Вы находитесь на странице: 1из 14

Numerical modeling of 3D turbulent free surface ow in natural waterways

Seokkoo Kang, Fotis Sotiropoulos

St. Anthony Falls Laboratory, Department of Civil Engineering, University of Minnesota, 2 Third Avenue SE, Minneapolis, MN 55414, USA
a r t i c l e i n f o
Article history:
Received 13 July 2011
Received in revised form 13 December 2011
Accepted 27 January 2012
Available online 8 February 2012
Keywords:
Free surface ow
Level set method
Immersed boundary method
Hydraulic structures
Natural waterways
Turbulence
a b s t r a c t
We develop a numerical model capable of simulating three-dimensional, turbulent free surface ows in
natural waterways. Free surface motion is captured by coupling the two-phase level set method and the
sharp-interface curvilinear immersed boundary (CURVIB) method of Kang et al. [1]. The model solves the
three-dimensional, incompressible, unsteady Reynolds-averaged NavierStokes (RANS) and continuity
equations in generalized curvilinear coordinates using a fractional step method extended to handle mul-
tiphase ows. Turbulence is modeled by a two-equation RANS model implemented in the context of the
CURVIB method. The accuracy of the level set method is veried by applying it to simulate two- and
three-dimensional sloshing problems, and the potential of the model for simulating real life, turbulent
free surface ows is demonstrated by applying it to carry out RANS simulation of ow past rock struc-
tures in a laboratory ume and ow in a eld scale meandering channel. The simulations show that
the method is able to accurately predict water surface elevation over complex hydraulic structures and
bathymetry, and capture the transition between subcritical and supercritical ows without any special
treatment.
2012 Elsevier Ltd. All rights reserved.
1. Introduction
Turbulent free surface ows are encountered in many hydraulic
and water resources engineering problems. Their understanding is
thus a critical prerequisite for designing stream and river restora-
tion projects and a broad range of hydraulic structures, such as
spillways, sh passages, weirs, etc. Even though free surface ows
have been studied for well over hundred years, their underlying
mechanisms are not yet fully understood especially in natural
waterways with embedded complex hydraulic structures. The pri-
mary challenge for modeling such ows arises from the fact that
the air/water interface is dynamic and its motion is coupled in a
non-linear manner with the instantaneous owelds in both the
air and water phases.
One- (cross-sectional averaged) and two-dimensional (depth-
averaged) models [24] have been widely used for modeling free
surface ows in open channels because of their computational sim-
plicity and expedience and the fact that they greatly simplify the
treatment of the air/water interface. Depth-averaged models, how-
ever, are based on the assumptions that the pressure distribution
across the depth is hydrostatic and that the vertical velocity
component of the ow is negligible, and cannot account for
three-dimensional effects in turbulent ows. Such assumptions
inherently limit the applicability of depth-averaged models to
relatively simple ows and preclude their application to tackle
turbulent free surface ows with complex hydraulic structures.
The objective of this paper is to develop a numerical model that
circumvents the inherent limitations of the depth-averaged mod-
els and is capable of simulating unsteady, three-dimensional
(3D), turbulent, free surface ow in complex open channels.
Methods that are suitable for fully, 3D simulations of free
surface ows can be classied into three broad categories: (1)
arbitrary LagrangianEulerian (ALE) methods; (2) front tracking
methods; and (3) front capturing methods.
In ALE methods only the water phase is simulated, and the com-
putational domain is discretized using a boundary-tted curvilin-
ear mesh that conforms at all times with the airwater interface.
Boundary conditions for the pressure and velocity elds at the
interface and the instantaneous shape of the interface are deter-
mined in a coupled manner by ensuring that: (1) the interface is,
at all times, a material surface (kinematic conditions); and (2)
the stresses at the air/water interface are continuous (dynamic
condition) [5]. ALE methods require the use of moving curvilinear
coordinates [5] since re-meshing is required at every time step to
ensure that the computational mesh conforms to the interface.
Consequently, such methods are inherently limited to situations
in which the interface does not break and its deformation is kept
bounded such that the connectivity of the surface-tted grid is
preserved at all times. Hence, violent free surface motions such
as wave breaking or splashing motions cannot be modeled.
Ramaswamy and Kawahara [6] used the ALE approach to solve
the two-dimensional (2D) sloshing problem on an unstructured
triangular mesh. Hodges and Street [7] employed an ALE
0309-1708/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
doi:10.1016/j.advwatres.2012.01.012

Corresponding author.
E-mail address: fotis@umn.edu (F. Sotiropoulos).
Advances in Water Resources 40 (2012) 2336
Contents lists available at SciVerse ScienceDirect
Advances in Water Resources
j our nal homepage: www. el sevi er. com/ l ocat e/ advwat r es
formulation in conjunction with a boundary-orthogonal curvilin-
ear grid to simulate 3D nite-amplitude surface waves in an open
channel with a at bed.
The front tracking method was rst introduced by Unverdi and
Tryggvason [8]. In this method both the air and water phases are
simulated and a closed material curve (or a closed surface for 3D
problems) is generated to represent the interface between the
two uids. This curve (or surface) is immersed in the background
computational domain and advected in a Lagrangian manner to
track the temporal evolution of the interface. This method solves
a single set of governing equations for the whole oweld [8].
The main advantage of the front tracking method is its inherently
non-diffusive nature, which enables simulating the temporal evo-
lution of the free surface as a sharp interface. This method can also
solve merging and breakup of the interfaces [9]. However, this
method can be very expensive computationally, especially in com-
plex 3D ows, because of the need to continuously re-mesh the
surface mesh used to track the interface. Front tracking methods
have been applied to a large range of multiphase ow problems,
and a comprehensive review of related literature can be found in
Tryggvason et al. [10]. To the best of our knowledge such methods
have yet to be applied to open channel ows.
Front capturing methods also simulate both phases but are
inherently Eulerian. The air/water domain is discretized using a
xed background grid within which the position of the interface
is captured by an appropriate scalar marker function dened on
the background grid. The main advantage of such methods, which
makes them very attractive for simulating complex hydraulic engi-
neering ows, is that no assumption needs to be made about the
connectivity of the air/water interface. Therefore, topologically
complex air/water transitions, such as wave breaking, air entrain-
ment, etc., can be modeled in a straightforward manner. Two
important methods in this category are the volume-of-uid
(VOF) method of Hirt and Nichols [11] and the level set method
of Osher and Sethian [12].
In the VOF method, the marker function is dened as the vol-
ume fraction of the water in each grid cell. The value of the marker
function is zero and one in the fully air and fully water cells,
respectively. This method requires the reconstruction of the inter-
face using the values of the marker function at neighboring cells to
prevent non-physical behavior of the interface [13]. For applica-
tions of the VOF method to simulate open channel ows using
the Reynolds-averaged NavierStokes (RANS) equations, the reader
is referred to Chen et al. [14] and Lin and Liu [15], for 2D cases, and
Ramamurthy et al. [16], for a 3D case.
In the level set method, on the other hand, the value of the mar-
ker function is dened as the distance from the closest interface.
Hence, the marker function is zero at the interface, positive in
the liquid phase, and negative in the gas phase. The advantage of
the level set method over the VOF method is that it does not re-
quire the reconstruction of the interface, which is not straightfor-
ward to implement in 3D unstructured or generalized curvilinear
grids. Yue et al. [17] solved 2D and 3D dam break problems over
a at bed with the level set method and could capture complex free
surface patterns such as air entrainment in the water and splashing
of the water surge front. Yue et al. [18] simulated free surface ow
over a laboratory-scale, simple-shaped dune using large-eddy sim-
ulation (LES) with the level set method. Carrica et al. [19] devel-
oped a single phase level set method and simulated wave
diffraction around a surface ship. Yang and Stern [20] simulated
wave-body interactions using a sharp interface immersed bound-
ary method and a level set method on Cartesian grids.
As the review of previous literature indicates, in spite of signif-
icant progress in modeling free surface ows only a handful of
studies have attempted to simulate 3D free surface ows in open
channels. This state of affairs could be attributed to the complex
geometries characterizing natural waterways as well as the dif-
culties with obtaining converged solutions in long and shallow do-
mains that are typical of open channel ows. Furthermore, and to
the best of our knowledge, no studies have been reported so far
attempting to simulate 3D free surface ows in complex open
channels. Stoesser et al. [21], for instance, have carried out 3D
RANS simulation of ow in a natural river with vegetated ood-
plains; however, the calculated free surface in their work appears
completely at across the transverse direction, and only varies in
the streamwise direction. Furthermore, the free surface tracking
method was not described in the paper and as such it is hard to
know as to whether the authors solved 1D, 2D or 3D free surface
tracking equations, or simply estimated the free surface elevations
using the pressure head calculated from a xed, at rigid lid
simulation.
This paper aims to develop a numerical model for solving 3Dtur-
bulent free surface ows in complex open channels with particular
emphasis on ows with complex embedded hydraulic structures,
such as, for example, boulders, groynes, bridge piers, stream-resto-
ration structures made of natural rocks, etc. The numerical method
is based on the method of Kang et al. [1] who developed an efcient
approach for carrying out LES and RANS simulation of turbulent
ows in complex natural waterways using the curvilinear im-
mersed boundary (CURVIB) method [22]. The method of Kang
et al. [1], however, treats the free surface as a xed, sloping rigid
lid whose shape is prescribed from experimental measurements.
In this paper, we remove this limitation of the method of Kang
et al. [1] by extending it to simulate turbulent free surface ows
in complex geometries using a fully-coupled, non-linear treatment
for the free surface. More specically, the fractional step method of
Kang et al. [1] is rst extended to simulate two-phase owand then
coupled with the level set method developed by Osher and Sethian
[12]. We select the level set method because, unlike most of the
other methods we reviewed above, it can be easily coupled with
the CURVIB method since no boundary-tted mesh is used to track
the interface and, thus, no re-meshing of the background CURVIB
grid is required. We chose the level set method instead of the VOF
method because in the VOF method the implementation of the uid
interface reconstruction in the 3D generalized curvilinear grids that
is employed by the CURVIB method can be very difcult and com-
putationally expensive. Even though the numerical formulation
we develop herein is general and in principle applicable to carry
out both RANS and LES of complex turbulent ows, only the RANS
formulation is put forth in this paper.
The paper is organized as follows. In Section 2 we present the
governing equations for two-phase, free surface ow and in Section
3 we describe the numerical method we develop for solving them.
In Section 4, we rst validate the level set method by applying it to
solve 2D and 3D sloshing problems. Subsequently we demonstrate
the predictive capabilities of the CURVIB-level set method in ows
with complex hydraulic structures by applying it to simulate tur-
bulent free surface ow in a straight laboratory ume with an
embedded stream-restoration structure made of an assembly of
natural rocks. Finally, we simulate the ow in a eld scale mean-
dering stream with complex bathymetry to demonstrate the po-
tential of the present method for solving real life free surface
ows in complex natural waterways. In Section 5 we summarize
our ndings and present the conclusions of this study.
2. Governing equations for two-phase free surface ow
2.1. Unsteady RANS equations
The governing equations are the 3D unsteady Reynolds-aver-
aged NavierStokes (URANS) and turbulence closure equations
24 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
for two immiscible and incompressible uids formulated using the
level set approach [12]. The governing equations are rst formu-
lated in Cartesian coordinates {x
i
} (where i = 1, 2, 3) and then trans-
formed fully (both the velocity vector and spatial coordinates are
expressed in curvilinear coordinates) in generalized, curvilinear
coordinates {n
i
} as required by the CURVIB method [1,22]). In this
section we present the resulting equations in compact tensor nota-
tion (repeated indices imply summation).
Let / denote the Reynolds-averaged level set function, which
measures the distance to the closest uid interface and is positive
in the water region, negative in the air region and zero at the inter-
face. Assuming for the time being that this function is known (see
Section 2.3 for the equations governing the evolution of this quan-
tity), the two-uid, level set form of the unsteady RANS equations
reads as follows (i, j = 1, 2, 3):
J
@U
j
@n
j
0; 1
1
J
@U
i
@t

n
i
l
J

@
@n
j
U
j
u
l

1
q/
@
@n
j
l/
g
jk
J
@u
l
@n
k
_ _ _

1
q/
@
@n
j
n
j
l
p
J
_ _

1
q/
@s
lj
@n
j
G
i
_
; 2
where: n
i
l
@n
i
=@x
l
are the transformation metrics; J is the jacobian
of the geometric transformation; u
i
is the ith component of the Rey-
nolds-averaged velocity vector in Cartesian coordinates; U
i

n
i
m
=J
_ _
u
m
is the contravariant volume ux; g
jk
n
j
l
n
k
l
are the compo-
nents of the contravariant metric tensor; q is the density; l is the
dynamic viscosity; p is the Reynolds-averaged pressure; and s
ij
is
the Reynolds stress tensor; G
i
is the acceleration due to gravity.
For a two-uid formulation, the density and viscosity of the
uid in the above equations are not constant. Rather, the uid
properties vary as function of /, transitioning smoothly across
the interface from the respective values in the water phase to those
in the air phase, as follows:
q/ q
air
q
water
q
air
h/; 3
l/ l
air
l
water
l
air
h/; 4
where h(/) is the smoothed Heaviside function [23] given as
follows:
h/
0 / < ;
1
2

/
2

1
2p
sin
p/

_ _
6 / 6 ;
1 < /;
_

_
5
where is a tunable parameter that determines the thickness of the
numerical smearing of the interface [23]. The value of is usually
set to be equal to the length of one to two grid spacings. The
smoothing of the interface achieved by introducing the Heaviside
function renders the density, velocity and pressure elds continu-
ous across the interface and prevents potential numerical
instabilities.
The Reynolds stress tensor s
ij
in Eq. (2) is modeled using the
Boussinesq hypothesis as follows:
s
ij
2l
t
S
ij

2
3
qkd
ij
; 6
where S
ij
is the Reynolds-averaged strain-rate tensor, l
t
is the dy-
namic eddy viscosity, k is the turbulence kinetic energy, and d
ij
is
the Kronecker delta. The eddy viscosity is calculated by solving
the URANS equations described in the subsequent section.
2.2. Turbulence closure equations
To close the URANS equations we employ the level set version
of the shear stress transport (SST) k-x (where x is the rate of en-
ergy dissipation per unit kinetic energy) model [24], which in gen-
eralized curvilinear coordinates reads as follows:
1
J
@k
@t

@
@n
j
kU
j

1
J
1
q/

P
1
J
b

kx
1
q/

@
@n
j
l/ r
k
l
t
/
g
jk
J
@k
@n
k
_ _
; 7
1
J
@x
@t

@
@n
j
xU
j

1
J
a

S
2

1
J
bx
2

1
q/

@
@n
j
l/ r
x
l
t
/
g
jk
J
@x
@n
k
_ _

1
J
21 F
1
r
x2
1
x
@n
l
@x
j
@k
@n
l
_ _
@n
l
@x
j
@x
@n
l
_ _
; 8
l
t
/
a
1
q/k
maxa
1
x;

SF
2

; 9
where

S
ij

S
ij
_
is the invariant measure of the strain rate and

P
is a limited production term given by the following equation:

P min l
t
/
@u
i
@n
k
@n
k
@x
j
@u
i
@n
l
@n
l
@x
j

@u
j
@n
l
@n
l
@x
i
_ _
; 10b

q/kx
_ _
: 10
F
1
and F
2
are blending functions dened by
F
1
tanh min max

k
p
b

xd
;
500l/
q/d
2
x
_ _
;
4q/r
x2
k
CD
kx
d
2
_ _ _ _
4
_
_
_
_
_
_
; 11
F
2
tanh max
2

k
p
b

xd
;
500l/
q/d
2
x
_ _ _ _
2
_
_
_
_
; 12
with
CD
kx
max 2q/r
x2
1
x
@k
@n
l
@n
l
@x
j
_ _
@x
@n
l
@n
l
@x
j
_ _
; 10
10
_ _
13
and d is the distance to the nearest wall. The closure coefcients of
the SST model are obtained by blending those of the k-x model, de-
noted as h
1
, and the standard k- model, denoted as h
2
, via the rela-
tion h = h
1
F
1
+ h
2
(1 F
1
). The coefcients are given by a
1
= 0.31, b

=
9/100, a
1
= 5/9, b
1
= 3/40, r
k1
= 0.85, r
x1
= 0.5, a
2
= 0.44, b
2
=
0.0828, r
k2
= 1, r
x2
= 0.856.
2.3. Level set equations
In order to model the motion of the free surface interface, the
level set function / is computed by the level set method proposed
by Osher and Sethian [12]. The governing equation for the time-
averaged motion of the interface of two immiscible uids (e.g. air
and water) curvilinear grids is written as
1
J
@/
@t
U
j
@/
@n
j
U
j;0
@/
0
@n
j
_ _
; 14
where hi denotes the Reynolds averaging and the superscript
0
de-
notes the temporal uctuation. Due to the Reynolds averaging of
the advection equation for /, a second order correlation term
involving instantaneous uctuations of / and U
j
appears in the right
hand side of the above equation. This term has been neglected from
the above equation in our model, assuming that the uctuations of
the level set function are small and their overall contribution to the
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 25
mean ow is negligible. The validity of this assumption, however,
has yet to be veried.
While Eq. (14) accurately advects the zero level set according to
the given velocity eld [12], the level set function away from the
gas/liquid interface is not necessarily a distance function, which
should always satisfy the relation of jr/j = 1. Keeping the level
set function as a distance function, at least within the several grid
cells near the interface, is very important in the level set method.
This is because the gradient of / can become large if the distance
function is not preserved, which leads to a loss of accuracy. In order
to preserve the distance function, Sussman et al. [25] suggested
solving the following reinitialization equation
@/
@s
S/
0
jr/j 1 0; 15
where @( )/@s is a pseudo-time derivative, which is driven to zero
during every physical time-step through iterations, /
0
is the level
set function at the beginning of the pseudo-time iteration, and S
is the smoothed sign function given by:
S/
0

1 /
0
P;
1 /
0
6 ;
/
0


1
p
sin
p/
0

_ _
otherwise:
_

_
16
The solutions of Eqs. (14) and (15) satisfy the distance function
requirement, but it does not guarantee conservation of mass. This
can lead to loss or gain of air/water volume. To circumvent this
problem, we solve the mass conserving reinitialization equation
proposed by Sussman and Fatemi [26], which adds the mass correc-
tion term to the right hand side of Eq. (15) as follows:
@/
@s
S/
0
jr/j 1 k
~
d/jr/j; 17
with
k
_
X
~
d/S/
o
1 jr/jdX
_
X
~
d
2
/jr/jdX
; 18
where X is the volume of an individual grid cell. The value of k is
constant within each grid cell.
3. Numerical methods
In this section, we present numerical methods for solving the
governing equations in the hybrid staggered/non-staggered curvi-
linear grids employed in the CURVIB method [1,22]. We also de-
scribe the CURVIB method and present the employed boundary
conditions.
3.1. Solution of the URANS equations
We extend the hybrid staggered/non-staggered grid fractional
step method proposed by Ge and Sotiropoulos [22] to multiphase
ow. To solve Eqs. (1) and (2), during the rst step of this method,
the momentumequations (Eq. (2)) are discretized in time using the
second-order CrankNicolson method:
1
J
U

U
n
Dt
Pp
n
; /
n

1
2
FU

; u

; /
n1
FU
n
; u
n
; /
n
; 19
where the superscript n denotes the previous time level, Dt is the
time step size, U is the vector of the contravariant volume ux, P
is the vector of the pressure gradient term, and F is the vector of
the right hand side of Eq. (2) containing terms other than the pres-
sure gradient term. The advection and diffusion terms are discret-
ized using the third-order WENO scheme [27] and the three-point
second-order central differencing scheme, respectively. Eq. (19) is
solved by the matrix-free NewtonKrylov method [1].
The intermediate velocity eld U

obtained by solving Eq. (19) is


not divergence-free yet and needs to be corrected to satisfy the
continuity equation. This is accomplished by formulating and solv-
ing the following Poisson equation for the pressure increment (or
pressure correction) P = p
n+1
p
n
:
J
@
@n
i
1
q/
n
i
l
J
@
@n
j
n
j
l
P
J
_ _ _ _

1
Dt
J
@U
j;
@n
j
: 20
Second-order central differencing scheme is used for calculating the
spatial derivatives in the above equation. The maximum number of
nonzero elements in a row of the above Poisson matrix is 19 for
fully nonorthogonal grids. Eq. (20) is solved by using the general-
ized minimal residual (GMRES) or conjugated gradient method with
the algebraic multigrid acceleration [1]. Following the solution of
the above equation, the pressure and contravariant volume uxes
are obtained as follows:
p
n1
p
n
P; 21
U
i;n1
U
i;
JDt
1
q/
n
i
l
J
@
@n
j
n
j
l
P
J
_ _
: 22
The so computed contravariant volume ux obtained from the
above equation is by construction divergence free. When evaluating
the left and right hand sides of Eqs. (20) and (22), respectively, one
needs to interpolate the values of q(/) from the cell centers to the
cell faces, which for the cell face in between (i, j, k) and (i + 1, j, k) can
be implemented as follows:
1
q/

i
1
2
;j;k

1
2
1
q/

i;j;k

1
q/

i1;j;k
_ _
: 23
The values in the other faces can be interpolated in the similar
manner.
3.2. Solution of the turbulence closure equations
The advection and diffusion terms in k-x equations (Eqs. (7)
and (8)) are discretized using the third-order WENO scheme [27]
and the three-point second-order central differencing scheme,
respectively. Eqs. (7) and (8) are discretized in time using the sec-
ond-order backward Euler scheme and solved by the fully implicit
matrix-free NewtonKrylov method [1].
3.3. Solution of level set equations
The equation for the interface motion (Eq. (14)) is discretized in
space by the third-order WENO scheme [27], and the discrete
equation is advanced in time using the second-order RungeKutta
method. The time step is therefore restricted by the CFL (Courant
FriedrichsLewy) condition. Our experience shows that the CFL
number less than 0.5 yields stable solutions.
The spatial derivative of / in the n
1
direction, for instance, can
be discretized as
@/
@n
1

i;j;k

/
i
1
2
;j;k
/
i
1
2
;j;k
Dn
1
: 24
The cell face values, /
i
1
2
;j;k
, in the above equation are calculated by
using the third-order WENO scheme [27]. /
i
1
2
;j;k
, for instance, is
computed as
/
i
1
2
;j;k

c
1
c
1
c
2
/
C
2

/
R
2
_ _

c
2
c
1
c
2

/
L
2

3/
C
2
_ _
; 25
where
26 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
c
1

2
3
1
/
C
/
R

2
; 26
c
2

1
3
1
/
L
/
C

2
27
and
/
L
; /
C
; /
R

i
1
2
;j;k

/
i1;j;k
; /
i;j;k
; /
i1;j;k
U
1
> 0;
/
i2;j;k
; /
i1;j;k
; /
i;j;k
U
1
6 0;
_
28
with
0
= 10
6
. Values of /
i
1
2
;j
1
2
;k
1
2
can be calculated in the same
way, and the temporal derivative term in Eq. (14) can be discretized
by RungeKutta methods.
The gradient term in the mass conserving reinitialization equa-
tion (Eq. (17)) is expressed in generalized curvilinear coordinates
as
jr/j

n
j
x
@/
@n
j
_ _
2
n
j
y
@/
@n
j
_ _
2
n
j
z
@/
@n
j
_ _
2

: 29
The spatial derivatives in Eq. (29) are discretized by the second-
order ENO scheme proposed by Sussman et al. [28]. In this study,
we extend the scheme of Sussman et al. [28] to generalized curvilin-
ear coordinates. The spatial derivative of / in the n
1
direction, for
instance, is given by
@/
@n
1

@/
@n
1

sgn/
0
/
i1;j;k
/
i;j;k
< 0 and
sgn/
0
/
i;j;k
/
i1;j;k
< sgn/
0
/
i1;j;k
/
i;j;k
;
@/
@n
1

sgn/
0
/
i;j;k
/
i1;j;k
> 0 and
sgn/
0
/
i1;j;k
/
i;j;k
> sgn/
0
/
i;j;k
/
i1;j;k
;
1
2
@/
@n
1

@/
@n
1

_ _
otherwise;
_

_
30
where sgn(/
0
) is dened as
sgn/
0

1 /
0
< 0;
0 /
0
0;
1 /
0
> 0
_

_
31
and
@/
@n
1

and
@/
@n
1

are dened as
@/
@n
1

/
i1;j;k
/
i;j;k

1
2
M/
i1;j;k
2/
i;j;k
/
i1;j;k
; /
i2;j;k
2/
i1;j;k
/
i;j;k
;
32
@/
@n
1

/
i;j;k
/
i1;j;k

1
2
M /
i1;j;k
2/
i;j;k
/
i1;j;k
; /
i;j;k
2/
i1;j;k
/
i2;j;k
_ _
:
33
In the above equations, M is dened by
Ma; b
a jaj < jbj;
b jbj 6 jaj:
_
34
Spatial derivatives along the n
2
and n
3
directions are calculated in
the same way as described above.
3.4. Boundary and initial conditions
A critical aspect for simulating free surface ows in open chan-
nel is specication of inow and outow conditions in a manner
that is consistent with the simulated ow regime. In this study,
we simulate cases for which the owat both the inowand outow
is subcritical (Fr < 1), where Fr is the Froude number. The following
boundary conditions are prescribed at the inlet:
u
i
u
inlet
i
x
i
; t;
@P
@n
0;
@/
@n
0 35
and outlet:
@u
i
@n
0; P 0; / /
outlet
x
i
; 36
where n denotes the direction normal to the boundary.
To initialize the simulation, we assume the pressure distribu-
tion to be hydrostatic and zero in the water and air phases,
respectively.
3.5. The CURVIB method
The curvilinear immersed boundary (CURVIB) method [22] can
handle complex geometries including complex embedded moving
boundaries with uidstructure interaction [2931]. The method
has been recently extended by Kang et al. [1] to carry out URANS
and LES of turbulent ows through natural river reaches and by
Khosronejad et al. [32,33] to simulate sediment transport and
scour phenomena in open channels with embedded hydraulic
structures. The details of the method, as they pertain to its applica-
tion to open channel ows, have been described extensively in pre-
vious publications by our group [1,32,33] and for the sake of
brevity only a brief description of the method is given herein.
In the CURVIB method the complex bathymetry of a natural
stream is discretized with an unstructured triangular mesh and
embedded in a background curvilinear grid used to discretize a
channel of regular cross-section that fully contains the streambed
bathymetry. The grid nodes of the curvilinear mesh are classied
based on their location relative to the immersed streambed
bathymetry as: nodes exterior to the ow domain, which are
blanked out of the computation; interior (or uid) nodes where
the governing equations are solved; and immersed boundary (IB)
nodes, which are located in the uid domain but in the immediate
vicinity of the immersed boundary. Boundary conditions for the
velocity components at the IB nodes are specied using a recon-
struction approach. The reconstruction is based either on linear
or quadratic interpolation along the local normal to the boundary
directions [1,22], when the grid is ne enough to resolve the
near-wall viscous layer, or a wall model [1], when the IB nodes
are located into the logarithmic layer. For coupled hydro-morpho-
dynamic simulations, the immersed streambed bathymetry is trea-
ted as a dynamically deforming boundary whose location is
determined by solving the Exner equation coupled with the gov-
erning equations for the ow (see [32] for details). In all previous
applications of the CURVIB method to open channel ows
[1,33,32] the free surface has been treated as a xed rigid lid. In
this work we remove this limitation of the method by coupling it
with the previously described level set approach to simulate
free-surface effects.
In order to implement the level set method in the context of the
CURVIB method, we need to reconstruct boundary conditions for
the level set function (/) at the IB nodes. This is accomplished by
setting the gradient of the level set function at the cell face in be-
tween the uid and IB nodes in the direction along an appropriate
curvilinear grid line to be zero. We illustrate this using the sche-
matic in Fig. 1. For instance, when the grid nodes (i, j, k) and
(i + 1, j,k) are the uid and IB nodes, respectively, the following
relation is assumed at the cell face (the cross symbol in Fig. 1) in
between the uid and IB nodes:
@/
@n
1

i1=2;j;k
0; 37
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 27
which implies the zero Neumann boundary condition along the cor-
responding grid line. The above equation at the IB node is used as
the boundary condition for adjacent uid node for solving the level
set equation (Eq. (14)) and reinitialization equation (Eq. (17)).
3.6. Solution procedure
The overall procedure for solving the governing equations is
summarized as follows.
1. /
n
, U
i,n
, p
n
, k
n
, x
n
are given from the initial condition or the
previous time step (i = 1, 2, 3).
2. The level set advection equation (Eq. (14)) is solved to obtain
/
n+1
.
3. /
n+1
is reinitialized by solving Eq. (17).
4. The velocity at the IB node is reconstructed (see [1] and Sec-
tion 3.5).
5. The momentum equations (Eq. (19)) are solved for U
i,
.
6. The Poisson equation (Eq. (20)) is solved for P.
7. Eq. (21) is solved for p
n+1
.
8. Eq. (22) is solved for U
i,n+1
.
9. SST turbulence model equations (Eqs. (7) and (8)) are solved
for k
n+1
and x
n+1
.
10. l
n1
t
is computed by Eq. (9).
11. Go to 1 to continue to the next time step.
4. Results
In this section, we present results aimed at: (1) validating the
accuracy of the free surface model; and (2) demonstrating its abil-
ity to simulate turbulent ows in open channels with complex
bathymetry. We rst present simulations of the 2D and 3D slosh-
ing of liquid inside a closed tank to verify the accuracy of the level
set method itself. Subsequently, we carry out URANS simulation of
ow past a complex submerged hydraulic structure in a laboratory
ume, and compare the computed water surface proles with
measurements to validate the level set-CURVIB method. Finally,
we demonstrate the applicability of the level set-CURVIB method
for turbulent free surface ow in natural waterways by carrying
out URANS simulation of ow in a eld-scale natural-like mean-
dering stream involving both subcritical and supercritical ow
regimes.
4.1. 2D linear sloshing in a rectangular tank
We consider the linear sloshing of inviscid liquid inside a rect-
angular tank. The computational domain is described in Fig. 2. The
initial free surface elevation is given as
fx D A
0
cos
2px
B
_ _
; 38
where D, A
0
and B denote the mean ow depth, the initial wave
amplitude and the width of the tank, respectively. The values of
D, A
0
and B are set to 1, 0.001 and 1, respectively. When the magni-
tude of the initial wave amplitude relative to the mean ow depth
(A
0
/D) is very small, the behavior of the sloshing motion can be de-
scribed by the linear wave theory and an analytic solution for the
wave amplitude at the center of the tank can be derived as follows
(see [34])
g x
B
2
; t
_ _
A
0
cosx
2
t cosk
2
x; 39
where k
m
mp=B; x
m

gk
m
tanhk
m
D
_
, and g is the acceleration
due to gravity oriented along the negative y-axis and it is set equal
to one (g = 1).
The initial velocity is zero everywhere, and the initial pressure
is hydrostatic in the liquid phase and zero in the gas phase. The
densities of the liquid and gas are set equal to 1 and 0.001, respec-
tively, and the viscosities are neglected. The free-slip boundary
conditions are applied at all boundaries. The number of grid points
in the horizontal and vertical directions are 101 and 101, respec-
tively. While the uniform grid spacing of Dx = 0.01 is employed
in the horizontal direction, stretched grids are used in the vertical
direction. The vertical grid points are clustered near y = 1, and the
minimum and maximum vertical grid spacing are 4 10
4
and
8 10
2
, respectively. The time step used for the computation is
Dt = 2.5 10
3
, and the value of is set to 6 10
4
.
Fig. 3 compares the time history of the computed wave ampli-
tude at the center of the tank with the analytic solution given by
Eq. (39). As seen, the analytic and computed wave amplitudes
are in excellent agreement with each other. The maximum error
of the liquid volume during the 6000 time steps of the computation
is 8 10
4
%.
To quantify the convergence rate of the numerical error for the
wave amplitude, the above problem was solved using two
Fig. 1. Schematic description of imposing the Neumann boundary condition for / at
the interface between the IB and uid nodes.
1
1
X
Y
(0,0)
0.25
A0
Liquid
Gas
Fig. 2. Schematic of the computational domain for the 2D sloshing problem.
28 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
additional grids consisting of 51 51 and 201 201 nodes in the
horizontal and vertical directions, respectively. On all grids the ra-
tio of the time step to the minimum grid spacing was kept constant
(Dt/Dx = 6.25). We dene the L2 norm of the numerical error of the
wave amplitude over the entire domain at each time step as
et
j

1
N
x

Nx
i1
gx
i
; t
j
g
a
x
i
; t
j

D
_ _
2

_
; 40
where N
x
denotes the number of grid nodes in the horizontal direc-
tion, and the subscript a indicates the analytic solution. Instanta-
neous spatial errors at each time step are calculated by the above
equation, and they are subsequently averaged for time T = 2.5 using
the following relation.
E
1
N
t

Nt
j1
fet
j
g
2

_
; 41
where N
t
= T/Dt. The spatio-temporally averaged error (Eq. (41)) is
plotted in Fig. 4 for the three different grids. It is seen that the con-
vergence rate is approximately 1.8 that is slightly lower than sec-
ond order. This is probably because the present level set method
treats the water surface as a diffused interface by introducing the
nite interface thickness using the parameter for numerical
stability.
4.2. 2D nonlinear sloshing in a rectangular tank
In this test case we consider the sloshing inside a rectangular
tank with a relatively larger initial wave amplitude compared to
the previous case. For a sufciently large wave amplitude to depth
ratio, the sloshing motion can no longer be accurately described by
the linear wave theory. With the initial free surface elevation given
by Eq. (38), Wu and Eatock Taylor [34] derived the nonlinear ana-
lytic solution of the wave amplitude (g) up to the second-order by
using the perturbation expansion method. According to this solu-
tion, g at the center of the tank is given by the following equation:
g
B
2
; t
_ _
g
1
B
2
; t
_ _
g
2
B
2
; t
_ _
; 42
where
g
1
B
2
; t
_ _
A
0
cosx
2
t cos
k
2
B
2
_ _
43
and
g
2
B
2
; t
_ _

1
8g
2x
2
A
0

2
cos2x
2
t
A
2
0
x
2
2
k
2
2
g
2
x
4
2
_ _
_

A
2
0
x
2
2
k
2
2
g
2
3x
4
2
_ _
cosx
4
t
_
: 44
The initial wave amplitude (A
0
) is set to 0.01, the initial velocity
is zero everywhere and the initial pressure is given as hydrostatic
and zero in the liquid and gas phases, respectively. The accelera-
tion due to gravity is directed along the negative y axis and its va-
lue is one (g = 1). The densities of the liquid and gas are set to 1 and
0.001, respectively, and the viscosities are neglected. The free-slip
boundary conditions are applied at all boundaries. The number of
grid points in the horizontal and vertical directions are 101 and
136, respectively. Uniform grid spacing of Dx = 0.01 is employed
in the horizontal direction, while stretched grid is used in the ver-
tical direction. The vertical grid points are clustered near y = 1, and
the minimum and maximum vertical grid spacing are 10
3
and
6.7 10
2
, respectively. The time step used for the computation
is Dt = 2.5 10
3
. The value of is set to 1.5 10
3
.
Fig. 5 compares the time history of the computed wave ampli-
tude at the center of the tank with the analytical solution given
by Eq. (42). It is evident from this gure that the wave amplitude
at the crests and troughs varies over time due to the nonlinear ef-
fect. Even for this more complex case, the overall agreement be-
tween the computation and the analytic solution is very good.
The maximum error of the liquid volume during the 6,000 time
steps of the computation is 3 10
3
%.
To quantify the convergence rate of the numerical error for the
wave amplitude, the above problem was solved on two additional
grids consisting of 51 68 and 201 271 nodes in the horizontal
and vertical directions, respectively. The ratio of the time step to
the minimum grid spacing was kept constant (Dt/Dx = 2.5). The
spatio-temporally averaged error (Eq. (41)) for this case is plotted
Time
W
a
v
e

a
m
p
l
i
t
u
d
e
0 2.5 5 7.5 10 12.5 15
-1.0E-03
-5.0E-04
0.0E+00
5.0E-04
1.0E-03
Fig. 3. Comparison of the computed and analytic wave amplitude at the center of
the tank for the 2D linear sloshing problem (symbol: computed solution, solid line:
analytic solution given by Eq. (39)). The computed solution is shown for every seven
time steps.
Nx
E
r
r
o
r
10
1
10
2
10
3
10
-8
10
-7
10
-6
slope = 1.8
Fig. 4. The convergence rate (slope) of the numerical error for the 2D linear
sloshing problem.
Time
W
a
v
e

a
m
p
l
i
t
u
d
e
0 2.5 5 7.5 10 12.5 15
-0.010
-0.005
0.000
0.005
0.010
Fig. 5. Comparison of the computed and analytic wave amplitude at the center of
the tank for the 2D nonlinear sloshing problem (symbol: computed solution, solid
line: analytic solution given by Eq. (42)). The computed solution is shown for every
ten time steps.
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 29
in Fig. 6 for the three different grids. As seen, the rst order conver-
gence rate is observed. The lower convergence rate compared to
the previous, linear case should be attributed to the fact that the
approximate analytic solution given by Eq. (42) involves approx-
imations including the rst-order truncation errors.
4.3. 3D sloshing in a rectangular tank
We consider the 3D sloshing of liquid in a tank width the
dimension of L L and the mean ow depth of D [35]. The initial
free surface elevation is of Gaussian shape given by
fx; y D g
0
x; y; 45
where g
0
is the initial free surface displacement above the still
water elevation (z = D)
g
0
x; y H
0
exp j x
L
2
_ _
2
y
L
2
_ _
2
_ _ _ _
: 46
H
0
is the initial hump height and j is the peak enhancement factor.
For the above initial free surface prole, Wei and Kirby [35] pro-
posed the following linear analytic solution of the free surface
displacement
gx; y; t

1
n0

1
m0
C
nm
e
ixnmt
cos
npx
L
_ _
cos
mpy
L
_ _ _ _
; 47
where i

1
p
and
C
nm

4
1 d
n0
1 d
m0
L
2
_
L
0
_
L
0
g
0
x; y
cos
npx
L
_ _
cos
mpy
L
_ _
dxdy: 48
Each of the (n, m) modes has a natural frequency that is given by
x
nm
gk
nm
tanhk
nm
D; 49
where
k
nm

p
L
_ _
2
n
2
m
2
: 50
In our computation, we use the following parameters: L = 20,
D = 1, j = 0.25 and H
0
= 0.1. The computational domain and the ini-
tial free surface prole are shown in Fig. 7. The initial velocity is zero
everywhere, and the initial pressure is givenas the hydrostatic pres-
sure and zero in the liquid and gas phases, respectively. The accel-
eration due to gravity is directed to the z axis and its value is
g = 9.8. The densities of the liquid and gas are set to 1,000 and 1,
respectively, and the viscosities are neglected. The free-slip bound-
ary conditions are applied at all boundaries. The number of grid
points in the x, y and z directions are 201, 201 and 41, respectively.
Uniform grid spacing of Dx = Dy = 0.1 is employed in the x and y
directions, while stretched grid is used in the z direction. The grid
points in the z direction are clustered near z = 1, and the minimum
and maximum vertical grid spacing are 2 10
2
and 9.5 10
2
,
respectively. The initial hump height (0.1) is resolved by approxi-
mately ve vertical grid nodes. The time step used for the computa-
tion is Dt = 10
3
and the value of is set to 3 10
2
.
Fig. 8 compares the time history of the computed wave ampli-
tude at the center of the tank with the analytic solution given by
Eq. (47). Although the time history of the wave amplitude exhibits
far more complex behavior compared to the previous 2D cases, the
agreement between the analytic and computed solution is excel-
lent. The maximum error of the liquid volume during the 60,000
time steps of the computation is 2 10
4
%, which is satisfactory.
Through the above three test cases, we demonstrated the accu-
racy of the level set method for free surface ow simulations. In
what follows, we apply the method to simulate real life turbulent
ow.
4.4. 3D free surface ow past a rock structure
We consider turbulent ow in a straight open channel with an
embedded rock structure, namely a cross vane typically used in
stream-restoration projects for preventing streambank erosion
and increasing ow diversity [36]. Experiments for this case were
carried out in a St. Anthony Falls Laboratory (SAFL) ume (see
Fig. 9). The SAFL ume is 0.9 m wide and 12 m long, and the
Nx
E
r
r
o
r
10
1
10
2
10
3
10
-7
10
-6
10
-5
slope = 1.0
Fig. 6. The convergence rate (slope) of the numerical error for the 2D nonlinear
sloshing problem.
Fig. 7. Schematic of the computational domain and the initial free surface prole
for the 3D sloshing problem (not to scale).
Time
W
a
v
e

a
m
p
l
i
t
u
d
e
0 10 20 30 40 50 60
-0.100
-0.050
0.000
0.050
0.100
Fig. 8. Comparison of the computed and analytic wave amplitude at the center
(10, 10) of the tank for the 3D sloshing problem (symbol: computed solution, solid
line: analytic solution given by Eq. (47)). The computed solution is shown for every
hundred time steps.
30 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
tailgate weir is adjusted to control the downstream water eleva-
tion. The mean water surface prole at several locations was mea-
sured using Massa ultrasonic distance sensor. While the
sidewalls are smooth, the channel bed is rough with the equivalent
sand grain roughness of 7 10
3
m. The cross vane is installed at
6 m downstream from the inlet of the ume. The ow discharge
is 3.81 10
2
m
3
/s, which corresponds to a Reynolds number of
4.08 10
4
and a Froude number of 0.19 based on the mean ow
depth (0.17 m) and velocity (0.24 m/s). The 3D geometry of the
cross vane in the ume was scanned using the high-resolution la-
ser topography scanner at sub-millimeter resolution. The mea-
sured topography was used to reconstruct the IB mesh for the
numerical simulation.
Fig. 10 shows the computational domain and the immersed
body reconstructed from the measured topography of the cross
vane. The x coordinate denotes the distance from the inlet of the
laboratory ume. The computational domain is 6 m long, 0.9 m
wide and 0.32 m high, and it is discretized with of 191 151
121 nodes in x, y and z directions, respectively. As seen in the g-
ure, the inlet and outlet of the computational domain are located
3 m and 9 m downstream of the inlet of the laboratory ume.
The unstructured IB surface mesh used to discretize the surface
of the cross vane consists of 21,970 triangular elements. The back-
ground grids are stretched along all three spatial directions. The
size of the grid spacing near the side walls (y = 0, y = 0.9 m) and
the channel bed (z = 0) are 6.8 10
4
m and 1.5 10
3
m, respec-
tively. The size of the vertical grid spacing near the free surface is
10
3
m. The maximum grid spacing in the transverse, vertical and
streamwise directions is 1.35 10
2
m, 1.8 10
2
m and
1.2 10
1
m, respectively. The computational grid near the cross
vane is shown in Fig. 11. While the discharge of water at the inlet
(x = 3 m) is xed in time, the free surface at the inlet is allowed to
move in the vertical direction as a result of the boundary condi-
tions described in Section 3.4 above. To preserve a constant water
discharge for temporally varying free surface elevation, at each
time step the velocity at the inlet is adjusted so that the volume
ux of the water matches the measured discharge (3.81
10
2
m
3
/s). At the outlet boundary (x = 9 m), the level set function
corresponding to the measured free surface elevation (z = 0.172 m)
is prescribed. The reference elevation is located at the bed of the
outlet (z = 0). The noslip velocity boundary condition is applied at
the sidewalls, and the following logarithmic velocity prole [37]
for rough wall is assumed near the channel bed.
u
t
u
s
2:5 log
d
wall
k
s
_ _
8:5; 51
where u
t
and u
s
denote the velocity component tangent to the wall
at the rst off-wall node and the wall shear velocity, respectively, k
s
(=7 10
3
m) is the size of equivalent sand grain roughness, and
d
wall
is the distance from the rst off-wall node to the wall. The wall
function for the smooth surface [1] is employed to reconstruct the
velocity components at the IB nodes near the cross vane. The den-
sity and dynamic viscosity of the water are 1000 kg/m
3
and
10
3
kg/(m s), and those of the air are 1.2 kg/m
3
and 1.8 10
5
kg/(m s), respectively. The acceleration due to gravity is set to
g = 9.8 m/s
2
. The time step used for the computation is Dt =
5 10
4
s, and the value of is set to 1.5 10
3
m. The computa-
tion was run for 100,000 time steps.
Fig. 12 shows the contour plots of the computed streamwise
velocity, free surface elevation and local Froude number at the
water surface. The local Froude number is calculated from the local
depth-averaged velocity and the ow depth. We can observe that
the mean streamwise velocity in the center region of the wake of
the cross vane is signicantly increased due to the effect of the rock
Fig. 9. Plan view of the cross vane installed in a St. Anthony Falls Laboratory ume
(image courtesy of Craig Hill). Flow is from left to right.
Fig. 10. 3D view of the computational domain with the initial free surface. Flow is from x to +x.
Fig. 11. Computational grids near the cross vane. Flow is from x to +x.
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 31
structures. We also see the increased free surface elevation in the
upstream of the structure and the regions of the low free surface
elevation directly above the rocks near the sidewalls. These low
free surface elevation demonstrate the ability of the method to re-
solve complex local effects as the ow transitions from subcritical
(Fr < 1) to supercritical (Fr > 1) as a result of the very shallow ow
depth near the sidewalls of the channel to which the rocks are at-
tached (see Fig. 12(c)). In Figs. 13 and 14, we compare the com-
puted water surface prole with the one measured along
transverse and streamwise directions, respectively. Both the com-
puted transverse and streamwise water surface proles show very
good agreement with the measured proles, thus clearly establish-
ing the ability of the model to correctly predict the 3D free surface
deformation due to the presence of an complex hydraulic
structure.
4.5. 3D free surface ow in a meandering stream
To demonstrate the ability of the method to simulate turbulent
free surface ows in natural streams with complex bathymetry, we
apply it to the meandering stream currently installed in the St. An-
thony Falls Laboratory Outdoor StreamLab (OSL), University of
Minnesota, Minneapolis, MN, USA. This research facility is a 40 m
by 20 m basin which has been congured into a sand-bed mean-
dering stream channel. More details of the OSL can be found in
Kang et al. [1]. Kang et al. [1] performed LES and URANS simulations
Fig. 12. Contour plots of the computed (a) streamwise velocity; (b) free surface elevation; and (c) local Froude number at the water surface (/ = 0). Flow is from x to +x.
32 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
of the ow in the OSL by prescribing the measured, xed water sur-
face at the top of the boundary, and the computed mean velocity
and turbulence kinetic energy agreed very well with the measure-
ments. In this study, we seek to predict the free surface elevation in
the OSL at the bankfull ow condition with the discharge of
2.85 10
1
m
3
/s. For this condition the channel was approxi-
mately 3 m wide and 0.3 m deep, and the Reynolds and Froude
numbers, based on the depth and velocity at the inlet, are approx-
imately equal to 10
5
and 0.4, respectively. The mean slope of the
channel bed was about 0.01. Bed topography in the OSL was col-
lected on a 0.01 m horizontal grid at sub-millimeter vertical accu-
racy using a laser distance sensor (Keyence LK-G series) and a
ultrasonic submersible transducer system (JSR Ultrasonics) (see
[33] for details).
Using the measured bathymetry data, we reconstructed the im-
mersed boundary surface and discretized it with an unstructured
mesh consisting of 472,736 triangular elements. Fig. 15 shows
the reconstructed IB mesh and the background computational
mesh. This gure shows the bed topography measured at the
Y [m]
E
l
e
v
a
t
i
o
n

[
m
]
E
l
e
v
a
t
i
o
n

[
m
]
E
l
e
v
a
t
i
o
n

[
m
]
E
l
e
v
a
t
i
o
n

[
m
]
0 0.3 0.6 0.9
0.15
0.16
0.17
0.18
0.19
Y [m]
0 0.3 0.6 0.9
0.15
0.16
0.17
0.18
0.19
Y [m]
0 0.3 0.6 0.9
0.15
0.16
0.17
0.18
0.19
Y [m]
0 0.3 0.6 0.9
0.15
0.16
0.17
0.18
0.19
(a) X=5.0 m
(c) X=6.2 m
(b) X=5.6 m
(d) X=6.8 m
Fig. 13. Comparison of the computed and measured transverse water surface proles (symbol: measurements, solid line: computation).
X [m]
X [m]
E
l
e
v
a
t
i
o
n

[
m
]
E
l
e
v
a
t
i
o
n

[
m
]
3 4 5 6 7 8 9
0.15
0.16
0.17
0.18
0.19
3 4 5 6 7 8 9
0.15
0.16
0.17
0.18
0.19
(a) Y=0.3 m
(b) Y=0.6 m
Fig. 14. Comparison of the computed and measured streamwise water surface
proles (symbol: measurements, solid line: computation).
Fig. 15. Plan view of the computational domain. Solid lines denote the boundaries of the background grid, contour plots show the immersed body reconstructed from the
measured bathymetry, and symbols show the water surface measurement locations. Flow direction is from right to left.
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 33
quasi-equilibrium state with the given constant discharge and
clearly illustrates the deposition of sediments along the inner bank
and the deep scour along the outer bank of the meander bend due
to the secondary ow. The curvilinear background grid (solid lines
in Fig. 15) is generated so that it fully contains the IB mesh and is
discretized with 51 96 191 nodes in the transverse, vertical
and streamwise directions, respectively. The grid spacing in the
transverse, vertical and streamwise directions is approximately
0.08 m, 0.01 m and 0.25 m, respectively. An unsteady uniform
velocity is given at the inlet boundary. As in the previous test case,
the free surface elevation at the inlet is allowed to move and the
inlet velocity is adjusted so that the volume ux in the water phase
of the inlet plane matches the measured discharge ( 2.85
10
1
m
3
/s) at each time step. At the outlet boundary, the level
set function corresponding to the measured water surface eleva-
tion (0.237 m) is prescribed. The reference elevation is the bed ele-
vation at the outlet. The channel bed of the OSL consists of large
size gravels and ne sands. In this computation, however, the effect
of the bed roughness is neglected because it is difcult to parame-
terize the effect of the heterogeneous roughness elements. More-
over, it is reasonable to assume that the free surface elevation in
a eld scale stream is largely determined by the large scale varia-
tion of the bed topography, and the effect of the bed roughness
should be expected to be relatively negligible. The wall model for
the smooth wall [1] is employed to interpolate the velocities at
the IB nodes. The acceleration due to gravity is set to g = 9.8 m/
Fig. 16. Contour plots of the computed (a) ow speed; (b) free surface elevation; and (c) local Froude number at the water surface (/ = 0). Flow is from right to left.
34 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336
s
2
. The time step used for the computation is Dt = 2 10
3
s, and
the value of is set to 1.5 10
2
m. The computation was run
for 90,000 time steps.
Fig. 16 shows the contour plots of the computed mean velocity
magnitude, free surface elevation and local Froude number at the
water surface. The local Froude number is calculated from the lo-
cal depth-averaged velocity and the ow depth. High velocity is
observed in the straight part of the channel where the cross sec-
tional area is small. As the ow enters the bend the high velocity
core impinges on the outer streambank. The model predicts the
superelevation of the water surface along the outer bank of the
two meander bends, and also predicts the transitions from sub-
critical (Fr < 1) to supercritical (Fr > 1) ows in the straight reach
of the channel caused by the change of the bathymetry. Fig. 17
compares the measured and computed free surface elevation at
the seven cross sections shown in Fig. 15. Fig. 17(a)(c) compares
the streamwise free surface prole along the inner bank, center
region and outer bank, respectively. As seen in these gures,
the model correctly predicts the higher elevations in the cross
Sections 35 along the outer bank, which is due to the superele-
vation in the bend. In all cross sections, the agreement is satisfac-
tory in spite of the coarse spatial resolution. This test case shows
that the level set-CURVIB method with a RANS turbulence model
is able to predict the free surface elevation over complex topogra-
phy with reasonable accuracy.
5. Conclusions
We developed and validated a novel numerical model for simu-
lating 3D, turbulent free surface ow in natural waterways with
complex hydraulic structures. The accuracy of the free surface
model was veried by solving the 2D and 3D sloshing problems
for which the analytic solutions are available. The ability of the
method to simulate ows past complex hydraulic structures was
demonstrated by applying it to carry out RANS simulation of tur-
bulent free surface ow past the highly complex rock structures in-
stalled in the laboratory ume. Finally, in order to demonstrate the
capability of the model for simulating ows in natural waterways,
RANS simulation was carried out for turbulent free surface ow
over the complex topography of a eld scale meandering channel.
The computed free surface elevations showed good agreement
with the measurements. The simulations of ow past rock struc-
tures (Section 4.4) and over the complex bathymetry (Section
4.5) also demonstrate that the method is able to capture the tran-
sition between subcritical (Fr < 1) and supercritical (Fr < 1) ows
without any special treatment.
A major conclusion of our work is that the level set-CURVIB
method we developed herein along with a RANS turbulence model
can capture the free surface elevation of complex hydraulic engi-
neering ows with good accuracy. This is an important nding as
it suggests a computationally expedient approach for carrying
out high-resolution LES of ows in natural waterways. In Kang
et al. [1] and Kang and Sotiropoulos [33] we carried out LES by pre-
scribing the free surface elevation from coarse experimental mea-
surements and treating it as a xed rigid lid. Our ndings suggest
that the need for using experimental data as input to the LES model
can be eliminated by rst running the present level set-CURVIB
method in RANS mode and prescribing the so-computed mean free
surface elevation as the rigid lid of the LES computation. Obviously
this approach does not take into account the instantaneous uctu-
ations of the free surface that can be resolved by the LES and is thus
only applicable in situations where such uctuations are small and
can be neglected. It should be noted, however, that the model we
developed herein is in principle applicable to carry out fully cou-
pled LES of free surface ows. The main modeling advance that
would be required in order to accomplish such undertaking is
the development of a subgrid scale model for the unresolved uc-
tuations of the level set function. All these are worthy topics for re-
search which we intend to pursue as part of our future work.
Acknowledgments
This work was supported by NSF grants EAR-0120914 (as part of
the National Center for Earth-Surface Dynamics) and EAR-0738726
and a grant from Yonsei University, South Korea Computational re-
sources were provided by the University of Minnesota Supercom-
puting Institute. We are grateful to Craig Hill for collecting the free
surface prole for the test cases presented in Sections 4.4 and 4.5.
References
[1] Kang S, Lightbody A, Hill C, Sotiropoulos F. High-resolution numerical
simulation of turbulence in natural waterways. Adv Water Resour
2011;34:98113.
Cross section
E
l
e
v
a
t
i
o
n

[
m
]
1 2 3 4 5 6 7
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
Cross section
E
l
e
v
a
t
i
o
n

[
m
]
1 2 3 4 5 6 7
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
Cross section
E
l
e
v
a
t
i
o
n

[
m
]
1 2 3 4 5 6 7
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
0.4
0.5
(a) outer bank
(b) channel center
(c) inner bank
Fig. 17. Comparison of the computed (blue line with circle) and measured (red line
with triangle) free surface elevation at the cross sections (gray line with square: bed
elevation). The locations of the cross sections are shown in Fig. 15. (For interpre-
tation of the references to colour in this gure legend, the reader is referred to the
web version of this article.)
S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336 35
[2] Garcia R, Kahawita RA. Numerical solution of the St. Venant equations with the
MacCormack nite-difference scheme. Int J Numer Methods Fluids
1986;6:25974.
[3] Molls T, Chaudhry MH. Depth-averaged open-channel ow model. J Hydraul
Eng 1995;121:45365.
[4] Yoon TH, Kang S-K. Finite volume model for two-dimensional shallow water
ows on unstructured grids. J Hydraul Eng 2004;130:67888.
[5] Lo DC, Young DL. Arbitrary LagrangianEulerian nite element analysis of free
surface ow using a VelocityVorticity formulation. J Comput Phys
2004;195:175201.
[6] Ramaswamy B, Kawahara M. Arbitrary LagrangianEulerian nite element
method for unsteady, convective, incompressible viscous free surface uid
ow. Int J Numer Methods Fluids 1987;7:105375.
[7] Hodges BR, Street RL. On simulation of turbulent nonlinear free-surface ows. J
Comput Phys 1999;151:42557.
[8] Unverdi SO, Tryggvason G. A front-tracking method for viscous,
incompressible, multi-uid ows. J Comput Phys 1992;100:2537.
[9] Shin S, Juric D. Modeling three-dimensional multiphase ow using a level
contour reconstruction method for front tracking without connectivity. J
Comput Phys 2002;180:42770.
[10] Tryggvason G, Bunner B, Esmaeeli A, Juric D, Al-Rawahi N, Tauber W, et al. A
front-tracking method for the computations of multiphase ow. J Comput Phys
2001;169:70859.
[11] Hirt CW, Nichols BD. Volume of uid (VOF) method for the dynamics of free
boundaries. J Comput Phys 1981;39:20125.
[12] Osher S, Sethian JA. Fronts propagating with curvature-dependent speed:
algorithms based on HamiltonJacobi formulations. J Comput Phys
1988;79:1249.
[13] Youngs DL. Time-dependent multi-material ow with large uid distortion. In:
Morton KW, Bianes MJ, editors. Numerical methods for uid dynamics. New
York: Academic Press; 1982.
[14] Chen Q, Dai G, Liu H. Volume of uid model for turbulence numerical
simulation of stepped spillway overow. J Hydraul Eng 2002;128:6838.
[15] Lin P, Liu PL-F. A numerical study of breaking waves in the surf zone. J Fluid
Mech 1998;359:23964.
[16] Ramamurthy AS, Qu J, Vo D. Numerical and experimental study of dividing
open-channel ows. J Hydraul Eng 2007;133:113544.
[17] Yue W, Lin C-L, Patel VC. Numerical simulation of unsteady multidimensional
free surface motions by level set method. Int J Numer Methods Fluids
2003;42:85384.
[18] Yue W, Lin C-L, Patel VC. Large eddy simulation of turbulent open-channel
ow with free surface simulated by level set method. Phys Fluids
2005;17:025108.
[19] Carrica PM, Wilson RV, Stern F. An unsteady single-phase level set method for
viscous free surface ows. Int J Numer Methods Fluids 2007;53:22956.
[20] Yang J, Stern F. Sharp interface immersed-boundary/level-set method for
wave-body interactions. J Comput Phys 2009;228:6590616.
[21] Stoesser T, Wilson CAME, Bates PD, Dittrich A. Application of a 3D numerical
model to a river with vegetated oodplains. J Hydroinform 2003;5:99112.
[22] Ge L, Sotiropoulos F. A numerical method for solving the 3D unsteady
incompressible NavierStokes equations in curvilinear domains with complex
immersed boundaries. J Comput Phys 2007;225:1782809.
[23] Osher S, Fedkiw R. The Level Set Method and Dynamic Implicit Surfaces. New
York: Springer-Verlag; 2002.
[24] Menter FR, Kuntz M, Langtry R. Ten years of industrial experience with the SST
turbulence model. Turbul Heat Mass Transf 2003;4:62532.
[25] Sussman M, Smereka P, Osher S. A level set approach for computing solutions
to incompressible two-phase ow. J Comput Phys 1994;114:14659.
[26] Sussman M, Fatemi E. An efcient, interface-preserving level set redistancing
algorithm and its application to interfacial incompressible uid ow. SIAM J
Sci Comput 1999;20:116591.
[27] Jiang G-S, Shu C-W. Efcient implementation of weighted ENO schemes. J
Comput Phys 1996;126:20228.
[28] Sussman M, Fatemi E, Smereka P, Osher S. An improved level set method for
incompressible two-phase ows. Comput Fluids 1998;27:66380.
[29] Borazjani I, Sotiropoulos F. Numerical investigation of the hydrodynamics of
carangiform swimming in the transitional and inertial ow regimes. J Expt Biol
2008;211:154158.
[30] Borazjani I, Sotiropoulos F. On the role of form and kinematics on the
hydrodynamics of self-propelled body/caudal n swimming. J Expt Biol
2010;213:89107.
[31] Borazjani I, Sotiropoulos F, Malkiel E, Katz J. On the role of copepod antenna in
the production of hydrodynamic force during hopping. J Expt Biol
2010;213:301935.
[32] Khosronejad A, Kang S, Borazjani I, Sotiropoulos F. Curvilinear immersed
boundary method for simulating coupled ow and bed morphodynamic
interactions due to sediment transport phenomena. Adv Water Resour
2011;34:82943.
[33] Kang S, Sotiropoulos F. Flow phenomena and mechanisms in a eld-scale
experimental meandering channel with a pool-rife sequence: insights gained
via numerical simulation. J Geophys Res 2011;116:F03011.
[34] Wu GX, Eatock Taylor R. Finite element analysis of two-dimensional non-
linear transient water waves. Appl Ocean Res 1994;16:36372.
[35] Wei G, Kirby JT. Time-dependent numerical code for extended Boussinesq
equations. J Waterway Port Coastal Ocean Eng 1995;121:25161.
[36] Rosgen DL. The cross vane, W-weir and J-hook structures: their description,
design and application for stream stabilization and river restoration. In:
Proceedings, Wetland Engineering and River Restoration Conference CD-ROM,
ASCE, Reston, VA; 2001.
[37] Schlichting H. Boundary layer theory. New York: McGraw-Hill; 1987.
36 S. Kang, F. Sotiropoulos / Advances in Water Resources 40 (2012) 2336

Вам также может понравиться