You are on page 1of 13

International Journal of Heat and Fluid Flow 31 (2010) 820832

Contents lists available at ScienceDirect

International Journal of Heat and Fluid Flow


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

Numerical and physical aspects in LES and hybrid LES/RANS of turbulent ow


separation in a 3-D diffuser
S. Jakirlic a,b,d,*, G. Kadavelil a,b, M. Kornhaas a,c, M. Schfer a,c, D.C. Sternel a,c, C. Tropea a,b,d
a

Department of Mechanical Engineering, Technische Universitt Darmstadt, Germany


Institute of Fluid Mechanics and Aerodynamics, Petersenstr. 30, D-64287 Darmstadt, Germany
c
Institute of Numerical Methods in Mechanical Engineering, Dolivostr. 15, D-64293 Darmstadt, Germany
d
Center of Smart Interfaces, Petersenstr. 32, D-64287 Darmstadt, Germany
b

a r t i c l e

i n f o

Article history:
Received 20 December 2009
Received in revised form 14 April 2010
Accepted 12 May 2010
Available online 6 July 2010
Keywords:
3-D separation
Wall-bounded ow
LES
Hybrid LES/RANS

a b s t r a c t
An incompressible fully-developed duct ow expanding into a diffuser, whose upper wall and one side
wall are appropriately deected (with the expansion angles of 11.3 and 2.56 respectively), and for
which reference experimental and DNS databases were provided by Cherry et al. (2008, 2009) and Ohlsson et al. (2009, 2010), was studied computationally by using a zonal hybrid LES/RANS (HLR) method,
proposed recently by Kniesner (2008) and Jakirlic et al. (2009). In addition a complementary Large-Eddy
Simulation (LES) method has been applied. The ow Reynolds number based on the height of the inlet
channel is Reh = 10,000. The primary objective of the present investigation was the comparative assessment of the computational models in this ow conguration characterized by a complex 3-D ow separation being the consequence of an adverse-pressure gradient evoked by the duct expansion. The focus of
the investigation was on the capability of different modelling approaches to accurately capture the size
and shape of the 3-D ow separation pattern and associated mean ow and turbulence features.
2010 Elsevier Inc. All rights reserved.

1. Introduction
Congurations involving 3-D boundary-layer separation are
among the most frequently encountered ow geometries in practice. Accordingly, the methods simulating them have to be appropriately validated using detailed and reliable reference databases.
However, the largest majority of the experimental benchmarks
being used for validating computational methods and turbulence
models relates to internal, two-dimensional ow congurations,
as e.g., ow in a 2-D diffuser (e.g., Obi et al., 1993), ow over a
backward-facing step and a forward-facing step, or ow over
fences, ribs, 2-D hills and 2-D humps mounted on the bottom wall
of a plane channel. In these examples it is assumed that the inuence of the side walls (according to Bradshaw and Wong (1972),
the minimum aspect ratio representing the ratio of the channel
height to channel width should be 1:10 in order to eliminate
the inuence of the side walls) is not felt at the channel midplane.
Consequently, within a computational framework, the spanwise
direction can be regarded as a homogeneous one, the fact enabling
the application of the periodic boundary conditions (even 2-D
computations when using the RANS approach). By doing so, the
3-D nature of the ow is completely missed: strong secondary mo* Corresponding author at: Institute of Fluid Mechanics and Aerodynamics,
Petersenstr. 30, D-64287 Darmstadt, Germany.
E-mail address: s.jakirlic@sla.tu-darmstadt.de (S. Jakirlic).
0142-727X/$ - see front matter 2010 Elsevier Inc. All rights reserved.
doi:10.1016/j.ijheatuidow.2010.05.004

tion across the inlet section of the channel induced by the Reynolds
stress anisotropy which is, as generally known, beyond the reach
of the eddy-viscosity RANS model group, complex 3-D separation
patterns spreading over duct corners (corner separation and corner
reattachment), etc.
These circumstances were the prime motivation for the recent
experimental study of the ow in a three-dimensional diffuser
conducted by Cherry et al. (2008, 2009). Such a diffuser conguration has also a high practical relevance. It mimics a diffuser
situated between a compressor and the combustor chamber in
a jet engine. Its task is to decelerate the ow discharging from
compressor over a very short distance to the velocity eld of
the combustor section. Typically a uniform inlet prole over
the diffuser outlet is desirable. Such a ow situation is associated
by a strong pressure increase. Cherry et al. provided a detailed
reference database comprising the pressure distribution along
the bottom non-deected wall, three-component mean velocity
eld and the streamwise Reynolds stress component eld within
the entire diffuser section. Recently Ohlsson et al. (2009, 2010)
have performed a complementary Direct Numerical Simulation
using a massively parallel high-order spectral element code.
The 3-D diffuser was meshed by approximately 172 million grid
points. In addition to the mean velocity eld, all six Reynolds
stress components were evaluated, as well as the surface
pressure distribution along the bottom wall. It should be noted
that in the experimental investigations two three-dimensional

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

diffusers with the same fully-developed channel inlet but slightly


different expansion geometries (the upper-wall expansion angle
is reduced from 11.3 to 9; the side-wall expansion angle is increased from 2.56 to 4) were investigated, Cherry et al. (2008).
Both diffuser ows are characterized by a three-dimensional
boundary-layer separation, but the size and shape of the separation bubble exhibit a high degree of geometric sensitivity to the
dimensions of the diffuser. The rst diffuser conguration
(Fig. 1), being also the ow geometry considered in the present
work, has served as a test case of the 13th and 14th ERCOFTAC
SIG15 Workshops on rened turbulence modelling, Steiner et al.
(2009) and Jakirlic et al. (2010b). In addition to different RANS
models, the LES and LES-related methods (different seamless
and zonal hybrid LES/RANS models; DES Detached Eddy Simulation) were comparatively assessed (http://www. ercoftac.org).
The latter schemes, hybridizing the RANS and LES methods
aimed at a reduction of spatial and temporal resolution, have recently experienced growing popularity in the Computational
Fluid Dynamics (CFD) community. Their goal is to combine the
advantages of both methods in order to provide a computational
procedure that is capable of capturing the large-scale eddy structures with a broader spectrum and the bulk ow unsteadiness
as encountered in the ows involving separation, but at affordable costs. Interested readers are referred to a relevant review
about hybrid LES/RANS methods from Frhlich and von Terzi
(2008). We mention here only their general classication into
two main groups: the zonal, two-layer schemes a RANS model
resolving the near-wall region is bridged at a distinct interface
with the conventional LES covering the outer layer (ow core)
and the seamless models, where a RANS-like model formulation, mimicking a sub-scale model, is applied in the entire ow
domain. The computational model examined in the present work
belongs to the former category. This zonal hybrid LES/RANS
model couples an eddy-viscosity-based RANS model in the wall
layer to an LES in the outer ow region. In addition to the exper-

821

imental and DNS database, the results obtained are comparatively assessed to a complementary LES simulation.
2. 3-D diffuser: case description
The diffuser shape, dimensions and the coordinate system are
shown in Fig. 1. The ow in the inlet duct (height h = 1 cm, width
B = 3.33 cm) corresponds to fully-developed turbulence (enabled
experimentally by a development channel being 62.9 channel
heights long). The L = 15h long diffuser section is followed by a
straight outlet part (12.5h long). Downstream of this the ow goes
through a 10h long contraction into a 1 in. diameter tube. The curvature radius at the walls transitioning between diffuser and the
straight duct parts are 6 cm. The bulk velocity in the inow duct
is Ubulk = Uinflow = 1 m/s in the x-direction resulting in the Reynolds
number based on the inlet channel height of 10,000. The origin of
the coordinates (y = 0, z = 0) coincides with the intersection of the
two non-expanding walls at the beginning of the diffusers expansion (x = 0). The working uid is water ( q = 1000 kg/m3 and
l = 0.001 Pas).
3. Computational method
The continuity and momentum equations governing the incompressible ow in the present 3-D diffuser conguration read:

@U i
0
@xi
@U i @U i U j
1 @P
@


sm st

@xj
@t
q @xi @xj ij ij



Here, smij 2mSij (with Sij 0:5 @U i =@xj @U j =@xi being the rate of
strain tensor) represents the viscous stress tensor, whereas the turbulent stress tensor stij is to be modelled. The rationale and the most

Fig. 1. Geometry of the 3-D diffuser considered (not to scale), Cherry et al. (2008).

822

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

important features of the turbulence models used in the present


work are outlined in the following sub-sections.
3.1. Computational models
3.1.1. Hybrid LES/RANS (HLR) model
The present hybrid LES/RANS formulation represents a zonal,
two-layer hybrid approach with a RANS model covering the
near-wall region and the LES model the remainder of the ow domain. Both methods share the same temporal resolution (time
step). The equations governing the velocity (Eq. (1)) eld operate
as the Reynolds-averaged NavierStokes equations in the near-wall
layer (with U i representing the ensemble-averaged velocity eld) or
as the ltered NavierStokes equations in the outer layer (with U i
representing the spatially ltered velocity eld). The turbulent
stress tensor stij in Eq. (1) representing either the sub-grid-stress
tensor (stij  sij 2mt Sij  skk dij =3) or the Reynolds-stress tensor
(stij  ui uj 2mt Sij  2kdij =3) is expressed in terms of the mean
strain tensor via Boussinesqs relationship. The equations governing
the velocity eld in the hybrid LES/RANS framework read:

"
!#

DU i
1 @p
@
@U i @U j

m m t


Dt
q @xi @xj
@xj @xi

The coupling of the instantaneous LES eld and the ensembleaveraged RANS eld at the interface is realized via the turbulent viscosity, which makes it possible to obtain solutions using one system
of equations. This means practically that the governing equations
are solved in the entire solution domain irrespective of the ow
sub-region (LES or RANS). Depending on the ow zone, the hybrid
model implies the determination of the turbulent viscosity mt either
from a near-wall ke RANS model: mt = Clflk2/e or from the subgrid-scale (SGS) model in the LES formulation (standard Smagorinsky model was applied, 1963): mt mSGS C S D2 jSj. The Smagorinsky constant CS takes the value of 0.1. D = (Dx  Dy  Dz)1/3
represents the lter width and jSj Sij Sij 1=2 the strain rate modulus.
The near-wall variation of the turbulent viscosity mt is obtained from
a two-equation, ke RANS model, implying solution of the transport
equations for kinetic energy of turbulence k and its dissipation rate e.
The near-wall and viscous damping functions and the relevant
source terms are presently modelled in line with the proposal of
Launder and Sharma (LS, 1974), with e representing a homogeneous part of the total viscous dissipation rate e ! ~e e  2m
1=2
1=2
@k =@xk @k =@xk , taking zero value at the wall (~ewall 0). This
near-wall model is a straightforward extension of the standard
high-Reynolds number ke model aimed at capturing the enhanced
viscosity inuence on turbulence in the immediate wall vicinity
(viscous sub-layer and buffer layer). Solving the equations for k
and e in the RANS sub-domain (being necessary for the turbulent
viscosity determination) requires appropriate boundary conditions
at the RANS-LES interface. The approach used presently is based on
the equality of the RANS interface values for k and e with corresponding SGS values. This condition, originating from the equivalence of the total stresses at both sides of interface (the RANS
total stress, similar as the LES one, comprises both resolved and
RANS
modelled parts) - sLES
tot stot , reduces to the equality of their modelled contributions by assuming the continuity of the resolved fractions across the interface (ifce), Temmerman et al. (2005):
2mSGS Sij 2mt Sij . Finally, the condition applied at the interface
implies the equality of the (modelled) turbulent viscosities:

mt;ifce jRANS-side mSGS;ifce jLES-side

esis assuming the equality between the production, dissipation


and energy ux through the cutoff and the analogy to the eddy-viscosity-based RANS modeling in line with the proposal of Mason
and Callen (1986):

kSGS

C S D2 jSj2
p ;
Cl

eSGS C S D2 jSj3

with Cl = 0.09. The algorithm being used to numerically ensure that


the condition expressed by Eq. (3) is fullled, resembles the wellknown procedure for setting the value of a variable at a computational node and is explained as follows. The RANS equations for k
and e are solved in the entire ow eld, but with the discretization
coefcients taking zero values in the LES sub-region (Ak,SGS = 0). By
appropriately manipulating
the corresponding
Pthe source
 terms in P
discretization equations
k Ak SP SGS UP;SGS
k Ak Uk;SGS SU;SGS
by taking SU,SGS = 1030USGS and SP,SGS = 1030, with USGS  kSGS, eSGS,
their numerical solution in the framework of the nite volume
method provides the interface values of the kRANS and e RANS, being
equal to the corresponding SGS values. By doing so, the boundary
condition at the LES/RANS interface (ifce), expressed by Eq. (3), is
implicitly imposed without introduction of any empirical transitional function. Both the RANS model and the SGS model are used
in their original form without any further adjustment or modication. This is illustrated in Fig. 2 displaying the variation of the modelled turbulent viscosity across the interface at two streamwise
positions x/h = 2 (inow duct) and x/h = 10 (diffuser section).
One can clearly see the damping of the RANS viscosity towards
the one of the SGS model. A gradual adjustment of the turbulent viscosities at both interface sides to each other ensures their smooth
transition. Fig. 2 also offers evidence that the adopted grid resolution for both HLR and LES is adequate (see corresponding discussion
in Section 3.3).
Unlike some seamless hybrid LES/RANS models, as e.g., Detached
Eddy Simulation (DES, Spalart et al., 1997; Spalart, 2009), the present zonal model treats the LES/RANS interface explicitly. Its position
in the ow eld plays an important role with respect to the quality
of the solution. The present hybrid LES/RANS model is equipped
with an algorithm providing an adaptive, ow-dependent interface
location, controlled by a parameter expressing the ratio of the modelled to the total kinetic energy of turbulence, averaged appropriately over all grid cells in the LES-region of interface (Jakirlic
et al., 2009). However, this procedure is not applied in the present
simulation. The interface position along all diffuser walls, that is
the RANS layer thickness, is presently estimated on the basis of
the introductory computations of the 3-D duct ow and the LES
results (see Section 3.3 dealing with the computational details).

RANS

LES

Because k and e are not provided within the LES sub-domain in


the case of the sub-grid-scale model of Smagorinsky, their SGS values are presently estimated utilizing the local equilibrium hypoth-

Fig. 2. Variation of modelled turbulent viscosity across the RANS/LES interface at


two selected streamwise locations in the diffuser conguration.

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

The present HLR model has been intensively validated in congurations of different geometrical complexity featuring ow separation (from sharp-edged and continuous curved surfaces) and
strong heating under the conditions of constant and variable uid
properties. Also, different sub-grid-scale (SGS) model formulations
in LES (besides the Smagorinsky model the transport model for the
residual kinetic energy from Yoshizawa and Horiuti (1985) has
been tested) were combined with different eddy-viscosity-based
RANS models. More details about the present HLR method are given in Jakirlic et al. (2006, 2009) and Kniesner (2008).

3.1.2. Large-Eddy Simulation (LES)


The sub-grid scale motion was modelled by the Smagorinsky
(1963) formulation (mSGS C 2S D2 jSj) utilizing the dynamic determination of the model coefcient C g C 2S proposed by Germano et al.
(1991). The determination of the coefcient Cg and the implementation of the dynamic procedure into the code FASTEST (Section
3.2) are briey outlined in the following. A detailed description
can be found in Ertem-Mueller (2003).
b > D to the
Application of a test-lter with the lter length D
ltered momentum equations leads to the sub-test-lter stress
tensor T test
in analogy to the sub-grid stress tensor sSGS
ij
ij :

cc
T test
Ud
i Uj  Ui Uj
ij

SGS
ij

1
test
b2 b c
T test
 dij T test
ij
kk 2C g D j Sj Sij 2C g aij
3
1
2
SGS
sSGS
 dij sSGS
ij
kk 2C g D jSjSij 2C g aij
3

^ are obtained by
The (test-) ltered quantities denoted with 
applying the test-lter to each control volume (CV) as the volumeweighted average of the CV value itself and its twenty-six (26)
neighbours. Beside the cell volumes additional weighting factors
are taken into account. The variable arrangement and the test-lter
with corresponding weighting factors are depicted in Fig. 3. The resolved part of the turbulent stresses Lij is computed in accordance
with Germano (1992):

cc
Lij T test
 sSGS
Ud
i Uj  Ui Uj
ij
ij

This identity is used to determine the model coefcient Cg


dynamically in conjunction with the instantaneous local ow conditions. Utilizing the model assumptions used for the Smagorinsky
formulation results in the following expressions:

8
9

Insertion of the above formulations in the identity (7) yields

dSGS
Lij 2C g atest
ij 2 C g aij

10

In order to compute the model coefcient the following


approximation
SGS
aSGS
C g ad
C gd
ij
ij

11

is utilized, being equivalent to the assumption that Cg is constant


within the volume of the test-lter (see Fig. 3). This leads to



SGS
Lij 2C g ad
 atest
2C g M ij
ij
ij

12

According to Germano et al. (1991) Eq. (12) can be contracted


with the resolved strain rate Sij resulting in Lij Sij 2C g M ij Sij . Since
Mij can become zero, Cg may be indeterminate. To overcome this
the least-squares method is used, as proposed by Lilly (1992). Application of the square error minimization algorithm leads to the following relation dening the preliminary model coefcient C g :

Ui Uj  Ui Uj

823

C g ~
x; t

Lij M ij
2M ij M ij

13

The determination of the model coefcient is carried out for


each SIMPLE (Section 3.2) iteration while computing the new time
level tn+1. To obtain a smooth behaviour of Cg and to avoid its wellknown oscillations, a relaxation of the coefcient C g in time with
the factor b 2 0; 1 is carried out utilizing its value from the previous time step at tn:

 n1 
~
C 
x; t n bC g ~
x; t n1
1  bC g ~
g x; t

14

Since C 
g can become negative, leading to a negative effective
viscosity and to unphysical behaviour and numerical instabilities,
the negative values of Cg are clipped. Furthermore, an upper bound
of Cg is introduced to ensure the model does not become too diffusive. This nally yields the model coefcient in the corresponding
expression for the turbulent viscosity enabling a converged solution at the new time level tn1

(a) control volume with neighbouring nodes

(b) Test filter

b with corresponding weighting factors (b).


Fig. 3. Variable arrangement (a) and applied test-lter D

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

824

i o
C g ~
x; t n1 min max C 
g ;0 ;1

15

3.2. Numerical method


The computational results were obtained using the in-house
code FASTEST (Flow Analysis by Solving Transport Equations Simulating Turbulence), which uses a nite volume method for blockstructured, body-tted, non-orthogonal, hexahedral meshes (FASTEST-Manual, 2005). Block interfaces are treated in a conservative
manner, consistent with the treatment of inner cell-faces. Cell
centred (collocated) variable arrangement and Cartesian vector
and tensor components are used. The equations are linearised and
solved sequentially using an iterative method. The velocitypressure coupling is ensured by the pressure-correction method based
on the SIMPLE algorithm which is embedded in a geometric multi-grid scheme with standard restriction and prolongation and an
ILU for smoothing, Durst and Schfer (1996). To avoid decoupling
of pressure and velocity on the collocated grid the selective interpolation method proposed by Rhie and Chow (1983) is applied. To enable simulations with a large number of control volumes the code is
strictly parallelised. The parallelisation is based on a domain
decomposition, which is directly related to the block structure of
the spatial discretisation. By using an automatic partitioning tool,
an optimal load balancing of the processors can be achieved, even

for complex geometries with hundreds of blocks, Sternel et al.


(2004). Communication between the processors is organised via
the Message Passing Interface (MPI) technique. The convective
transport of all variables was discretised by a second-order, central
differencing scheme for LES and DES. In the case of the HLR method
some upwinding is used for k and e equations by applying the socalled ux blending technique. Time discretisation was accomplished applying the Crank-Nicolson scheme.
3.3. Computational details
3.3.1. LES
The solution domain comprising a part of the development duct
(5 h), the diffuser section (15 h), the outlet straight duct (12.5 h) as
well as the convergent part (9 h), is meshed with 3.55 Mio. grid
cells: Nx  Ny  Nz = 408  64  136, Fig. 4. Two simulations with
and without the SGS model have been performed (only the results
obtained by using the SGS model are shown here). The wall boundary layers are resolved with y+ values of O(1). The off-wall resolution is close to that used in HLR simulation, see Figs. 2, 5 and 7
and corresponding discussion. According to the experimental setup
of Cherry et al. (2008) the fully-developed turbulent channel ow
has been computed with respect to the inow generation. These inlet data are generated by a simultaneously running periodic channel ow simulation of a channel (Nx  Ny  Nz = 48  64  136)

Fig. 4. Solution domain and computational mesh used in the LES simulation (every 5th grid line is plotted).

Fig. 5. Near-wall resolution expressed in wall units along the upper and lower walls in the central xy plane (z/B = 1/2) in the HLR simulation.

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

825

Fig. 6. Interface position in the HLR simulation corresponding to y+  50.

with the same cross-section as the diffuser inlet; small gure in


Fig. 4 (see also Fig. 8). To allow the ow through the diffuser to
inuence the ow eld in the development channel a part (5 channel heights) of this channel has been modelled in front of the diffuser. The turbulent ow elds in a cross-section in the periodic
channel are copied to this inlet location.
We are aware of the fact that the 3-D streamwise-periodic
channel of length 4.5h used for the inow generation might be
too short, keeping in mind the spatial extent of the characteristic
vortex structures, which is in general larger (due to the secondary
currents) than in a channel ow with the spanwise homogeneity.
Furthermore, Nikitin (2008) argued that an auxiliary streamwiseperiodic simulation might not be suitable since it causes a spatial
periodicity, which is not physical for turbulent ows. Let us recall
that the solution domain in the DNS of Ohlsson et al. (2010) comprises the inow development duct of 63h length, accounting even
for the transition of the initially laminar inow. The present simplication of the numerical setup is especially pertinent to the hybrid
LES/RANS method, since the overall aim is to improve the efciency (lower computational costs) and applicability to complex
geometries. In order to achieve the same basis for mutual comparison of the presently employed LES and HLR, both methods used
the same inow conditions, i.e. the same inow duct length.
The emphasis of the present study was on the time-averaged
ow eld (see Section 4) which was in reasonable agreement

with the reference databases, retroactively justied the simplied


inow generation. We mention here also the comparable inow
generation of Schneider et al. (2010), successfully applied over
a channel length of only 3 channel heights. The temporal resolution adopted corresponds to a non-dimensional time step of
Dt = 0.0001Ubulk/h = 0.01 (normalized by the inlet channel parameters Ubulk = 1 m/s and h = 1 cm). This leads to the maximum CFL
number being substantially smaller than one. As a comparison
the value of the viscous time scale m=U 2s in the inow region corresponds to a dimensionless time step of Dt = 0.028 (the value of friction velocity at the bottom wall of the inow channel at z/B = 0.5
obtained by DNS was about 0.063 this values can be extracted
from Fig. 10-left; signicantly lower values were documented in
the diffuser section). Averaging was performed for approximately
80,000 time steps. The simulations were carried out on 16 IBM
Power 5 CPUs with a load balancing efciency of 100%.
3.3.2. HLR
The inow data were, in analogy to the present LES, generated
by a precursor simulation of the fully-developed duct ow (with
the auxiliary periodic duct length of 6.5h) using the respective
models on the grid of Nx  Ny  Nz = 78  62  134 cells. The
solution domain for the HLR simulation comprised a part of the
development duct (5 h), the diffuser section (15 h) and the
straight outlet duct (12.5 h), Fig. 1. At its outlet cross-section

Fig. 7. Proles of the kinetic energy of turbulence at two selected streamwise locations (x/h = 2 inlet duct and x/h = 10 diffuser section) obtained by the present HLR
method (vertical lines denote the LES/RANS interface position).

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

826

Fig. 8. The instantaneous velocity eld obtained by LES (small gure on the upper left illustrates the turbulent inlet data generation through a separate inow duct simulation
utilizing a simultaneously running periodic channel simulation).

(a) experiment:

(b) HLR:

0.1 m/s

0.1 m/s

Fig. 9. Velocity vectors in the yz plane in the inow duct.

the convective outow conditions were applied. No-slip boundary


conditions were applied at the walls. The grid applied contained
Nx  Ny  Nz = 224  62  134 cells (approximately 1.86 Mio. grid
cells in total; the number of grid cells in the cross-plane yz as
well as of those cells in the diffuser section correspond to the
number applied in the LES simulation). A dimensionless time step
of Dt 0:00028U bulk =h 0:028 was used for the computations,
ensuring a CFL number less than unity throughout the entire
solution domain (CFLmax  0.76). This is of the same order of
magnitude as the viscous time scale m=U 2s (see discussion related
to the LES-relevant computational details). The near-wall resolution corresponds to Dy+ < 0.8, Dx+  10100 and Dz+  220
(along the xy plane at z/B = 1/2) for the HLR simulation, Fig. 5.

Similar resolution was achieved on the side walls (not shown


here). The nal position of the interface, separating the near-wall
(RANS; black area in Fig. 6) and off-wall (LES; gray area in Fig. 6)
sub-regions, corresponds to the dimensionless distances from the
bottom/upper and side walls y+, z+  50. This position was determined in accordance to the preliminary computations of the
fully-developed ow in the inow duct as well as the results
obtained by the pure LES simulations. It should be noted that
the symmetry plane in the central vertical plane of the inow
duct is at y+  300 (see position of the velocity maximum in
Fig. 10-left). Accordingly, the wall layer thickness covered by
the RANS model amounts to one sixth of the half-height of the
inow duct. Averaging was performed for approximately 80,000
time steps. The simulations were carried out on the same computer as LES but using only seven processors. One of the important issues when employing a hybrid LES/RANS method is the
computational time used relative to the complementary LES simulation. Such an analysis could not be performed straightforwardly in the present work keeping in mind the differences in
the grid size, model structure (HLR model employs the standard
Smagorinsky model and LES its dynamic version) and number
of processors. But for instance, in a fully-developed channel ow
(the analysis was conducted by computing the ow at Res = 640),
the present hybrid model needs about 1030% more time
(depending on the version of the low-Reynolds number, twoequation ke model covering the RANS layer) if using the same
grid, see Kniesner (2008).
Further information about the suitability of the grid resolution
can be gained from Figs. 2 and 7 by analysing the ratio of the turbulent SGS viscosity to molecular viscosity and the SGS contribution to total turbulence kinetic energy at the locations across the
inow duct (x/h = 2) and the interior of the diffuser (x/h = 10).
Although extracted from the HLR simulation, these results could
serve as the basis for discussion about the grid resolution assessment also for the LES simulation. Fig. 2 illustrates the SGS viscosity

Fig. 10. Semi-log plots of axial velocity component across the inow duct (x/h = 2) and at a cross-section in the interior of the diffuser (x/h = 10).

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

Fig. 11. Surface pressure distribution at the bottom wall.

being substantially smaller than its molecular counterpart in the


LES sub-region. Fig. 7 demonstrates clearly the contribution of
the kSGS (obtained from Eq. (4)) to the total k-value being less than
10% everywhere in the LES sub-region (compatible with the one in
the pure LES simulation); the RANS layer thickness corresponds
approximately to 0.1h at both streamwise locations (the vertical
lines denote the LES/RANS interface position). One can also see
the fraction of the modelled turbulence in the RANS layer being
between 20% and 25% of the total turbulence kinetic energy. The
exception is the portion of the upper wall in the interior of the
diffuser section being affected by separation, where almost the
entire turbulence kinetic energy originates from the resolved
uctuations.
4. Results and discussion
Fig. 8 displaying the instantaneous velocity eld obtained by LES,
provides a rst impression about the ow topology. A slice through
the computational domain and the periodic channel is depicted,
showing the contours of the instantaneous streamwise velocity
component at an arbitrary instant of time. The ow in the inow duct
complies with the fully-developed turbulence underlying approximately the equilibrium conditions (see Fig. 10 and corresponding
discussion). Increased ow perturbations start already at the corner
between the at walls of the inow duct and the sloped diffuser
walls, causing the deceleration of the boundary layer and a deformation of the mean velocity prole. The boundary layers along the
ared walls interact and separate nally in the corner region under
the inuence of the adverse-pressure gradient. The separation bubble grows and gradually occupies the entire upper diffuser wall.
Fig. 8 illustrates clearly the unsteady nature of the diffuser ow,
characterized by apping motion of the separated shear layers and
multiple separating and reattaching regions.
A selection of the computational results along with the experimental and DNS data is displayed in Figs. 915. The presentation
and analysis of the results obtained has been conducted with respect to ow in the inow duct, the size and shape of the ow separation pattern and associated mean ow and turbulence features.
4.1. Inow duct
Unlike the ow through a circular pipe, the ow in a duct with
rectangular cross-section is no longer unidirectional. It is characterized by a strong secondary motion with the velocity components being perpendicular to the axial direction, Fig. 9. This
secondary ow transporting momentum into the duct corners is
characterized by jets directed towards the duct walls bisecting
each corner with associated vortices at both sides of each jet. This
secondary current resembles Prandtls ow of the second kind

827

(possible only for turbulent ows) induced by the Reynolds stress


anisotropy. Indeed, the Reynolds stress gradients cause the generation of forces which induce the normal-to-the-wall velocity
components in the secondary ow plane. Accordingly, correct capturing of anisotropic turbulence in the inow duct is an important
prerequisite for a successful computation of the diffuser ow. Fig. 9
displays the time-averaged velocity vectors in the cross-plane yz
located at x/h = 2 obtained experimentally and computationally
by the HLR model. Despite the lower intensity of the secondary
motion, whose largest velocity has the magnitude corresponding
approximately to <(12)% of the axial bulk velocity (Ubulk = 1 m/
s), its inuence on the ow in the diffuser is signicant. The
qualitative agreement with respect to the secondary ow topology
discussed above is obvious.
Fig. 10-left depicts the semi-log plot of the axial velocity
component across the central plane (z/B = 0.5) of the inow duct
at x/h = 2. The inow conditions correspond clearly to those
typical for an equilibrium ow. The velocity prole shape obtained
by DNS follows closely the logarithmic law, despite a certain
departure from it. This departure, expressed in terms of a slight
underprediction of the coefcient B in the log-law (U+ = ln (y+)/
j + B; B = 5.2), can also be regarded as a consequence of the
back-inuence of the adverse-pressure gradient evoked by the ow
expansion. The pressure coefcient evolution, displayed in Fig. 11,
reveals an appropriate pressure increase already in the inow duct
(x/h 6 0). The present computational results exhibit a certain overprediction of the velocity in the logarithmic region. It resembles a
typical outcome in the case of a somewhat coarser grid resolution.
One would assume a sufciently ne grid in this region according
to the ratio of the turbulent viscosity to the molecular viscosity
(Fig. 2) which is less than 1. On the other hand, the prole of the
kinetic energy of turbulence (Fig. 7-left) reveals a somewhat higher
level of the modelled turbulence in this attached ow region. All
these issues contributed to an appropriate underprediction of the
friction velocity Us, serving here for the normalization U+ = U/Us.
Whereas the LES returned a smooth prole, the velocity prole
obtained by the present HLR method exhibits a certain bump in the
region around the LES/RANS interface (y+  50). It represents an
outcome typical for the application of a zonal hybrid LES/RANS
method. The relatively high turbulent viscosity in the RANS layer
strongly damps the uctuations, suppressing their recovery in
the LES sub-region until some distance behind the interface. Such
an anomaly can be remedied by generating some additional uctuations and superimposing them onto the LES ones (see Jakirlic et al.
(2010a) for specic details). This so-called forcing procedure was
not used in the present work. This is justied by the fact that the
buffer layer associated with the interface region is limited only
to the immediate wall vicinity in the inow duct, and does not signicantly affect other quantities in this region (see e.g., smooth
behaviour of the kinetic energy prole in Fig. 7-left) or the remainder of the ow domain (see e.g., smooth evolution of the velocity
and kinetic energy proles at x/h = 10 across the interface in Figs.
7-right and 10-right; both proles displaying a behaviour, typical
of a ow affected by an adverse-pressure gradient underprediction of the logarithmic law and strong enhancement of the turbulence intensity corresponding to a substantial mean ow
deformation in the streamwise direction).
4.2. Diffuser section
Fig. 11 illustrates the evolution of the surface pressure
coefcient along the bottom non-expanding wall in the diffuser
conguration, representing the most important integral characteristic. The surface pressure distribution along the bottom wall was
evaluated from the expression C P p  pref =0:5qU 2bulk ; the reference pressure was taken at the position x/L = 0.05. The pressure

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

x/h=15

x/h=12

x/h=8

x/h=5

x/h=2

828

Experiment

LES

HLR

Fig. 12. Comparison between experimentally and computationally obtained iso-contours of the axial velocity eld in the cross-planes yz at selected streamwise locations
within the diffuser section (the thick line denotes the zero-velocity line).

curve exhibits a development typical of ow in diverging ducts.


The pressure decrease in the inow duct is followed by a steep
pressure increase already at the very end of the inow duct and
especially at the beginning of the diffuser section. The transition
from the initial strong pressure rise to its moderate increase occurs
at x/L  0.3, corresponding to the position where about 9% of the
entire cross-section is occupied by the ow reversal (see e.g.,
Fig. 13). Onset of the separation zone causes a certain contraction
of the ow cross-section, leading to a weakening of the deceleration intensity and, accordingly, to a slower pressure increase. The
region characterized by a monotonic pressure rise was reached in
the remainder of the diffuser section. The results obtained by both
computational methods display this correct tendency, but some
quantitative differences are present. The result obtained by applying the HLR method exhibit very good agreement with the experimental data, apart from the slight underprediction in the ow

region affected by a steep pressure gradient. The LES results are


typical of a higher pressure gradient in the region x/h = 35 (consistent with an enhanced ow deceleration in the lower half of the
diffuser section, Fig. 14) resulting in an oveprediction of the CP value over the entire diffuser length.
Fig. 12 shows the contour plots of the axial velocity component
at ve streamwise cross-sectional areas indicating the evolution of
the ow separation pattern. The recirculation zone development
displayed here can be analyzed in parallel with the quantitative
information about the fraction of the diffuser cross-sectional area
occupied by the reversed ow depicted in Fig. 13. The adversepressure gradient is imposed onto the intersecting boundary layers
along at walls upon entering the diffuser section. According to the
experimental investigation the boundary layers along all walls are
of comparable thickness. The separation starts immediately after
the beginning of the diffuser section at x/L = 0 (x/h = 0), see.

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

Fig. 13. Fraction of the cross-sectional area occupied by the ow reversal.

Fig. 13. The onset of separation pattern is located in the upper-right


diffuser corner, built up from the deected side wall and the top
wall, see e.g., the position x/h  2 (x/L = 0.13) in Fig. 12. Initial
growth of this corner bubble reveals its spreading rate along two
sloped walls being approximately of the same intensity, see position x/h = 5. As the adverse-pressure gradient along the upper wall
outweighs signicantly the one along the side wall due to substantially higher angle of expansion, 11.3 vs. 2.56, the separation zone
spreads gradually over the entire top wall surface, see position x/
h = 8. One notes a strong three-dimensional nature of the separation pattern. The maximum occupation of the diffuser cross-sectional area by the ow reversal, around 22% (Fig. 13), is
documented at the position x/h = 1215 (x/L = 0.81.0). The thickness of the ow reversal zone (its dimension in the normal-to-wall
direction) is almost constant over the diffuser width in this region,
resembling approximately a 2-D pattern. After this position the
intensity of the back-ow weakens. The experimental results point
out the reattachment region being located around x/L  1.4 within
the straight outlet duct, Fig. 13.
The results obtained by the LES and HLR methods exhibit qualitatively close agreement with the previously described spatial
development of the three-dimensional back-ow region, both in
size and shape. This is also the case when comparing with the
DNS database (Fig. 3 in Ohlsson et al. (2010)). As in the experimental conguration, the separation starts in the upper-right corner.
Whereas the LES simulation results in a corner bubble covering
both deected walls (streamwise position x/h = 5 in Fig. 12), in
close agreement with the experimental ndings with respect to
shape but not size, the HLR model response to the adversepressure gradient onset is manifested in an asymmetrical corner
bubble occupying mostly the upper wall. The size of the cross-sectional area occupied by the latter separation zone is smaller than
the experimentally detected one. This outcome is reected in the
lower fraction of the back-ow region compared to the throughow area within the rst half of the diffuser section, Fig. 13.
Whereas the corner bubble predicted by LES preserves its symmetrical shape growing equally in both coordinate directions in the
vertical cross-plane yz at x/h = 8 (x/L = 0.53; one notes the appearance of a smaller bubble in the upper left corner disconnecting
separation zone), the HLR simulation results in the back-ow zone
similar in shape to the experimental one. The total amount of the
cross-sectional area affected by the ow reversal is about 10%
and agrees well with the experimental outcome at this position,
Fig. 13. Both computational methods yield a separation zone
spreading over the entire upper wall from the streamwise location
x/h = 810 (x/L = 0.50.7). Contrary to the experimental ndings
the computationally obtained separation zone exhibits an asymmetric form with a non-constant thickness over the diffuser duct

829

width. The position of the maximum reverse ow cross-section


(about 20%) is reached at x/h = 1215 (x/L = 0.81.0) in good agreement with experiment. After this location the fraction of the crosssectional area covered by the separated ow follows closely the
experimentally determined results. Analogous to the experiment
the ow reattaches at position x=L  1:41:5 in the straight outlet
part of the diffuser conguration. In summary it can be stated that
despite certain deviations the topology of the separation region,
with respect to its onset, shape and size, is reasonably well reproduced by both computational methods.
Good overall agreement obtained by HLR and LES is further conrmed in Figs. 14 and 15 illustrating the axial velocity and streamwise stress component proles at 14 streamwise locations situated
in all characteristic ow regions attached ow in the inow duct,
separation point, recirculation zone, reattachment and recovery
region within the diffuser conguration in three vertical
cross-planes corresponding to the spanwise locations z/B = 1/4,
z/B = 1/2 and z/B = 3/4 (B = 3.33 cm).
Both the mean velocity and turbulence elds have been mapped
three-dimensionally, enabling a detailed quantitative insight into
the entire ow domain. In addition to the reference experiments,
the present computational results are analysed along with the
DNS database from Ohlsson et al. (2010). Fig. 14 displays the velocity eld development typical for the ow in an expanding duct. The
bulk ow exhibits deceleration, leading to an asymmetry of the
velocity prole. The effect of the adverse-pressure gradient is especially visible in the ow region along the upper expanding wall.
The velocity prole approaches gradually the form characterizing
a separating ow, exhibiting regions of the zero velocity gradient
(at separation and reattachment points) and prole inection.
The through-ow, that is the ow in the positive streamwise direction, is characterized by a spreading dictated by the pressure gradient arising from the geometry expansion. Accordingly, the
position of the reduced velocity maximum is gradually shifted towards the upper wall, eventually reaching the center (y/h  2) of
the straight outlet channel (with the height 4h). In this post-reattachment zone the velocity prole exhibits a fairly attened form,
being almost symmetric. The consequence of the velocity prole
attening is a continuous monotonic decrease of the wall shear
stress.
The velocity prole evolution is similar in all three longitudinal
vertical planes. The specic differences are related to the vicinity of
both bottom and upper walls, especially the latter upper wall,
where the ow passes regions of three-dimensional ow reversal
(see Fig. 12 and corresponding discussion). In Figs. 15, the development of the streamwise turbulence intensity is shown. The lowest
turbulence intensity is situated in the region coinciding with the
mean velocity maximum ow zone with approximately zero
velocity gradient along the entire diffuser section. The Reynolds
stress proles exhibit their highest values in the regions with the
most intensive ow deformation. These are the near-wall layer in
the attached ow regions and the ow zone along the shear layer
bordering the recirculation zone. The peak of the turbulence intensity originating from the boundary layer at the top inow duct wall
increases initially, after the strong rise in the pressure eld is imposed (see pressure coefcient development in Fig. 11), and weakens slightly after ow transition to the second part of the diffuser
section, being characterized by a decreasingly adverse-pressure
gradient. The streamwise turbulence intensity in the outlet duct
is uniformly distributed over the cross-section.
The present computational results are in close agreement with
both reference data sets, experimental and DNS. This holds for the
individual velocity and Reynolds stress proles but also for the
ow development as a whole. However, some discrepancies should
not be overlooked. The largest ones with respect to the velocity
eld are found in the region aligned with the mean dividing

830

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

Fig. 14. Evolution of the axial velocity proles in the vertical plane xy at three spanwise locations z/B = 1/4, 1/2 and 3/4.

streamline. This is especially the case with the LES simulations.


Further deviations, particularly with respect to the Reynolds stress
proles, are noticeable in the core ow within the transitional region between the diffuser section and the straight outlet channel

(x/h P 15), where the ow is exposed to a relaxation. Both present


simulations were performed on the grids with the cells clustered
adequately upon approaching the solid walls. The wall-closest
numerical nodes are situated deeply in the viscous sub-layer along

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

831

Fig. 15. Evolution of the turbulent streamwise stress component proles in the vertical plane xy at three spanwise locations z/B = 1/4, 1/2 and 3/4.

all duct walls. The adopted expansion factor led to appropriately


ne near-wall grid resolution and a somewhat coarser grid in the
ow core, the latter feature representing probably the reason for
the afore mentioned deviation.

Similar deviations of the LES simulations from the reference


data are observed in the region characterized by high turbulence
intensity. Both computational methods complement each other
in the core ow region. The basic differences originate from

832

S. Jakirlic et al. / International Journal of Heat and Fluid Flow 31 (2010) 820832

the near-wall treatment. Whereas the hybrid LES/RANS method


employs an elaborate near-wall RANS model which solves additional transport equations for k and e for the velocity and length
scale determination in the turbulent viscosity formulation, the
LES method extracts both representative scales from the velocity
eld and the grid spacing. As elaborated earlier, the wall boundary layer is covered by an irregular, highly non-uniform grid, the
fact being particularly applicable with respect to the streamwise
and spanwise resolutions. The LES models assume a cutoff wave
number well situated in the inertial sub-range, a condition not
fullled in the immediate wall vicinity characterized by lowReynolds numbers, where production and dissipative ranges lie
almost immediately next to each other. This situation is consistent with a very narrow inertial region.
Unlike the present LES, the near-wall RANS model works well
under conditions of non-uniform, anisotropic grids with higher
aspect ratio. Generally speaking, an appropriate grid renement
leading eventually to the improved agreement would be easily
manageable; keeping in mind a fairly low ow Reynolds number
(Reh = 10,000) a moderate increase in the number of grid cells
would probably be sufcient to accomplish this task. However,
as the primary objective was the validation of the hybrid LES/
RANS method designed to function under the condition of a
coarser spatial resolution in such a complex ow conguration
featured by three-dimensional separation, the presently used
grid was intentionally adopted.

5. Conclusions
Two eddy-resolving modelling approaches: Large-Eddy Simulation (LES) and a zonal hybrid LES/RANS scheme (HLR) were used to
predict the ow in a 3-D diffuser, measured recently by Cherry et
al. (2008, 2009) and computed by means of Direct Numerical Simulation by Ohlsson et al. (2009, 2010), aiming at a comparative
analysis of their features and performance. The latter method,
developed by the authors, merging a low-Reynolds-number ke
Reynolds-Averaged NavierStokes (RANS) model with Large-Eddy
Simulation (LES) in the framework of a two-layer scheme, was
the focus of the present investigation, keeping in mind that the basic motivation behind the HLR scheme is to obtain results on a
coarser grid, whose quality is comparable with that of a well-resolved LES. In order to check its feasibility, an intensive validation
has been performed. The ow conguration exhibited a 3-D recirculation pattern was thought to represent a suitable benchmark.
Good results were obtained by both computational models with
respect to the characteristics of the duct ow expanding into a diffuser section, the consequent separation ow region (onset, shape
and size), the mean velocity eld and associated integral parameters (pressure distribution), as well as the turbulence quantities.
More importantly, a close agreement was achieved between the
hybrid LES/RANS model and both experimental and DNS reference
databases with respect to the near-wall behaviour of the timeaveraged quantities, demonstrating its high potential for computing such complex, wall-bounded turbulent ows.

Acknowledgements
The work of G. Kadavelil and M. Kornhaas is nancially
supported by the Deutsche Forschungsgemeinschaft (DFG) within
the German Collaborative Research Center (SFB 568) Flow and
Combustion in Future Gas Turbine Combustion Chambers. Our
special thanks goes to the authors of the reference results J. Eaton
and E. Cherry (experimental database) and J. Ohlsson (DNS database) for making their data available.

References
Bradshaw, P., Wong, F.Y.F., 1972. The reattachment and relaxation of a turbulent
shear layer. J. Fluid Mech. 52 (1), 113135.
Cherry, E.M., Elkins, C.J., Eaton, J.K., 2008. Geometric sensitivity of threedimensional separated ows. Int. J. Heat Fluid Flow 29, 803811.
Cherry, E.M., Elkins, C.J., Eaton, J.K., 2009. Pressure measurements in a threedimensional separated diffuser. Int. J. Heat Fluid Flow 30, 12.
Durst, F., Schfer, M., 1996. A parallel block-structured multigrid method for the
prediction of incompressible ows. Int. J. Numer. Methods Fluids 22, 549565.
Ertem-Mueller, S., 2003. Numerical efciency of implicit and explicit methods with
multigrid for Large Eddy Simulation in complex geometries. Dissertation.
Technische Universitt Darmstadt.
FASTEST-Manual, 2005. Institute of Numerical Methods in Mechanical Engineering,
Department of Mechanical Engineering, Technische Universitt Darmstadt,
Germany.
Frhlich, J., von Terzi, D., 2008. Hybrid LES/RANS methods for the simulation of
turbulent ows. Prog. Aerosp. Sci. 44, 349377.
Germano, M., 1992. Turbulence: the ltering approach. J. Fluid Mech. 238, 325336.
Germano, M., Piomelli, U., Moin, P., Cabot, W.H., 1991. A dynamic subgrid-scale
eddy viscosity model. Phys. Fluids A 3 (7), 17601765.
Jakirlic, S., Kniesner, B., aric, S., Hanjalic, K., 2006. Merging near-wall RANS models
with LES for separating and reattaching ows. In: ASME Joint USEuropean Fluids
Engineering Summer Meeting: Symposium on DNS, LES and Hybrid RANS/LES
Techniques, Paper No. FEDSM2006-98039, Miami, FL, USA, July 1720.
Jakirlic, S., Kniesner, B., Kadavelil, G., Gnir, M., Tropea, C., 2009. Experimental and
computational investigations of ow and mixing in a single-annular combustor
conguration. Flow Turbul. Combust. 83 (3), 425448.
Jakirlic, S., Kniesner, B., Kadavelil, G., 2010. A method for interface-turbulence
forcing in hybrid LES/RANS simulations. In: 8th Int. ERCOFTAC Symp. on
Engineering Turbulence Modelling and Measurements, Marseille, France, June
911.
Jakirlic, S., Kadavelil, G., Sirbubalo, E., von Terzi, D., Breuer, M., Borello, D., 2010.
Report on 14th ERCOFTAC Workshop on Rened Turbulence Modelling.
September 18, 2009. Sapienza University of Rome, ERCOFTAC Bulletin, No.
84, 2010.
Kniesner, B., 2008. Ein hybrides LES/RANS Verfahren fr konjugierte Strmung,
Wrme- und Stoffbertragung mit Relevanz zu Drallbrennerkongurationen (A
hybrid LES/RANS method for conjugated ow, heat and mass transfer with
relevance to swirl combustor congurations). PhD Thesis, Technische
Universitt Darmstadt. <http://www.tuprints.ulb.tu-darmstadt.de/950/>.
Launder, B.E., Sharma, B.I., 1974. Application of the energy dissipation model of
turbulence to the calculation of ow near a spinning disc. Lett. Heat Mass
Transfer 1, 131138.
Lilly, D.K., 1992. A proposed modication of the Germano subgrid-scale closure
method. Phys. Fluids A 4 (3), 633635.
Mason, P.J., Callen, N.S., 1986. On the magnitude of the subgrid-scale eddy
coefcient in large-eddy simulation of turbulent channel ow. J. Fluid Mech.
162, 439462.
Nikitin, N., 2008. On the rate of spatial predictability in near-wall turbulence. J.
Fluid Mech. 614, 495507.
Obi, S., Aoki, K., Masuda, S., 1993. Experimental and computational study of
turbulent separating ow in an asymmetric plane diffuser. In: Proc. 9th Symp.
on Turbulent Shear Flows, Kyoto, Japan, August 1619, pp. 305.1305.5.
Ohlsson, J., Schlatter, P., Fischer, P.F., Henningson, D.S., 2009. DNS of threedimensional separation in turbulent diffuser ows. In: Advances in Turbulence
XII. Proceedings of the 12th EUROMECH European Turbulence Conference, vol.
132. Springer Proceedings in Physics, Marburg. ISBN: 978-3-642-03084-0.
Ohlsson, J., Schlatter, P., Fischer, P.F., Henningson, D.S., 2010. DNS of separated ow
in a three-dimensional diffuser by the spectral-element method. J. Fluid Mech.
650, 307318.
Rhie, C.M., Chow, W.L., 1983. A numerical study of the turbulent ow past an
isolated airfoil with trailing edge separation. AIAA J. 21, 15251532.
Schneider, H., von Terzi, D., Bauer, H.-J., Rodi, W., 2010. Reliable and accurate
prediction of three-dimensional separation in asymmetric diffusers using largeeddy simulation. ASME J. Fluids Eng. 132 (3), 031101031111.
Smagorinsky, J., 1963. General circulation experiments with the primitive
equations: I. The basic experiment. Mon. Weather Rev. 91 (3), 99164.
Spalart, P.R., 2009. Detached-Eddy simulation. Annu. Rev. Fluid Mech. 41, 181202.
Spalart, P.R., Jou, W.-H., Strelets, M., Allmaras, S., 1997. Comments on the feasibility
of LES for wings and on a hybrid RANS/LES approach. In: 1st AFOSR Int. Conf. on
DNS and LES, Columbus, OH, USA, August 48.
Steiner, H., Jakirlic, S., Kadavelil, G., aric, S., Manceau, R., Brenn. G., 2009. Report on
13th ERCOFTAC Workshop on Rened Turbulence Modelling. September 2526,
2008. Graz University of Technology, ERCOFTAC Bulletin, No. 79, pp. 2429.
Sternel, D.C., Junglas, D., Martin, A., Schfer, M., 2004. Optimisation of partitioning
for parallel ow simulation on block structural grids. In: Topping, B.H.V., Mota
Soares, C.A. (Eds.), Proc. of the 4th Int. Conf. on Engineering Computational
Technology, Stirling: Paper 93. Civil-Comp Press, Stirling.
Temmerman, L., Hadziabdic, M., Leschziner, M.A., Hanjalic, K., 2005. A hybrid twolayer URANS-LES approach for large eddy simulation at high Reynolds numbers.
Int. J. Heat Fluid Flow 26, 173190.
Yoshizawa, A., Horiuti, K., 1985. A statistically-derived subgridscale kinetic energy
model for the large-eddy simulation of turbulent ows. J. Phys. Soc. Jpn. 54,
28342839.

You might also like