)

(3.1c)
(3.2a)

(3.2b)

(3.2c)

Here, subscripts u and l designate the variables corresponding to upper and lower layers,
respectively. In other subscripted variables, symbols appearing as o, s and b denote the
position at interface, surface and bed, respectively. Here, hk, uk and vk represent filtered water
depth, streamwise and spanwise layered average velocity, respectively in kth layer;
average filtered turbulent stress at ith face in jth direction in kth layer;

and

is layered
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

24

density of water; so= channel bed slope in streamwise direction; and g = gravitational
acceleration.

Bed shear stresses

and, shear stresses

and flow velocities (uo, vo, wo) at horizontal

interface of the two layers appearing in governing equations (Eqs. 3.1-3.2) is expressed
empirically which is similar to that adopted in Chau and Jin (1995).

The filtered bed shear stresses are represented empirically similar to depth averaged method as
;
where,

(3.3)

is the local friction flow velocity, and is defined as


(3.4)

where,
,
(3.5)
and total depth of water, h=hu+hl ; and cf , the coefficient of bed resistance, can be determined
using
(3.6)
where, n is the Mannings roughness coefficient.
Here, horizontal interface is considered at the top of lower layer.

The filtered shear stresses at interface between the two layers are approximated empirically as
;
where,

(3.7)
; and

is the mixing layer thickness at interface which can be

approximated as Prandtl mixing length, lo, defined as


(3.8)
where, k is the Von Karman constant generally equal to 0.41.

The filtered surface shear stress terms, i.e.,

and

are not taken into account as the effect of

wind conditions have been assumed negligible in present model.

25

The filtered horizontal interfacial velocities in streamwise and spanwise directions can be
expressed approximately (Chau and Jin, 1995) as
,

(3.9)

Here, vertical interface velocity, wo, considered positive upward, resulting due to convective
exchange between the layers, is determined from the layer-averaged continuity equation (Eq.
3.1a or 3.2a) of either of the layers after obtaining flow velocities and water surface levels of
respective layers.

Eqs. (3.1) and (3.2) can be written in divergence or strong conservative form as
(3.10)
where Uh, Eh, Gh, and Sh are row vectors, defined as

In Eq. (3.10), Uh is the state vector, Eh and Gh are the vectors representing flux terms and Sh is
the vector representing source term.

26

3.2.2 Governing Equations for Solute Transport

According to Fick's laws of diffusion developed by Adolf Fick in 19th century (Mehrer and
Stolwijk, 2009), the solute flux due to diffusion is proportional to the solute concentration
gradient, and the rate of change of solute concentration at a point in space is proportional to the
second derivative of its concentration with space.
Mathematically, for molecular processes, flux is specified by Ficks law, and diffusion equation
can be written in Cartesian coordinate in two-dimension as
(3.11)
where, Dx and Dy are the diffusion coefficients in x- and y-directions, respectively. In case of
molecular diffusion, Dx=Dy=D.
The processes of advection, caused due to mean motion of the fluid, and diffusion are separate,
however, additive in nature. The rate of mass transport by the velocity through a unit area in
perpendicular direction (x-direction) is given by the quantity (u.c). The total rate of mass
transport per unit area in perpendicular direction, sum of advection and diffusive fluxes, in two
dimensions is represented as
(3.12)
While substituting Eq. (3.12) into the equation for conservation of mass, we obtain
(3.13)
Eq. (3.13) is generally referred to as advection-diffusion equation.
Using the above analogy and following Ficks law with negligible molecular motion in
comparison to turbulent motion, two layered two-dimensional solute transport equations for inert
solute in a compound channel can be written as
Upper layer:

27

(3.14)
Lower layer:

(3.15)
where,
ck

concentration in layers (kg/m3), k= u, l


SGS eddy diffusivity (m2/s) in ith direction of kth layer

sk

quantity of inert solute (kg/m2/s) in kth layer

Ke

rate of exchange of solute concentration from one layer to other (per sec)

The layered average transport equation in conservative form can be written as


(3.16)
where, Uc, Ec, Gc, and Sc are row vectors defined as

Combining Eq. (3.10) and Eq. (3.16) [i.e. all equations containing hydrodynamic and transport
terms together], final form of governing equation in divergence form can be expressed as
(3.17)
where, U, E, G, and S are row vectors. U is the state vector, E and G are the vectors representing
flux terms and S is the vector representing source term; and can be expressed as

28

3.3 METHOD OF SOLUTION

The governing equations expressed in two layered two dimensional form (Eq. 3.17) are a set of
non-linear, hyperbolic partial differential equations, and their solutions can be obtained using the
numerical methods. Finite volume method, among the other available numerical methods, is
suitable choice, for the solutions of non-linear partial differential equations as they can be
implemented with ease even for complex unstructured domains. In present study, finite volume
method, with explicit scheme, is used for solutions of the governing equations. For discretisation
of various terms of the governing equations of both layers, both in space and time, method of
lines approach is employed. This approach has lot of flexibility as flux integrals are evaluated
29

using grid volume technique and the resulting time-dependent equations are solved from a
known initial values. Further, for numerical approximations, different accuracy level of spatial
and temporal derivatives can be adopted. Brief description of finite volume method,
discretisation of governing equations, turbulent closure schemes, appropriate boundary
conditions and stability consideration of numerical scheme are described in following
paragraphs.

3.3.1 Finite Volume Method

Finite volume method (FVM) utilises conservation laws through integral formulation of NavierStokes equations. It was first employed by McDonald in 1971 for simulation of 2-D inviscid
flows. The governing equations, in such method, are discretized by dividing physical space into a
number of non-rectangular control volumes as shown in Fig. 3.2. The surface integral flux terms
are approximated by the sum of fluxes crossing individual faces of control volume. The accuracy
of spatial discretisation depends on a particular scheme with which fluxes are evaluated.

Fig. 3.2 Representation of physical domain in model

The FVM is based on direct discretisation of conservation laws such as mass, momentum and
energy which enable it to compute correctly even weak solution of the governing equations. In
present study, cell-centred approach of FVM is used. In this approach, entire flow domain is first
30

divided into a set of cells. Each cell is identified by the corresponding centre point and the flow
quantities are stored at the centroid of grid cells.

Fig. 3.3 shows a non-orthogonal elemental cell (i,j) and its immediate neighbours. The governing
equations are integrated using finite volume technique on each of cells covering the whole flow
domain. After the application of Gauss divergence theorem, Eq. (3.17) can be written in integral
form, as under,
(3.18)
where, F = flux terms at cell centre.
The 1st term represents the integral of time evolution of function over the area of cell. The 2nd
term is the total normal flux through cell boundaries and the 3rd term represents the integral of
source quantities over the volume of grid cells.

Fig. 3.3 Non-orthogonal grid cell and its neighbours

The scalar product (F.n) can be expressed in terms of components in X and Y directions as
F.n = E.nx + G.ny

(3.19)

where, nx, ny are X- and Y- components of unit vector n respectively.


The unit vectors normal to cell wall (considered as positive along outward facing) are shown in
Fig. 3.4.

31

Fig. 3.4 Control volume of grid cell (i,j)

Assuming the vector U to be uniform over a 2-D grid cell with unit depth, Eq. (3.18) can be
written as
(3.20)
where,

is the area of finite volume cell.

The surface integral in Eq. (3.20) is approximated by sum over the four walls of grid cell as
(3.21)
where, drr is length of each wall of grid cell that contours the cell (i,j), and Fr is numerical flux
through the cell face r that contour the cell (i,j)

In order to approximate surface integral by summing over all four walls of grid cells, geometrical
quantities of grid cell in terms of its surface area, its face area and unit normal vectors are
calculated.

In present study, 2-D flow problem is considered on XY plane having symmetric with respect to
Z-direction. Because of symmetry, and in order to obtain physical units for flow quantities, depth
of each grid cell is set equal to one resulting into the grid cell as two dimensional with its volume
equal to its area.
32

Area of a grid cell is calculated using Gaussian formula as


(3.22)
In language form, area of a grid cell is the summation of product of opposite diagonals taken as
X- and Y-coordinates of the vertices, alternatively.

In 2-D, faces of control volume are given by straight lines and therefore, unit normal vector is
constant along them. During approximation process, fluxes at cell face are integrated based on
the product of area of face A and the corresponding unit normal vector n and the face vector s.
(3.23)
Face vectors of the grid cell are given by the relations
(3.24a)
(3.24b)
(3.24c)
(3.24d)
The unit normal vector at face m is obtained from Eq. (3.23) as
(3.25)
where,

While coding, face vectors

and

are only computed and stored for each grid cell. In order to

save memory and reduce the number of operations, the face vectors

and

are taken from the

respective neighbouring grid cell with reverse sign.

3.3.2 Discretisation of Governing Equations

The procedure for discretisation of each term of governing equations is followed separately as
discussed in following paragraphs:

33

(i) Discretisation in Space and Time


For the solution of partial differential governing equations, i.e., Eq. 3.17, method of lines
approach was employed for each layer. In order to discretise governing equations, physical
domain for each layer is considered as per Fig. 3.2. The 2nd term, i.e., surface integral of Eq.
(3.18) is approximated by sum of fluxes crossing the faces of grid cell. It is usually supposed that
flux is constant along the face and evaluated at midpoint of the face. The source term, i.e., 3rd
term of Eq. (3.18) is generally assumed to be constant inside the grid cell.
For a fixed grid cell, time derivative of conservative variable

can be expressed as
(3.26)

where, Ap is area of Pth grid cell.

(ii) Continuity Equation


Continuity equation is approximated for each layer as follows
(3.27)
where, dsr represents length of rth face surrounding grid cell (i,j) and

is the flow variable

defined at face centre of grid cell (i,j).

(iii) Convective flux terms


Fig. 3.5 shows grid cell (i,j) where left and right sides of faces are indicated in both X and Y
directions. In cell-centred scheme, linear interpolation of flow variables is carried out at each
side of cell face based on information available at cell centroid. The left and the right state values
are utilised to compute convective flux through the cell face.

Evaluation of numerical flux at a face of grid cell (i,j) between nodes (i+1,j) and (i,j) is carried
out as
(3.28)
where, is the positive coefficient. FR = f(UR) is the flux computed using the information from
the right side of cell face and, FL = f(UL) is the flux computed using the information from the left
side of cell face.

34

Fig. 3.5 Left and right state at cell face [(i+1/2,j) & (i,j+1/2)]

The UR and UL are obtained using a MUSCL (Monotone Upwind Scheme for Conservation
Laws) approach as
,
and

(3.29)

where, subscript (i,j) is the value at node (i,j) and subscript (i+1/2,j) is the value at interface
between nodes (i,j) and (i+1,j). There are several ways to determine

and

using

slope limiter procedures (Yee, 1989; Alcrudo et al., 1992). The slope limiter procedure is
generally used to suppress non-physical oscillation of the solution. The minmod limiter
procedure has been adopted in present study. According to this,

and

(3.30)

where, minmod function is approximated as


minmod (a,b) = a,

if |a| < |b| and a.b > 0

b,

if |a| > |b| and a.b > 0

0,

if a.b < = 0

Positive coefficient in Eq. (3.28) is determined using the maximum value (for all the grid
points) of the largest eigen value of Jacobian of the system of equations (Nujic, 1995) such as
i=1 to Ni

(3.31)
35

and

j=1 to Nj

where
in which Ni, Nj are the total number of grid points in X- and Y-directions respectively. Vi,j is the
resultant velocity and

for 2D cases.

(iv) Viscous flux terms


In order to obtain consistent spatial discretisation for viscous flux, grid cells are generally chosen
to be the same as for convective fluxes. However, viscous flux, Fv are included in governing
equations as
(3.32)
where, Fc is the convective flux
Viscous flux along with related stress terms are evaluated from neighbouring variables averaged
at each face of grid cell. The Greens theorem approach is, generally, used to accomplish the
computation which requires construction of an additional grid cell known as auxiliary grid cell
(AGC). The derivative terms appearing in diffusive flux of momentum equation of each layer are
evaluated at each cell face using Greens theorem over the flow variables in an auxiliary control
volume (surrounded with dashed line) constructed by joining the midpoints of the edges of
adjacent grid cells corresponding to streamwise and spanwise directions, see Fig. 3.6.

(a) Streamwise direction


(b) Spanwise direction
Fig. 3.6 Computational cells showing auxiliary control volume (dashed lines) used for
calculations of derivative terms of diffusive flux
36

While computing diffusive flux, the non-uniform nature of grid cell is treated by adopting an
interpolation scheme based on averaging of the flow variable of the grid cell with the area of the
diagonally opposite cell (Sharma and Eswaran, 2003). These are subsequently summed up as the
contribution to the viscous fluxes as given in Eq. (3.32).

Derivative in X-direction is approximated as


(3.33)
where, A is the area of AGC, is the flow variables at face centre of AGC and
are the components of face vector of AGC computed based on geometrical shapes.
Expansion of Eq. (3.33) gives
(3.34)
where value of variables at the face of AGC is obtained as

The interpolation formula adopted herein is based on the averaging procedure as


(3.35a)
(3.35b)
where, 23 and 21 are the interpolated flow variables at the corners of face 2 of cell P used in
the computation of derivative in streamwise direction for diffusive flux, Ai,j is the area of cell P,
i,j is the flow variable at the centre of cell P. Similar approach of interpolation has also been
followed in calculation of derivatives in spanwise direction.
(v) Source terms
The source term is integrated over cell area as under
(3.36)
where Sp represents source terms at centre of grid cell (i,j) and Ap is area of grid cell (i,j).

37

3.3.3 Expansion of Turbulent Stress Terms

(i) Turbulent terms in Hydrodynamic equations


Stress tensor,

, designates stresses at ith face in jth direction of governing equations,

representing turbulence terms can be replaced as


(3.37)
where, subscript l represents laminar flow and the last term is the subgrid scale (SGS) stress
tensor.
Here, ui u j can be expressed as
(3.38)
Averaging over the area of grid cell (for 2-D case) and neglecting infinitesimal terms, we obtain
(3.39)
The second term appeared in right hand side of Eq. (3.39) represents the turbulence terms which
can be expressed as
(3.40)
where,

is the SGS eddy viscosity at kth layer approximated using turbulence closure schemes

such as zero equation, one equation, two equations, large eddy simulation (LES) and direct
numerical simulation (DNS). In present study, LES approach is adopted for closure of turbulence
terms appearing in governing equations for both layers. Two schemes, viz. Standard
Smagorinsky (SS) and dynamic subgrid scale model (DSGS) model under large eddy simulation
(LES) technique have been used in present study.
(ii) Turbulent terms in Transport equations
Considering convective terms of transport equations in X-direction as
(3.41)
Averaging over the area of grid cell and neglecting small terms, Eq. (3.41) can be written as
(3.42)
where subscript k represents the kth layer of flow domain

38

Similarly, in Y-direction, averaged convective terms can be written as


(3.43)

In order to model correlations between turbulent fluctuation of velocity and scalar quantity
transport turbulence terms, viz.

and

appearing in right hand side of Eqs. (3.42) and

Eq. (3.43), eddy diffusivity coefficient, t is introduced using Boussinesq approach similar to that
adopted in hydrodynamic equations as
(3.44)
and,
(3.45)
where,
(3.46)
and,

is Schmidt number, generally, varies between 0.5 to 1.0 (Djordjevic, 1993; Lin and
is turbulent eddy viscosity in ith direction at kth

Shiono, 1995; Simoes and Wang, 1997), and


layer.

Turbulent eddy viscosity, in X and Y directions, can be expressed as


(3.47a)
and,

(3.47b)

3.4 TURBULENCE CLOSURE SCHEME

Turbulence is a frenzied process in nature in which flows are generally indiscriminate and
therefore, in order to model such a system it is necessary to employ statistically averaged flow
variables. Shears stress is generally generated due to friction from the bed and banks of flow
domain resulting into development of turbulence and change in the mean flow field. The
modelling of turbulent flows requires proper treatment for the length and time scales over which
turbulent eddies occur.
39

The turbulence-modelling research is based on the eddy-viscosity concept given by Boussinesq


which, in turn, closely related to the mixing length hypothesis proposed by Prandtl. The mixing
length model is generally specified as an algebraic model of turbulence. First, complete model of
turbulence was introduced by Kolmogorov (Wilcox, 1994).

3.4.1 Large Eddy Simulation (LES)

Flow variables are generally separated into resolved and unresolved parts while using Large
Eddy Simulation (LES) in investigating the turbulent eddies in compound open channel flows
(Lesieur et al., 2005). The resolved parts (large scale) affecting the turbulent diffusion of
momentum are computed through the numerical solution of the conservation equations whereas
the unresolved parts (small scale) are modelled using different subgrid models. The LES
technique has been utilised successfully in capturing the unsteady nature of large eddies. It has
also been proved that LES can model the flow behaviours better than the Reynolds Averaged
Navier-Stokes (RANS) approach. Details of subgrid models can be found elsewhere including in
Lesieur et al. (2005). In LES generally central schemes are applied and the subgrid scale (SGS)
term is treated with explicit model.

3.4.2 Subgrid Scale (SGS) Modelling

Time-dependent equations for large-scale motion are generally resolved directly, however,
small-scale turbulence, that cannot be resolved with the chosen numerical grid, must be
approximated by a model, known as subgrid scale (SGS) model. In this approach, large scales
turbulence is solved using time-dependent equations along with suitable numerical grids and
small scale turbulence is approximated through modelling technique.

(i) Standard Smagorinsky method (SS Model)


Smagorinsky model is first and most widely used SGS model, which is based on the concept of
an eddy viscosity relating small-scale part of SGS stresses,

, to strain rate

velocity field. The Smagorinsky model for kth layer can be expressed as
40

of the large scale

(3.48)
The SGS viscosity term,

appearing in Eq. (3.48) can be approximated as


(3.49)

where,

is the strain rate; and

is the grid volume and C

is the SGS coefficient, hereafter called as model parameter or coefficient. The parameter C is a
calibration parameter and its optimal value depends upon type of flow, the Reynolds number and
the discretisation scheme. Smagorinsky model having the limitation that, for

, it is

strictly dissipative and does not allow for opposite flow.

Similarly, expanding Eqs. (3.47) in terms of flow velocities in streamwise and spanwise
directions, can be expressed as
(3.50a)
and,

(3.50b)

(ii) Dynamic subgrid scale method (DSGS Model)


In standard Smagorinsky (SS) model, constant value of C is applied for the whole flow domain
and in order to model the flows near the solid walls, the damping factor is used. However, in
LES, several potential difficulties do occur in case of shear flows and flow near the solid walls
due to selection of a constant Smagorinsky coefficient, C. A dynamic model for C was proposed
by Germano et al. (1991) to account for asymptotic behaviour of the subgrid scale stresses
existing near the walls. In this model, coefficient C is allowed to vary locally to adjust eddy
viscosity in the flow field. The procedure is started initially by specifying a nominal value of C,
and, eventually at the end of each time step, a new value of C at each grid cell is determined
based on the flow field. Thus, the value of coefficient C is continuously modified with time as
simulation proceeds.

41

The dynamic model is developed by the introduction of a test filter with a length scale larger
than original filter and, generally, a factor of two is adopted for the same. The stresses obtained
due to test filter are referred to as subtest scale (STS) stresses, and are modelled similar to that of
the SGS stresses. All parameters are calculated separately for each layer before averaging C
value for lower and upper layers in spanwise direction.

The

anisotropic

part

of

the

SGS

stress

at

grid-filter

scale

is

given

by

and that of test-filter scale is given by Tij, as under (Eqs.3.51)


(3.51a)
(3.51b)
where

and

are unknowns. Eqs. (3.51) are modeled by using scale-invariance, as


(3.52a)

and

(3.52b)
where,

and

refers to the strain rate tensors at the grid- and test- filter levels for each layer

respectively. The magnitude of strain rates

and

can be written as

(3.53)

(3.54)

It is worth mentioning here that grid-filtered strain rates can be obtained locally from nonequilibrium conditions, as mentioned above; and test-filtered strain rates can be determined
directly from gradient of test-filtered velocity field, which can, in turn, be obtained by explicit
application of test-filter on grid-filtered velocity field.

The unknown SGS stress at each filter level can be related by Germano identity as
(3.55)

42

where, Lij is the resolved turbulent stress for each layer, and upon substituting Eq. (3.51) and Eq.
(3.52) into Eq. (3.55), an expression relating Lij and other known variables are obtained with
model coefficient C being the only unknown (Eq.3.56).
(3.56a)
where,

(3.56b)

Since, the tensor expressions, i.e., Eq. (3.56) leads to five independent equations containing C, its
solution can be obtained using a least square minimization approach as (Eq.3.57)
(3.57)
where, < - > represents the spatial averaging in homogenous directions so as to stabilize the
computations following the summation convention of the repeated indices. In this study,
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 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 the
flow field is examined, and a new value of C for each grid cell is determined. Therefore, the
parameter C is continuously updated as solution proceeds with time. In present study, dynamic
model has been developed by introduction of a test filter, replacing the grid filter with a length
scale of two times larger than that the length scale of the grid filter. The stresses in this range are
referred to as subtest scale (STS) stresses, and are modelled separately for lower and upper layers
which is similar to that of SGS stresses. In order to calculate the subtest scale (STS) stresses, the
test-filtered quantities are determined using the grid filtered variables. Among the several
available approaches, the discrete trapezoidal filter approach has been adopted for an efficient
implementation of test-filtered quantities, as they are used to accommodate with underlying
rectangular grid structure (Premnath et al., 2009). The procedure of filtering involves application
of one dimensional filter individually in each of the two coordinate directions in both the layers.
For detailed description, the work of Premnath et al. (2009) can be referred.
43

In present study, Lij and Mij are calculated separately for lower and upper layers at each grid cell
before taking their summative average in streamwise direction. In order to keep positive value of
C,

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 streamwise direction.
Thus, the same value of C is obtained at grid cells in spanwise direction for both the layers.
However, at each grid cell for rest of the floodplain portion in spanwise direction at upper layer,
the value of C is obtained based on summative average in streamwise direction.

Test Filters
There are several approaches available to obtain the test-filtered quantities directly from gridfiltered variables. Discrete trapezoidal filter approach is one of them applied in this work as they
naturally accommodate with the underlying non-uniform grid structure and thereby allowing an
efficient implementation. They involve applying one-dimensional filter successively in each of
two coordinate directions to obtain test filtered values

from

at grid cell (i,j) as follows:


(3.58a)
(3.58b)

In order to calculate test-filter quantities from the grid-filtered variables, one layer of ghost cell
around the model domain is considered whose values are interpolated from the respective inner
grid cells.

3.5

INITIAL AND BOUNDARY CONDITIONS

The finite volume grid is chosen such that the boundaries of the flow domain coincide with the
cell faces of the control volume cells. Boundary conditions are specified at all the outer cell faces
of the flow domain. Four types of boundaries (i) no flow boundary, i.e., solid wall boundary, (ii)
inflow boundary, i.e., upstream boundary (iii) outflow boundary, i.e., downstream boundary and
(iv) free surface boundary, i.e., top boundary are encountered in present study. In order to close
the problem and solve the governing equations suitable initial and boundary conditions are
44

required to be specified. The initial conditions determine the state of the flow at the time t = 0.
Further, for a faster final solution, initial condition should be chosen closer to the likely solution
and therefore, it is important that the initial solution should satisfy the governing equations. The
proper imposition of boundary condition is crucial as the accuracy of the solution depends on a
proper physical and numerical treatment of boundaries. In addition to that, the robustness of the
model and the convergence speed of the solution are considerably affected by the boundary
conditions. The treatment of various boundaries of the flow domain is presented in following
paragraphs.

(i) Upstream boundary


Generally, at inlet or upstream boundary, all flow variables are prescribed for calculation of
convective flux terms. The diffusive flux terms are usually unknown, and they are approximated
using known boundary values of the flow variables and one-sided approximations for the
gradients. In the present work, Dirichlet boundary condition is applied at the upstream end of
lower and upper layers. Layered average discharge, calculated using the respective crosssectional area of the layer, is imposed at upstream boundary in lower and upper layers. The value
of water depth in respective layers is extrapolated from the respective inner grids in such a way
that flow should remain in subcritical state at the boundary. Further, random perturbations upto
5% of the mean bulk velocity were superimposed in flow velocities in both the directions.

(ii)

Downstream boundary

Outlet boundaries should be as far downstream of the region of interest as possible. The flow
should be directed out of the domain over the entire outlet cross-section and be parallel. The
simplest approximation at the outlet boundary is that of zero gradients along the grid lines. For
convective flux, a first order upwind approximation is used implicitly. For getting higher
accuracy, higher-order and one-sided finite difference approximations of the derivatives at the
outlet boundary can be used. Both convective and diffusive fluxes can be expressed in terms of
the variable values at inner nodes. In present study, Neumann condition is imposed at the
downstream boundaries for both lower and upper layers, i.e.,
(3.59)

45

and

(3.60)

where, h, u and v are flow variables in kth layer.

(iii) Sidewall boundary condition


In boundary layers, close to the walls, the normal gradients in the flow variables become
extremely large as the wall distance reduces to zero. A large number of mesh points close to the
wall is required to resolve these gradients. Furthermore, as the wall is approached, the turbulent
fluctuations are suppressed and viscous effects become important in the region, known as the
viscous sub-layer. Standard turbulence models is not valid near the wall, hence, special wall
modelling procedures is required in the vicinity of the wall.
Standard wall functions are based on the assumption that the first grid point off the wall is
located in the universal law-of-the-wall or logarithmic region. This allows avoiding the
resolution of very thin viscous sub layer, leading to a reduction of the number of cells. On the
other hand, standard wall function formulations are difficult to handle, because it has to be
ensured that the grid resolution near the wall satisfies the wall function requirements. The most
popular way to impose boundary conditions at solid walls for flow problems is the reflection
technique, where an extra row of ghost cells is added behind the wall, see Fig. 3.7. The
description of wall boundary conditions adopted in present study is enumerated in following
paragraphs:

Fig.3.7 Model grid along with ghost cells


46

Partial slip condition


As explained above, a wall function is applied to avoid the necessity of using extremely small
grids near solid wall (Yuan et al., 1988). The log law is assumed to hold near the wall for time
averaged velocity (Launder and Spalding 1974). Generally, a wall-wake velocity profile
(Eq.3.61) represent wall boundary layer at different skin friction cases (Ghose and Kline, 1978;
Bardina et al., 1981).
(3.61)
where,

, k=0.41, and A1=5.3, adopted based on results of

Nezu and Rodi (1986) for open channels flow. Here, up represents tangential velocity component
along wall boundary as shown in Fig. 3.8.

Fig. 3.8 Diagram of partial slip boundary condition


The wall shear velocity in kth layer,

, is represented as
(3.62)

where, w is wall shear stress, and over bar represents time-averaged value.
In order to resolve time-averaged streamwise velocity, Up, near wall boundary, similarity rule
suggested by Shiono et al. (2003) is used in present study as
(3.63)
and,

(3.64)

47

Here, up and

are the instantaneous and time-averaged tangential velocity respectively at grid

cell near sidewall; and

is the time-averaged shear stress at wall.

Above hypothesis associate wall stresses with the streamwise velocity at half cell distance from
the wall. In order to calculate slip velocity at given instant, time-averaged velocity,
calculated at y/2 away from the wall. Mean wall shear velocity,

is

, is calculated using Eq.

(3.62). Instantaneous shear stress, , is calculated using Eq. (3.63) and Eq. (3.64). Eddy viscosity
is approximated using SGS Reynolds stresses under LES scheme.
Finally, velocity gradient,

, at y/2 away from the wall is calculated using


(3.65)

where

has negligibly small value.

Finally, slip velocity ub at the wall boundary of kth layer is evaluated from up and

as
(3.66)

Flow variables at the centre of ghost cell are extracted from the known slip velocity at wall
boundary. Based on known flow variables at the centre of ghost cell, derivatives of diffusive flux
are calculated and imposed as boundary condition at the side wall.

Boundary conditions for solute transport model were considered in such a way that there was no
loss of material through the solid boundaries of flow domain, i.e., by considering solute
concentration in terms of

(iv)

in the governing equations (Eqs.3.14 and 3.15).

Free Surface boundary condition

A rigid free-slip condition has been implemented at the free-surface.

(v) Initial conditions


Initial conditions are required at time, t=0 to start the transient computations. At time, t=0 the
flow depth, the velocity components and the concentrations for all the cells are specified as the
initial conditions. Ideally, initially, the flow depth should be equal to zero, however, being a two
layer model, a water depth upto the top face of lower layer is assumed in lower layer and a small
48

water depth is assumed in upper layer which lies over the lower layer at time t=0. This
assumption simplifies the overcoming of the numerical singularity in computation.

3.6 PREDICTOR-CORRECTOR SCHEME

In order to solve the hyperbolic partial differential governing equations, a two-step predictorcorrector scheme developed by Nujic (1995) for solving shallow water flow equations was used.
The procedure adopted in the scheme was based on Total Variation Diminishing (TVD) and
Essentially Non-Oscillating (ENO) schemes. The high resolution scheme was later on modified
by Singh and Bhallamudi (1997) for two-dimensional case to account for the non-zero source
term in the continuity equations. This scheme, unlike many other classical second order accurate
schemes, such as the MacCormack scheme, is non-oscillatory even when sharp gradients in the
flow variables are present. The main advantages of this scheme are its simplicity, ease of
implementation and more robust than the classical MacCormack scheme. Also, the explicit
scheme is second-order accurate in both space and time. The results obtained by using this
method are satisfactory for many fluid applications. The brief description of the method is
included in following paragraph:
(i)

Predictor part

Predicted value of vector

at unknown time level, t+t, is determined using following

discretisation
(3.67)

In which, superscripts p and * refer to value at known time level, t, and predicted value at
unknown time level, t+t, respectively. Here, t is computational time step. Eq. (3.17) should be
interpreted component wise for the vector U. Eq. (3.67) gives the predicted values of flow
variables, viz., h, u and v for both the layers at unknown time level, p+1, at each node (i,j) of
flow domain.

49

(ii)

Corrector part

The vector U at unknown time level, p+1, and node (i,j), i.e.,

, is computed using the

predicted values and the values at known time level, p.

(3.68)
Following Singh and Bhallamudi (1997),
determined from

and

and

using the same

(which are needed to compute Fp) are


and

which are already determined

in predictor step (Eq.3.69). The procedure cited above results in better numerical stability.
(3.69a)
(3.69b)
Computation of

and

in Eq. (3.68) is similar that of F and S in Eq. (3.67) except that the

predicted values of U are used instead of U values at time level t.

3.7

STABILITY CONDITIONS

The proposed finite volume scheme is an explicit scheme and, therefore, computational time
step, t is computed dynamically at every time step using the Courant-Friedrichs-Lewy (CFL)
condition proposed by Alcrudo et al. (1993) (Eq.3.70).
(3.70)
Herein, dr(i,j) are the whole set of distance between every centre point of grid cell (i,j) and those
of its four adjacent cells. The t is chosen dynamically in numerical model, such that Eq. (3.70)
is satisfied at all nodes of lower and upper layers. The (Vi,j)k is the resultant velocity at grid cell
(i,j) in kth layer and (hi,j)k is the water depth at grid cell (i,j) in kth layer.
The complete methodology including implementation of selected numerical method in
estimation of flow variables of compound channel using developed SS and DSGS models is
shown through a flow chart, see Fig. 3.9. In Fig. 3.9, T represents the last time step of simulation.
50

Fig. 3.9 Flowchart of developed SS and DSGS models

51

3.8 CONCLUDING REMARKS

In this chapter, governing equations of fluid flow and solute transport in a rectangular compound
open-channel in the form of a two-layer system have been presented. Different terms of
governing equations, such as convective, diffusive, source are discretised following the method
of lines and finite volume approach. The approaches followed in present study are second
accurate both in time and space. Turbulent stress terms are dealt with turbulence closure schemes
of standard Smagorinsky (SS) and dynamic subgrid scale (DSGS) model under large eddy
simulation (LES) technique. The discretised equations have been solved using two steps
predictor-corrector approach. Boundary conditions for the treatment of open and sidewall
boundaries of model domain along with stability conditions useful for a numerical model are
discussed in detail. At the end, various steps involved in the development of the models (SS and
DSGS models), in present study, are sequenced through a flowchart.

52

Chapter 4

Simulation of Hydrodynamic Characteristics: Results and


Discussions

4.1 GENERAL

The hydrodynamic model developed based on governing equations presented in Subsection 3.2.1
in Chapter 3, has been used to simulate the flow conditions in wide symmetric and narrow
asymmetric compound channels reported in the literature (Fraselle et al., 2008; Shiono and Feng,
2003). The observed experimental data of these channels have been used to assess performance
of the hydrodynamic model. In addition to this, two cases of laboratory experiments undertaken
by Tominaga and Nezu (1991) conveying shallow and deep flows in an asymmetric compound
channel is also considered to assess the performance of hydrodynamic model. In hydrodynamic
model, two turbulence closure schemes under LES technique, viz. standard Smagorinsky model
(SS model) and dynamic subgrid scale model (DSGS model) are employed. Also, the
comparative performance of SS model and DSGS model using compound channels data are
presented.

4.2 EXPERIMENTAL DATA

4.2.1 Symmetric Compound Channel

For assessment of performance of SS model proposed in present study, laboratory experiments


undertaken by Fraselle et al. (2008) on symmetric compound open-channel were used.
Schematic diagram of symmetric compound open-channel used by Fraselle et al. (2008) in their
53

experiments is included in Fig. 4.1, wherein position of upper and lower layers along with their
horizontal interface is also depicted. Total length (L) and width (B) of symmetric compound
channel were 10 m and 1.2 m respectively. The brief description of channel geometry along with
key flow characteristics employed during the experiments are in included in Table 4.1. The
reported flow measurements were undertaken at a test section 0.6L downstream of the channel
inlet. Further, Fraselle et al. (2010) reported observed depth-averaged primary flow velocity in
aforesaid experiments while comparing the results of Telemac 2-D for two turbulence schemes,
i.e., k- and Elder.

Fig. 4.1 Schematic of symmetric compound channel (Fraselle et al., 2008)

4.2.2 Asymmetric Compound Channel

The experimental set up consisting of 20 m long tilted narrow asymmetric compound channel
was developed by Shiono and Feng (2003). The schematic diagram of asymmetric compound
channel and the position of upper and lower layers along with horizontal interface is depicted in
Fig. 4.2. The dimensions of the channel along with key flow characteristics of observed data is
included in Table 4.1. In experiment, sophisticated measuring equipment such as laser Doppler
anemometer (LDA) was used to measure flow velocities at a test section 14 m downstream of
channel inlet. The experiments were also performed by Tominaga and Nezu (1991) in a tilting
laboratory flume of asymmetric compound section having 12.5 m length and 0.4 m width (see
Fig. 4.3). Channel bed was constructed with a painted iron plate and the sidewalls were made of
glass.

54

Flow velocity were measured using a two-color fiber-optic laser Doppler anemometer (FLDA) at
a test section 7.5 m downstream from the channel entrance after the establishment of fully
developed uniform flow conditions.

Fig. 4.2 Schematic of asymmetric compound channel (Shiono and Feng, 2003)

Experiments were undertaken for different flow depth ratios; out of which two cases, viz. Exp.S2 and Exp.S-3 of flow depth ratios 0.5 and 0.25 respectively, were referred herein for validation
of proposed hydrodynamic model. In both the cases, width of main channel as well as floodplain
of the channel were each equal to 0.2 m. Schematic diagram of the compound channel and the
position of upper and lower layers, and horizontal interface are depicted in Fig. 4.3. Channel
section and flow characteristics employed in the experiments are presented in Table 4.1.

Table 4.1 Compound channel geometries and hydraulic characteristics of experimental data
Upper lower
Channel type

depth

Mean
bulk
Slope
flow vel.
S0

Depth
ratio,
rh

Channel

hu

width
bl

depth
hl

(m)

(m)

(m)

(m)

(m/s)

x10-3

1.2

0.025

0.4

0.0506

0.369

1.0

3.0

wide/
shallow

0.2

0.055

0.1

0.055

0.237

0.5

2.0

narrow/
deep

Exp.
S-2

0.4

0.04

0.2

0.04

0.349

0.64

2.0

normal/
deep

Exp.
S-3

0.4

0.06

0.2

0.02

0.288

0.56

4.0

normal/
shallow

Symmetric
(Fraselle et al.,
2008)
Asymmetric
(Shiono and Feng,
2003)
Asymmetric
(Tominaga
and Nezu,
1991)

width
bu

Lower layer

55

(hu hl )
]
hu

/flow
type

4.3 HYDRODYNAMIC MODEL SIMULATION RESULTS

As mentioned earlier, two turbulence closure schemes, viz. standard Smagorinsky and dynamic
SGS are employed in the proposed hydrodynamic model. Model simulations were carried out for
compound open-channel experiments reported in Section 4.2 and their corresponding results are
presented in following subsections:

(a) Experimental case S-2 (Exp.S-2)

(b) Experimental case S-3 (Exp.S-3)


Fig. 4.3 Schematic of asymmetric compound channel (Tominaga and Nezu, 1991)
4.3.1 Standard Smagorinsky Model (SS model)

4.3.1.1 Symmetric compound channel

First of all, grid dependence test were performed for given flow conditions in symmetric
compound open-channel (Fraselle et. al., 2008) having length to width ratio equal to 8.3 (wide
channel). Thereafter, sensitivity analysis was undertaken for two important parameters employed
56

in the hydrodynamic model. Among the two parameters; one is SGS coefficient, C, used as
model parameter in turbulent schemes and other is a constant, , appearing during the
discretisation of fluxes of governing equations. Based on the optimised number of grid cells and
the values of parameters, the hydrodynamic model under standard Smagorinsky turbulence
closure scheme (SS model) has been validated for primary flow velocities.

(a) Grid dependence test


A set of runs with different aspect ratios (x/y=) 1, mixed (i.e. 1-1.3), and 1.33 for symmetric
compound channel were undertaken for grid dependence test. The root mean square error
(RMSE) of simulated depth averaged primary flow velocities with respect to the observed flow
velocities were calculated for each run as listed in Table 4.2.

Table 4.2 Grid cells used in wide symmetric compound channel for grid dependence test and
verification of SS model

Channel
type

Symmetric

500x60

500x20

0.02x0.02

1.0

RMSE
of
primary
flow
velocity
(m/s)
0.028

500x70

500x26

0.02x(vary)

1-1.3

0.026

500x80

500x32

0.02x0.015

1.33

0.026

Run
No.

Grid cells used in model domain


(Longitudinal x Transverse)
Total Nos.
Spacing
Upper
Lower
(m)
layer
layer

Aspect
ratio
(x/y)

Other
characteristic

Re = 5.5x104
Fr = 0.43

The grid cells giving least RMSE equal to 0.026, i.e., Run No.2, were chosen for further
simulation of other flow parameters. For symmetric compound channel, 500x70 grid cells in
upper layer and 500x26 grid cells in lower layer were used in streamwise x spanwise directions
respectively. In streamwise direction, uniform grid spacing was used whereas, in spanwise
direction, non-uniform grid spacing near side walls was adopted in upper and lower layers of the
channel.
Root mean square error (RMSE) is a frequently used measure of differences between values
predicted by a model and observed in an experiment. It, generally, represents sample standard
deviation of differences between predicted and observed values. The RMSE, expressed in unit, is

57

a good measure of accuracy, but only to compare errors of different models for a particular
variable and not between variables, as it is scale-dependent and quantifies spread of residuals.
Expected value of RMSE is zero indicating good performance of model. The RMSE is defined as
(4.1)
where, Ue is the experimentally observed longitudinal flow velocity (m/s)
Us is the model computed longitudinal flow velocity (m/s)
N is number of observations in spanwise direction

(b) Sensitivity analysis


The sensitivity analysis were performed for two important parameters, namely, C and (Alpha)
used in foregoing hydrodynamic model. Here, C is the turbulence scheme parameter, generally,
varies from 0.065 to 0.2 depending upon the nature of flow and size of the channel. The
(Alpha) is a positive coefficient used in calculation of numerical flux at cell face between two
nodes, and its value normally varies from 0.0 to 1.0 (Nujic, 1993). In present study, both the
parameters were varied within their extreme limits for five combinations as given in Table 4.3.
The depth averaged primary flow velocity profile for each combination is shown in Figs. 4.4 (ab).
Table 4.3 Combination of values of parameters C and
Sl. No.

1
2
3
4
5

0.1
0.065
0.2
0.1
0.1

(Alpha)
0.5
0.5
0.5
0.2
1.0

From Fig. 4.4(a), for =0.5 and C varying from 0.065 to 0.2, it is seen that there is a little
variation exists in simulated primary flow velocity in the vicinity of floodplain walls, whereas in
Fig. 4.4(b), for C=0.1 and varying from 0.2 to 1.0, small variation occurs in flow velocities at
the junction of the main channel and the floodplain. Based on least root mean square error
(RMSE), optimum values of C (=0.1) and (=0.5) were used further for model validation and
extraction of other simulated flow parameters in the channel.
58

Primary flow velocity (m/s)

C=0.1

C=0.2

C=0.065

0.5
Fp
0.4
0.3
0.2

Half of Mc

0.1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Half channel width (m)


Fig.4.4 (a) Variation of primary flow velocity of symmetric compound channel (SS model) with
C (Alpha, =0.5)
(Fp=Floodplain, Mc=Main channel)

Primary flow velocity (m/s)

Alpha=0.5

Alpha=1.0

Alpha=0.2

0.5
Fp

0.4
0.3
0.2

Half of Mc

0.1
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Half channel width (m)


Fig. 4.4 (b) Variation of primary flow velocity of symmetric compound channel (SS model) with
Alpha, , (C=0.1)
(c) Model simulation
The developed hydrodynamic model has been used to simulate the flow conditions in the
symmetric compound channel as per the parameters listed in Table 4.1 and optimal grid
59

conditions derived from grid dependency test (Table 4.2). Near solid walls both in lower and
upper layers, finer grid cells were adopted considering the limiting criteria (y+>30) based on loglaw rule. The value of y+=47 (see, Eq.3.61) has been adopted in foregoing simulation. The
schematic diagram of computational cells in upper and lower layers used in the simulation is
shown in Fig. 4.5.

Fig. 4.5 Computational grid cells and boundaries in symmetric compound channel for SS model
The Mannings roughness coefficients for lower and upper layers were considered as 0.01 and
0.013 respectively and turbulence coefficient, C=0.1 was adopted for entire model domain. The
model was run for steady state discharge of 0.01855m3/s and total water depth of 0.0756m.
Being an explicit model, computational time steps were dynamically calculated based on
60

stability criterion; and after stabilizing the uniform flow condition, a time step of the order of
0.002 sec was found adequate for Courant number equivalent to 0.25. Flow was treated as fully
developed when successive bed shear stress plots had converged at about 75000 time steps. The
simulation was further continued for 25000 time steps to obtain time averaged results which are
presented in following paragraphs

(i) Primary flow velocity


Longitudinal flow velocities obtained from SS model were compared with experimental data
observed by Fraselle et al. (2008) and reported by Fraselle et al. (2010), see Fig. 4.6. From Fig.
4.6, it is observed that velocities obtained from SS model are reasonably in good agreement
(RMSE for Uda=0.026 m/s) with experimental data except in floodplains near the junction of the
main channel and the floodplain. The primary flow velocities are slightly under predicted in the
main channel while reverse trend is observed in the floodplain. It is also observed from the figure
that, due to shallow flows in the channel, only one inflection point occurs near the junction of the
main channel and the floodplain. Flow field of longitudinal flow velocities in upper and lower
layers of the channel are shown in Fig. 4.7. From Fig. 4.7, it is apparent that velocity in
floodplain near solid wall is the minimum and gradually increases upto maximum near the
midpoint in main channel.

Exp. (Fraselle et al. 2010)


30

SS model

Half of Mc

Fp

25
RMSE for Uda=0.026 m/s

Uda/U*

20
15
10

5
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Half of channel width (m)


Fig. 4.6 Comparison of primary flow velocity in symmetric compound channel (SS model)
61

(a) Upper layer

(b) Lower layer

Fig. 4.7 Computed flow field in symmetric compound channel (SS model)

At junction of the main channel and the floodplain in upper layer, a sharp rise in gradient of flow
velocity is seen due to significant difference in primary flow velocities exist between the main
channel and the floodplains which can be explained based on calculations of percentage
difference between primary flow velocities of two layers with respect to mean bulk velocity of
channel

, see Fig. 4.8.

16
Mc
12
8
4
0
-4
0.4

0.5

0.6

0.7

0.8

Channel width at interface (m)


Fig. 4.8 Percentage difference between primary flow velocities of two layers with respect to
mean bulk velocity of wide symmetric compound channel (SS model)

62

In Fig. 4.8, ul and uu represent layer averaged primary flow velocities in lower and upper layers,
respectively, and Ub (=Q/A) represents the mean bulk flow velocity in the same direction. From
Fig. 4.8, it is seen that maximum difference between primary velocities of two layers, of the
order of +12%, is noticed near the junction of the main channel and the floodplain.

It is also observed that, in rest of the interface, the velocity difference between two layers is very
small to the tune of -3%. The change in sign (from ve at mid section of interface to +ve at
junction edges) of velocity difference suggests that considerable reduction in primary velocities
in upper layer may be due to the effect of side walls of main channel and shallower flow depth in
floodplains. Such decrease in primary flow velocities (velocity-dip) near the junction was also
reported by Tominaga and Nezu (1991), and described its occurrence due to secondary currents
that transport low-speed momentum from the junction edge towards the free surface. The higher
velocity in lower layer vis--vis upper layer in the main channel near the junction, clearly
signifies the influence of floodplain in modifying the flow characteristics in the main channel.

(ii) Bed shear stress


The bed shear stress is a useful parameter in hydraulic and river engineering as it governs solute
and sediment transport in open-channels. Spanwise normalised bed shear stress distribution was
calculated based on layered average simulated flow variables, see Figs. 4.9(a-b). From the
figure, it is observed that bed shear stress increases rapidly as the junction of the main channel
and floodplain is approached. Shear stress due to side walls of main channel is comparatively
large, indicating significant effect of the wall in retarding the flows at the junction edges in
floodplains. From Fig. 4.9, it is noticed that maximum bed shear stress (1.35

N/m2) has been

simulated near the junction of main channel and floodplain. The bed shear stress decreases
drastically in the main channel to 0.602

N/m2 and increases to a maximum value of 1.1

N/m2

at the mid span of the main channel. Such sudden decrease in bed shear stress from maximum
value in the floodplain to a minimum value in the main channel at the junction is responsible for
the vortex like structures at the same location. Similar bed shear stress pattern are also observed
in isolines plot shown in Fig. 4.9(b).

63

1.6

Half of Mc

Fp

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)

Test section (m)

(a) Bed shear profile

6.05
6
5.95
0

0.1

0.2

0.3

0.4

0.5

0.6

Half of channel width (m)


(b) Isolines plot
Fig. 4.9 Spanwise bed shear stress in wide symmetric compound channel (SS model)

(iii) Secondary velocities at horizontal interface


Fig. 4.10(a) shows interface flow velocities in spanwise and vertical directions along horizontal
interface between lower and upper layers of channel. Significant existence of spanwise and
vertical components of flow velocity is seen at the junctions of the main channel and the
floodplains, see Fig. 4.10(a). In present SS model, the maximum magnitude of secondary
current generated in the compound channel is of the order of 1.1% (see Fig. 4.10(c)) of its mean
bulk velocity (Ub). It is worth mentioning that secondary currents reported herein are computed
from the layer averaged velocities of the channel.
64

Velocity at Interface (m/s)

spanwise (v0)
0.004
0.003
0.002
0.001
-1E-17
-0.001
-0.002
-0.003
-0.004

Vertical (w0)
Mc

0.4

0.5
0.6
0.7
Channel width at Interface (m)
(a) Velocities profile

0.8

Test section (m)

6.05

5.95
0.4

0.5

0.6

0.7

0.8

Channel width at Interface (m)

(b) Isolines plot

(c) Vector plot


Fig. 4.10 Secondary flow velocities at horizontal interface in wide symmetric compound channel
(SS model)
65

However, in literature (Nezu and Rodi, 1986; Tominaga et al., 1989; Tominaga and Nezu, 1991;
Joung and Choi, 2008), the maximum magnitude of the secondary currents observed (point
observation) near the interface in various channels were reported in the range of 2.5-5% of Ub. It
indicates that the strength of secondary currents is under predicted in SS model. It is evident
from the figure that both velocities attain their maximum values near the junction of the main
channel and floodplain. Spanwise interface velocity shows gradual rise in magnitude from zero
to its maximum value (0.0027 m/s) near the junctions.

The rise in magnitude of the vertical interface velocity is sharp before attaining its maximum
(0.0029 m/s) value little away from junction edges. The interfacial vertical velocity is +ve at both
ends of the main channel. It is also evident from the figure (see Fig. 4.10a) that both the interface
velocities become zero at midpoint of the channel. At left end of main channel wall, interfacial
span-wise velocity is +ve while, at right end, it is ve. Isolines plot of secondary current
magnitude at interface is shown in Fig. 4.10(b). The sign of interfacial spanwise flow velocities
represents direction of secondary current at interface. It is clearly seen from the figure (see Fig.
4.10c) that secondary currents are clockwise at left end and anticlockwise at right end of main
channel. Secondary circulation pattern are opposite in nature in both half of the lower layer with
zero exactly at mid width of the channel, indicating that secondary interface velocities are
directed towards the junction of the main channel and the floodplains. Such phenomenon can
also be seen in vector plots (see Fig. 4.10c) of secondary currents, indicating the existence of
vertical interface velocity directed towards upper layer. This implies that transfer of momentum
generally occurs towards floodplains in a symmetric compound channel.

(iv) Shear stress at horizontal interface


Variation of shear stress at interface (0) of upper and lower layers is presented in Fig. 4.11(a-b).
From profile plot of interfacial shear stress (see Fig. 4.11a), it is seen that it is symmetrical about
mid section of the main channel with peaks (0.38 N/m2) occur at both ends of the interface. It is
also seen from Fig. 4.11, that there is sharp rise in magnitude of interfacial shear stress at both
ends near junctions of the main channel and the floodplains. Such sudden rise in interfacial shear
stress at the junctions is due to mass and momentum transfer taking place from lower layer to
upper layer at these regions of the channel. The similar trend is also reported in isolines plot
66

wherein densely spaced isolines near the junctions indicate sharp rise in interfacial shear stress at
the junction of the main channel and floodplain, see Fig. 4.11(b).

0.5
Mc

0 (N/m2)

0.4

0.3
0.2
0.1
0
0.4

0.5

0.6

0.7

0.8

Channel width at Interface (m)


(a) Profile plot

Test section (m)

6.05

5.95
0.4

0.45

0.5

0.55

0.6

0.65

0.7

0.75

0.8

Channel width at Interface (m)


(b) Isolines plot
Fig. 4.11 Variation of shear stress at horizontal interface in wide symmetric compound channel
(SS model)
(v) SGS turbulent shear stress in upper layer
The SGS turbulent shear stresses in upper layer were computed using layered average flow
velocities in the channel. The SGS turbulent shear stress (

; see Eq.3.40) in upper layer is

normalised with square of mean shear velocity (U*) and the resulting profile is plotted along
width of upper layer, see Fig. 4.12(a). In the figure, peak values (0.052 U*2), with opposite in
sign are simulated at the junctions in upper layer of the compound channel. This signifies

67

reversal of flow along with transfer of momentum at these regions. In rest of the upper layer
width, the value of (

) is reduced nearly equal to zero. Vortices due to circulation of

opposite nature at the junctions are also depicted in isoline plots in Fig. 4.12(b).

0.06
Fp

Mc

Fp

0.04
0.02
0
-0.02

-0.04
-0.06
0

0.2

0.4

0.6

0.8

1.2

Channel width (m)

Test section (m)

(a) Profile plot


6.05
6
5.95
0

0.2

0.4

0.6

0.8

1.2

Upper layer width (m)

(b) Isolines plot


Fig. 4.12 Variation of SGS turbulent shear stress in upper layer of wide symmetric compound
channel (SS model)
(vi) SGS Turbulent shear stress in lower layer
Similar to upper layer, normalised SGS turbulent shear stress (

; see Eq.3.40) profile and

contour plots for lower layer of the channel is shown in Fig. 4.13(a-b). These values were
calculated from the layered average flow velocities of lower layer. From Fig. 4.13(a-b), it is seen
that distribution of turbulent shear stress is same and opposite in nature about the midpoint in the
lower layer, and maximum values (0.01 U*2) are observed near junctions edges where transfer of
momentum generally takes place. Isolines plot of the SGS turbulent shear stresses in lower layer
68

(see Fig. 4.13b), confirms circulation forming at the junction and related processes of momentum
transfer.

0.02
Mc

0.01

-0.01
0.4

0.5

-0.02

0.6

0.7

0.8

Lower layer width (m)


(a) Profile plot

Test section (m)

6.05

5.95
0.4

0.5

0.6

0.7

0.8

Lower layer width (m)

(b) Isolines plot


Fig. 4.13 Variation of SGS turbulent shear stress in lower layer of wide symmetric compound
channel (SS model)
4.3.1.2 Asymmetric Compound Channel

For validation of hydrodynamic SS model developed in present study, experimental data of


asymmetric compound open-channels flow, extracted from Shiono and Feng (2003) and
Tominaga and Nezu (1991), were used. Shiono and Feng (2003) conducted laboratory
experiments for a constant discharge in an asymmetric compound channel having length to width
ratio equal to 100 (narrow channel). The experimental results of Shiono and Feng (2003) were
69

used by Shiono et al. (2003) for validation of their numerical models. Tominaga and Nezu
(1991) conducted several laboratory experiments in an asymmetric compound channel having
length to width ratio equal to 31.25 (normal channel), under various steady state discharges. Two
experimental cases, namely, S-2 and S-3 falling in the category of deep and shallow flows,
respectively are considered in present study. Their experimental results were also used by several
researchers for validation of their numerical models (Kara et al., 2012). Similar to symmetric
compound channel, firstly, grid dependence test was undertaken. Thereafter, hydrodynamic SS
model with optimal grid has been used for simulation of flow characteristics in asymmetric
compound channel used by Shiono and Feng (2003) and Tominaga and Nezu (1991). The
detailed description of such simulation is described in following paragraphs:

I. Narrow asymmetric compound channel


(a) Grid dependence test
For grid dependence, a set of runs with different aspect ratios (x/y=) 0.5, 1 and 2 for narrow
asymmetric compound channel were undertaken using SS model. The root mean square error
(RMSE) of primary flow velocities with respect to observed primary flow velocities were
calculated for each run as listed in Table 4.4.

Table 4.4 Grid cells used in narrow asymmetric compound channel for grid dependence test and
verification of SS model
Grid cells used in model domain
RMSE
Run
(Longitudinal x Transverse)
Aspect
of
Other
Channe No.
ratio
primary characteristic
Total Nos. (in
l type
flow
layer)
Spacing
(x/y) velocity
(m)
Upper
Lower
Asymmetric
(Shiono
& Feng,
2003)

2000x10

2000x5

0.01x0.02

0.5

0.03

2000x20

2000x10

0.01x0.01

1.0

0.021

2000x40

2000x20

0.01x0.005

2.0

0.023

Re = 3.7x104
Fr = 0.23

From Table 4.4, it is seen that RMSE is minimum for grid corresponding to 2000 x 20
(streamwise X spanwise) in upper layer and 2000 x 10 (streamwise X spanwise) in lower layer.
However, the grid corresponding to Run-3 (RMSE=0.023 m/s) has been used, in present study,
for further simulation to enable better resolution near the wall of the model domain. Uniform
70

grid spacing has been adopted in streamwise direction, however, non-uniform grid spacing was
adopted in spanwise direction for better capturing of the flow near the wall.

(b) Model simulation


The SS model developed in present study has been tested for asymmetric channel geometry for
flow characteristics enumerated in Table 4.1 using the computational grid cells mentioned in
sub-section of grid dependence test. Near the solid walls, both in lower and upper layers, finer
grid cells were adopted following the log-law rule. The Mannings roughness coefficeint for both
lower and upper layers were considered equal to 0.012 and Smagorinsky coefficient, C=0.1 was
adopted for the entire flow domain. Model was run for a constant discharge with total water
depth 0.11 m. Here, also, computational time step was dynamically calculated and, after attaining
of uniform flow condition, a time step was found to be of the order of 0.0019 sec for Courant
number equivalent to 0.25. Similar to symmetric compound channel, flow was treated as fully
developed when successive bed shear stress plots had converged. The simulation was further
continued for 30000 time steps to get the time averaged results.

(i) Primary flow velocity


Longitudinal flow velocity obtained from the SS model has been compared with experimental
data reported by Shiono et al. (2003), see Fig. 4.14. The depth-averaged flow velocities are
normalised with mean friction velocity (U*).
From the figure, it is apparent that computed flow velocities (RMSE for Uda=0.019 m/s) in both
the main channel and the floodplain regions follow the trends of the measured data, though some
quantitative differences remain near the junction region of the channel. Further, it is seen that
model over predicts the flows in the main channel and under predicts the same in floodplain. It is
also observed that two inflection points occur near the junction of the main channel and the
floodplain. According to Nezu et al. (2004), deep flows generally exhibit two very close
inflection points at the junction of the main channel and the floodplain and, based on such a
phenomenon, they inferred that horizontal vortices produce complicated 3-D structures near the
junction region in combination with the secondary currents.

71

Exp. (Shiono et al. 2003)

SS model

25

Uda/U*

20

15
Mc

10

Fp
RMSE=0.019 (for Uda)

5
0
0

0.05

0.1

0.15

0.2

Channel width (m)


Fig. 4.14 Comparison of primary flow velocity in narrow asymmetric compound channel (SS
model)
At junction of the main channel and the floodplain in upper layer, a sharp rise in gradient of flow
velocity is seen due to difference in primary flow velocities existing between the main channel
and the floodplains which can be explained based on calculation of percentage difference
between primary velocities of two layers with respect to mean bulk velocity of channel
, see Fig. 4.15.

From the figure it is seen that maximum difference of the order of -14% occurs near the junction
of the main channel and the floodplain. It can also be observed that velocity difference is
gradually rising from -3% at common wall of lower and upper layers up to the junction edge.
The -ve sign of velocity difference shows that primary velocity in upper layer is more than that
of lower layer throughout the channel span. Further, higher ve value of velocity difference near
the junction edge of the main channel and floodplain suggests that the effect of side wall and
floodplain on the primary velocity in upper layer is minimal. And, also, the higher velocity at
upper layer near the junction edge confirms /causes the presence of two inflection points.

72

0
Mc

-2
-4
-6
-8
-10
-12
-14

-16
0

0.02

0.04

0.06

0.08

0.1

Channel width at interface (m)


Fig. 4.15 Percentage difference between primary flow velocities of two layers with respect to
mean bulk velocity of narrow asymmetric compound channel (SS model)

While comparing the shallow flow condition (see, Fig. 4.8 for symmetric compound channel)
with deep flow condition (Fig. 4.15), clearly demonstrates that flow conditions in the floodplain
play dominant role in affecting the flow conditions in the main channel for shallow flow
conditions.

The situation is reversed for deep flow condition wherein the flow velocity in upper layer is
consistently higher from lower layer in the main channel.

(ii) Bed shear stress


Spanwise bed shear stress distribution was computed using simulated results of SS model as
shown in Figs. 4.16(a-b). From the figure, it is observed that maximum value (1.203

N/m2) of

bed shear stress occurs in the floodplain near the junction of the main channel and the floodplain.
Also, another peak of 1.198

N/m2 has been observed in the main channel. The formation of

vortex like structures are clearly seen on both sides of junction in the isolines plot of bed shear
stress, see Fig. 4.16(b).

73

1.4
1.2
1
0.8
0.6
Fp

Mc

0.4
0.2
0
0

0.05

0.1

0.15

0.2

Channel width (m)


(a) Profile plot

Test section (m)

14.05

14

13.95
0

0.05

0.1

0.15

0.2

Channel width (m)

(b) Isolines plot


Fig. 4.16 Spanwise bed shear stress in narrow asymmetric compound channel (SS model)

(iii) Secondary velocity at horizontal interface


Fig. 4.17(a) shows interface flow velocities in spanwise and vertical directions along horizontal
interface of lower and upper layers of asymmetric compound channel. In the channel, see Fig.
4.17(a), existence of spanwise and vertical components of flow velocity is seen at the junctions
of the main channel and the floodplains. In SS model, the magnitude of secondary current
74

generated in the channel is of the order of 0.015% (see Fig. 4.17c) of its mean bulk flow velocity
(Ub). It is worth mentioning that secondary currents reported herein are computed from the layer
averaged velocities of the channel. It is seen that, in the channel, the strength of secondary
currents is highly under predicted by the model. It represents only the qualitative value of the
currents.

Vertical (w0)

5E-05
Mc
4E-05
3E-05
2E-05
1E-05
0

0.02

0.04

0.06

0.08

Channel width at Interface (m)


(a) Velocities profile
14.05

Test section (m)

Velocity at Interface (m/s)

Spanwise (v0)

14

13.95
0

0.02

0.04

0.06

0.08

Channel width at Interface (m)

(b) Isolines plot


75

0.1

0.1

(c) Vector plot


Fig. 4.17 Secondary flow velocities at horizontal interface of narrow asymmetric compound
channel (SS model)
However, it is evident from the figure that both the velocities attain their maximum values near
the junction region. Spanwise interface velocity shows gradual rise in magnitude before attaining
its maximum value (0.00003 m/s) very close to the junction. The rise in magnitude of the vertical
interface velocity is gradual before attaining its maximum (0.000039 m/s) little away from
junction point. It is also evident from the figure (see Fig. 4.17a) that both the interface velocities
become zero at left end of main channel. At junction, both the interfacial velocities are +ve in
nature which represents direction of secondary current at interface. It is clearly seen from the
figure (see Fig. 4.17c) that secondary currents are anticlockwise at the junction of the main
channel and floodplain. From Fig. 4.17(b), also, it is inferred that spanwise velocity at interface
in the channel increases from zero to its maximum value at the junction of the channel. However,
vertical interface velocity, in such case, is having negligibly smaller value near the far end of the
channel and rises to maximum value near the junction of the channel. Such phenomenon can also
be seen in vector plots (see Fig. 4.17c) of secondary currents, indicating the existence of vertical
interface velocity directed towards upper layer.

76

(iv) Shear stress at horizontal interface


Variation of shear stress at interface (0) between upper and lower layers is presented in Figs.
4.18(a-b).

0.4
Mc
0 (N/m2)

0.3
0.2
0.1
0

0.02

0.04

0.06

0.08

0.1

Channel width at Interface (m)


(a) Profile plot

Test section (m)

14.05

14

13.95
0

0.02

0.04

0.06

0.08

0.1

Channel width at Interface (m)

(b) Isolines plot


Fig. 4.18 Variation of shear stress at horizontal interface of narrow asymmetric compound
channel (SS model)
77

From profile plot of interfacial shear stress (see Fig. 4.18a), it is seen that peak value (0.33 N/m2)
occurs at the right end of interface near the junction of the main channel and the floodplain. It is
seen that there is gradual rise in its magnitude from left end of interface to the junction edge.
However, as junction is approached, the rise in interfacial shear is sharp. Such behaviour of rise
towards the junction is due to existence of momentum transfer from lower layer to upper layer at
foregoing locations of the channel. Further, formation of densely spaced isolines at the ends of
interface, near the junction, can be clearly seen in contour plot of interface shear stress (see Fig.
4.18b).

(v) SGS turbulent shear in upper layer


The SGS turbulent shear stress in upper layer (

; see Eq.3.40) is normalised with square of

shear velocity (U*) and the resulting profile across width of upper layer is shown in Fig. 4.19(a).
From the figure, maximum value (0.073 U*2 N/m2) of turbulent shear stress can be observed near
left wall of upper layer in the main channel whereas, at junction of the main channel and the
floodplain, negligibly small value is observed.

0.08

Fp

Mc

0.06
0.04
0.02
0
-0.02
-0.04
-0.06
-0.08
0

0.05

0.1
Channel width (m)
(a) Profile plot

78

0.15

0.2

Test section (m)

14.05

14

13.95
0

0.04

0.08

0.12

0.16

0.2

Upper layer width (m)

(b) Isolines plot


Fig. 4.19 Variation of SGS turbulent shear stress in upper layer of narrow asymmetric compound
channel (SS model)

Apart from this, near the right wall of upper layer in floodplain, a shear stress value of the order
of 0.058U*2 N/m2 has been simulated in the SS model. It is also seen that nature of shear stresses
at both the ends are opposite to each other, see isolines plot of turbulent shear stress in Fig.
4.19(b).

(vi) SGS turbulent shear in lower layer


Similar to upper layer, normalised SGS turbulent shear stress (

; see Eq.3.40) profile and

contour plots for lower layer of the channel is shown in Fig. 4.20(a-b).
From the shear stress profile (Fig. 4.20a), it is seen that a peak shear stress value (0.07 U*2 N/m2)
occurs near the left wall of lower layer whereas, second peak value (0.04U*2 N/m2), opposite in
nature, is simulated near the junction of the main channel and the floodplain where transfer of
momentum is affected. In isolines plot (see Fig. 4.20b), circulation formed at both the ends of
lower layer is seen whereas, in upper layer plot (see Fig. 4.20b), such formation is absent at the
junction of the main channel and floodplain. The model under predicts the magnitude of shear
stresses, however, qualitatively it represents flow features in the compound channel.

79

0.08

Mc

0.06
0.04
0.02
0
-0.02
-0.04

-0.06
0

0.02

0.04

0.06

0.08

0.1

Lower layer width (m)


(a) Profile plot

Test section (m)

14.05

14

13.95
0

0.02

0.04

0.06

0.08

0.1

Lower layer width (m)


(b) Isolines plot
Fig. 4.20 Variation of SGS turbulent shear stress in lower layer of narrow asymmetric compound
channel (SS model)

80

II. Normal asymmetric compound channel


The flow characteristics in asymmetric compound channel having length to width ratio equal to
31.2 (normal channel) as reported by Tominaga and Nezu (1991) (see Table 4.1) were simulated
using developed 2-D two layered model proposed in present study. The description of simulated
key flow variables is included in following paragraphs:

(a) Grid dependence test


The grid dependence test was undertaken with a set of runs with different aspect ratios (x/y=)
1.0, 1.3 and 2.0 for normal asymmetric compound channel. The root mean square error (RMSE)
of primary flow velocity with respect to observed primary flow velocity were calculated for each
run, see Table 4.5.
Table 4.5 Grid dependence test for normal asymmetric compound channel under SS model

Channel
type

Asymmetric

Run
No.

Grid cells used in model domain


(Longitudinal x Transverse)
Total Nos. (in
Spacing
layer)
(m)
Upper
Lower

Aspect
ratio

RMSE of
primary flow
velocity

(x/y)

(m/s)

625x25

625x13

0.02x0.02

1.0

0.01

625x31

625x16

0.02x0.015

1.3

0.009

625x48

625x24

0.02x0.01

2.0

0.008

For simulation of asymmetric compound channel, number of grid cells under Run No.3 with
RMSE=0.008 m/s were adopted wherein 625x48 grid cells in upper layer and 625x24 grid cells
in lower layer were used in streamwise x spanwise directions respectively. It is to be noted that,
in streamwise direction, uniform grid spacing was selected whereas in spanwise direction nonuniform grid spacing was adopted in upper and lower layers of both the channels.

(b) Sensitivity analysis


Sensitivity of coefficients C and (Alpha) of developed SS model were examined. Both the
parameters were varied to their limits and five combinations were formed (see Table 4.3).

81

Similar flow conditions and channel geometries of the channel were adopted for each
combination. The primary flow velocity profile across width of asymmetric compound channel
obtained for each combination is shown in Figs. 4.21(a-b).

Primary flow velocity (m/s)

C=0.1
0.5

C=0.2

Mc

C=0.065
Fp

0.4
0.3
0.2

0.1
0
0

0.1

0.2

0.3

0.4

Channel width (m)


Fig. 4.21 (a) Variation of primary flow velocity of normal asymmetric compound channel
(Exp.S-2: SS model) with C (Alpha, =0.5)

Primary flow velocity (m/s)

Alpha=0.5

Alpha=1.0

Alpha=0.2

0.5
Mc

Fp

0.4
0.3
0.2
0.1
0
0

0.1

0.2

0.3

0.4

Channel width (m)


Fig. 4.21 (b) Variation of primary flow velocity of normal asymmetric compound channel
(Exp.S-2: SS model) with Alpha, (C=0.1)

82

From Fig. 4.21(a), for =0.5 and C varying from 0.065 to 0.2, it is seen that there is a little
variation exists in simulated primary flow velocity in the region of floodplain as well as in the
vicinity of walls, whereas in Fig. 4.21(b), for C=0.1 and varying from 0.2 to 1.0, significant
variation is observed in main channel and floodplain of the channel. However, in rest of the
channel, negligibly small effect can be observed due to variation in aforesaid parameters. In
further analysis, an average value of C=0.1 and =0.5 have been taken to simulate the flow
parameters.

(c) Model verification


The SS model developed in present study was tested for normal asymmetric compound channel,
see Table 4.1, using 625 computational grid cells in streamwise direction and, 24 and 48
computational grid cells, in spanwise direction for lower and upper layers respectively. Near
solid walls, both in lower and upper layers, finer grid cells were adopted considering the limiting
criteria (y+>30) based on log-law rule. For deep flows, i.e., Exp.S-2, y+=41 and for shallow
flows, i.e., Exp.S-3, y+=35 is adopted following Eq. (3.61).
Mannings roughness coefficient for deep flows for both lower and upper layers was considered
as 0.01 whereas for shallow flows 0.01 and 0.013 were adopted for lower and upper layers
respectively. The Smagorinsky coefficient, C=0.1 was adopted for entire model domain for both
the cases. The time step was found to be of the order of 0.002 sec for Courant number equivalent
to 0.25 for both the cases. Flow was treated as fully developed when successive bed shear stress
plots had converged (at approximately 50h/U* seconds). Here, h/U* represents eddy turn-over
time (Kara et al., 2012). The simulation was further continued for 12h/U* seconds to get time
averaged results which are presented in following paragraphs.

(i) Primary flow velocity


The simulated longitudinal flow velocities using SS model have been compared with
experimental results of Exp.S-2 and Exp.S-3 (Tominaga and Nezu, 1991), see Fig. 4.22(a-b). The
depth-averaged flow velocities, here, are normalised with shear velocity (U*).

83

From Fig. 4.22, it is observed that velocities are reasonably in good agreement (RMSE for
Uda=0.019 m/s and 0.017 m/s for deep and shallow flows respectively) except near the junction
of the main channel and the floodplain. In deep flows, two inflection points are simulated, see
Fig. 4.22(a), whereas in shallow flows, only one inflection point occurs near the junction of the
main channel and the floodplain, see Fig. 4.22(b).

Exp.S-2 (Tominaga & Nazu 1991)


30

SS model

Mc

Fp

Uda/U*

25
20
15
10
RMSE for Uda =0.019 m/s

5
0
0

0.1

0.2

0.3

0.4

Channel width (m)


(a) Deep flows (Exp.S-2)

Exp.S-3 (Tominaga & Nezu 1991)


30

Mc

Uda/U*

25

SS model
Fp

RMSE for Uda =0.017 m/s

20
15
10
5
0
0

0.1

0.2
Channel width (m)

0.3

0.4

(b) Shallow flows (Exp.S-3)


Fig. 4.22 Comparison of primary flow velocity in normal asymmetric compound channel (SS
model)
84

For both the cases, at junction of the main channel and floodplain, a rise in gradient of flow
velocity is seen due to significant difference in primary flow velocities existing between the main
channel and floodplain which can be explained based on calculation of percentage difference
between primary velocities of two layers with respect to mean bulk velocity of channel
, see Fig. 4.23(a-b).

0
Mc
-2
-4
-6
-8
-10
0

0.05

0.1

0.15

0.2

Channel width at interface (m)


(a) Deep flows (Exp.S-2)

6
Mc

4
2
0
-2
-4

-6
0

0.05

0.1

0.15

0.2

Channel width at interface (m)


(b) Shallow flows (Exp.S-3)
Fig. 4.23 Percentage difference between primary flow velocities of two layers with respect to
mean bulk velocity of normal asymmetric compound channel (SS model)
85

From Fig. 4.23(a-b), it is seen that the maximum difference has been found to be of the order of
-9% for deep flows (Fig. 4.23a) and +4% for shallow flows (Fig. 4.23b) near the junction of the
main channel and the floodplain. In deep flows (Exp.S-2, Fig. 4.23a) condition, the primary flow
velocity in upper layer has been found higher than lower layer throughout the main channel.
However, such difference is increased at the junction of main channel and floodplain. The -ve
sign of velocity difference shows that primary velocity in upper layer is more than lower layer.
Further, higher ve value of velocity difference near the junction edge of the main channel and
floodplain suggests that the effect of side wall and floodplain on the primary velocity in upper
layer is minimal. Also, the higher velocity in upper layer near the junction edge confirms the
presence of two inflection points. On the other hand, in shallow flows (Exp.S-3, Fig. 4.23b)
condition, the primary velocity in upper layer is higher than lower layer at left end, and, the trend
is getting reversed at the junction edge of the main channel and floodplain. In other words, the
effect of floodplain is dominant in affecting the flow characteristics of main channel in shallow
flow condition vis--vis deep flow condition.

(ii) Bed shear stress


Spanwise bed shear stresses distribution were computed for deep and shallow flows cases using
simulated results of SS model as shown in Figs. 4.24(i) and 4.24(ii). Spanwise bed shear stress is
non-dimensionalised with respect to its average value ( ), and compared with experimental data
(Tominaga and Nezu, 1991), see Figs. 4.24 (i-a) and 4.24(ii-a).

Exp.S-2 (Tominaga & Nezu 1991)

SS model

1.6
1.2
0.8
Mc

Fp

0.4

RMSE for b=0.027 N/m2

0
0

0.1

0.2
Channel width (m)

(a) Bed shear stresses profile


86

0.3

0.4

Test section (m)

7.55

7.5

7.45
0

0.1

0.2

0.3

0.4

Channel width (m)

(b) Contour plot


(i) Deep flows case (Exp.S-2)

Exp.S-3 (Tominaga & Nezu 1991)

SS model

1.6
RMSE for b=0.027 N/m2

1.2
0.8
0.4

Mc

Fp

0
0

0.1

0.2
Channel width (m)

0.3

0.4

(a) Bed shear stresses profile


Test section (m)

7.55

7.5

7.45
0

0.1

0.2

0.3

0.4

Channel width (m)

(b) Contour plot


(ii) Shallow flows case (Exp.S-3)
Fig. 4.24 Comparison of bed shear stresses in normal asymmetric compound channel (SS model)

87

From Figs. 4.24(i) and 4.24 (ii), it is seen that bed shear stresses are in general agreement
(RMSE for

=0.027 N/m2 for both deep and shallow flows) with experimental results except at

few places in the main channel where it is over-predicted the experimental values. The minor
differences between experimental and numerical results can be attributed to the layered average
assumption being applicable in the main channel. Further, in deep flows case, peak value (1.2 )
is observed at two places across the channel width, i.e., at the junction of the main channel and
the floodplain, and at mid section of the main channel. From the peak, its value decreases
towards the walls of both main channel and floodplain, and attains a minimum value of the order
of 0.6 . In shallow flows, the peak (1.55 ) occurs at mid section of the main channel, and a
minimum value of the order of 0.5

is found at walls of floodplain. Figs. 4.24(i-b) and 4.24(ii-

b) show isolines of bed shear stresses at test section of the channel in which formation of vortex
like structures are clearly visible near the junction of the main channel and the floodplain.

(iii) Secondary velocities at horizontal interface


Figs. 4.25(i-a) and 4.25(ii-a) show interface flow velocities in spanwise and vertical directions
across horizontal interface between lower and upper layers of the compound channel for deep
and shallow flow conditions respectively.

Velocity at Interface (m/s)

Spanwise (v0)

Vertical (w0)

0.0004
Mc

0.0003
0.0002
0.0001
0
0

0.05

0.1

0.15

Channel width at Interface (m)


(a) Secondary velocity profile

88

0.2

Test section (m)

7.55

7.5

7.45
0

0.05

0.1
Channel width at Interface (m)

(b) Isolines plot

(c) Vector plot


(i) Deep flows (Exp.S-2)

89

0.15

0.2

Velocity at Interface (m/s)

Spanwise (v0)

0.0008

Vertical (w0)

Mc

0.0006
0.0004
0.0002
0
0

0.05

0.1

0.15

0.2

Channel width at Interface (m)

(a) Secondary velocity profile

Test section (m)

7.55

7.5

7.45
0

0.05

0.1
Channel width at Interface (m)

(b) Isolines plot

90

0.15

0.2

(b) Vector plot


(ii) Shallow flows (Exp.S-3)
Fig. 4.25 Secondary velocity at horizontal interface in normal asymmetric compound channel
(SS model)
From Figs. 4.25(i-a) and 4.25(ii-a), it is observed that spanwise and vertical components of
interface velocity are prominent at the junctions of the main channel and floodplain. The
magnitude of secondary current generated in the channel is of the order of 0.12% of Ub (for deep
flows, Fig. 4.25(i-c)) and 0.23% of Ub (for shallow flows, Fig. 4.25(ii-c)). It is worth mentioning
that secondary currents, though under predicted, are based on calculation from layer averaged
velocities of the channel. Thus, it represents only qualitative value of the currents across the
interface. From Figs. 4.25(i) and 4.25(ii), it is evident that both the interfacial velocities attain
their peak near the junction of the main channel and floodplain. Spanwise interface velocity
shows gradual rise in magnitude before attaining its maximum value (0.00026 m/s for deep flows
and 0.0007 m/s for shallow flows) at the junction edge. The rise in magnitude of vertical
interface velocity is gradual up to 0.15 m of channel width but rises sharply before attaining its
maximum (0.00035 m/s for deep flows and 0.00039 m/s for shallow flows) little away from the
junction edge. At the junction, both the interfacial velocities are +ve in nature which represents
91

direction of secondary current at interface. It is clearly seen from Figs. 4.25(i-c) and 4.25(ii-c)
that secondary currents are anticlockwise at right end of the main channel which follows the
trends reported by Tominaga and Nezu (1991). The variation of secondary currents across the
width of main channel at horizontal interface is also indicated in their respective isolines plots,
i.e., Figs. 4.25(i-b) and 4.25(ii-b) respectively. From Figs. 4.25(i-a) and 4.25(ii-a), it is clearly
seen that spanwise interfacial velocity in shallow flow condition is higher than deep flow
condition, particularly at junction of the main channel and floodplain.

(iv) Shear stress at horizontal interface


Variation of shear stress at interface (0) between the upper and the lower layers for shallow and
deep flows are depicted in Figs. 4.26(i) and 4.26(ii). From Figs. 4.26(i-a) and 4.26(ii-a), it is seen
that peak value (0.32 N/m2 for deep flows and 0.08 N/m2 for shallow flows) occurs at junction
edge of the main channel and floodplain. It is also seen that it gradually rises from left wall of the
main channel to the junction of the main channel and the floodplain. But, in case of shallow
flows, near the junction, the interface shear stress shows sudden fall and then sudden rise in its
value. The variation of interfacial shear stresses, as described above, is due to the transfer of
mass and momentum occurring from lower to upper layer at these locations of the channel. The
sudden rise in interfacial shear stress at the junction is also corroborated in isolines plots of
densely spaced isolines near the junction, see Figs. 4.26(i-b) and 4.26(ii-b).

0.4
Mc
0 (N/m2)

0.3
0.2
0.1

0
0

0.05

0.1

0.15

Channel width at Interface (m)


(a) Interfacial shear stresses profile
92

0.2

Test section (m)

7.55

7.5

7.45
0

0.05

0.1

0.15

0.2

Channel width at Interface (m)

(b) Isolines plot


(i) Deep flows (Exp.S-2)

0.1
Mc

0 (N/m2)

0.08
0.06
0.04
0.02
0
0

0.05

0.1
Interface width (m)

(a) Interfacial shear stresses profile

93

0.15

0.2

Test section (m)

7.55

7.5

7.45
0

0.05

0.1

0.15

0.2

Channel width at Interface (m)

(b) Isolines plot


(ii) Shallow flows (Exp.S-3)
Fig. 4.26 Variation of shear stresses at horizontal interface in normal asymmetric compound
channel (SS model)
(v) SGS turbulent shear stress in upper layer
The SGS turbulent shear stress in upper layer (

; see Eq.3.40) is normalised with square of

shear velocity (U*) and the resulting profile plot across upper layer width is shown in Figs.
4.27(i-ii).
0.015
Fp

Mc
0.01
0.005
0
-0.005

-0.01
0

0.1
0.2
0.3
Upper layer width (m)
(a) Profile plot
94

0.4

Test section (m)

7.55

7.5

7.45
0

0.1

0.2

0.3

0.4

Upper layer width (m)

(b) Isolines plot


(i) Deep flows case (Exp.S-2)

0.02
0.01
0
-0.01
-0.02
-0.03
-0.04
-0.05

Mc

0.1

Fp

0.2
0.3
Upper layer width (m)

0.4

(a) Profile plot


Test section (m)

7.55

7.5

7.45
0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

Upper layer width (m)

(b) Isloines plot


(ii) Shallow flows case (Exp.S-3)
Fig. 4.27 Variation of SGS turbulent shear stress in upper layer of normal asymmetric compound
channel (SS model)
From Figs. 4.27(i) and 4.27(ii), the peak value for deep flows (+0.012 U*2) occurs near the left
wall of upper layer in the main channel whereas, in shallow flows case, peak value (-0.045 U*2)
95

occurs at the junction edge. The variation of SGS shear stresses in upper layer clearly signifies
the effect of walls in modifying the pattern of shear stresses in the channel. The isolines plots
(Figs. 4.27(i-b) and 4.27(ii-b)) also indicate the existence of circulation at the junction and both
ends of upper layer. The significant sudden drop in shear stress at the junction of shallow flow
clearly signifies the major influence of floodplain in flow characteristics in the main channel and
transfer of momentum at the junction of the main channel and floodplain.

(vi) SGS turbulent shear stress in lower layer


Similar to upper layer, normalised SGS turbulent shear stress (

; see Eq.3.40) profile and

isolines plots for lower layer of the channel is shown in Figs. 4.28(i-ii).
0.015
Mc

0.01
0.005
0
-0.005
-0.01
-0.015
0

0.05
0.1
0.15
Lower layer width (m)

0.2

(a) Profile plot

Test section (m)

7.55

7.5

7.45
0

0.05

0.1

Lower layer width (m)

(b) Isolines plot


(i) Deep flows case (Exp.S-2)
96

0.15

0.2

0.04
0.03
0.02
0.01
0
-0.01
-0.02
-0.03
-0.04

Mc

0.05
0.1
0.15
Lower layer width (m)

0.2

(a) Profile plot

Test section (m)

7.55

7.5

7.45
0

0.05

0.1

0.15

0.2

Lower layer width (m)

(b) Isolines plot


(ii) Shallow flows case (Exp.S-3)
Fig. 4.28 Variation of SGS turbulent shear stress in lower layer of normal asymmetric compound
(SS model)
From Figs. 4.28(i-a), it is seen that first peak (0.011 U*2) occurs near the left wall of lower layer
whereas second peak (-0.01U*2), opposite in nature, is seen near the junction of the main channel
and floodplain for deep flows case. In shallow flows, one peak value (-0.03 U*2) occurs near the
junction of the main channel and floodplain where transfer of momentum generally takes place.
The second peak (0.01U*2), opposite in nature, occurs near the left wall of lower layer in the
main channel due to existence of the wall. In isolines plots (see Fig. 4.28(i-b) and 4.28(ii-b)),
97

circulations formed at both ends of lower layer confirms processes of mass and momentum
transfer from these regions of the channel.

4.3.2 Dynamic Subgrid Scale Model (DSGS model)

The hydrodynamic model developed with dynamic SGS turbulence closure has been validated
using the data of both types of compound channels, i.e., deep and shallow flows. The same sets
of compound channel data are used in validation of DSGS model as were used in Section 4.3.1
for validation of SS model. In all the cases, same computational grid cells are adopted.

In standard Smagorinsky (SS) model, constant value of C was applied for the whole flow domain
and in order to model the flows near the solid walls, damping factor was used. However, in LES
several potential difficulties do occur in case of shear flows and flow near the solid walls due to
selection of a constant Smagorinsky coefficient. A dynamic model for C, as proposed by
Germano et al. (1991) to account for asymptotic behaviour of the subgrid scale stresses existing
near the walls, has been used in DSGS model. In dynamic model, coefficient C is allowed to
vary locally to adjust eddy viscosity in the flow field.

4.3.2.1 Symmetric compound channel


(i) Spanwise variation of model parameter C
Figure 4.29 shows spanwise variation of depth averaged model parameter C in symmetric
compound channel flow (Fraselle et al, 2008) which was determined locally during the model
simulation while adjusting the level of eddy viscosity in the flow field.

From the figure, it is seen that, at sidewalls of the channel, the value of C reduces to minimum as
eddy viscosity is minimal close to the walls due to occurrence of anisotropy turbulence (Rodi,
1980). Further, away from the wall in the floodplain (Fp), its value rises steeply in symmetric
channel reaching a maximum value (0.138), and then, decreases slowly before attaining its
minimum value (0.02) near the wall of lower layer.

98

0.16

Fp

Half of Mc

0.2
0.3
0.4
Half of channel width (m)

0.5

0.12
0.08
0.04
0
0

0.1

0.6

Fig. 4.29 Variation of SGS coefficient across wide symmetric compound channel (DSGS model)
In the main channel of symmetric compound channel, the value of model parameter rises again,
and regains to its second peak at mid of the channel. Being a symmetric compound channel,
variation of the model parameter would also be symmetric. The parameter varies within the
range of 0.02 to 0.138 throughout the channel during the model simulations (Fig. 4.45) after
attaining steady state flow condition. The variation of model parameter in both parts of the
channel is in the recommended range of Rodi (1980), i.e., C 0.1 . Thus, it is seen in the model
simulation that variation of model parameter C is highly dependent on grid spacing of the flow
domain. The variation of model parameter C across the width of compound channel, underlines
the importance of taking dynamic SGS parameter instead of taking a constant value throughout
the simulation.
The depth averaged flow characteristics in the main channel and floodplain of corresponding
channels using DSGS model are described in flowing paragraphs:

(ii) Primary flow velocity


Longitudinal flow velocity obtained using developed DSGS model has been compared using
experimental data of Fraselle et al. (2010) as shown in Fig. 4.30. The depth-averaged flow
velocities were normalised using shear flow velocity (U*). From the figure, it is observed that
velocities obtained from present model are reasonably in good agreement (RMSE for Uda=0.024
m/s) with experimental data except in floodplains near the junction of the main channel and the

99

floodplain where model under predicted flow velocities. It is also observed from the figure that
due to shallow flows in channel only one inflection point occurs near the junction of the main
channel and the floodplain.

Exp. (Fraselle et al. 2010)

30

RMSE for Uda=0.024 m/s

25
Uda/U*

DSGS model

20
15

10
Half of Mc

Fp

5
0
0

0.1

0.2

0.3

0.4

0.5

0.6

Half of channel width (m)


Fig. 4.30 Comparison of primary flow velocity in wide symmetric compound channel (DSGS
model)
16
12

8
4
0
-4
0.4

0.5

0.6

0.7

0.8

Channel width at interface (m)


Fig. 4.31 Percentage difference between primary flow velocities of two layers with respect to
mean bulk velocity of wide symmetric compound channel (DSGS model)
At junction of the main channel and the floodplain in upper layer, a sharp rise in gradient of flow
velocity is seen due to significant difference in primary flow velocities existing between the main
channel and the floodplains which can be explained based on calculation of percentage
100

difference between primary velocities of two layers with respect to mean bulk velocity of
channel

, see Fig. 4.31.

From the figure, it is seen that the difference is found to be similar to that of obtained in SS
model case. The pattern of primary velocity prediction from DSGS model is similar to SS model
(see Fig. 4.6), however, former predicts the flow characteristics slightly better than latter model.

(iii) Bed shear stress


Bed shear stress is useful hydraulic parameter in river engineering as it governs the solute
transport in open-channels. Spanwise bed shear stress distribution using DSGS model was
calculated based on layered average simulated flow variables as is shown in Figs. 4.32(a-b).
From the figure, it is observed that, on floodplain, bed shear stress increases rapidly (1.37
N/m2) as the junction of the main channel and the floodplain is approached. Further, it is seen
that the variation of simulated bed shear stress across the channel width using DSGS model is
marginally more in comparison to that obtained in SS model case. The side walls of the main
channel have considerable influence on shear stress distribution as they significantly retards the
flow in the floodplain. Similar, trend was also reported earlier by Joung and Choi (2008).

1.6
Fp

1.4

Half of Mc

1.2
1
0.8
0.6
0.4

0.2
0

0.1

0.2

0.3

0.4

Half of channel width (m)


(a) Bed shear profile

101

0.5

0.6

Test section (m)

6.05
6
5.95
0

0.1

0.2

0.3

0.4

0.5

0.6

Channel width (m)


(b) Isolines plot
Fig. 4.32 Variation of spanwise bed shear stress in wide symmetric compound channel (DSGS
model)
The bed shear stress decreases drastically in the main channel at junction of main channel and
floodplain, and, then, increases gradually to attain another peak (1.085
the main channel. Such drastic variation (from 0.6

N/m2 to 1.37

N/m2) at midpoint of
N/m2) in the bed shear

stresses at junction edge is due to formation of vortex like structure as can be seen clearly in
isolines plot of Fig. 4.32(b).

(iv) Secondary velocities at horizontal interface


Fig. 4.33(a) shows interface flow velocities in spanwise and vertical directions along horizontal
interface of lower and upper layers of the channel. Similar to SS model, existence of significant
vertical component of flow velocity is seen at the junctions of the main channel and the
floodplains (see Fig. 4.33a). Spanwise interface velocity shows gradual rise in magnitude before
attaining its maximum value (0.0027 m/s) very close to the junction. The rise in magnitude of the
vertical interface velocity is sharp before attaining its maximum (0.0035 m/s) little away from
junction point. It is clearly seen from the figure (see Fig. 4.33c) that generated secondary
currents are clockwise at left end and anticlockwise at right end of the main channel. From Fig.
4.33(b), similar to SS model, it can be inferred that spanwise velocity at interface in the channel
increases from zero to its maximum value at the junction of the channel. On other hand, vertical
interfacial velocity is having negligibly smaller value near the middle reach of the channel and
rises to maximum values near the junctions of the main channel and floodplain, see Fig. 4.33.
The vertical velocity component in DSGS model (Fig. 4.33a) is slightly higher than
corresponding values in SS model (see Fig. 4.10a).

102

Velocity at Interface (m/s)

Spanwise (v0)

Vertical (w0)

0.004
0.003
0.002
0.001
0
-0.001
-0.002
-0.003
-0.004

Mc

0.4

0.5

0.6

0.7

0.8

Channel width at Interface (m)


(a) Profile plot
Test section (m)

6.05

5.95
0.4

0.5

0.6

0.7

0.8

Channel width at Interface (m)

(b) Isolines plot

(c) Vector plot


Fig. 4.33 Secondary flow velocities at horizontal interface of wide symmetric compound channel
(DSGS model)
103

(v) Shear stress at horizontal Interface


Variation of shear stress at interface (0) of upper and lower layers, similar to SS model, is
presented in Figs. 4.34(a-b). From profile plot of interfacial shear stress (see Fig. 4.34a),
symmetrical behaviour of shear stresses distribution is seen about the midpoint of main channel
with maximum value (0.38 N/m2) occurs at the ends of interface. A sharp rise in magnitude of
interfacial shear stress is reported at the junction of main channel and floodplain. Existence of
large shear stresses at the junctions is due to presence of significant circular pattern in aforesaid
locations. Such trend is also reported in isolines plots of interfacial shear stresses wherein
densely spaced isolines are seen in the region of junction of main channel and floodplain, see
Fig. 4.34(b).
0.5
Mc

0 (N/m2)

0.4
0.3
0.2
0.1
0
0.4

0.5

0.6

0.7

0.8

Channel width at Inteface (m)


(a) Profile plot
Test section (m)

6.05

5.95
0.4

0.5

0.6

0.7

0.8

Channel width at Interface (m)

(b) Isolines plot


Fig. 4.34 Variation of shear stress at horizontal interface of wide symmetric compound channel
(DSGS model)
104

(vi) SGS turbulent shear at upper layer


The variation of SGS turbulent shear stresses at upper layer (

; see Eq.3.40), normalised

with square of shear velocity (U*), is shown in Fig. 4.35(a). From the figure, it is clearly seen
that maximum values (0.05 U*2) of SGS shear stresses occur at the junction of main channel and
floodplain. The SGS shear stresses at both the junctions and floodplains are opposite in nature
due to transfer of momentum across the floodplains from the main channel.

0.06
Fp

Mc

Fp

0.2

0.4
0.6
0.8
Upper layer width (m)

0.04
0.02
0
-0.02
-0.04
-0.06
0

1.2

Test section (m)

(a) Profile plot

6.05
6
5.95
0

0.2

0.4

0.6

0.8

1.2

Upper layer width (m)

(b) Isolines plot


Fig. 4.35 Variation of SGS turbulent shear stress in upper layer of wide symmetric compound
channel (DSGS model)

The isolines plot (Fig. 4.35b) also depicts the formation of vortices due to existence of opposite
circulation pattern at both junctions of the symmetric compound channel. The variation of SGS
105

turbulence shear stresses across upper layer simulated from SS model (Fig. 4.12a) and DSGS
model (Fig. 4.35a) is due to the turbulence coefficient C varying across the channel width.

(vii) SGS Turbulent shear stress in Lower layer


Similar to upper layer, normalised SGS turbulent shear stresses for lower layer (

; see

Eq.3.40) was also computed and plotted in Fig. 4.36(a-b).

0.02

Mc

0.01
0
-0.01
-0.02
0.4

0.5

0.6
0.7
Lower layer width (m)

0.8

(a) Profile plot


Test section (m)

6.05

5.95
0.4

0.5

0.6

0.7

0.8

Lower layer width (m)

(b) Isolines plot


Fig. 4.36 Variation of SGS turbulent shear stress in lower layer of wide symmetric compound
channel (DSGS model)

From the figure, it is seen that distribution of turbulent shear stress is symmetrical about the
midpoint of the lower layer, and maximum values (0.015U*2), opposite in nature, are observed
near junctions of the main channel and floodplain where transfer of momentum in the floodplains
106

are affected. The isolines plot (Fig. 4.36b) confirms the formation of circulation pattern at either
end of lower layer wherein transfer of momentum takes place from main channel to floodplain.
In comparison to SS model (see Fig. 4.13a), SGS turbulent shear stress at lower layer was
simulated marginally higher in DSGS model (see Fig. 4.36b) due to variation of turbulence
parameter C across the channel width.
4.3.2.2 Asymmetric compound channel
(i) Spanwise variation of model parameter C
Fig. (4.37) shows spanwise variation of model parameter, C, in asymmetric compound channel
flow. The values of C were determined locally during the model simulation while adjusting the
level of eddy viscosity in flow field.

0.16

Mc

Fp

0.12
0.08
0.04
0
0

0.05

0.1

0.15

0.2

Channel width (m)


Fig. 4.37 SGS coefficient profile in narrow asymmetric compound channel (DSGS model)
From the figures, it is seen that, at sidewall of the channel, the value of C reduces to its minimal
value of 0.028, and, then, increases away from the wall to attain a peak of 0.145 in the main
channel. The value of C reduces rapidly to a lower value (0.028) near the junction of the main
channel and floodplain. In floodplain, it again rises to attain second peak of the order of 0.08.
The spanwise variation of C has been found to be in the range of 0.028 to 0.145. The variation of
model parameter C across the width of compound channel, underlines the importance of taking
dynamic SGS parameter instead of taking a constant value throughout the simulation.
107

(ii) Primary flow velocity (Uda)


Longitudinal flow velocity obtained from DSGS model in asymmetric compound channel is
compared with experimental data of Shiono and Feng (2003), see Fig. 4.38. In Fig. 4.38, depthaveraged flow velocities are normalised with mean friction velocity (U*).

Exp.(Shiono et al. 2003)

DSGS model

25
Mc

Fp

Uda/U*

20
15
10
RMSE for Uda=0.019 m/s

5
0
0

0.05

0.1
Channel width (m)

0.15

0.2

Fig. 4.38 Comparison of primary flow velocity in narrow asymmetric compound channel (DSGS
model)
0
Mc

-2
-4

-6
-8
-10
-12

-14
0

0.02

0.04

0.06

0.08

0.1

Asymmetric channel width at Interface (m)


Fig. 4.39 Percentage difference between primary flow velocities of two layers with respect to
mean bulk velocity of narrow asymmetric compound channel (DSGS model)

108

From the figure, it is apparent that computed flow velocities (RMSE for Uda=0.019 m/s) in both
main channel and floodplain follow the trend of the measured data, though slight differences
between simulated and experimental values remain near the junction of the channel. Similar to
SS model, it is seen that model over predicts the flows in the main channel and under predicts the
same in the floodplain.

The depth-averaged primary velocities in lower layer (ul) and upper layer (uu) were computed
separately. The differences between ul and uu, expressed as

, have been found

significant at the junction of the main channel and floodplain, see Fig. 4.39. Here, Ub represents
mean bulk velocity in the channel. From the figure, it is seen that the differences between ul and
uu have been found to be of the order of 13% near the junction of the main channel and the
floodplain. Similar order of differences between the velocities, ul and uu, have been observed in
SS model, also.

(iii) Bed shear stress


The bed shear stress distribution across the asymmetric compound channel width was obtained
from simulated results of DSGS model, see Fig. 4.40.

1.6

Mc

Fp

1.4
1.2
1
0.8
0.6
0.4
0.2
0
0

0.05

0.1
Channel width (m)

(a) Bed shear profile


109

0.15

0.2

Test section (m)

14.05

14

13.95
0

0.05

0.1

0.15

0.2

Channel width (m)

(b) Isolines plot


Fig. 4.40 Variation of bed shear stress in narrow asymmetric compound channel (DSGS model)

From the figure, it is observed that a maximum value (1.39

N/m2) occurs in the floodplain near

the junction of the main channel and floodplain. In comparison to SS model, more bed shear
stresses have been simulated in the channel. Formation of vortex like structure, at the junction of
the main channel and floodplain, is clearly seen in Fig. 4.40(b), showing isolines plot of bed
shear stresses in the channel.

(iv) Secondary velocity at horizontal interface


Fig. 4.41(a) shows interfacial flow velocities in spanwise and vertical directions along horizontal
interface of lower and upper layers of asymmetric compound channel. It is clearly seen from the
figure (see Fig. 4.41c) that secondary currents are anticlockwise at right end of the main channel.
From Fig. 4.41(b), it is inferred that spanwise velocity at horizontal interface in the channel
increases from zero to its maximum value (0.00002 m/s) at the junction of the channel. However,
vertical interface velocity, in such case, is having negligibly smaller value near the far end of the
channel and rises to maximum value (0.000027 m/s) near the junction of the channel. The
existence of secondary currents at the junction of main channel and floodplain is clearly seen in
isolines plot of velocity vector, see Fig. 4.41(b). In comparison to SS model (see Fig. 4.17), the
110

spanwise and vertical flow velocities at interface are simulated lesser in magnitude in DSGS
model (Fig. 4.41).

Spanwise (v0)

Vertical (w0)

Velocity at Interface (m/s)

0.00003

Mc

0.000025
0.00002
0.000015
0.00001
0.000005
0
0

0.02

0.04

0.06

0.08

0.1

Channel width at Interface (m)


(a) Velocities profile

Test section (m)

14.05

14

13.95
0

0.02

0.04

0.06

0.08

Channel width at Interface (m)

(b) Isolines plot


111

0.1

(c) Vector plot


Fig. 4.41 Secondary flow velocities at horizontal interface of narrow asymmetric compound
channel (DSGS model)

(v) Shear stress at horizontal Interface


Variation of shear stress at interface (0) of upper and lower layers is presented in Figs. 4.42(a-b).

0.4
Mc

0 (N/m2)

0.3
0.2
0.1
0
0

0.02

0.04

0.06

Channel width at Interface (m)


(a) Profile plot
112

0.08

0.1

Test section (m)

14.05

14

13.95
0

0.02

0.04

0.06

0.08

0.1

Channel width at Interface (m)

(b) Isolines plot


Fig. 4.42 Variation of shear stress at horizontal interface in narrow asymmetric compound
channel (DSGS model)
From Fig. (4.42a), it is seen that peak of interface shear stress (0.3 N/m2) occurs at the junction
of the main channel and floodplain. Variation of interface shear stresses across interface width
observed in DSGS model was similar to that of SS model. Isolines plot of interface shear stresses
at test section are shown in Fig. (4.42b). Existence of high interfacial shear stress at the junction
of main channel and floodplain is due to transfer of momentum from main channel to floodplain
at the same region.

(vi) SGS turbulent shear in upper layer


The SGS turbulent shear stress in upper layer (

; see Eq.3.40) is normalised with square of

shear velocity (U*) and the resulting profile across width of upper layer is shown in Fig. 4.43(a).
From the figure, maximum value (+0.027 U*2) of turbulent shear stress is simulated near left wall
of upper layer in the main channel, and a value of -0.01U*2 has been simulated at the junction of
main channel and floodplain. The reverse trend of shear stresses in upper layer is reported in the
floodplain due to transfer of flow and momentum from main channel to floodplain. In
comparison to SS model (see Fig. 4.19), more distributed variation of SGS turbulent shear

113

stresses in upper layer are simulated in DSGS model (Fig. 4.43) due to dynamic variation of
model coefficient C in latter across the width of the channel.

0.04
Mc

Fp

0.02

0
-0.02
-0.04
0

0.05

0.1

0.15

0.2

Upper layer width (m)


(a) Profile plot

Test section (m)

14.05

14

13.95
0

0.05

0.1

0.15

0.2

Upper layer width (m)

(b) Isolines plot


Fig. 4.43 Variation of SGS turbulent shear stress in upper layer of narrow asymmetric compound
channel (DSGS model)
(vii) SGS turbulent shear in lower layer
Similar to upper layer, normalised SGS turbulent shear stress (

; see Eq.3.40) profile and

isolines plots across the width of lower layer of the channel were obtained as shown in Fig. 4.44
(a-b).

114

0.04

Mc
0.02

-0.02

-0.04
0

0.025

0.05

0.075

0.1

Lower layer width (m)


(a) Profile plot

Test section (m)

14.05

14

13.95
0

0.02

0.04

0.06

0.08

0.1

Lower layer width (m)

(b) Isolines plot


Fig. 4.44 Variation of SGS turbulent shear stress in lower layer of narrow asymmetric compound
channel (DSGS model)
From the figure, it is seen that a peak (+0.021 U*2) occurs near the left wall of lower layer
whereas a second peak, opposite in nature, (-0.02U*2) is found near the junction of the main
115

channel and the floodplain where transfer of momentum generally takes place. Further, in
comparison to SS model (see Fig. 4.20), more distributed variation of SGS turbulent shear
stresses in lower layer are simulated in DSGS model (Fig. 4.44).

4.4 CONCLUDING REMARKS

The performance of hydrodynamic models such as SS model and DSGS model developed using
turbulence closure schemes under LES approach were assessed for turbulent flows in a wide
symmetric and a narrow asymmetric compound channels. In addition, a normal asymmetric
compound channel conveying shallow and deep flows was also considered. The grid dependence
tests have been undertaken to arrive best comparisons of grids before simulations of hydraulic
characteristics in symmetric and asymmetric compound channels using developed SS and DSGS
models. The sensitivity analysis of model parameters C and has been undertaken, and it is
found that average values of C and equals to 0.1 and 0.5 respectively are adequate for
simulation of hydraulic characteristics using developed model. The SS model has been able to
correctly simulate the primary velocity in symmetric and asymmetric compound channels with
RMSE in the range 0.017 m/s to 0.026 m/s. Significant difference in primary velocity of lower
and upper layers

has been simulated at the junctions of main channel and

floodplain in a wide symmetric channel with shallow flow conditions. The situation is reversed
for a narrow asymmetric compound channel with deep flow conditions where primary velocity
of upper layer is consistently higher than lower layer, and significant difference
has been simulated at the junctions of the main channel and floodplain in the order of -14%. The
simulated trend is also reported for shallow vis--vis deep flow conditions in a normal
asymmetric compound open channel. The simulated bed shear stresses in symmetric and
asymmetric compound channels for shallow and deep flow conditions have shown sudden
changes at junctions of the main channel and floodplain due to existence of significant secondary
currents at that location. The existence of secondary currents (combination of spanwise and
vertical velocity components along the horizontal interface) at the junction of the main channel
and floodplain is responsible for transfer of momentum from main channel to the floodplain.
116

Shear stresses at horizontal interface of both the layers, and averages shear stresses in both upper
and lower layers are simulated, and the same give clear signature of significant secondary
currents at the junctions of the main channel and floodplain. The model parameter, C, has been
found to have significant variations across the channel width which depends on the variation of
grid sizes and flow conditions. The DSGS model, while varying the value of C across the width,
has been found to give marginally better results vis--vis SS model in simulating hydraulic
characteristics such as flow velocities, turbulent shear stresses in compound channel flows.

117

Chapter 5

Simulation of Solute Transport: Results and Discussions

5.1 GENERAL

The transport characteristics of inert solute in lower and upper layers, and along the horizontal
interface between the layers are investigated in compound open channel flows using dynamic
subgrid scale (DSGS) closure scheme in large eddy simulation (LES) model. The performance of
transport model coupled with hydrodynamic 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.

5.2 EXPERIMENTAL DATA

The solute measurements in symmetric and asymmetric compound channels carried out by
Freselle et al. (2010) and Shiono and Feng (2003) respectively, are used for validation of the
performance of DSGS model in respective compound open-channels. Schematics of symmetric
and asymmetric compound cross-sections are included in Figs. 4.1 and 4.2 respectively in
Chapter 4.

5.2.1 Symmetric Compound Channel

Channel geometry and flow characteristics of a wide symmetric compound channel, used for
laboratory experiment of solute transport by Fraselle et al. (2008), are presented, see Fig. 4.1 and
Table 4.1, in Chapter 4. Two sets of experimental data for solute transport, as reported by
118

Fraselle et al. (2008), in symmetric compound channel were used in the present study. In first
experiment, named as Exp.I-1, injection point of solute as point source was located at mid of the
main channel at y=0.6 m and 0.0375 m above the main channel bed, and in the second
experiment, named as Exp.I-4, the solute was injected as line source near the junction of main
channel and floodplain at y=0.415m in a vertical line along depth of the main channel. In both
the experiments, solute was injected into the flow at longitudinal distance x=2 m from the
channel inlet and measuring sections were considered at four different cross-sections along
streamwise direction located at downstream of the channel inlet, i.e., at a distance of 3 m, 4 m, 6
m and 9 m from the inlet section across the channel depth. Diluted common table salt (NaCl) at
flow rate of 1 g/s was applied as tracer material. For respective measurements of concentration in
the channel, specially designed conductivity meter (Fraselle et al., 2008) were used. Dye
injection and measurement locations in experimental runs in symmetric compound channel are
detailed in Table 5.1.

5.2.2 Asymmetric Compound Channel

For solute transport experiments in a narrow asymmetric compound open-channel, hydraulic


characteristics and channel geometry are given in Fig. 4.2 and Table 4.1 in Chapter 4. In three
sets of experiments for solute transport (Shiono and Feng, 2003), the injection points of solute
were located along streamwise direction at x=13 m downstream of channel inlet near the free
water surface (i.e., at 0.108 m above the main channel bed) at y=0.05 m in case Exp.C-1, at
y=0.1 m in case Exp.C-2 and at y=0.15 m in case Exp.C-3. The measuring section was also
located along the streamwise direction at 1 m downstream of the injection point (i.e. 14 m from
the upstream section) at the same level as of injection point, i.e., near the free surface (i.e. at
0.108 m above the channel bed). The tracer material Rhodamine 6G was applied and
sophisticated measuring equipment such as laser induced fluorescence (LIF) was used to
measure the concentration. Dye injection and simulated locations in numerical mode runs in
asymmetric compound channel are detailed in Table 5.1.

119

Table 5.1 Solute injection and measurement locations in simulation experiments in wide
symmetric and narrow asymmetric compound channels
Channel

Channel
type

length,
L

Exp.
runs

(m)

Symmetric
(Fraselle et
al., 2008)

mid of lower
layer

Cs & Qs
12.55 g/l
0.0797 l/s

I-4

0.415

mid of both
lower &
upper layers

Cs & Qs
16.74 g/l
0.0597 l/s

C-1
C-2

0.05
0.1

mid of upper
layer

Cs & Qs
2500 ppb
54 ml/min

10.0

20.0

Injection
concentration
(Cs) & flow
rate
(Qs)

0.6

I-1

Asymmetric

(Shiono
and Feng,
2003)

Solute injection locations in


numerical model
*
D/S of
Y
Vertical
position along
inlet
channel
depth
(m)
(m)

2.0

C-3

13.0

0.15

Simulated
locations d/s
of channel
inlet
(m)

3, 4, 6 & 9

14.0

* Depth averaged results reported from developed DSGS model

5.3 TRANSPORT MODEL SIMULATION RESULTS

The solute transport model coupled with hydrodynamic model developed for governing
equations for solute transport, presented in Subsection 3.2.2 of Chapter 3, is used to simulate the
transport of passive materials in wide symmetric and narrow asymmetric compound channels
reported in the literature (Fraselle et al., 2008; Shiono and Feng, 2003).

As the transport model proposed in present study is two layered two dimensional in nature, the
vertical locations of solute injection and observation points in respect of each numerical
experimental set were placed at mid depth in respective layers of both channels. Turbulent
Schmidt number, t , was considered as 0.5 for the compound channel. After attaining the steady
state condition, model simulations were continued further for about 25000 time steps in
symmetric and 30000 time steps in asymmetric compound open-channels for acquiring time
averaged results.
120

5.3.1 Symmetric Compound Channel

In the numerical experiments of solute transport, grid spacing similar to that of hydrodynamic
model (see Subsection 4.3.1.1a in Chapter 4), i.e., streamwise x spanwise, 500x70 grid cells in
upper layer and 500x26 grid cells in lower layer for symmetric channel, has been adopted. The
predicted layered average solute concentrations were converted into depth averaged value for
comparing the same with experimental data. The data for symmetric compound channel were
available in terms of their depth averaged value (Fraselle et al., 2008; Fraselle et al., 2010).

Solute concentration profile


In order to compare simulated transport results in symmetric compound channel with
experimental data, unit of solute concentration was expressed in terms of gram per litre (g/l). It is
observed in Fig. 5.1(a), at 3 m downstream from the inlet, that experimental results (Fraselle et
al., 2008) agrees well with simulated results of the present model in both sides of the flood plain
with marginal difference in the main channel (RMSE=0.031 g/l). Similarly, in Fig. 5.1(b), for
location at 4.0 m downstream from inlet (Frasellle et al., 2008), the simulated results follow
closely the experimental observations in the floodplain with slight difference in the main channel
(RMSE=0.021 g/l). Similar performance of the DSGS model are reported for the observations
available at 6 m and 9 m downstream of inlet locations with RMSE=0.014 g/l and RMSE=0.01
g/l respectively, see Figs. 5.1(c-d). Thus, from Figs. 5.1(a-d), it is apparent that performance of
the model is better at downstream locations vis--vis upstream locations due to better mixing of
pollutants at former locations. Figs. 5.1(a-d) and Figs. 5.3(a-c) show comparison of spanwise
variation of computed and experimental results (Fraselle et al., 2008) of depth averaged
concentration at various observations due to injection of solute in experiments I-1 and I-4
respectively of symmetric compound channel. The differences in measured and simulated values
are due to the fact that simulations were carried out in two layered 2-D model by injecting the
solute at mid points of lower and upper layers depths while the injection points for experiments
I-1 and I-4 were different from above locations in physical experiments (Fraselle et al., 2008).

121

(a) Location at 3 m downstream from inlet

(b) Location at 4 m downstream from inlet

122

(c) Location at 6 m downstream from inlet

(c) Location at 9 m downstream from inlet


Fig. 5.1 Comparison of depth-averaged solute concentration at different locations downstream of
inlet in wide symmetric compound channel for Exp.I-1

123

In experiments, the solute was injected as point source in Exp.I-1 at a vertical distance of 0.0375
m above the main channel bed while in Exp.I-4, solute was injected as a line source along the
depth of the whole main channel as mentioned in Table 5.1. In Figs. 5.3(a-c), significant
differences in distribution of solute concentration between the computed and the experimental
results are also due to under estimation of secondary currents by the present model near the

Upper layer width (m)

junction edge of the channel.

1.2
0.8
0.4
0
2

Channel length from injection point (m)

Lower layer width (m)

(a) Upper layer

0.8
0.6
0.4
2

Channel length from injection point (m)


(b) Lower layer

Fig. 5.2 Isolines plot of depth-averaged solute concentration (g/l) in layers of wide symmetric
compound channel for Exp.I-1

Figs. 5.2(a-b) and 5.4(a-b) show the simulated concentration distributions, both in upper and
lower layers, for experiments I-1 and I-4 respectively. From both the figures, it is evident that
mixing of solute is significant in upper layer even though the solute was injected in lower layer
(Exp.I-1). The existence of vertical velocity component causes higher rate of distribution of
124

concentrations vertically vis--vis in horizontal direction. Similar observations were also


reported by Sauvaget (1985).

(a) Location at 4 m downstream from inlet

(b) Location at 6 m downstream from inlet

125

(c) Location at 9 m downstream from inlet

Upper layer width (m)

Fig. 5.3 Comparison of depth-averaged solute concentration at different locations downstream of


inlet in wide symmetric compound channel for Exp.I-4
1.2
0.8
0.4
0
2

2.5

3.5

4.5

5.5

6.5

6.5

Channel length from Injection point (m)

Lower layer width (m)

(a) Upper layer

0.8
0.6
0.4
2

2.5

3.5

4.5

5.5

Channel length from Injection point (m)

(b) Lower layer


Fig. 5.4 Isolines plot of depth-averaged solute concentration (g/l) in layers of wide symmetric
compound channel for Exp.I-4
126

5.3.2 Asymmetric Compound Channel

In the numerical experiments of solute transport in asymmetric compound channel, grid spacing
similar to that of hydrodynamic model (see Subsection 4.3.1.2 in Chapter 4), i.e., streamwise x
spanwise, 2000x40 grid cells in upper layer and 2000x20 grid cells in lower layer were adopted.
The predicted depth averaged solute concentrations in upper layer are compared with
experimental data. The experimental data is available in literature as point observed value, i.e., at
the level of 0.105 m above the channel bed (Shiono and Feng, 2003; Shiono et al., 2003).

Solute concentration profile


Figs. 5.5(a), 5.6(a) and 5.7(a) show the simulated concentration profiles across the width
(spanwise) of the channel for experiments C-1, C-2 and C-3 respectively in narrow asymmetric
compound channel. From the figures, it is seen that simulated results follow the trend of
experimental data except near the junction of the main channel and floodplain.

Fig. 5.5(a) Solute concentration profile in narrow asymmetric compound channel for Exp.C-1

127

The injection point in the numerical experiment has been located at 13 m downstream of inlet
channel at midpoint of upper layer. The differences between observed and simulated
concentration profiles for all these experimental sets are due to, (a) vertical locations of injection
point in physical and numerical experiments are different, (b) measurements of concentration in
physical experiments are located at the same vertical positions as location of injection points
while the simulated results are reported for depth averaged concentration of upper layer, (c) The
existence of the 3-D effects in the narrow channel, are not well reproduced in 2-D depth
averaged model.
Figs. 5.5(b), 5.6(b) and 5.7(b) show the simulated depth-averaged concentration isolines, both for
upper and lower layers, for experiments C-1, C-2 and C-3 respectively. From the figures, it is
apparent that concentration of solute is predominant in upper layer only as the injection point is
located in upper layer and existence of vertical velocity, though magnitude is very small, at the

Upper layer width (m)

interface.

0.2
0.1
0
13

13.5

14

14.5

Channel length from Injection point (m)

Lower layer width (m)

(i) Upper layer

0.1
0.05
0
13

13.5

14

14.5

Channel length from Injection point (m)

(ii) Lower layer


Fig. 5.5(b) Isolines plot of solute concentration (ppb) in layers of narrow asymmetric compound
channel for Exp.C-1

128

Upper layer width (m)

Fig. 5.6(a) Solute concentration profile in narrow asymmetric compound channel for Exp.C-2

0.2
0.1
0
13

13.5

14

14.5

Channel length from Injection point (m)

Lower layer width (m)

(i) Upper layer

0.1
0.05
0
13

13.5

14

14.5

Channel length from Injectiuon point (m)

(ii) Lower layer


Fig. 5.6(b) Isolines plot of solute concentration (ppb) in layers of asymmetric compound channel
for Exp.C-2

129

Upper layer width (m)

Fig. 5.7(a) Solute concentration profile in narrow asymmetric compound channel for Exp.C-3

0.2
0.1
0
13

13.5

14

14.5

Channel length from injection point (m)

Lower layer width (m)

(i) Upper layer

0.1
0.05
0
13

13.5

14

14.5

Channel length from Injection point (m)

(ii) Lower layer


Fig. 5.7(b) Isolines plot of solute concentration (ppb) in layers of narrow asymmetric compound
channel for Exp.C-3
130

5.4 CONCLUDING REMARKS

The performance of solute transport model developed in present study using dynamic SGS
scheme has been assessed for experimental data of symmetric and asymmetric compound
channels from literature. Transport process has been investigated through injection of inert
pollutant at various locations in compound open-channels. The model performs better for
pollutant injected at other than shear layer locations. Significant differences in distribution of
solute concentration between the computed and the experiments results are due to under
estimation of secondary currents near the junction of main channel and floodplain. However, the
performance of the model is better at downstream locations as compared to upstream locations
due to better mixing of pollutants in the said locations. Further, the presence of vertical velocity
component at horizontal interface causes higher rate of distribution of concentrations vertically
as compared to in horizontal directions.

131

Chapter 6

Conclusions

6.1 GENERAL

A two-layer 2-D hydrodynamic and solute transport model for turbulent flows in straight
compound channel is developed using turbulence closure schemes based on LES approach. The
turbulent closure schemes employed are standard Smagorinsky (SS) scheme and dynamic
subgrid scale (DSGS) scheme. The governing equations are solved using finite volume method
based on a two step predictor-corrector scheme. The developed model is tested with the
experimental data of symmetric and asymmetric straight compound channels from literature. The
major findings of foregoing study are presented in the following paragraphs:

6.2 HYDRODYNAMIC MODEL

a) The model result on flow parameters like interfacial shear stress and vertical flow
velocity gives insight about the overall momentum exchange between the main channel
and the floodplain of straight compound open-channel.
b) Based on sensitivity analysis of model parameters C and , it is found that average
values of C and equals to 0.1 and 0.5 respectively are adequate for simulation of
hydraulic characteristics using developed model.

132

c) The significant difference in layered average primary flow velocities between lower and
upper layers, i.e., of the order of 10-14% of mean bulk velocity (Ub), has been simulated
near the junction of the compound channels.

d) Variations of flow variables in symmetric and asymmetric straight compound channels in


terms of bed shear stress distributions (between 1.15

to 1.35

at the junction) and

spanwise and vertical velocities are analyzed at the horizontal interface between the
lower layer and the upper layer. The analyses revealed the existence of horizontal
vortices near the junction of the main channel and the floodplain where momentum
exchange takes place.

e) The spanwise distribution of shear stresses at interface of symmetric and asymmetric


compound channel, varying between 0.08 N/m2 to 0.4 N/m2, reveals the dominance of
shear stresses at the junction.

f) The developed model has been able to simulate satisfactorily the depth averaged primary
velocity in compound open channels with RMSE in the range 0.017 m/s to 0.026 m/s.
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.

g) The maximum secondary currents are generated (between 0.02% of Ub to 1.1% of Ub)
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 secondary currents, bed and interfacial shear stresses
indicate the formation of vortices near the junction of the main channel and the
floodplains.

h) 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. The model parameter, C, depends on the
variation of grid sizes and flow conditions.
133

i) The DSGS model has been found to give marginally better results vis--vis SS model in
simulating hydraulic characteristics such as flow velocities, turbulent shear stresses in
compound open channel flows.

6.3 SOLUTE TRANSPORT MODEL

a) The DSGS model has been able to predict the peak solute concentrations irrespective of
point of injection in the compound channel cross-section. However, spanwise distribution
of concentration across the junction of the main channel and the floodplain, depends upon
the magnitude of secondary currents.

b) The developed model has been able to simulate the concentration of solute in both
symmetric (Exps.I-1, I-4) and asymmetric (Exp.C-1, C-2, C-3) compound open channels
with RMSE ranging from 0.01 g/l to 0.05 g/l and 0.73 ppb to 1.19 ppb respectively.

c) The presence of vertical velocity component causes higher rate of distribution of


concentrations vertically vis--vis in horizontal direction.

d) The performance of developed model is better at downstream locations as compared to


upstream locations due to better mixing of pollutants in said locations.

6.4 FUTURE SCOPE OF WORK

a) Performance of present model can be improved further by adding additional terms at the
junction of the main channel and floodplain, responsible for generating secondary
currents at those regions.

134

b) Empirical relations can be modified for efficient turbulent exchange across the horizontal
interface between the layers.

c) Performance of hydrodynamic model can be considered for density driven flows while
giving due consideration to the density of fluids.

d) The developed model can be made more realistic by incorporating pressure gradient
terms using SIMPLE like procedure.

e) Active pollutants term can be added in governing equations for model simulation to
tackle real world problems.

135

References
Abbott, M.B. (1979). Computational Hydraulics: Elements of the theory of free surface flows,
Ashgate, Boston.
Ackers, P. (1993). Flow formulae for straight two-stage channels, J. Hyd. Res., 31(4), 509
531.
Alcrudo, F. and Navarro, P.G. (1993). A high resolution Godunov type scheme in Finite
volumes for the 2D shallow water equations, Int. J. Num. Meth. Fluids, 16, 489-505.
Bardina, J., Lyrio, A., Kline, S.J., Ferziger, J.H. and Johnston, J.P.(1981). A prediction method
of planar diffuser flows, J. Fluids Engg., 103, 315-321.
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.
Celik, I.B. (1999). Introductory Turbulence Modeling. Lectures Notes, West Virginia University
Mechanical & Aerospace Engineering Dept.
Chatila, J.G. and Townsend, R.D. (1998). Modelling of pollutant transport in compound openchannels, Canadian Water Res. J., 23(3), 259-271.
Chau, K.W. and 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. Meth. Fluids, 21,
1087-1107.
Chow, V.T. (1959). Open Channel Hydraulics, MacGraw-Hill, NewYork.
Cokljat, D. and Younis, B. (1995). Second order closure study of Open channel flows, J. Hyd.
Engg., ASCE, 121(2), 94-107.
Craft, T.J., Suga, K. and Launder, B.E. (1993). Extending the applicability of eddy viscosity
models through the use of deformation invariants and non-linear elements, Proc 5th Int. Symp.
on Refined Flow Modelling and Turbulence Measurements, 125132.
Cunge, J. (1975). Two dimensional modelling of flood plains, in: Mahmood K. and Yevjevich
V., eds., Unsteady flow in open channels, Water Resources Publications, Fort Collins, Colorado,
705-762.
Djordjevic, S. (1993). Mathematical model of unsteady transport and its experimental validation
in compound open channel flow, J. Hyd. Res., 31(2), 229248.

136

Fernandes, J.N., Leal, J.B. and Cardoso, A.H. (2014), Longitudinal development of compound
channel flows, 3rd IAHR European Congress, Porto, Portugal.
Fraselle, Q., Arnould, T., Lissoir, X., Bousmar, D. and Zech, Y. (2008). Investigating diffusion
and dispersion in compound channels using low cost tracer, Proc. River Flow 2008 conf. on
Fluvial Hydraulics, Cesme, Izmir, Turkey, September 3-5, 1, 529-538.
Fraselle, Q., Bousmar, D., and Zech, Y. (2010). Diffusion and dispersion in compound
channels: The role of large scale turbulent structures, Proc. 1st European Congress, IAHR,
Ediburge, UK, May 4-6, No.FPIVb.
Germano, M., Piomelli, U., Moin, P. and Cabot, W. H. (1991). A dynamic subgrid scale eddyviscosity model Physics of Fluids, A3(7), 17601765.
Ghose, S. and Kline, S.J. (1978). The computation of optimum pressure recovery in twodimensional diffusers, J. Fluid Engg., 100, 419-426.
Ghosh, S. N. and Jena, S. B. (1971). "Boundary shear distribution in open compound channel",
Proc. I.C.E, 49.
Ghosh, S. N. and Mehta, P.J. (1974). "Boundary shear distribution in a channel with varying
roughness distribution", Proc. I.C.E., 57.
Hosoda, T., Sakurai, T., Kimura, I., and Muramoto, Y. (1999). 3-D computations of compound
channel flows with horizontal vortices and secondary currents by means of non-linear k-
model, J. Hydros. Hyd. Engg., JSCE, 17(2), 87-96.
Jin, H.S., Egashira, S. and Liu, B.Y. (1998). Characteristics of meandering compound channel
flow evaluated with two-layered, 2-D method, J. Hydros. Hyd. Engg., 16(1), 87-96.
Joung, Y. and Choi, S.U. (2008). Investigation of twin vortices near the interface in turbulent
compound open-channel flows using DNS data, J. Hyd. Engg., ASCE, 134(12), 1744-1756.
Kara, S., Stoesser, T. and Sturm, T.W. (2012). Turbulence statistics in compound channels with
deep and shallow overbank flows, J. Hyd. Res., 50(5), 482-493.
Kawahara, Y., Fukuoka, S., Ikeda, S., Pender, G. and Shimizu, Y. (2008). Chapter 4: Computer
simulation, Flow and sediment transport in compound channels: The experiences of Japanese
and UK research, S. Ikeda & I.K. McEwan (Eds.), IAHR Monograph Series, 167-234
Knight, D.W and Demetriou, J.D. (1983). Floodplain and main channel flow interaction J.
Hyd. Engg., ASCE, 109(8), 1073-1092
Knight, D.W. and Hamed, M.E., (1984). Boundary shear in symmetrical compound channels,
J. Hyd. Engg., 110(10), 1412-1430

137

Launder, B.E. and Spalding, D.B. (1974). The numerical computation of turbulent flows,
Comp. Meth. Applied Mech. Engg., 3, 269-289.
Lesieur, M., Metais, O. and Comte, P. (2005). Large eddy simulation of turbulence. Cambridge
University Press, Cambridge.
Lin, B. and Shiono, K. (1995). Numerical modelling of solute transport in compound channel
flows, J. Hyd. Res., 33(6), 773-788.
McKeogh, E.J. and Kiely, G. (1989). Experimental study of the mechanisms of flood flow in
meandering channels, Proceedings of the 23rd Int. Ass. Hydraulics Res. (IAHR), Congress,
Ottawa.
Mehrer, H. and Stolwijk, N.A. (2009). Heroes and highlights in the history of diffusion,
diffusion-fundamentals.org, 11(1), 1-32.
Myers, W.R.C. and Elsawy, E.M. (1975). "Boundary shear in channel with floodplain", J.
Hydraulic Div., ASCE, 101(7), 933-946.
Myers, W.R.C. (1978). "Momentum transfer in a compound channel", J. Hydraulic Res., IAHR,
16(2), 139-150.
Nadaoka, K. and Yagi, H. (1998). "Shallow-water turbulence modeling and horizontal largeeddy computation of river flow", J. Hydraulic Engg., ASCE, 124(5), 493-500.
Naot, D. and Rodi, W. (1982). Calculation of secondary currents in channel flow, J. Hydraulic
Div., ASCE, 108(8), 948968.
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. (1996). Experimental and numerical study on 3-D turbulent structures in compound
open-channel flows. Flow modelling and Turbulence measurements, edited by C.J. Chen et al.,
Balkema, 65-74.
Nezu, I. and Nakayama, T. (1997). Space-time correlation structures of horizontal coherent
vortices in compound channel flows by using particle-tracking velocimetry, J. Hydraulic Res.,
IAHR, 35(5), 191- 208.
Nezu, I. and Rodi, W. (1986). Open channel flow measurements with a laser Doppler
anemometer, J. Hydraulic Eng., 112(5), 335-355.
Nezu, I., Sanjou, M. and Goto, K. (2004). Transition process of coherent vortices in depthvarying unsteady compound open-channel Flows. Shallow Flows (ed. G.H. Jirka & W.S.J.
Uijttewaal), Balkema, 251-258.

138

Nugroho, E.O. and Ikeda, S. (2007). Comparison study of flow in a compound channel:
Experimental and numerical method using large eddy simulation SDS-2DH model, ITB J.
Engg. Sc., 39B(2), 67-97.
Nujic, M. (1995) Efficient implementation of non-oscillatory schemes for the computation of
free-surface flows, J. Hydraulic Res., IAHR, 33(1), 101-111.
Othman, F. and Valentine, E.M. (2006). Numerical modelling of the velocity distribution in a
compound channel, J. Hydrol. Hydromech., 54(3), 269279
Patel, V.C. (1965). Calibration of the Preston tube and limitations on its use in pressure
gradients, J. Fluid Mech., 23(1), 185-208.
Pender, G. and Manson, J.R. (1994). Developments in three-dimensional numerical modelling
of river flows, Proc. 2nd Int. Conf. River Flood Hyd., 22-25 March, York, UK, Wiley
Pezzinga, G. (1994). Velocity distribution in compound channel flows by numerical
modelling, J. Hydraulic Engrg., 120(10), 11761198.
Prandtl, L. (1925). Bericht ber Untersuchungen zur ausgebildeten Turbulenz, Z. Angew.
Math, Meth., 5, 136-139.
Prandtl, L. (1932). Recent Results of Turbulent Research, translation by National Advisory
Committee for Aeronautics, TM No. 720 (originally published in German in 1932).
Premnath, K.N., Pattison, M.J. and Banerjee, S. (2009). Generalized lattice Boltzmann equation
with forcing term for computation of wall-bounded turbulent flows, Physical Review, E79 (2,
Part 2).
Price, R.K. (1975). "A mathematical model for river developments Theoretical development",
Report No. INT 127, December.
Rajaratnam, N. and Ahmadi, R.M. (1970). "Interaction between main channel and floodplain
flows", J. Hyd. Div., ASCE, 105(5), 573-588.
Rajaratnam, N. and Ahmadi, R.M. (1981). "Hydraulics of channels with floodplains", J. Hyd.
Res., IAHR, 19(1), 43-60.
Rodi, W. (1980). Turbulence Models and Their Application in Hydraulics, Delft, Netherland,
IAHR.
Sanjou, M., Nezu, I. and Itai, K. (2010). Space-time correlation and momentum exchanges in
compound open-channel flow by simultaneous measurements of two-sets of ADVs, River Flow
2010 - Dittrich, Koll, Aberle and Geisenhainer (eds), Bundesanstalt.

139

Sauvaget, P. (1985). "Dispersion in rivers and coastal Waters - 2. Numerical computation of


dispersion". Developments in hydraulic research-3, P. Novak (ed.), Elsvier, Applied Science,
London, 39-78.
Sellin, R.H.J. (1964). A laboratory investigation into the interaction between flow in the
channel of a river and that of its floodplain, La Houille Blanche, 7, 793-801.
Sharma, A. and Eswaran, V. ( 2003). Computational fluid flow and heat transfer, Edited by K.
Muralidhar and T. Sundararajan under the auspices of IIT Kanpur Series, Narosa, India.
Shi, J., Thomas, T.G. and Williams, J.J.R. (2000). Coarse resolution large-eddy simulation of
turbulent channel flows, Int. J. Num. Meth. Heat and Fluid Flow, 11(1), 25-31.
Shiono, K. and Knight, D.W. (1989). Transverse and vertical measurements of Reynolds
stress in a shear region of a compound channel, Proc. 7th Int. Symp. Turbulent Shear
Flows, Stanford, USA.
Shiono, K. and Feng, T. (2003). Turbulence measurements of dye concentration and effects of
secondary flow on distribution in open channel flows, J. Hyd. Engg., 129(5), 373-384.
Shiono, K., Scott, C.F. and Kearney, D. (2003). Predictions of solute transport in a compound
channel using turbulence models, J. Hyd. Res., 41(3), 247-258.
Simoes, F.J.M. and Wang, S.S.Y. (1997). Numerical prediction of three dimensional mixing in
a compound channel, J. Hyd. Res.. IAHR, 35(5), 619642.
Singh, V. and Bhallamudi, S.M. (1997), Hydrodynamic modelling of basin irrigation, J. Irrig.
Drain. Engg., ASCE, 123(6), 407-414.
Smagorinsky, J. (1963). General circulation experiments with the primitive equations: the
basics experiments. Monthly Weather Review, 91, 99-164.
Sofialidis, D. and Prinos, P. (1998). Compound open channel flow modeling with non-linear
low - models, J. Hyd. Engg., ASCE, 124(4), 253-262.
Sofialidis, D. and Prinos, P. (1999). Numerical study of momentum exchange in compound
open channel flow, J. Hyd. Engg., ASCE, 125(2), 152-165.
Stocchino, A. and Brocchini, M. (2010). Horizontal mixing of quasi-uniform, straight,
compound channel flows, J. Fluid Mech., 643, 425-435
Stocchino, A., Besio, G., Angiolani, S. and Brocchini, M. (2011). Lagrangian mixing in straight
compound channels, J. Fluid Mech., 675, 168-198.
Tang, X. and Knight, W. (2008). Lateral depth-averaged velocity distributions and bed shear in
rectangular compound channels, J. Hyd. Engg., ASCE, 134(9), 1337-1342.
140

Thomas, T.G. and Williams J.J.R. (1995), Large eddy simulation of turbulent flow in an
asymmetric compound open channel, J. Hyd. Res., 33(1), 27-41.
Tominaga, A., Ezaki, K., Nezu, I. and Nakagawa, H. (1989). Three-dimensional turbulent
structure in straight open channel flows, J. Hydraulic Res., 27, 149173
Tominaga, A. and Nezu, I. (1991). Turbulent structures in compound open-channel flow, J.
Hyd. Engg., ASCE, 117(1), 2140.
Von Karman, T.H. (1930). Mechanische aehnlichkeit und turbulenz, Proc. Third Intern.
Congress for Appl. Mech., Stockholm.
Wilcox, D.C. (1994) Turbulence modelling for CFD, DCW Industries, Inc., California.
Wood, I.R. and Liang, T. (1989). Dispersion in an open channel with a step in the crosssection, J. Hyd. Res., 27(5), 587601.
Xie, Z., Lin, B. and Falconer, R.A. (2013). Large-eddy simulation of the turbulent structure in
compound open-channel flows, Adv. Water Res., 53, 66-75.
Yang, K., Nie, R., Liu, X. and Cao, S. (2013). Modelling depth-averaged velocity and boundary
shear stress in rectangular compound channels with secondary flows, J. Hyd. Engg., ASCE,
1329(1), 76-83.
Yee, H. (1989). A class of high resolution explicit and implicit shock-capturing methods,
NASA Report N89-25652.
Yuan, M., Song, C.S. and He, J. (1988),Numerical analysis of turbulent flow in a twodimensional non-symmetric plane-wall diffuser, J. Fluids Engg,, 113, 210-215.
Zarrati, A.R., Jin, Y.C. and Karimpour, S. (2008). A semi-analytical model for shear stress
distribution in simple and compound open channels, J. Hyd. Engg., ASCE, 134(2), 205-215.
Zheleznyakov, G.V. (1971). "Interaction of channel and floodplain streams", Proc. 14th Int.
Congr., IAHR, Paris.

141

Research Publications
1. Development of two-layered model for compound open-channel flow Journal of
Hydraulic Engineering, ISH, 2014, Vol. 20, Issue 3, pp.250-262; published by Taylor and
Francis, UK.
2. Modelling of solute transport in a two-layered compound open channel flows by
using dynamic SGS closure scheme submitted to the Journal of Hydrodynamics,
Elsevier. (Paper submitted in May 2016 and reported under review).
3. Flow characteristics at horizontal interface of asymmetric compound open
channel Proceedings of 21st International Conference on Hydraulics, Water Resources
and Coastal Engineering, HYDRO-2016 International, CWPRS, Pune, Dec.8-10, 2016.
4. Hydrodynamic characteristics of flows in a two-layered compound open channels
by using dynamic SGS model National Conference on Water Resources & Flood
Management with special reference to Flood Modelling, October 14-15, 2016. SVNIT,
Surat, pp.WRF-24-1 to WRF-24-16.
5. Development of two-layered model for compound open channel flow Proceedings
of National Conference on Hydraulics and Water Resources, HYDRO-2012, IIT
Bombay, India, Dec.7-8, 2012, pp.395-404.

142

You might also like