You are on page 1of 10

AIAA JOURNAL

Vol. 49, No. 12, December 2011

Near-Wall Formulation of the Partially Averaged


NavierStokes Turbulence Model

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

Branislav Basara
AVL List GMBH, 8010 Graz, Austria
Sinia Krajnovi
Chalmers University of Technology, 412 96 Gothenburg, Sweden
Sharath Girimaji
Texas A&M University, College Station, Texas 77843
and
Zoran Pavlovic
AVL List GmbH, 2000 Maribor, Slovenia
DOI: 10.2514/1.J050967
The variable-resolution partially averaged NavierStokes bridging strategy is applied to the four-equation k-"--f
turbulence model. In this approach, the popular two-equation model is enhanced with an additional transport
equation for the velocity scale ratio  and an equation for the elliptic relaxation function f for the purpose of improved
near-wall behavior. By using the elliptic relaxation technique to model the wall blocking effect, the new four-equation
partially averaged NavierStokes model retains the simplicity of the previous two-equation partially averaged
NavierStokes versions but signicantly improves predictions in the near-wall region. The proposed partially
averaged NavierStokes k-"--f model is evaluated in a turbulent channel ow and ow around a three-dimensional
circular cylinder mounted vertically on a at plate. The results clearly show benets of the improved near-wall
modeling and extend partially averaged NavierStokes applicability to a broader range of smooth bluff-body
separated ows.
Superscript

Nomenclature
aij
C, c
fe
fk
k
P
p
Re
sk , se , sz
t
U
u
u
V
v 2
"


t

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

normalized anisotropy tensor


model constants with various subscripts
unresolved-to-total-dissipation-rate ratio
unresolved-to-total-kinetic-energy ratio
turbulent kinetic energy
production of k
pressure
Reynolds number based on friction velocity
model constants
time
resolved mean velocity
velocity uctuation
friction velocity
instantaneous mean velocity
wall-normal velocity scale
dissipation rate of k
velocity scale ratio, v2 =k
molecular kinematic viscosity
turbulent eddy viscosity

I. Introduction
HILE the notion of an accuracy-on-demand turbulence
computation paradigm is many decades old, the origin of the
current arbitrary cutoff length scale closure modeling trend can be
traced to the work of Speziale [1]. This bridging closure model is
purported to provide the best possible physical delity on any given
numerical grid while varying seamlessly in form and function
between the Reynolds-averaged NavierStokes (RANS) model and
direct numerical simulation (DNS). With a similar objective, Spalart
et al. [2] proposed the idea of combining RANS and large eddy
simulation (LES) in different zones of a single ow calculation to
circumvent the deciencies of each method while maximizing
individual benets giving rise to the detached eddy simulation (DES)
(Spalart et al. [2]). A widely accepted terminology for categorizing
these models has not yet evolved, and this paper will follow the
classication guideline presented in Sagaut et al. [3]. The superset of
all these smart models will be called the variable-resolution (VR)
methods. The VR methods are composed of two important subsets,
bridging (Speziale [1]) and zonal (Spalart et al. [2]) methods.
Currently, each approach is evolving along distinctly different lines.
These VR models/approaches ll a crucial void in practical
turbulence computational tool box, which, before the advent of these
approaches, offered only two choices: the computationally inexpensive but possibly inaccurate RANS on one hand and the accurate
but computation-intensive LES/DNS on the other. Currently, several
zonal strategies, also called hybrid RANS/LES models, are in use,
with the main distinctions among them being 1) the rules according
to which the RANS and LES zones are partitioned and 2) the manner
in which the two zones interact and interface with one other. On the
bridging model side, after some initial variations (Lakshmipathy and
Girimaji [4]), different proposals are gravitating toward a common
turbulence model closure form. Two of the most developed bridging
methods are the partially integrated transport model (PITM)
(Chaouat and Schiestel [5]) and partially averaged NavierStokes

Subscript

= normalized quantity by inner variables

= unresolved quantity

Received 30 September 2010; revision received 18 May 2011; accepted for


publication 31 May 2011. Copyright 2011 by the American Institute of
Aeronautics and Astronautics, Inc. All rights reserved. Copies of this paper
may be made for personal or internal use, on condition that the copier pay the
$10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood
Drive, Danvers, MA 01923; include the code 0001-1452/11 and $10.00 in
correspondence with the CCC.

Chief CFD Developer, Department of Advanced Simulation Technologies, Hans List Platz 1; branislav.basara@avl.com.

Associate Professor, Department of Applied Mechanics, Division of Fluid


Dynamics; sinisa@chalmers.se.

Professor, Department of Aerospace Engineering, 607B H.R. Bright


Building; girimaji@aero.tamu.edu. Associate Fellow AIAA.

Analysis Engineer, Department of Advanced Simulation Technologies,


Trg Leona Stuklja 5; zoran.pavlovic@avl.com.
2627

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

2628

BASARA ET AL.

(PANS) (Girimaji et al. [68]] approaches. The PITM closure


derivation is based on energy spectrum decomposition and the PANS
model is obtained from a parent RANS closure by applying the
averaging-invariance principle. Yet, the nal closure models forms in
the two methods are very similar lending corroboration for each
approach. Between PANS and PITM research, several important
bridging model closure attributes have been established: consistency
with the Kolmogorov zeroth law (Chaouat and Schiestel [9]); correct
xed-point behavior and energetics as a function of cutoff wave
number (Girimaji et al. [8]); and recovery of the Smagorinsky model
in the limit of cutoff approaching the dissipative scale (Chaouat and
Schiestel [10] and Fadai-Ghotbi et al. [11]). It has been shown that a
posteriori PANS computed unresolved eddy viscosity and kinetic
energy are consistent with a priori specications. Furthermore, the
PANS unresolved turbulent transport model has been developed
from rst principles (Murthi et al. [12]). At this stage of development,
it is reasonable to claim that the theoretical underpinnings of PANS/
PITM bridging approaches are adequately established. From a
practical viewpoint, both bridging and zonal models appear useful,
but some important deciencies remain.
One of the areas in which both the bridging and zonal models
require crucial improvements is the near-wall behavior in complex
ows. High-delity near-wall modeling is important for many
industrial applications, especially aerodynamic attached and separated ows. The closure model delity requirements for at plate
boundary layers and sharp-edge separation are less stringent than for
smooth bluff-body separated ows. In sharp-edge ows, separation
location is uniquely determined by geometry, whereas, in smooth
bluff bodies, ow features such as pressure gradients and streamline
curvature are responsible for separation. For bluff-body ows, a very
ne numerical grid is needed for a wall-resolved LES calculation and
most industrial computations cannot meet the stringent requirements. The bridging methods of PITM and PANS offer computational relief. It has been established elsewhere that PANS/PITM
perform adequately well at reasonable resolution for simple at plate
boundary layers (e.g., see PITM calculations in Chaouat and
Schiestel [5]) and sharp-edge separated ows (e.g., PANS calculations of Jeong and Girimaji [13] and Song and Park [14]). While
PANS, in its current form, can be considered to be adequate for
simple attached boundary layers and sharp-edge ow separation, its
performance in smooth bluff-body separated ows needs much
improvement (Lakshmipathy and Girimaji [15]). For such ows,
current PANS and other VR methods predict late separation in ow
around two-dimensional and three-dimensional (3D) cylinders, as
shown in the works of Girimaji [7] and Krajnovi and Basara [16],
despite the fact that some ne-scale ow structures are captured
accurately. In summary, there exists a clear and imminent need to
improve the near-wall performance of PANS model for smooth bluffbody separation.
The objective of this paper is to enhance, beyond the level
incumbent in current PANS and PITM, the near-wall physical delity
of the bridging closure by including three additional physical
features: 1) rendering the near-wall eddy viscosity a function of wallnormal uctuations rather than the total kinetic energy; 2) taking into
account the inviscid wall blocking effect; and 3) incorporating low
Reynolds effects prevalent in the viscous and buffer sublayers.
Toward this end, we develop the PANS k-"--f turbulence model by
transposing the enhanced near-wall description incumbent in the
RANS k-"--f model (Hanjalic et al. [17]).
The RANS k-"--f model is a variant of the v2 -f model (Durbin
[18]), the difference being that a transport equation for the wallnormal velocity scale ratio  (v 2 =k, where k is turbulent kinetic
energy) is included rather than for the velocity scale v 2 . Additionally,
an elliptic relaxation equation for f, which is a parameter closely
related to the pressure-strain redistribution term, is solved. In this
approach, elliptic relaxation physics (Durbin [18]) is retained to
sensitize the wall-normal turbulence intensity v2 to the inviscid wall
blocking effect. The f equation is more robust at the wall as
fwall  y2 . In addition, the source term in the f equation is the
production P of turbulent kinetic energy, whereas the corresponding
term in the v2 equation is the less tractable dissipation ". There are

advantages to be gained from a numerical viewpoint as well:  is


clearly bounded in the interval (0, 2), whereas v2 is not. Importantly,
this model can be used in conjunction with the universal wall
approach, which combines the integration up to the wall with wall
functions as shown in Popovac and Hanjalic [19] and Basara [20,21].
All these attributes make the k-"--f closure approach ideally suited
for serving as the basis of near-wall PANS enhancement.
The reminder of the paper is arranged as follows. In Sec. II, the
PANS k-"--f model is presented from rst principles. The
numerical implementation and computation issues are presented in
Sec. III. In Sec. IV, PANS results are compared with available data for
1) turbulent channel ow and 2) ow past a nite cylinder mounted
vertically on a at plate. Finally, Sec. V concludes with a short
discussion.

II.

Mathematical Model

Consider the decomposition of a turbulent velocity Vi eld into


two parts by an arbitrary lter:
Vi  U i  u i

(1)

where Ui and ui are the resolved and unresolved elds. If the lter is
commutative with the spatiotemporal differential operator, then the
resolved eld evolves according to Germano [22]:
@Vi ; Vj 
@U
1 @p
@ 2 Ui
@Ui
 Uj i 


@x

@x
@x
@x
@t
j
j
i
j @xj

(2)

This is called the PANS equation (Girimaji [7]). Here, Vi ; Vj  is


the generalized second moment (Germano [22]) and represents the
effect of the unresolved motion on the resolved eld. At this stage,
there are several choices for the model form of the constitutive
relationship. The various options in order of decreasing sophistication are: 1) unresolved stresses can be obtained from an evolution
equation, as in Reynolds stress closure models; 2) a Boussinessq
constitutive relation can be used and the requisite length and velocity
scales be obtained from evolution equations; and 3) the algebraic
Smagorinsky, dynamic or otherwise, can be used. Since the
purported use of the PANS method is at cutoffs in the large inertial
scales, we deem Smagorinsky closure to be too elementary. While
solving evolution equations for the unresolved stresses may be
possible, we choose the Boussinessq constitutive relation in keeping
with computational uid dynamics (CFD) industrial practice. For
further discussions on this choice, the reader is referred to Girimaji
[7]. We now proceed with the Boussinessq constitutive model for the
second-order central moment:
2
Vi ; Vj   2u Sij  ku ij
3

(3)

where ku and "u are the unresolved kinetic energy, and the dissipation
and the eddy viscosity of unresolved scales is given by Girimaji [7] as
u  c

k2u
"u

(4)

The resolved stress tensor is given by


Sij 



1 @Ui @Uj

2 @xj
@xi

(5)

The model equations for the unresolved kinetic energy ku and the
unresolved dissipation "u are required to close the system of equation
given previously. These equations are derived in Girimaji [7] and we
only present the nal form:

2629

BASARA ET AL.



@k
@ u @ku
@ku
 Uj u  Pu  "u 
@xj ku @xj
@xj
@t


@"
"
"2
@ u @"u
@"u
 u
 Uj u  C"1 Pu u  C"2

@xj
ku
ku @xj "u @xj
@t

(7)

fk
C  C"1 ;
f" "2

ku  k

fk2
;
f"

"u  "

fk2
f"
(8)

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

where the unresolved-to-total ratios of kinetic energy and dissipation


are written, respectively, as
k
fk  u ;
k

"
f"  u
"

(11)

We can derive the unresolved velocity scale ratio u  v2u =ku by


using Eq. (11):
Du Dv2u =ku  Dfv2 v2 =fk k fv2 1 Dv2  fv2 v 2 Dk




Dt
Dt
fk k Dt
fk k2 Dt
Dt
(12)

The model coefcients are



 C"1 
C"2

v 2
Dv 2
 kf  "
k
Dt

(6)

(9)

Girimaji [7] showed that by varying these parameters the implied


cutoff can be placed in any part of the spectrum including the
dissipation range. The PANS asymptotic behavior goes smoothly
from RANS to DNS with decreasing fk . The PANS energetics,
characterized by the unresolved production-to-dissipation ratio,
behaves as required at intermediate cutoff length scales. The
production-to-dissipation ratio varies gradually from RANS value to
DNS value as fk is reduced from one to zero (Girimaji et al. [8]).
The unresolved-to-total-kinetic-energy ratio fk depends upon the
grid spacing as follows (Girimaji and Abdul-Hamid [23]):
 2=3
1

k3=2
;

(10)
fk  p
c 
"
where  is the geometric-average grid cell dimension (  x
y z 1=3 ) and  is the integral length scale of turbulence. When
the grid size is much smaller than the local Taylor length scale, more
and more of the turbulent kinetic energy can be resolved. This is
similar in spirit to DNS, wherein the grid spacing should be of the
same size of the Kolmogorov scale to guarantee adequate resolution.
It is important to note that the fk   relation is an inequality. The
implication is that the model cutoff can be larger than the numerical
grid but not smaller. While the model derivation is performed with
the assumption that fk is a spatiotemporal constant, in practice fk is
allowed to vary slowly in space and time for most efcient use of the
grid. Basara et al. [24], introduce fk as a dynamic parameter in the
computational procedure. Dynamic change in fk is permissible
provided that its time scale of change is smaller than those of
turbulence quantities.
One of the most important challenges in all VR methods is the
accurate description of near-wall turbulence behavior. In zonal
methods, the log law is not generally recovered if the transition from
RANS to LES occurs in the near-wall equilibrium region. Further
treatment of this handshake (RANSLES interface) zone is
required to recover reasonable results. On the contrary, with the
bridging methods, the closure models vary seamlessly and all aspects
of near-wall behavior in the parent RANS model can be reasonably
transferred to the corresponding PANS closure without any
additional effort (Chaouat and Schiestel [5]). Most of the PANS
models in the literature are derived from two-equation parent RANS
models. While these RANS models capture the log law and other
elementary aspects of near-wall behavior, they do not adequately
embody more sophisticated aspects of boundary layers that are
crucial for computing smooth bluff-body separation. This inadequacy is reected in the corresponding PANS/PITM model as well.
In an effort to further improve the near-wall PANS performance, in
this paper we develop a four-equation PANS model from the RANS
k-"--f model, which is known to exhibit excellent near-wall
characteristics.
A derivation of the PANS k-"--f model can be started by
introducing the velocity scale equation from the RANS v2 -f model:

In the preceding equation, the unresolved-to-total-kinetic-energy


and unresolved-to-total-velocity-scale ratios are
fk 

ku
;
k

fv2 

v 2u
v 2

(13)

To close the PANS velocity scale ratio equation, we use Girimajis


[7] derived relation (see also Girimaji et al. [6]) for the total
production of kinetic energy, which yields
Pu  "u  fk P  ";

) P

1
"
P  "u   u
fk u
f"

(14)

Now, from Eq. (12), one can arrive at the following u equation:



f2
 Pu  "u  "u
Du fv2
(15)

f  u P  v f  u

fk
fk
k
fk
k
f"
Dt
Assumingf"  1, Eq. (15) can be written in its simplied form as


Du
 fu  u Pu  u "u 1  fk 
ku
ku
Dt

(16)

where fu  fv2 =fk f. Note that f"  1 implies "u  ", so we
assume that our numerical meshes support the cutoff in the energy
containing scales and inertial range. It is also clear that Eq. (16) gets
its RANS form for fk  1. Then, from the RANS form of the elliptic
relaxation function,
kf  22  "b22 ;

b22 

v 2 2

k 3

(17)

where 22 is the pressure-rate-of-strain term. Hanjalic et al. [17] used


the quasi-linear pressure-strain model of Speziale et al. [25] for the
elliptic relaxation function f in the k-"--f model and therefore the
same approach is used here; see next. It follows that
kfu  k

fv2
f2
f  v 22  "b22 
fk
fk

(18)

Here, fu is the PANS counterpart of the RANS parameter f.


Further, taking into the consideration that
22  "b22 u  fv2 22  "b22 

(19)

we obtain
fu 

1
  "b22 u
ku 22

(20)

For the turbulent transport model and following Girimaji [7], we


use the so-called zero-transport model, which assumes that the
resolved uctuations do not contribute to the net transport:




@ t @ku
@k
@ u @ku
 Uj  U j  u 
@xj k @xj
@xj @xj ku @xj




@ t @"u
@"u
@ u @"u

 Uj  Uj 

@xj " @xj
@xj @xj "u @xj




@ t @u
@
@ u @u
 Uj  U j  u 
(21)
@xj  @xj
@xj @xj u @xj
where

2630

BASARA ET AL.

ku  k

fk2
;
f"

"u  "

fk2
;
f"

u  

fk2
f"

(22)

Therefore, the complete PANS k-"--f model is given by the


following set of equations:


k2
@ku
@k
@ u @ku
 Uj u  Pu  "u 
u  C u u
@xj ku @xj
"u
@t
@xj


2
@"u
@"u
"u
"
@ u @"u
 u
 C"1 Pu  C"2 
 Uj
@t
@xj
ku
ku @xj "u @xj

C"2
 C"1  fk C"2  C"1 

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967



@u
@


@ u @u
 Uj u  fu  u Pu  u "u 1  fk  
@xj u @xj
@t
@xj
ku
ku



1
P
2
L2u r2 fu  fu 
c  c2 u u 
(23)
Tu 1
3
"u

where constants C , c1 , c2 , and C"2 are equal to 0.22, 0.4, 0.65, and
1.9, respectively. Lu is the length scale and Tu is the time scale
dened by using unresolved kinetic energy:
 1=2 
 3 1=4 

 3=2
k

ku

Lu  CL max
; C
Tu  max u ; C
"
"
"
"
(24)
where constants CL and C are equal to 0.36, and 6.0, respectively.
Note again that
p "u  " or f"  1. Note also that C"1 
1:41  0:045= u . This represents the four-equation PANS
k-"--f model with enhanced near-wall attributes. It must be noted
that three (ku -"u -u ) of the four variables are obtained by solving
transport equations and the fourth variable fu is obtained from a
Poisson equation.
Next, we enhance the model further by including the hybrid wall
treatment to account for the low Reynolds number effects. This
entails combining integration up to the wall with wall functions.
Smoothing functions that blend two formulations together are known
by different names: automatic, hybrid, or compound wall treatments
(see Popovac and Hanjalic [19] and Basara [20,21]). The blending
formula for the quantities specied at the cell next to the wall is
given as
   e  t e1=

(25)

where  is the viscous and t the fully turbulent value of the variables:
wall shear stress, production, and dissipation of the turbulent kinetic
energy. The function  is given as


0:01y 4
1  5y

(26)

where y is the normalized distance to the wall. For further details see
Basara [20,21].
Overall, the new PANS model developed in this paper includes
three critical effects not accounted for in the original PANS closures.
The term u accounts for the strong anisotropy in the Reynolds stress
near the wall. In the new model, the near-wall eddy viscosity scales
with the wall-normal uctuations rather than the overall kinetic
energy. The parameter fu takes into consideration the wall blocking
effect. The hybrid implementation accounts for the low Reynolds
number effects. These enhancements are purported for improving the
PANS behavior in smooth bluff-body separation ows.

III. Numerical Procedure


The PANS k-"--f model is implemented in the CFD code AVL
FIRE (AVL FIRE Manual version 2009.1 [26]). The brief numerical
basis of this code is given next.
The discretization of the governing differential equations is
obtained using a cell-centered nite volume approach. In the method
used here, the governing equations are integrated, term by term, over

the polyhedral control volumes. Such discretization practice was


explored in other publications starting with Demirdzic and
Muzaferija [27] and continuing with Ferziger and Peric [28], Marthur
and Murthy [29], Basara et al. [30,31], etc. Hence, the method has
been applied and proven on various applications. The method rests
on the integral form of the general conservation law. For a grid cell P
with the volume Vol surrounded by its neighbors Pj , and with the
outward surface (cell-face) vectors A  Ak ik , the discretized control
volume equation can be written as
nf

nf

nf

X
X
X
d
P VolP P  
Cj 
Dj  sVol
sAk Ak j
 P VolP 
dt
j1
j1
j1
(27)
where Cj and Dj are convective and diffusion transport through the
face j, respectively; nf is the number of cell faces; and sVol
are

volumetric and sAk surface source terms.
All dependent variables, such as momentum, pressure, density,
turbulence kinetic energy, dissipation rate, velocity scale ratio, and
passive scalar are evaluated at the cell center. The cell-face-based
connectivity and interpolation practices for gradients and cell-face
values are introduced to accommodate an arbitrary number of cell
faces. A second-order midpoint rule is used for integral approximation and a second-order linear approximation for any value at the
cell face. The cell gradients can be calculated by using either the
Gauss theorem or a linear least-square approach. The convection is
solved by a variety of differencing schemes, namely upwind, central
differencing, MINMOD (Sweby [32]), and AVL SMART (Przulj and
Basara [33]). In this work, beside central differencing scheme also
MINMOD was used, as this scheme is often used like a compromise
between accuracy and convergence properties. Usually MINMOD
differencing scheme or similar schemes, e.g., AVL SMART (Przulj
and Basara [33]), are used in conjunction with the RANS calculations especially when applied on the complex industrial ows.
However, when the unresolved turbulence, as modeled with the
PANS method, is small, then the role of differencing schemes is
getting to be of higher importance. See the next section for the
comparisons.
The time derivative [see Eq. (27)] is discretized by two implicit
schemes, namely the rst-order accurate Euler (two levels) scheme
and the second-order accurate three time level scheme, which is due
to its accuracy used in all calculations shown in this paper.
The outcome of all presented previously is a set of algebraic
equations: one for each control volume and for each transport
equation. Thus, for a computational domain with M control volumes,
a system of M N algebraic equations needs to be solved for N
dependent variables . Considering the nonlinearity and coupling of
equations, they are solved by the segregated SIMPLE-like algorithm
of Patankar and Spalding [34] (see also Ferziger and Peric [28] and
Demirdzic and Muzaferija [27]). The SIMPLE algorithm effectively
couples the velocity and pressure elds by converting a discrete form
of the continuity equation into an equation for the pressure
correction. The pressure corrections are then used to update the
pressure and velocity elds so that the velocity components obtained
from the solution of momentum equations satisfy the continuity
equation.

IV.

Results and Discussions

The fundamental premise of PANS is now well proven in a variety


of canonical ows such as 1) square cylinder (Song and Park [14] and
Jeong and Girimaji [13]), 2) backward-facing step (Frendi et al. [35]
and Lakshmipathy and Girimaji [4,36]), and 3) circular cylinder
(Lakshmipathy and Girimaji [15]). These studies demonstrate that
PANS performance in ows with sharp-edge separation (square
cylinder and backward-facing step) is much better than in the case of
smooth bluff-body (circular cylinder) separation. The four-equation
PANS developed in this study is aimed toward enhancing the model
performance in smooth bluff-body ows. To assess the model in the
absence of sharp-edge separation, we consider two test cases in this

BASARA ET AL.

2631

study: 1) planar turbulent channel ow and 2) ow past a nite


vertically mounted circular cylinder. The rst is a simple canonical
ow whereas the second contains complex ow features encountered
in many practical applications. In this section, we will establish that
the enhanced four-equation PANS model captures the physics of the
channel ow very closely and provides solutions of engineering
utility in the second test case.

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

A. Channel Flow
1. Description of the Test Case and Numerical Details

The channel ow with Re  650, based on the wall friction


velocity u , the channel half-width , and kinematic viscosity , was
computed with the PANS k-"--f model. DNS data of Iwamoto et al.
[37] were used for comparisons. Three computational grids were
employed: 1) 64 32 24, 2) 64 64 48, and 3) 64 128 96
computational cells (see Fig. 1). In the streamwise direction, the
periodic boundary conditions with the given mass ow were applied,
and in the spanwise direction periodic/cyclic boundaries were used.
All computational grids are compressed near the wall providing the
wall nondimensional distance y between 0.5 for the nest grid up to
2 for the coarsest grid.
2.

Comparisons with Direct Numerical Simulation Data

Figure 2 shows comparisons for the predicted mean velocity


prole as obtained with the PANS k-"--f and DNS. The PANS mean
and higher-order statistics are calculated from the resolved and
unresolved elds employing appropriate averaging. The ner the
grid, the larger degree of uctuations and longer is the duration over
which averaging is performed. Calculations on ner numerical grids
improve the PANS model performance in the viscous sublayer and
the buffer layer. An important point to note is that the coarse grid
appears to agree marginally better with DNS data than ner grids in
the log law region. This is to be expected as the coarse PANS is

Fig. 1 Channel ow: 64  64  48 computational cells (the medium


grid).

Fig. 2 Computed normalized mean velocity in a channel. Nonlled


symbols: DNS, Iwamoto et al. [37].

Fig. 3 Computed normalized rms velocity in a channel. Nonlled


symbols: DNS, Iwamoto et al. [37].

identical to RANS (Basara [21]) in this region and RANS is tuned to


yield the correct log law behavior for the mean velocity. The proles
from higher-resolution computations are certainly within acceptable
accuracy in the log law region. It must be emphasized that this feature
of coarse grid yielding marginally improved behavior in the log law
region is also seen in the two-equation models and that observation is
also attributed to the same reason that RANS is precisely tuned to
reproduce the log law velocity eld. However, Rahman and Siikonen
[38] recently suggested that the performance of the hybrid wall
(Basara [20,21]) and compound wall (Popovac and Hanjalic [19])
treatments can be further improved with a minor modication (for
further details see Rahman and Siikonen [38]).
Predicted proles of rms velocity uctuations are shown in Fig. 3.
In the coarse grid simulations, most of the contribution toward the
rms comes from the unresolved kinetic energy calculation and in the
ner grids the contributions from the uctuating resolved eld is
more dominant. Fairly good agreement is achieved on all grids. The
spatial variation of different rms components and their differences
are well described.
Figure 4 shows averaged subgrid scale kinetic energy, total kinetic
energy, and their ratio as calculated with the PANS on the 64
32 24 mesh. Near the wall, all of the kinetic energy content is in the
modeled part as very ne grid is required to resolve any of the
uctuating motion. Away from the wall, the resolved energy
progressively increases and, indeed, dominates the modeled portion
of the energy. The resolved ow structures shown in the other gures
correspond to this region of the ow. Finally, as we approach the
center of the channel, the total kinetic energy becomes lower as
production diminishes. In this region a coarse grid is employed and
consequently the modeled fraction of the kinetic energy increases

Fig. 4 Computed averaged subgrid scale kinetic energy, averaged total


scale kinetic energy and their ratio (64  32  24 computational cells).

2632

BASARA ET AL.

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

Fig. 7 Computational domain.

Fig. 5 Channel ow: PANS predicted an invariant map (64  64  48


computational cells).

marginally. It is evident from this gure that PANS performs as


required accomplishing the objective of a VR model. It is important
to note that other methods such as DES also attempt to achieve such a
distribution of kinetic energy between resolved and modeled parts.
A better indication of the model performance can be obtained by
plotting the invariants of Reynolds stress anisotropy on the so-called
Lumley invariant map. The normalized anisotropy tensor is aij 
ui uj =k  2=3ij and its two independent invariants are 2  A2 =8
and 3  A3 =24, where A2  aij aji and A3  aij ajk aki . The behavior
of the invariants in the wall-normal direction provides the strongest
indication of the delity of the model calculation. The PANS
(medium grid: 64 64 48) invariant distribution in the wallnormal direction is plotted on the Lumley triangle in Fig. 5. The
Reynolds stresses are two-componential near the wall and reach a
peak level of anisotropy as we move toward the interior in the wallnormal direction. Beyond the peak, the Reynolds stresses attain an
axisymmetric state which prevails over most of the channel interior.
The agreement with DNS calculations is excellent. This agreement
clearly indicates that the model results capture the ow physics quite
accurately. Instantaneous isosurfaces of the second invariant of the
velocity gradient as computed by PANS, on different numerical
grids, are shown in Figs. 6a6c. Note again from Eq. (10) that the
ratio between unresolved and total kinetic energy fk will be

decreased with the grid renement and consequently ner structures


are captured. In summary, the results show that the PANS k-"--f
model does not suppress turbulence uctuations and at the same time
provides a good agreement with DNS data on any numerical grid.
B. Finite Vertically Mounted Circular Cylinder
1. Description of the Test Case

The complex ow chosen for validation of the present approach is


that of a nite cylinder mounted vertically on a at plate. This ow
was used in the previous experimental investigation of Park and Lee
[39] and LESs by Krajnovi [40,41] and Afgan et al. [42]. Freestream
inlet velocity U0  10 m=s and diameter D  0:03 m give a
Reynolds number of approximately 2 104. At this Reynolds
number, the ow in the midsection of the cylinder (i.e., similar to that
of the innite cylinder) experiences transition along the free-shear
layer characterized by the formation of the transition eddies
(Zdravkovich [43]), which is challenging to predict (Krajnovi
[40,41]). Furthermore, the ow consists of three different ow
regimes: the impingement ow at the bottom resulting in horseshoe
vortices, Karman vortex shedding along the cylinder, and the ow
over the cylinders free end producing complicated ow on the free
end. It is worth mentioning that the previous PANS with k-" model
with dynamic unresolved-to-total-kinetic-energy ratio fk (Krajnovi
and Basara [16]) failed to predict this ow correctly. This was due to
the high values of dynamic fk near the wall resulting from the mesh
resolution used in the calculations. Thus, the application of the new
PANS approach with more advanced RANS turbulence model is
relevant.
2.

Fig. 6 An instantaneous isosurface of the second invariant of the


velocity gradient as predicted by the PANS k-"--f model on different
grids: a) coarse, b) medium, and c) nest grid.

Numerical Details

A test section of 24D 20D 28D (W H L) was used in


simulation (see Fig. 7). The average turbulent intensity at the inlet of
the wind tunnel used in the experiments of Park and Lee [39] was less
than 0.08%. A uniform velocity prole constant in time was thus
used as the inlet boundary condition in present PANS. The inlet
velocity prole was made to produce the boundary-layer thickness of
20 mm corresponding to approximately 7% of the cylinders
diameter at the location of the cylinder (x  0), similar to the
experimental ow condition. The homogeneous Neumann boundary
condition was used at the downstream boundary. The lateral surfaces
and the ceiling were treated as slip surfaces using symmetry
conditions. Hybrid wall treatment and a no slip boundary condition
were applied on the channel oor and the surface of the body.
Numerical accuracy was established by making three PANS
calculations on different computational grids containing 4, 13.5, and
21.5 million nodes. Three multiblock hexahedral meshes were used
in the simulations. A grid topology was made using several O and C
grids in order to concentrate most of the computational cells close to
the cylinder. The large number of cells close to the surface is needed
in order to resolve laminar boundary layer on the front part of the
cylinder. The maximum CourantFriedrichsLewy (CFL) number
was smaller then one for all time steps.
The ne and the medium computational grids have a wall-normal
resolution of n < 1, 12 < s < 25 in the streamwise direction
and 3 < l < 20 in the direction parallel with the surface of the

2633

BASARA ET AL.

4.

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

Fig. 8 An instantaneous isosurface of the second invariant of the


velocity gradient as predicted by PANS: a) MINMOD differencing
scheme and b) CD differencing scheme.

body and normal to the streamwise direction (the mean l is around
10). This resolution is valid on entire cylinder except at the free end
where n reaches locally 1.5. Here, n  nhu it =, s 
shu it =, l  lhu it =, and hu it is the time-averaged friction
velocity. The time step was 5 105 s, giving a maximum CFL
number of approximately one. All simulations were rst run during
time t  t Uo =D  112 corresponding to four ow sthrough
passages through the tunnel. Afterwards, the ow was averaged
during t  333 (20,000 time steps).
3.

Choice of Numerical Scheme

VR approaches present an interesting challenge to numerical


schemes. In coarse-resolution regions, the VR model computation is
similar in character to RANS and the choice of the numerical scheme
is made with computational robustness as the desired attribute. Most
RANS and, even, DES approaches typically employ second-order
upwind schemes such as MINMOD scheme due to numerical
problems associated with the more precise central difference (CD)
scheme. In the ne resolution regions, the VR computational
requirements are more like those of LES. Accuracy and ability to
resolve ne ow features become the preferred numerical qualities
and CD scheme is usually chosen. Here, we perform PANS
computations using both MINMOD and CD numerical schemes and
contrast their results. Figure 8 shows the second invariant of the
velocity gradient Q from PANS using two different numerical
schemes. With the CD scheme more ne-scale ow structures are
resolved for a given grid resolution in all regions from the horseshoe
vortex at the bottom, KelvinHelmholtz (KH) vortices along the
cylinder to the ow on the free top and the wake ow. Furthermore
the ow structures from PANS using CD scheme are in better
agreement with those found in previous LES (Krajnovi [40]) and
therefore all PANS simulation results presented in the following text
are obtained using the CD scheme.

Fig. 9

Comparisons with Experimental Data

The experimental data of Park and Lee [39,44,45] consist of


surface pressure (Cp) proles at four positions along cylinders. First,
we show here the results obtained in Krajnovi and Basara [16]
where the previous PANS k-" fails to predict the separation from the
curved surfaces of the circular cylinder in line with the RANS k-"
model in conjunction with the standard wall functions (see Fig. 9).
Note that the standard mesh resolution yields fk  1 near the wall.
Therefore, RANS equations are solved in that region. The results
presented in Fig. 9 are typical of two-equation RANS and PANS
calculations in smooth-body separation ows. Such behavior is the
main motivation for enhancing PANS using RANS k-"--f model as
the basis.
Figure 10 shows the comparison of the PANS results from the
three different computational grids against the experimental data.
Contrary to previous PANS models, simulations using the new PANS
k-"--f model (at all grids levels) yield similar Cp proles at the
lower (z=D  0:3) and middle portions of the cylinder (z=D  0:5)
and they are all in good agreement with experimental data (Figs. 10a
and 10b). The agreement deteriorates near the top quarter of the
cylinder even as the differences among the three meshes become
signicant (Fig. 10c), especially in the proximity of the free end. As
discussed in Krajnovi [40], the complex interaction between the
Karman ow around the cylinder and the ow over the cylinders free
end requires a high degree of resolution. This is the reason for large
Cp difference among the various meshes near the cylinders free end.
In this region, only the ne PANS results are in good agreement with
the experimental data (Fig. 10d). The strong inuence of the
computational grid on the results at position z=D  0:917 (Fig. 10d)
is consistent with the LES ndings of Krajnovi [40]. Overall, the
performance of the present four-equation PANS computations is
superior to that of the standard two-equation (k-") PANS model
results of Krajnovi and Basara [16].
Experimental investigation by Park and Lee [39] has shown the
existence of vortex shedding with a frequency of 47 Hz
corresponding to dimensionless frequency of St  0:141 (St being
Strouhal number, St  f D=U). Previous LESs by Krajnovi [40]
and Afgan et al. [42] have predicted shedding frequencies of St 
0:148 (Krajnovi [40]) and 0.132 (Afgan et al. [42]). The present
PANS simulation predicts the aerodynamic coefcients as shown in
Fig. 11a. The power spectral density of the side force coefcient
(Fig. 11b) shows the dominant frequency at St  0:146, which is in
agreement with previous LES results and experimental data.

5.

Flow Structures

Flow around the free end of a cylinder has been studied by several
authors using both experimental and numerical techniques (Park and
Lee [39,44,45], Krajnovi [40,41], and Afgan et al. [42]). Previous
investigations have shown the existence of two tip vortices and one
large recirculation region originating at the leading edge of the
cylinder. The tip vortices form due to the interaction between
the Karman ow along upper part of the cylinder and the ow over

Predicted surface pressure coefcient at location a) z=D  0:30 and b) z=D  0:50 with the PANS k-" model.

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

2634

BASARA ET AL.

Fig. 10 Surface pressure coefcient, as predicted by the PANS k-"--f model, at different locations: a) z=D  0:3, b) z=D  0:50, c) z=D  0:75, and
d) z=D  0:917.

the free end. Figure 12a shows two isosurfaces of streamwise


vorticity component obtained from the present PANS calculations.
This shows the existence of the two counter-rotating tip vortices,
consistent with ndings from previous studies. In addition, vortex
cores were calculated using EnSight postprocessing software which
uses algorithms based on techniques outlined by Sujudi and Haimes
[46]. The vortex core together with the streamlines projected on the
center plane show the large recirculation region on the top of the
cylinder and arch vortex in the near wake region.
Figure 12b shows the time-averaged ow in the near wake region
of the upper part of the cylinder. The vortex cores in the near wake
show that the legs of the arch vortex are inclined in the streamwise
direction with respect to the cylinder axis, consistent with LES
computations of Krajnovi [40,41] and Afgan et al. [42]. The timeaveraged arch vortex is a result of the downwash process behind the

cylinder and is stronger in the upper half-segment. The downwash


process was previously studied by Krajnovi [40,41] using LES.
Figure 13 shows a comparison of the instantaneous pressure
isosurfaces from LES (Krajnovi [40]) and present PANS. The
agreement is quite reasonable. These ow structures include vortices
that originate from KH instabilities along the cylinder. The KH
vortices are inclined in the streamwise direction near the free end of
the cylinder. The near wake ow structures in the downwash process
are also inclined in respect to the cylinders plane. The PANS
computation captures all these ow features. The only difference
between LES and PANS is that the latter produces less resolved
structures indicative of more eddy viscosity. Here, it is worth noting
that the LES (Krajnovi [40,41]) and PANS computations are
performed using different computational codes. The LES simulation
employs a structured code while the AVL FIRE used for PANS

Fig. 11 Aerodynamic coefcients as predicted by PANS: a) drag and side force time histories for nondimensional time normalized with the inlet velocity
and cylinders diameter and b) power spectral density of the side force coefcient signal.

2635

BASARA ET AL.

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

Fig. 12 Isosurface of streamwise component of vorticity and vortex cores showing the tip vortices: a) tip vortices colored with streamwise vorticity and
b) streamwise projected on center plane and vortex cores.

The overall conclusion that can be drawn from this work is that the
new PANS k-"--f is a useful VR approach for complex bluff-body
ows with multiple ow features. This paper represents an important
step toward the development of bridging VR methods for practical
applications.

References
Fig. 13 Isosurface of low-pressure p  0:55: a) LES (Krajnovi [40])
and b) PANS.

simulations is an unstructured code. The difference in the numerical


schemes between the two codes could cause some of the differences
in the observed ow structures.

V.

Conclusions

The VR PANS model, employed thus far in literature exclusively


in two-equation form, has been enhanced to include the following
near-wall improvements: 1) a transport equation for velocity scale
ratio, 2) a wall-blockage effect by solving an elliptic relaxation
equation, and 3) a capability to be integrated to the wall using
blending functions. This paper develops the new VR ku -"u -u -fu
model derived from the parent RANS k-"--f model by employing
the PANS methodology. The RANS k-"--f model is an advanced
eddy viscosity closure that is especially designed for improved
physical delity in the near-wall region. The superior near-wall
attributes of the parent RANS model are systematically transferred to
the PANS closure.
Results for the channel ow obtained with the PANS k-"--f are in
a good agreement with DNS data on all meshes. The mean velocity
prole, the rms velocity uctuations, and the invariants of Reynolds
stress anisotropy are accurately predicted. Calculations show that,
with the grid renement, ner structures are captured and turbulence
uctuations are not suppressed for any numerical grid used in this
work.
The new model is also applied to a complex 3D ow around a tall
surface mounted circular cylinder (Park and Lee [39]). This ow
exhibits several different streamline patterns including complicated
3D features near the cylinders free end, transition in the shear layer,
Karman vortex shedding along the cylinder, and formation of
multiple horseshoe vortices around the junction of the cylinder with
the ground. Previous LES computations (Krajnovi [40]) have
shown that these ow features cannot be easily captured even with a
large number of computational cells. The four-equation PANS
k-"--f calculations reproduce nearly all of the unsteady ow
features around the cylinder in good agreement with experimental
observations (Park and Lee [39]) and LESs (Krajnovi [40,41] and
Afgan et al. [42]). The dominant vortex shedding frequency is in
good accord with the experimental value. All mean features of the
ow are captured accurately. The predicted average pressure
coefcient is in good agreement with measurements at all positions.
Even the ner ow structures are captured with reasonable precision.

[1] Speziale, C. G., Computing Nonequilibrium Flows with TimeDependent RANS and VLES, Proceedings of the 15th International
Conference on Numerical Methods in Fluid Dynamics, edited by P.
Kutler, J. Flores, and J. Chattot, Lecture Notes in Physics, Springer,
New York, 1996, pp. 123129.
[2] Spalart, P. R., Jou, W. H., Strelets, M., and Allmaras, S. R., Comments
on the Feasibility of LES for Wings, and on a Hybrid RANS/LES
Approach, Advances in DNS/LES: Proceedings of the First AFOSR
International Conference on DNS/LES, Ruston, LA, 1997.
[3] Sagaut, P., Deck, S., and Terracol, M., Multiscale and Multiresolution
Approaches in Turbulence, Imperial College Press, London, 2006.
[4] Lakshmipathy, S., and Girimaji, S., Partially-Averaged NavierStokes
Method for Turbulent Flows: k-! Model Implementation, AIAA
Aerospace Sciences Meeting and Exhibit, Reno, NV, 2006, p. 119.
[5] Chaouat, B., and Schiestel, R., A New Partially Integrated Transport
Model for Subgrid-Scale Stresses and Dissipation Rate for Turbulent
Developing Flows, Physics of Fluids, Vol. 17, No. 6, 2005, pp. 119.
[6] Girimaji, S., Srinivasan, R., and Jeong, E., PANS Turbulence Models
for Seamless Transition Between RANS and LES: Fixed Point Analysis
and Preliminary Results, ASME Paper FEDSM2003-45336, 2003.
[7] Girimaji, S., Partially-Averaged NavierStokes Model for Turbulence:
A Reynolds-Averaged NavierStokes to Direct Numerical Simulation
Bridging Method, Journal of Applied Mechanics, Vol. 73, No. 3, 2006,
pp. 413421.
doi:10.1115/1.2151207
[8] Girimaji, S., Jeong, E., and Srinivasan, R., Partially Averaged Navier
Stokes Method for Turbulence: Fixed Point Analysis and Comparison
with Unsteady Partially Averaged NavierStokes, Journal of Applied
Mechanics, Vol. 73, No. 3, 2006, pp. 422429.
doi:10.1115/1.2173677
[9] Chaouat, B., and Schiestel, R., From Single-Scale Turbulence Models
to Multiple-Scale and Subgrid-Scale Models by Fourier Transform,
Theoretical and Computational Fluid Dynamics, Vol. 21, No. 3, 2007,
pp. 201229.
doi:10.1007/s00162-007-0044-3
[10] Chaouat, B., and Schiestel, R., Progress in Subgrid-Scale Transport
Modelling for Continuous Hybrid Nonzonal RANS/LES Simulations,
International Journal of Heat and Fluid Flow, Vol. 30, No. 4, 2009,
pp. 602616.
doi:10.1016/j.ijheatuidow.2009.02.021
[11] Fadai-Ghotbi, A., Friess, C., Manceau, R., and Boree, J., A Seamless
Hybrid RANS-LES Model Based on Transport Equations for the
Subgrid Stresses and Elliptic Blending, Physics of Fluids, Vol. 22,
No. 5, 2010, p. 22-41.
doi:10.1063/1.3415254
[12] Murthi, A., Reyes, D., Girimaji, S., and Basara, B., Turbulent
Transport Modelling for PANS and Other Bridging Closure
Approaches, Proceedings of the Fifth European Conference on CFD
(ECCOMAS CFD 2010), edited by J. C. F. Pereira, and A. Sequeira,
Lisbon, Portugal, 2010.
[13] Jeong, E., and Girimaji, S. S., Partially-Averaged NavierStokes
(PANS) Method for Turbulence Simulations: Flow past a Square

2636

[14]

[15]

[16]

Downloaded by Universite de Sherbrooke on November 25, 2012 | http://arc.aiaa.org | DOI: 10.2514/1.J050967

[17]

[18]
[19]

[20]

[21]

[22]
[23]
[24]

[25]

[26]
[27]

[28]
[29]

[30]

BASARA ET AL.

Cylinder, Journal of Fluids Engineering, Vol. 132, Dec. 2010,


pp. 121203-1121203-11.
doi:10.1115/1.4003153
Song, C., and Park, S., Numerical Simulation of Flow past a Square
Cylinder Using Partially-Averaged NavierStokes Model, Journal of
Wind Engineering and Industrial Aerodynamics, Vol. 97, No. 1, 2009,
pp. 3747.
doi:10.1016/j.jweia.2008.11.004
Lakshmipathy, S., and Girimaji, S. S., Partially-Averaged Navier
Stokes PANS Method for Turbulence Simulations: Flow past a Circular
Cylinder, Journal of Fluids Engineering, Vol. 132, Dec. 2010,
pp. 121202-1121202-9.
doi:10.1115/1.4003154
Krajnovi, S., and Basara, B., Numerical Simulations of the Flow
Around a Tall Finite Cylinder Using LES and PANS, Proceedings of
the iTi Conference on Turbulence, edited by J. Peinke, M. Oberlak, and
A. Tallamelli, Progress in Turbulence 3, Springer, Bertinoro, Italy,
2008.
Hanjalic, K., Popovac, M., and Hadziabdic, H., A Robust Near-Wall
Elliptic-Relaxation Eddy-Viscosity Turbulence Model for CFD,
International Journal of Heat and Fluid Flow, Vol. 25, No. 6, 2004,
pp. 10481051.
Durbin, P. A., Near-Wall Turbulence Closure Modelling Without
Damping Functions, Theoretical and Computational Fluid Dynamics,
Vol. 3, 1991, pp. 113.
Popovac, M., and Hanjalic, K., Compound Wall Treatment for RANS
Computation of Complex Turbulent Flows and Heat Transfer, Flow,
Turbulence and Combustion, Vol. 78, No. 2, 2007, pp. 177202.
doi:10.1007/s10494-006-9067-x
Basara, B., An Eddy Viscosity Transport Model Based on Elliptic
Relaxation Approach, AIAA Journal, Vol. 44, No. 7, 2006, pp. 1686
1690.
doi:10.2514/1.20739
Basara, B., A Nonlinear Eddy Viscosity Model Based on Elliptic
Relaxation Approach, Fluid Dynamics Research, Vol. 41, No. 1, 2009,
pp. 121.
doi:10.1088/0169-5983/41/1/012403
Germano, M., Turbulence: The Filtering Approach, Journal of Fluid
Mechanics, Vol. 238, No. 1, 2006, pp. 325336.
doi:10.1017/S0022112092001733
Girimaji, S., and Abdul-Hamid, K. S., PartiallyAveraged Navier
Stokes Model for Turbulence: Implementation and Validation, AIAA
Paper 2005-0502, Reno, NV, 2005.
Basara, B., Krajnovi, S., and Girimaji, S., PANS vs. LES for
Computations of the Flow Around a 3D Bluff Body, Proceedings of
ERCOFTAC 7th International Symposium: ETMM7, Vols. 23,
Lymassol, Cyprus, 2008, pp. 548554.
Speziale, C. G., Sarkar, S., and Gatski, T. B., Modeling the PressureStrain Correlation of Turbulence, an Invariant Dynamical Systems
Approach, Journal of Fluid Mechanics, Vol. 227, June 1991, pp. 45
272.
AVL AST, FIRE Manual, Ver. 2009.1, AVL List GmbH, Graz, Austria,
2009.
Demirdzic, I., and Muzaferija, S., Numerical Method for Coupled
Fluid Flow, Heat Transfer and Stress Analysis Using Unstructured
Moving Meshes with Cells of Arbitrary Topology, Computer Methods
in Applied Mechanics and Engineering, Vol. 125, Nos. 14, 1995,
pp. 235255.
doi:10.1016/0045-7825(95)00800-G
Ferziger, J. H., and Peric, M., Computational Methods for Fluid
Dynamics, SpringerVerlag, New York, 1996.
Marthur, S. R., and Murthy, J. Y., A Pressure Based Method for
Unstructured Meshes, Numerical Heat Transfer, Vol. 31, No. 2, 1997,
pp. 195215.
doi:10.1080/10407799708915105
Basara, B., Employment of the Second-Moment Turbulence Closure
on Arbitrary Unstructured Grids, International Journal for Numerical

[31]

[32]

[33]
[34]

[35]

[36]

[37]

[38]
[39]

[40]

[41]

[42]

[43]
[44]

[45]

[46]

Methods in Fluids, Vol. 44, No. 4, 2004, pp. 377407.


doi:10.1002/d.646
Basara, B., Alajbegovic, A., and Beader, D., Simulation of Single- and
Two-Phase Flows on Sliding Unstructured Meshes Using Finite
Volume Method, International Journal for Numerical Methods in
Fluids, Vol. 45, No. 10, 2004, pp. 11371159.
doi:10.1002/d.733
Sweby, P. K., High Resolution Schemes Using Flux Limiters for
Hyperbolic Conservation Laws, SIAM Journal on Numerical Analysis,
Vol. 21, No. 5, 1984, pp. 9951011.
doi:10.1137/0721062
Przulj, V., and Basara, B., Bounded Convection Schemes for
Unstructured Grids, AIAA Paper 2001-2593, Anaheim, CA, 2001.
Patankar, S. V., and Spalding, D. B., A Calculation Procedure for Heat,
Mass and Momentum Transfer in Three-Dimensional Parabolic Flows,
International Journal of Heat and Mass Transfer, Vol. 15, No. 10, 1972,
pp. 15101520.
Frendi, A., Tosh, A., and Girimaji, S. S., Flow past a Backward Facing
Step: Comparison of PANS, DES and URANS Results with Experiments, International Journal of Computational Methods, Vol. 8, No. 1,
2006, pp. 2338.
doi:10.1080/15502280601006207
Lakshmipathy, S., and Girimaji, S. S., Extension of Boussinessq
Turbulence Constitutive Relation for Bridging Models, Journal of
Turbulence, Vol. 8, No. 31, 2007, pp. 121.
doi:10.1080/15502280601006207
Iwamoto, K., Suzuki, Y., and Kasagi, N., Reynolds Number Effect on
Wall Turbulence: Toward Effective Feedback Control, International
Journal of Heat and Fluid Flow, Vol. 23, No. 5, 2002, pp. 678689.
doi:10.1016/S0142-727X(02)00164-9
Rahman, M. M., and Siikonen, Compound Wall Treatment with LowRe Turbulence Model, International Journal for Numerical Methods
in Fluids, Vol. 66, No. 5, 2011, pp. 118.
Park, C-W., and Lee, S-J., Flow Structures Around a Finite Circular
Cylinder Embedded in Various Atmospheric Boundary Layers, Fluid
Dynamics Research, Vol. 30, No. 4, 2002, pp. 197215.
doi:10.1016/S0169-5983(02)00037-0
Krajnovi, S., Flow Around a Surface-Mounted Finite Cylinder:
A Challenging Case for LES, Notes on Numerical Fluid Mechanics
and Multidisciplinary Design, Vol. 97, 2008, pp. 305315.
doi:10.1007/978-3-540-77815-8_32
Krajnovi, S., Flow Around a Tall Finite Cylinder Explored by Large
Eddy Simulation, Journal of Fluid Mechanics, Vol. 676, 2011,
pp. 294317.
doi:10.1017/S0022112011000450
Afgan, I., Moulinec, C., Prosser, R., and Laurence, D., Large Eddy
Simulation of Turbulent Flow for Wall Mounted Cantilever Cylinders of
Aspect Ratio 6 and 10, International Journal of Heat and Fluid Flow,
Vol. 28, No. 4, 2007, pp. 561574.
doi:10.1016/j.ijheatuidow.2007.04.014
Zdravkovich, M. M., Flow Around Circular Cylinders, Vol 1:
Fundamentals, Oxford Univ. Press, Oxford, England, U.K., 1997.
Park, C-W., and Lee, S-J., Free End Effects on the Near Wake Flow
Structure Behind a Finite Circular Cylinder, Journal of Wind
Engineering and Industrial Aerodynamics, Vol. 88, Nos. 23, 2000,
pp. 231246.
doi:10.1016/S0167-6105(00)00051-9
Park, C-W., and Lee, S-J., Effects of Free-End Corner Shape on Flow
Structure Around a Finite Cylinder, Journal of Fluids and Structures,
Vol. 19, No. 2, 2004, pp. 141158.
doi:10.1016/j.juidstructs.2003.12.001
Sujudi, D., and Haimes, R., Identication of Swirling Flow in 3-D
Vector Fields, AIAA Paper 95-1715, San Diego, CA, 1995.

A. Tumin
Associate Editor

You might also like